Ionian marine terraces of southern Italy: Insights into the Quaternary

0 downloads 0 Views 2MB Size Report
Jul 14, 2010 - evidence for an important relative sea level fall [Tropeano et al., 2002]. ... Teodoro terrace, while both La Petrulla and Piano San Nicola sites to the Policoro terrace. ...... Moretti, I., and L. Royden (1988), Deflection, gravity.
Click Here

TECTONICS, VOL. 29, TC4005, doi:10.1029/2009TC002625, 2010

for

Full Article

Ionian marine terraces of southern Italy: Insights into the Quaternary tectonic evolution of the area R. Caputo,1 M. Bianca,2 and R. D’Onofrio2 Received 28 October 2009; revised 27 January 2010; accepted 26 February 2010; published 14 July 2010.

[1] New detailed morphotectonic analyses of a well

exposed flight of marine terraces along the Ionian coast of southern Italy has been carried out. The area represents a key transect for investigating the middle‐ late Quaternary evolution of the Southern Apennines chain‐foredeep‐foreland geodynamic system. A major result of the research is the reconstruction of a virtually complete 3D geometry of the marine surfaces along a coastal sector of ca. 70 km, which (1) documents the occurrence of 18 paleo‐shorelines and (2) provides evidence for a strong regional uplift affecting the investigated area. Following a systematic critical review of literature relating to geochronological data, integrated with a morphogenetic model based on the interaction between tectonic uplift and eustatic sea level changes, the different terraces are correlated to as many highstand sea level peaks, dating the highest/ oldest terrace to ca. 600 ka (MIS 15). The vertical and horizontal distribution of the terraces show a general convergence of the paleo‐shorelines toward NNE, which indicates a decreasing trend in differential uplift in that direction ranging from almost 2 mm/a in the southwestern sector to about 0.2 mm/a in the northeastern sector. Detailed mapping and 3D reconstruction also emphasize the partitioning of the area into three distinct sectors characterized by different tilting rates. This behavior is likely caused by the combined role and activity of three major tectonic structures working at different scales and rates including (1) the reactivation of an out‐of‐sequence thrust, (2) sliding along the basal detachment of the external Apennines wedge and (3) a lithospheric‐scale duplexing (crustal or deeper). As a major conclusion, within the external sector of the Southern Apennines chain and its foredeep, regional shortening and a contractional tectonic regime persisted throughout the whole Quaternary and it is probably still active. Citation: Caputo, R., M. Bianca, and R. D’Onofrio (2010), Ionian marine terraces of southern Italy: Insights into the Quaternary tectonic evolution of the area, Tectonics, 29, TC4005, doi:10.1029/2009TC002625.

1

Department of Earth Sciences, University of Ferrara, Ferrara, Italy. Di.S.G.G., University of Basilicata, Potenza, Italy.

2

Copyright 2010 by the American Geophysical Union. 0278‐7407/10/2009TC002625

1. Introduction [2] Orogenic activity along the external sectors of the Southern Apennines (Figure 1a) is well documented for the Lower and part of the Middle Pleistocene, when it has been suggested that a major geodynamic re‐arrangement occurred in the wider region [e.g., Cinque et al., 1993; Patacca and Scandone, 2004] (see Appendix A). On the other hand, along the axial sector of the Apennines chain, but mainly west of the present‐day water divide, several normal faults have been forming since (Middle‐)Late Pleistocene and are still active nowadays as documented by both the historical and instrumental seismicity which affects the upper crustal level [Castello et al., 2005; Chiarabba et al., 2005; Di Bucci et al., 2006] and which is suggested by break‐out data from boreholes [Montone et al., 1999]. [3] Following this premise, in the external (i.e., northeastern) sector of the orogenic wedge along the boundary with the Bradanic Foredeep, it is commonly assumed that thrust activity has completely ceased [Patacca and Scandone, 2001]. In order to challenge this general opinion, we investigated the Ionian coastal sector between the regions of northern Calabria and Puglia, which represents a key‐area regarding the mid‐to‐late Quaternary geodynamic evolution of the Southern Apennines. Indeed, its coastline (corresponding to the Gulf of Taranto) (Figure 1b) trends NE–SW, i.e., roughly perpendicular to (1) the most external thrusts of the chain, (2) the axis of the Bradanic Foredeep and (3) the western sector of the outcropping Apulian platform commonly considered as the present‐day foreland (Figure 1b). In particular, we focus on the most evident morphological features occurring in this area, namely the already well‐ known step‐like sequence of Quaternary marine terraces and relative paleo‐shorelines, which is considered to be the result of the interaction between uplift, glacio‐eustatic sea level changes [Brückner, 1980a; Lajoie, 1986; Bosi et al., 1996] and, of course, later continental geomorphic processes. [4] The sequence of the Ionian raised marine terraces and paleo‐shorelines occurring along the whole coastline of the Gulf of Taranto has been studied by several authors since the second half of the 19th century [Fuchs, 1874], although most of the published works focus only on limited portions and sectors of the entire succession [e.g., Vezzani, 1967; Boenzi et al., 1976; Parea, 1986; Amato et al., 1997; Mastronuzzi and Sansò, 2002; Bianca and Caputo, 2003; Zander et al., 2006]. In the context of this short review, it is worth mentioning the work of Brückner [1980a, 1982], which represents the first and unique regional‐scale detailed investigation of the whole terrace sequence. [5] The main goals of our research are twofold. First, we aimed at obtaining a new, detailed and complete map of the

TC4005

1 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 1. (a) Synthetic geodynamic map of Italy. (b) Location map of the investigated area (box) showing the buried external thrust front of the Apennines (modified from Agenzia per la Protezione dell’Ambiente e per i servizi Tecnici [2004]). Large oblique hatching indicates the area undergoing upper crustal extension, while the dense oblique hatching represents the deeper ‘active’ fold‐and‐thrust‐ belt characterized by compressive‐transpressive regime. Vertical hatching represents the outcropping Apulian platform. entire sequence of marine terraces and associated paleo‐ shorelines along the whole coastal sector between Calabria and Puglia. This was achieved following a systematic remote sensing analysis of air‐photos, integrated by systematic field‐work devoted to checking all the detected morphological features and locating them precisely on maps. The obtained data were interpreted by uniform methods of geomorphological analysis in order to obtain an accurate reconstruction of the lateral continuity of the terraces and, therefore, to solve the problems of geometric and numerical fit between the different orders of terraces that generally arise along the overlapping sectors of the diverse maps proposed by previous researchers. In particular, our efforts were directed toward verifying the lateral continuity of the different terraces, where these are widely eroded by the major NW–SE trending fluvial valleys (e.g., Sinni, Agri, Cavone, Basento, Bradano and Lato rivers). It is important here to stress that this research does not represent a review of previous morphological works and a ‘simple’ check of published maps of the terrace flight, but it is based on, and intended as, an ex‐novo detailed mapping of the entire region covering an area of about 1500 km2. In this regard, all data graphically presented in Figures 2, 3, 4 and 5 and extensively discussed in the text should be considered original.

[6] The second major target of this research is aimed at providing new insights relative to the recent geodynamic evolution of this key‐area of the Southern Apennines within the central Mediterranean framework. This goal is based on the reconstructed space distribution of the paleo‐shorelines and their chronological correlations, which suggest a prevailing and persistent compressional tectonic regime during mid‐late Quaternary.

2. Quaternary Marine Terraces and Paleo‐shorelines [7] During the last decades, the ever‐increasing knowledge of coastal morphogenic processes has made marine terraces the most recognizable, widespread and scientifically reliable tools to determine, both qualitatively and quantitatively, the vertical movements that have affected the tectonically active coastal regions during the Quaternary [e.g., Bordoni and Valensise, 1998; Ferranti et al., 2006]. Indeed, the sequence of Quaternary marine terraces occurring along the investigated 70 km‐long Ionian coastal transect (Figures 1 and 2) is made up of a set of both stratigraphical and morphological elements that show a good identification on the field, a well‐known original geometry, and a high preservation potential with respect to the time scale of the

2 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Figure 2. Marine and fluvial terraces mapped within the investigated area. See Figure 1 for location. Note that some of the narrowest surfaces, like Metaponto 2 (near Sinni and Agri rivers) and San Teodoro 2 (near Rocca Imperiale village), are hardly visible on map for graphical reasons, but they are clearly shown in the longitudinal projection profile A‐A′ (Figure 4). Small black circles represent the sites for which ages are available and critically analyzed in the text (see also Table 1). The dashed area SE of Laterza represents a large‐scale deep‐seated landslide. 3 of 24

TC4005

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 3. Selected topographic profiles perpendicular to the coast line and to the inner edges of the mapped marine terraces. Dotted lines represent correlated inner edges. See inset for location of profiles. The vertical and horizontal scales are graphically shown and are the same for all profiles. processes responsible for their formation and subsequent deformation and/or obliteration. This chapter mainly focuses on the geometry and evolution of marine terraces, while their chronology is discussed at length in the following chapter.

[8] As mentioned before, we use the well‐accepted model based on the correlation between a complete set of both stratigraphical and morphological features, which form a flight of marine terraces, and a set of highstands reached by the sea level during the interglacial and interstadial sub-

4 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 4. Longitudinal projection profile of the marine terraces mapped in the investigated area and traced parallel to the present‐day coast line. See also Figure 2 for projection orientation. (top) The raw data; (bottom) the colored dotted lines correlate the marine terraces across the major valleys, which in turn are represented by the black V‐shaped dashed lines. The black circles are the sites for which absolute ages are available from the literature and lengthily discussed in the text. Note that both Ponte del Re and San Teodoro II sites belong to the San Basilio terrace; both San Teodoro I and La Maddalena sites to the San Teodoro terrace, while both La Petrulla and Piano San Nicola sites to the Policoro terrace. Accordingly, each couple of sites should be coeval. stages, commonly labeled as MIS (Marine Isotope Substage) in the reference isotopic curves [e.g., Martinson et al., 1987; Shackleton, 1987; Chappell et al., 1996]. In particular, one of the fundamental assumptions in this model is the correlation between the inner edge of each terrace and one of the interglacial‐interstadial peaks of the relative sea level (RSL) curve. [9] In the general stratigraphical framework of the broader area [Pieri et al., 1996; Sabato, 1996; Tropeano et al., 2002], the investigated sequence of marine terraces is considered, as a whole, a regressive, generally coarsening upward, sedimentary body consisting of sand and conglomerate levels unconformably overlying the Pliocene‐Middle Pleistocene Argille subappennine Formation largely outcropping within the Bradanic Foredeep. In particular, within the

Metaponto area, these regressive coastal deposits show a downward shifting of detached wedges, therefore providing evidence for an important relative sea level fall [Tropeano et al., 2002]. Detailed sedimentary analyses carried out in three key areas (Taranto, Pisticci and Nova Siri) (Figures 1 and 2) applying the principles of the sequential stratigraphy [Grippa, 2006, 2007; Cilumbriello et al., 2007], clearly document that each terrace is characterized by a sedimentary sequence formed by paralic and beach deposits. [10] The morphological approach of this research mainly focused on the survey of the surfaces of the marine terraces and especially on their geographical limits. Indeed, any uplifted marine terrace generally shows a sub‐planar, slightly sloping offshore geometry (3°–5°). The surface is bounded uphill by the inner edge, downhill by the outer

5 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 5. Enlarged view of a small sector of the investigated area (Figure 2) showing (1) the details of the mapping; (2) the physical lateral continuity of the marine terraces with the corresponding (and coeval) fluvial terraces and (3) the progressive northward coast‐parallel restraining and final disappearing of some marine terraces associated with interstadial peaks (Marconia 2, San Teodoro 3, Metaponto 3 and Metaponto 2). The inner edges of these terraces geometrically merge with lower ones (both in map view and altitude) (see also Figure 4) because eroded during the subsequent highstand sea level and masked by the younger marine terrace [Anderson et al., 1999]. edge and laterally by fluvial incisions. Within the investigated area, the flat geometry of the younger/lower terraces is locally altered by several well‐preserved systems of sand dunes that are generally parallel to the coastline. [11] The coast‐parallel lateral continuity of the terrace surfaces is repeatedly interrupted by the valleys of both main rivers and secondary tributaries that form the hydrographic network of the area. It is noteworthy that the present‐day network began to form only during Middle Pleistocene due to the seaward river lengthening following the progressive entrenching of the marine sediments that were continuously uplifted and definitively brought into continental conditions [Brückner, 1980a, 1982; Tropeano et al., 2002]. Although the fluvial morphology is not a major topic of this paper, the good physical continuity especially for the younger marine terraces with their corresponding, genetically correlated and coeval fluvial terraces, which are discontinuously preserved along the lower valley slopes of the main rivers, is to be emphasized here, as it represents further evidence of the crucial and synergic role played by regional uplift and eustatic processes in the morphological evolution of this coastal area [Brückner, 1980b; Bianca and Caputo, 2003; Caputo and Bianca, 2009]. The physical lateral continuity between marine and fluvial terraces, as

clearly shown in Figure 5, is also a confirmation that the observed morphological scarps described here are not associated with a unique erosional/depositional surface cut and displaced by major normal faults as recently suggested [Bentivenga et al., 2004]. At this regard, the only mappable ‘normal’ sliding surface, in reality a large‐scale roto‐ translational landslide, affects a limited area near Laterza and it is discussed in more detail at the end of this chapter. [12] Additionally, the physical lateral continuity and geometrical correlation between marine and fluvial terraces dismisses the hypothesis of associating the latter morphological features with lowstand sea level periods as recently suggested by Westaway and Bridgland [2007] (see also discussion from Caputo and Bianca [2009]), therefore supporting the morphological evolutionary model proposed in this paper. [13] The most evident effect of the morphological evolution of the river valleys is that the older/higher marine terraces have been eroded more than the younger/lower ones, so that their present‐day coast‐parallel extent systematically decreases from the coastline to the innermost area, i.e., from the younger/lower terraces to the older/higher terraces. A further argument suggesting that the higher the terrace, the older the age is provided by pedogenetic and mineralogical

6 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

studies showing that weathering intensity (viz. age) increases with altitude [Brückner, 1980a]. [14] The other morphological element we focused on, namely the inner edge (viz. paleo‐shoreline or paleo‐cliff) of the marine terraces, is geometrically represented by a linear (in map view) (Figure 2) concave‐upwards knickpoint (in topographic profiles) (Figure 3) formed by the junction of the seaward slightly sloping surface of the terrace and the toe of the scarp separating the older (viz. previously uplifted) upper terrace surface. Therefore, the higher the slope angle of the scarp separating two terraces, the clearer the morphological evidence of the inner edge and, finally, the sharper its graphic representation on the geomorphological map is. [15] In the first phase of remote sensing analysis, we used different suites of stereographic pairs of air photos of various years and scales in order to always obtain both the best image quality and the suitable elevation exaggeration. The latter is crucial in order to better localize those inner edges developed at the toe of scarps characterized by a low slope angle. In a second phase, in order to check the geomorphological and stratigraphical reliability of the surveyed linear and planar features, all data obtained by the remote sensing analysis underwent a detailed and systematic ground control based on intensive field surveys. Following this multidisciplinary approach, the marine terraces and paleo‐ shorelines were thus mapped on I.G.M. (Italian Military Geographical Institute) topographic maps at scale 1:25,000, with principal and secondary contour lines at 25 m and 5 m, respectively. In order to minimize the errors in mapping the inner edges and especially in attributing their present‐day altitude, during the field‐work we took great care to determine the precise degree (or amount) of scarp degradation. This process might have accumulated a colluvial wedge at the base of the scarp (viz. debris slope) and hence partially mask the original knickpoint geometry. This aspect was carefully checked during the field work by thoroughly inspecting the numerous road‐cuts and natural entrenches. Accordingly, the estimated maximum error made when determining the elevation of a paleo‐shoreline is in the order of just a few meters, as derived from the accuracy given by the equidistance of 5 m of the secondary contour lines of the topographic maps. On the other hand, a higher centimetric‐ scale precision, based for example on GPS instruments, would be of little or no practical advantage for the purpose of this research. In conclusion, over the entire investigated area we directly checked in the field and mapped inner edges for an aggregate length of almost 1000 km, which represents the core of the created database in terms of points with full x‐y‐z coordinates. [16] The whole investigated terrace area is about 70 km‐ long and in the central sector up to 25 km‐wide from the coast line (Figure 2). The area is characterized by the occurrence of 18 orders of marine terraces, which have been named with as many local toponyms reported in all topographic maps of I.G.M. With regard to this, we purposely avoided using a sequential numbering, either increasing upwards or downward, for two main reasons. First, it will be shown that in different sectors of the investigated coastal area, the number of marine terraces is different because

TC4005

some of the terraces have been progressively destroyed or covered by younger ones. As a consequence, if numbering could be of practical use when a short coastal sector or just a single coast‐perpendicular profile is considered, the lack of continuous numerical sequence in most of the investigated sectors would likely increase the confusion among the readers. Second, if a new terrace, which for any reason was not preserved within the study area, will be possibly mapped in a contiguous coastal sector would oblige a complete re‐numbering. Third, although in terms of mapped terraces similarities exist with older researches, our new mapping and the consequent new numbering would have probably produced misunderstandings in comparing our results with previous ones. [17] Numerous topographic profiles (a selection of which is shown in Figure 3), roughly perpendicular to the sequence of paleo‐shorelines, have been carried out across the study area, thus emphasizing the vertical trend of the staircase‐like arrangement of the terrace surfaces, in order (1) to check and/or confirm the real vertical separation of the surveyed marine terraces and (2) to achieve a sharper determination of the elevation of each paleo‐shoreline. [18] As observed in map view and topographic profiles (Figures 2 and 3), the width of the terraces slightly varies along strike, but this geometric parameter in particular may strongly differ from terrace to terrace. In the context of the conceptual model proposed for the formation of marine terraces and basically due to the fact that all involved geological processes are time‐dependent, the width of each surface mainly depends on (1) lithological and textural characteristics of shore‐related deposits representing the terrace stratigraphic sequence, (2) persistence of the corresponding highstand sea level [e.g., Trenhaile, 1987; Anderson et al., 1999], which is also a function of uplift‐ rate, but especially (3) on the relative elevation, differential timing and persistence of the subsequent highstand sea level [e.g., Anderson et al., 1999]. It is commonly assumed that the longer the highstand sea level, the wider the marine surface. In contrast, Anderson et al. [1999] clearly show that the closer in time and elevation the immediately following highstand sea level is, and the longer its activity, the more the upper/old terrace is eroded at its base and this process progressively decreases the original width of the older uplifted surface. Accordingly, the widest surface of a flight of marine terraces should be not automatically associated with MIS 5.5 highstand as widely accepted in the scientific community. As further discussed in the following sections, it is also worth noting that some of the narrowest surfaces progressively disappear northward as it is clearly shown in Figure 5. [19] In order to emphasize the vertical distribution of the mapped inner edges, we performed a 70 km‐long longitudinal projection, parallel to the coastline, of the whole set of paleo‐shorelines, with the aim of assessing the deviations from their originally horizontal trend (Figure 4). The projection shows a clear fan‐shape, NNE‐wards‐converging geometry of all the paleo‐shorelines, therefore documenting that the cumulative amount of uplift of this coastal area progressively decreases toward the northeast.

7 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Table 1. Available Absolute Ages Based on Different Dated Material and Different Analytical Methods Like U‐Th, Amino Acid Racemization, Apparent Parabolic Kinetics, Accelerated Mass Spectrography and Infrared Stimulated Luminescencea Site

Terrace

Age

Method

Material

References

La Petrulla La Petrulla Piano San Nicola Fosso Marzoccolo La Maddalena San Teodoro I Ponte del Re San Teodoro II

Policoro Policoro Policoro Metaponto 3 San Teodoro San Teodoro San Basilio San Basilio

63.0 ± 3.0 55.2 ± 4.6 54.2 ± 6.5 >42 (42 ka that represents a lower time boundary. Accordingly, because (1) the Policoro terrace, which corresponds to the MIS 3.3 (see Piano San Nicola and La Petrulla sites) is immediately above, (2) the Metaponto terrace is locally at least 10 m lower (i.e., younger) and probably Holocene [Brückner, 1980a], the Metaponto 3 terrace containing the Fosso Marzoccolo site is chronologically constrained and must be correlated to an early MIS 3.1 (Tables 1 and 2). 3.4. La Maddalena Site [29] La Maddalena site is the southernmost site and located about 2 km southeast from Rocca Imperiale (Calabria) (Figure 2) on the San Teodoro terrace here characterized by a narrow surface whose inner edge is locally at an elevation of ca. 120 m a.s.l. The sampled terrace deposit consists of coarse littoral gravels that show a fossil content represented by remains of thick Glycymeris sp. shells. Amato et al. [1997] analyzed three sampled shells by the AAR method, obtaining a mean D/L ratio value of 0.34 ± 0.01. For this sector of the Mediterranean, this ratio represents an intermediate value between the Aminogroups C and E [Hearty et al., 1986], which have been associated with MIS 5.1(−5.3) and 5.5, respectively. However, following the above mentioned Belluomini et al. [2002] empirical relationship, the calculated age is 70.3 ± 3.2 ka. Given that the San Teodoro terrace is higher and obviously older than the Policoro surface (see La Petrulla and Piano San Nicola sites) (Figures 2 and 4), but the site does not contain any Senegalese fauna, the paleo‐coast line is probably correlated to a late MIS 5.1 (Tables 1 and 2). 3.5. San Teodoro I Site [30] The geological section of San Teodoro I site is close to La Petrulla site and has also been investigated by Zander et al. [2006] (Figure 2). The analyzed stratigraphic sequence is associated with the San Teodoro terrace, whose surface is at ca. 42 m a.s.l. in this sector. Several mainly sandy samples were collected from different stratigraphic levels and dated by both SAR and MAA protocols of the IRSL technique. We concentrated, as we did at the La Petrulla site on the samples belonging to the ‘main gravel layer’ of the terrace deposit which can be interpreted as the HST of the whole stratigraphic sequence [Brückner, 1980a]. The two protocols applied to coarse‐grained feldspars yielded ages of 81.4 ± 7.2 and 59.8 ± 5.5 ka by SAR and MAA, respectively. This range of ages and its mean value of 70.6 ka satisfactorily agree with the age obtained from La Maddalena site, also due to the compelling fact that the two sites must be coeval, being associated with the same marine terrace as mapped directly in the field (Figures 2 and 4). In conclusion, the correlation between the San Teodoro terrace with the late MIS 5.1 seems to be confirmed (Tables 1 and 2). 3.6. Ponte del Re Site [31] The Ponte del Re site is the northernmost dated site within the investigated area and is located on the left flank

TC4005

of the Lato River valley, about 5 km from the Ionian coast (Figure 2). From a sedimentary point of view, the site is represented by a regressive marine sequence consisting of littoral upward‐coarsening sands covered by a few‐meters‐ thick pebbly deposits. This sequence belongs to the San Basilio terrace, whose surface is at about 40 m a.s.l. at the Ponte del Re site and whose local inner edge is at about 50 m a.s.l. By means of the AAR method, Dai Pra and Hearty [1992] analyzed some samples of Glycimeris, collected in the level of fine sands outcropping at about 24 m a.s.l., showing a D/L ratio of 0.43 ± 0.03, which is commonly attributed to the Aminogroup E tentatively associated with MIS 5.5 [Hearty et al., 1986]. However, using again the empirical relationship between D/L ratios and U‐Th absolute ages [Belluomini et al., 2002], the absolute age calculated from the published D/L value is 99.4 ± 9.7 (Table 1). We therefore correlate the San Basilio terrace with an early MIS 5.3 (Tables 1 and 2). A possible objection to this relatively young age could be raised because of the occurrence of a rich fossil fauna interpreted as Senegalese association [Boenzi et al., 1985; Caldara, 1987], which is also traditionally correlated to the last interglacial (MIS 5.5). However, it has been documented that Senegalese species probably invaded the Mediterranean Sea either earlier (e.g., MIS 7) or later during the whole MIS 5, thus including MIS 5.3 and possibly 5.1 [Hillaire‐Marcel et al., 1996; Zazo, 1999; Zazo et al., 2003]. Moreover, the few samples of Strombus bubonius Lamark collected at the site belong to a ‘mixed fossil assemblage’ with a “modest autochthonous component” and come from a conglomeratic layer including rock boulders [Caldara, 1987]. Therefore, it is at the least questionable whether these fossils are coeval with the deposits or rather they could be considered as ‘simple’ pebbles eroded from an higher (i.e., older) terrace (G. Mastronuzzi, personal communication, 2008). Accordingly, the chronological correlation suggested for the San Basilio terrace (MIS 5.3) is not in contrast with the paleontological constraint as far as the latter is weak. Another argument commonly used by supporters of the MIS 5.5 age of this site is the local width of the terrace, which is much larger than the immediately higher Marconia surface (our suggested MIS 5.5 terrace). At this regard, we recall few basic morphological concepts showing up that the final width of each marine surface mainly depends on (1) the persistence of the corresponding highstand sea level, which is also a function of the uplift‐rate, (2) the relative elevation, differential timing and persistence of the subsequent highstand sea level and (3) the lithological and textural characteristics of the shore‐related deposits representing the terrace stratigraphic sequence and the substratum. Accordingly, the longer the highstand sea level, the wider the marine surface which forms is, while the closer in time and elevation the immediately following highstand sea level is, and the longer its activity, the more the upper terrace is eroded at its base and this process progressively decreases the original width of the older uplifted surface. Therefore, the common assumption of associating the largest terrace with the major interglacial peak is, at the best, not always correct. Our original and systematic mapping eventually proved this is a common place and sometimes an error of researchers working only on

10 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Figure 6 11 of 24

TC4005

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

a single coast‐perpendicular profile. The latter observation becomes obvious when mapping large coastal sectors. For example, within the investigated area and in the specific case of the Ponte del Re site, first, the width‐ratio between the San Basilio and Marconia terraces is inverted south of the Cavone River and, second, the San Basilio terrace is certainly not the widest surface and its width is systematically exceeded by the Bernalda terrace and locally even by the Policoro, Upper Montalbano and Metaponto terraces (Figure 2). In conclusion, the terrace width does not seem a crucial criterion to infer or constrain the chronological correlation between terraces and corresponding MISs. 3.7. San Teodoro II Site [32] The San Teodoro II site has also been investigated by Zander et al. [2006] (Figure 2) and it is represented by a geological section (ca. 60 m a.s.l.) cutting across the marine and continental deposits of the San Basilio terrace, whose inner edge is locally at 80 m. Also in this case, we concentrate on the chronological data related to the upper 2 m of the ‘main gravel layer‘ of the terrace, namely the HST of the stratigraphic sequence [Brückner, 1980a]. The absolute IRSL feldspar ages range from 82.3 ± 6.0 to 69.7 ± 7.9 and from 83.7 ± 7.4 to 69.5 ± 7.9 by SAR and MAA protocols, respectively. Although the mean ages (76.0 and 76.6 ka) are slightly older than those obtained at the San Teodoro I site, which belongs to the immediately lower San Teodoro terrace, they are not sufficiently old to fit an early MIS 5.3 as inferred from the above discussed Ponte del Re site and even much younger than a supposed MIS 5.5 age of the San Basilio terrace. In any case, the detailed mapping provides the physical constraint that the two sites are coeval because belonging to the same marine surface (Figures 2 and 4). Accordingly, it is unlikely that the northern site is even younger than already assumed. In agreement with Zander et al. [2006] who suggest the possibility of systematic underestimates in their IRSL ages, we maintain the chronological correlation between the San Basilio terrace and an early MIS 5.3 (Tables 1 and 2). 3.8. RSL Curves Correlations [33] Following the inferred correlations between some of the terraces (Metaponto 3, Policoro, San Teodoro and San Basilio) and specific MIS events (Table 2), in this section we extend the discussion to the other marine terraces mapped within the investigated area (Figures 2 and 3), trying to correlate the corresponding inner edges with as

TC4005

many highstand sea level peaks. A critical analysis of the differences occurring between the many RSL curves published in the international literature has been recently presented (see Caputo [2007] and references therein for the considered RSL curves). In particular, concerns have been raised regarding the timing and the relative altitude/depth of the major interglacial and interstadial peaks, while the following crucial features have been emphasized. First of all, with regard to the age‐height assessment of the major interglacial peaks (MIS 5.5, 7.5, 9.3 and 11.3), important differences, up to ca. 18 ka and more than 35 m, can occur in the different curves (Figure 6). Second, the position (height and age) and even the number of the major interstadial peaks markedly differ from curve to curve. For example, the height of the peaks corresponding to MIS 5.1 and MIS 6.5 varies as much as 65 m (Figure 6). Third, if the uplift‐rate is for example 1 mm/a, only one RSL curve [Shackleton, 2000] predicts the possibility that MIS 3.3 may have left some coastal morphological evidence (viz. a terrace). Similarly, a marine terrace associated with MIS 6.5 could be theoretically documented only if based on three curves out of fourteen (Figure 6). Fourthly in correspondence with the interstadial and/or interglacial periods, some of the curves present a couplet with peaks sometimes differing more than 15 ka; when referring to that MIS, which age should an end‐user of RSL curves apply? [34] Following the above criticisms and lacking any specific and scientifically valid criterion for choosing one curve instead of others, but being necessary for the morphotectonic analysis of marine terraces the reference to a RSL curve, we decided to consider many curves (14) among the most recently published and the most common and widely accepted by the scientific community (Figure 6). Although this procedure qualitatively resembles a statistical approach, the use of a rigorous statistical method is unlikely due to the complex and multiparameter nature of the different data sets and especially because there is no control over the degree of uncertainty associated with the diverse curves and even along the different segments of the single curves. For these reasons, in Figure 6 the proposed space‐time correlation between inner edges (short horizontal arrows along the ordinate axis) and highstand sea level peaks is graphically represented as a range of values (long inclined arrows). [35] Based on (1) the critical analysis of all absolute ages available in literature, as discussed in the previous section, and notwithstanding the uncertainties associated with the different dating techniques (OSL‐IRSL, U‐Th, AMS‐14C),

Figure 6. Chronological correlations between observed marine terraces and highstand sea level peaks. Fourteen RSL curves are stacked and represented as gray lines; they have been selected among the most recently published and the most common and widely accepted in the scientific community (modified from Caputo [2007, and references therein]). Numbers in italics indicate the principal marine isotopic stages. The present‐day altitude of the observed inner edges is represented by the short horizontal arrows along the ordinate axis, while the suggested chronological correlation is marked by a couple of inclined arrows for including age and sea level depth uncertainties. Terraces that disappear along the coast‐parallel profile (Figures 2, 3 and 4), due to progressively decreasing uplift‐rate, are emphasized by dashed lines. The inclination of the correlating arrows gives a measure of the long‐term uplift‐rate (see graphical clinometer for reference and Table 3). (a–f) The section locations from where the six diagrams have been obtained are represented in Figure 4. In the legends, terraces not appearing in the specific section are marked in light gray. Question marks and dotted arrows in Figure 6f refer to terraces affected by deep‐seated landsliding. 12 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

(2) the standard assumption that all observed and mapped terraces are correlated with a highstand sea level and (3) the fact that when a terrace formed well below the present‐day sea level and/or it was not sufficiently and rapidly uplifted, it was therefore masked and/or obliterated during the subsequent highstand event(s), it is possible to chronologically constrain the whole sequence of marine terraces and tentatively date all corresponding inner edges. The complete chronological correlations between the marine terraces mapped along the Ionian coast (Figure 2) and the major interglacial/interstadial highstand sea level peaks are shown in Figure 6 and are synthesized in Table 2, where the approximate age of each terrace is also indicated rounded to the 5 ka. [36] It is worth mentioning that different chronological correlations have been proposed in the literature [e.g., Boenzi et al., 1976; Brückner, 1980a, 1980b; Caldara, 1987; Dai Pra and Hearty, 1992; Bentivenga et al., 2004; Westaway and Bridgland, 2007; Cilumbriello et al., 2008]. However, it is also crucial to stress that in the few papers presenting original chronological data, the proposed correlations between (1) laboratory numerical results, (2) sampled terraces and (3) a specific MIS have been always attempted only for limited coastal sectors and sometimes just for the specific sampled terrace or site. As discussed at length in this chapter, this procedure has often led to wrong chronological inferences based (though uncritically largely accepted in the scientific community) on the erroneous morphological correlation between terraced surfaces. For example, San Nicola and Ponte del Re sites absolutely do not belong to the same terrace (see Figure 4) as stated by Dai Pra and Hearty [1992] (and subsequently assumed by many authors) and hence they simply cannot be coeval. 3.9. Further Chronological Constraints and Validation [37] Following the above tentative chronological correlations, the highest mapped marine terrace (Sivilia) is ca. 600 ka old (MIS 15). This inference is in good agreement with all paleogeographic reconstructions of the Bradanic Foredeep [e.g., Pieri et al., 1996; Sabato, 1996; Tropeano et al., 2002]. Indeed, the internationally known stratigraphic section of Montalbano Jonico, which has been proposed as a Global Stratotype Section and Point (GSSP) [Abbate et al., 2002; Ciaranfi and D’Alessandro, 2005], is located in this sector of the tectonic trough and within the investigated area (Figure 2). Based on nannofossil biostratigraphy, sapropel and oxygen isotope stratigraphy and magnetostratigraphy, it has been definitely documented that the top of the well‐ exposed, continuous and extended succession of hemipelagic deposits (Argille subappennine Formation) is here as young as early Middle Pleistocene (early “Ionian” stage) and well beyond the Brunhes‐Matuyama boundary (780 ka) [Ciaranfi et al., 1996, 2001; Maiorano et al., 2004, 2008; Stefanelli et al., 2005; Joannin et al., 2008]. As a matter of fact, this large amount of data pose a conclusive chronological ante‐quem constraint for the beginning of the coastal deposition [Cilumbriello et al., 2007] and hence for the terracing process. Taking also into account that (1) the top of the Montalbano Jonico hemipelagic stratigraphic succession

TC4005

is truncated by an erosional surface [e.g., Ciaranfi et al., 1996, 2001] and (2) the estimated water depth for the hemipelagic deposits is 250–350 m [D’Alessandro et al., 2003; Maiorano et al., 2008], marine sedimentary conditions likely persisted for a longer period after the 780 ka‐ boundary and therefore the oldest marine terrace within the investigated area is certainly younger than the above mentioned chronological constraint. This is in perfect agreement with the inferred ca. 600 ka age of the highest Sivilia terrace. [38] A further stratigraphic confirmation of the overall chronological correlations here proposed is provided by the relative abundance of volcanic minerals within the uppermost/oldest terrace deposits and their progressive decrease in the lower/younger ones [De Marco, 1990]. The provenance of this mineralogical assemblage is from the nearby Monte Vulture volcano whose activity is well documented during Middle Pleistocene [Cortini, 1975]. In particular, the youngest stratovolcano build‐up occurred between 0.66 and 0.58 Ma [Bonadonna et al., 1998], while the intense erosion of the volcanic apparatus took place up to the 300–400 ka. On the other hand, within the investigated area (near Pomarico) (Figure 2), tuffite layers have been documented in the Argille subappennine Formation [De Marco, 1990; Ciaranfi et al., 1996, 2001], but they are completely lacking in the marine terraced deposits. In summary, tephrochronology suggests that marine conditions still occur during the major eruptive phases (up to 660–580 ka) and therefore validates the lower chronological boundary for the suggested correlations and particularly between the oldest Sivilia terrace and MIS 15 (570–590 ka) (Figure 6 and Table 2). [39] Notwithstanding all the above arguments supporting the chronological correlation summarized in Table 2, we attempted a sort of validation on the proposed age of the flight of marine terraces by assuming a priori that the correlations presented in this paper are wrong. We just keep for granted that each marine terrace corresponds to a highstand sea level event. Accordingly, there are basically two possibilities: (1) for each coast‐perpendicular section the age of the marine terraces increases with their altitude or (2) it does not and the chronological sequence could jump up and down along the flight of marine terraces. We could exclude the latter case because the pedogenic alteration and the colluvial evolution at the top of the terraces, which are functions of time, increase with altitude [Brückner, 1980a, 1980b]. Moreoveor, horizontal‐to‐vertical seismic ratio (HVSR) and downhole measurements carried out on several terraces between the coast and Pisticci village (Figure 2) document a decrease of the seismic amplification effect and an increase of the shear wave velocity with increasing altitude of the terrace [Mucciarelli et al., 2007]. Accordingly, also these geophysical results can be straightforwardly associated with the progressive compaction and induration that naturally occur during aging of the terraced deposits. [40] As regards the other possibility (i.e., the terraces are chronologically ordered), in order to find a chronological correlation significantly different and really alternative to the one proposed here (Table 2) we should assume that for various (unknown) reasons some of the highstand sea level peaks did not leave any morphological evidence. As a

13 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Table 3. Estimated Long‐Term Uplift‐Rates for Six Selected Sectors Along the Investigated Areaa Sector

Mean Long‐Term Uplift Rate (mm/a)

a b c d e f

1.7–1.8 1.2–1.3 0.9–1.0 0.7–0.8 0.5–0.6 0.4

a

See Figure 4 for location.

consequence, beyond the need of explaining the geological mechanism for the lacking evidence of a specific terrace, the omission of one or more interglacial‐interstadial peaks implies also the need of correlating the highest mapped terraces with much older highstand sea level events, therefore exceeding the maximum oldest age of ca. 600 ka, which is fairly well constrained. Accordingly, a wrong conclusion implies a nul hypothesis and demonstrates the incorrect assumption.

4. Regional Tectonics and Geodynamics [41] On the basis of (1) vertical and areal distribution of mapped marine terraces (Figures 2, 3, 4 and 5), (2) their chronological correlation with as many highstand sea level events (Figure 6 and Table 2), and (3) depth/height of sea level during corresponding eustatic peaks (Figure 6) [see also Caputo, 2007, and references therein], it is thus possible to calculate long‐term uplift‐rates affecting the investigated area during mid‐late Quaternary (Table 3). [42] As shown in Figure 4, the whole sequence of marine terraces is clearly tilted NE‐wards. Accordingly, different coastal sectors have undergone differential amounts of cumulative uplift and uplift‐rates, which were both progressively increasing toward the southwest. As a direct consequence of this differential behavior, the probability that marine terraces associated with minor interstadial peaks have escaped the subsequent transgression and consequential erosion similarly increases toward the southwest as well as the number of marine terraces observed in the field. This phenomenon well explains why, for example, south of the Sinni River, 14 inner edges have been recognized and correlated to the MIS 7 to MIS 1 events (last ca. 240 ka) (Table 2), while north of the Bradano River, only 7 terraces could be mapped for the same time span (Figures 2, 3 and 4). As above mentioned, some of the terraces are observed only in the southern sectors, while toward the north they progressively reduce the altitude difference with respect to the immediately lower marine surface, therefore merging with it (Figures 2 and 5). North of these branch points, the morphological evidence of the older terrace disappears, having been completely eroded and obliterated by the younger sequential system tract. This is the case of terraces Metaponto 2 (morphologically not distinguishable north of the km 20 in Figures 4 and 5), Metaponto 3 (km 30), San Teodoro 2 (ca. km 10), San Teodoro 3 (km 25), Marconia 2 (ca. km 20) and Gaudella (km 52).

TC4005

[43] Following the above discussion and especially due to the differential uplift that characterizes the whole coastal region, it is thus more meaningful to estimate an average value for selected profiles, as represented in Figure 6. The average (in time) long‐term uplift‐rate in the six given coast‐perpendicular profiles varies from 1.8 to ca. 0.4 mm/a, from sector (a) to (f), respectively (Table 3). The observed altimetric and velocity differences are far outside the degree of uncertainty of all parameters considered in the calculation and hence are likely to be meaningful and relevant for mid‐ late Quaternary tectonic and geodynamic inferences. [44] Although they do not represent the major topic of this paper, also fluvial terraces have been carefully mapped along the major valleys providing useful information about the tilting that affected the southernmost sector of the investigated area. For example, along the Sinni Valley characterized by the highest uplift‐ and tilting‐rate values (Figures 4 and 6), morphological evidence of paleo‐alluvial surfaces has been recognized exclusively along the southern slope of the valley. This strongly asymmetric distribution is symptomatic of a similarly strong differential uplift that affected the two valley flanks. As a consequence, the phenomenon caused (1) the continuous northward tilting of the 2–3 km‐wide valley floor, (2) the preferential lateral erosion at the base of the northern valley slope and hence (3) the progressive removal of any morphological evidence of past alluvial plains on the hydrographic left side of the river, therefore preserving wide fluvial remnants only on the southern side. The lack of a similar asymmetric fluvial behavior along the northern valleys is also in agreement with the proposed evolutionary model of differential uplift and regional tilting. [45] Beyond the self‐evident coast‐parallel variation of the cumulative uplift and uplift‐rate, a more careful inspection of Figure 4 allows to distinguish three major sectors, each characterized by relatively uniform tilting (i.e., differential uplift‐rate) (Figure 7) and separated by two marked changes in both uplift‐rate and tilting‐rate. For example, the mean coast‐parallel slope of the Lower Montalbano terrace (MIS 7.5, 225–240 ka) is 1.27%, 0.51% and 0.28% in the three sectors, respectively. [46] Due to the regional scale of observed and mapped features and the inland dimension of the investigated area (Figure 2), these three sectors probably correspond to as many NW–SE trending zones parallel to the Southern Apennines chain‐foredeep‐foreland system (Figure 1) and characterized by the activity of different tectonic structures. [47] In order to understand the mechanism(s) and process (es) that caused this coast‐parallel variable behavior, we briefly recall the major tectonic elements that played a role in the region during mid‐late Quaternary: (1) the subducting Adria Plate characterized by a flexed geometry [Ricchetti and Mongelli, 1980; Doglioni et al., 1994; Argnani et al., 2001]; (2) the basal detachment separating the Apennines orogenic wedge from the subducting plate, running roughly on top of the Apulian carbonate succession and characterized by a convex‐upwards shape [Doglioni et al., 1996; Pieri et al., 1997; Scrocca et al., 2003; Finetti, 2005; Finetti et al., 2005]; (3) several minor contractional structures

14 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 7. Simplified profile of the marine terraces showing the major longitudinal tri‐partition of the inner edges. For the sake of simplicity, this segmented trend is emphasized for only three terraces, though the same behavior can be observed for the remaining surfaces. Arrows mark the corresponding knickpoints. Dashed lines represent the inner‐edges from field data, while dotted lines the intravalley correlations (see also Figure 4 for specific terrace names).

affecting the external sector of the orogenic wedge (Figure 1) [Sella et al., 1988; Senatore et al., 1988; Menardi Noguera and Rea, 2000; Patacca and Scandone, 2001; Lentini et al., 2002]. [48] It should also be noted that the first change in behavior between the southwestern and central sectors occurs between the Sinni and Agri rivers in correspondence with the most external sub‐emergent Rotondella‐Valsinni thrust that affects the orogenic wedge (Figure 1) [Bonardi et al., 1988; Bigi et al., 1992; Patacca and Scandone, 2001]. On the other hand, the boundary between the central and the northeastern sectors roughly corresponds to the vertical projection of the fault‐tip of the most external blind thrust (between the Cavone and Basento rivers), which represents the above mentioned basal detachment (Figure 1) [Consiglio Nazionale Delle Ricerche–Progetto Finalizzato Geodinamica, 1991; Patacca et al., 1993; Meletti et al., 2000]. [49] Having emphasized that, we interpret the differential uplift and tilting as due to the combined role played by three major tectonic structures working at different scales, with different rates and affecting partially different areas (Figure 8). First, in the southwestern sector, part of the observed superficial tilting is due to the progressive growth (amplification and tightening) of a broad ramp‐anticline and/or fault‐propagation fold associated with the out‐of‐ sequence (re‐)activation of the Rotondella‐Valsinni blind thrusts and its offshore extension (Figure 1) [Senatore et al., 1988; Hippolyte et al., 1994; Patacca and Scandone, 2001; Lentini et al., 2002]. For the sake of simplicity and from now on, we refer to this as the fold‐related process (Figure 8a). Within the investigated area, this 10–15 km wide morphological sector corresponds to the external northeastern flank of such a fold whose total width (across strike) is ca. 30–40 km [e.g., Bonardi et al., 1988; Bigi et al., 1992; Patacca and Scandone, 2001]. With the available morphotectonic information, it is not possible to exclude (but neither specifically support) an oblique‐slip, likely left‐

lateral, component of motion along these blind fault(s) as suggested by Del Ben et al. [2008] and Ferranti et al. [2009]. The amount of maximum cumulative uplift estimated for the hinge zone of the fault‐related fold is ca. 220 m for the last 225–240 ka (Lower Montalbano terrace) (Figure 9c). However, assuming that the out‐of‐sequence faulting‐folding has been active since ca. 600 ka (viz. the whole terrace sequence) with a roughly constant rate, the total uplift caused solely by the blind thrusting could have locally reached the 560 m along the crest of the anticline as represented in Figure 8a. [50] The second tectonic structure playing a role in the investigated area is represented by slip activity along the basal detachment (Figure 8b). Several geological, seismic, gravimetric and magnetotelluric profiles across the broader area [e.g., Ricchetti and Mongelli, 1980; Pieri et al., 1997; Menardi Noguera and Rea, 2000; Van Dijk et al., 2000; Scrocca et al., 2003; Finetti, 2005; Patella et al., 2005; Finetti et al., 2005] show that, the external sector of the basal detachment has a concave‐downward curved‐ shape (Figure 8b). As a consequence of displacement along such thrust geometry, the hanging wall top is tilted toward the foreland, thus causing a differential uplift at the earth’s surface. In this case, the activated tectonic structure is several tens of kilometers wide in a downdip direction, therefore involving a comparable dimension of orogenic wedge. From now on, this is referred to as the wedge‐rotation process (Figure 8b). The preserved terraces provide information only for the last ca. 280–295 ka (Upper Montalbano terrace) and allow to estimate an uplift of ca. 150 m (Figure 9b). Also in this case, assuming that this structure was active with a roughly constant slip‐rate for the last ca. 600 ka, the maximum cumulative uplift at the southwestern edge of the profile would have been of 250 m as reported in Figure 8b. In order to quantify displacement along the basal detachment, a simple arc‐shape geometry could be considered (Figure 8b). Assuming a curvature radius of 100 km, a

15 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Figure 8. Conceptual models to explain the differential uplift, and consequent tilting, within the different sectors of the investigated area: (a) fold‐related process, (b) wedge‐rotation process and (c) lithospheric duplexing process. For each process, the causative tectonic structures are represented by thick lines, while the thin dashed lines represent a pre‐deformation reference frame. The gray areas are indicative of the affected rock volumes and thus emphasize the different scale of the three geological processes and associated tectonic structures. Both vertical movements at the surface and slip are not to scale with the depth of the tectonic structures. The amount of cumulative uplift at the southwestern boundary of the investigated area (560, 250 and 550 m) is tentatively inferred for each mechanism by assuming that it was characterized by a steady state regime during the last ca. 600 ka. See text for explanations. 16 of 24

TC4005

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 9. Uplift and tilting contributed by the three different tectonic structures and associated processes: (a) lithospheric duplexing, (b) wedge‐rotation and (c) fold‐related. See text for discussion.

fault slip of only 0.6–0.8 km is sufficient to produce, across a distance of 30–40 km on top of the hanging wall block (Figures 7 and 9b), a differential uplift of about 250 m. Even considering a curvature radius of 200 km, the necessary slip for the same amount of differential uplift would, in any case, be less than 1.3–1.7 km (Figure 8b). Again, assuming a stationary regime for the last ca. 600 ka along the basal detachment, the estimated long‐term slip‐rate (dip‐slip component only) ranges between 1.0 and 2.8 mm/a. These values are 1+ order lower than in nearby active subduction zones, like the Calabrian and the Hellenic arcs [e.g., Jackson and McKenzie, 1988; Jackson, 1994; Le Pichon et al., 1995; Hollenstein et al., 2008; Serpelloni et al., 2007; Ganas and

Parsons, 2009], and satisfactorily agree with the relatively slow‐convergence collisional setting that characterized this sector of the Southern Apennines during mid‐late Quaternary. [51] The third tectonic mechanism claimed for explaining the regional tilting affecting the investigated area must occur on a much larger scale, necessarily involving (at least) the crust of the subducting Adria Plate and possibly the lithospheric mantle. We thus propose a deep‐seated duplexing process (Figure 8c) progressively replacing further subduction due to, and as a consequence of, the late Pliocene‐Early Pleistocene involvement of thick continental lithosphere in the Southern Apennines fold‐and‐thrust system [e.g.,

17 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Casnedi et al., 1982; Ricchetti et al., 1988; Patacca and Scandone, 1989; Doglioni et al., 1994, 1996; Menardi Noguera and Rea, 2000]. Enduring convergence between Apennines and Dinarides [Bertotti et al., 2001; Basili and Barba, 2007] and persisting compressional‐transpressional conditions during Quaternary [Philip, 1987; Console et al., 1989; Pieri et al., 1997; Turco and Zuppetta, 1998; Doglioni et al., 1999a; Mantovani et al., 2002; Hollenstein et al., 2003; Catalano et al., 2004; Scrocca et al., 2005; Serpelloni et al., 2005, 2007; Viti et al., 2006; Bennett et al., 2008; Del Ben et al., 2008; Nardo, 2007; Ferranti et al., 2008, 2009] likely caused an external and deeper jump of the convergence‐related shear zone, therefore shifting large crustal volumes of Adria Plate from the footwall to the hanging wall block. This process caused a progressive thickening of the crust [Doglioni et al., 1999b; Chiarabba et al., 2008; Steckler et al., 2008] and a partial duplexing as supported by geophysical information at depth and suggested by different authors [Menardi Noguera and Rea, 2000; Van Dijk et al., 2000; Mazzotti et al., 2000; Scrocca et al., 2005; Finetti et al., 2005; Solarino and Cassinis, 2007]. In the following, we thus refer to it as the lithospheric duplexing process (Figure 8c). The amount of uplift and especially of differential uplift (viz. tilting) induced at the surface strongly depends on the general shape, dimension and depth of the suggested tectonic structure. In particular, the corresponding amount of causative fault slip and overall shortening depend on the internal texture of the duplex and the dimensions of the single horses [e.g., Boyer and Elliot, 1982; Tanner, 1991; Davison, 1994]. The duplex system could have formed either at the base of the carbonate sequence within the evaporitic succession or, more likely, close to the crust‐mantle transition where the rheological contrast could have enhanced the strain concentration. In order to quantify this mechanism, as a first approximation we could tentatively consider an arc‐shape geometry similar to that of the wedge‐rotation process. As suggested by Doglioni et al. [1994, 2007] and Argnani et al. [2001] for this sector of the Adria Plate, the curvature radius for this intralithospheric shear‐zone could range between 400 and 800 km. Accordingly, the maximum differential uplift of about 550 m interpolated for the MIS 15 terrace (570– 590 ka) (Figure 9a) would be produced by a cumulative shear displacement of 2.0–4.6 km (Figure 8c). In the case of a stationary regime, the estimated long‐term shear‐rate at depth would thus range between 3.3 and 7.8 mm/a. [52] It is worth mentioning that the superficial inferred area affected by this lithospheric‐scale process could be much wider than the 70 km‐long investigated zone. In particular, it probably extends further NNE‐wards, offshore the Adriatic coast of Puglia, where the cumulative Middle‐ Late Pleistocene uplift tapers to zero (Figure 9a). The latter observation is in good agreement with many morphological data showing a substantial vertical stability, or very low uplift rates, along the NE Puglia coastal sector [e.g., Bordoni and Valensise, 1998; Mastronuzzi and Sansò, 2002; Ferranti et al., 2006; Mastronuzzi et al., 2007]. [53] Alternative tectonic and geodynamic models have been proposed in the literature for explaining the recent

TC4005

southern Italy uplift, like a slab‐rupture rebound, erosionally induced rebound, lithospheric doming or mantle wedge [e.g., Spakman et al., 1993; Westaway, 1993; Hippolyte et al., 1994; Patacca et al., 1993; Amato and Montone, 1997; Bordoni and Valensise, 1998; Doglioni et al., 1999a]. Although the present paper is not intended to be a review on the topic, we briefly analyze the mantle wedge model as a possible mechanism to explain the regional scale uplift and tilting observed across the investigated area. In order to better understand this process, we recall a few key geodynamic aspects. First, the rigid Adria continental lithosphere entered the Apennines subduction factory during Pliocene. Second, in the meantime, the Dinarides system characterized by an opposed vergence was progressively approaching. Third, the flexural effects of the two subductions, namely the peripheral bulging and the slab retreat, started mutually interfering and eventually locked each other. At this regard, in the lithosperic profiles reconstructed across the investigated area [Moretti and Royden, 1988; Doglioni et al., 1994, 2007; Argnani et al., 2001], the two slabs have a unique common peripheral bulge that obviously cannot further develop. As a consequence of persisting convergence between Apennines and Dinarides, the eastward mantle flow [Doglioni, 1990, 1991] counteracting on the Adria lithosphere progressively started, sometimes during the Pleistocene, to intrude between the two plates and to form a wedge of Tyrrhenian mantle. Although this model has been successfully proposed for unravelling the geodynamics of the Central Mediterranean [e.g., Doglioni et al., 1999b; Panza et al., 2007] and it can potentially induce both regional uplift and tilting, the large areal misfit of the possible superficial effects (mainly centered along the Apennines chain) with respect to the mapped marine terraces suggest that this process has unlikely played a direct role within the investigated area. Notwithstanding, this mechanism could have played an important indirect role for the two larger‐scale processes (lithospheric‐duplexing and wedge‐rotation) providing them the necessary rear‐push during their Middle‐Late Pleistocene activity.

5. Discussion and Concluding Remarks [54] The present research is essentially grounded on a systematic, completely new, mapping of a wide coastal sector (ca. 1500 km2) between northern Calabria and Puglia (southern Italy) allowing to recognize and reconstruct in detail the 3D geometry of a flight of marine terraces. Based on (1) the general trend of the whole terrace sequence (Figures 2, 3 and 4), (2) the critically revised chronological correlations (Table 2) and (3) the relatively constant (for each coast‐perpendicular profile) uplift‐rates during the last 600 ka (Figure 6), it is thus possible to suggest the occurrence, or better the concurrence, of three major tectonic structures operating at different scales and with different rates (fold‐related, wedge‐rotation and lithospheric duplexing). In a geological time‐scale perspective, the contemporaneous activity of the three structures could be considered as a typical effect of the strain partitioning that always characterizes large rock bodies undergoing deformation. The

18 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

TC4005

Figure 10. Simplified geological section crossing the southern Italy chain‐foredeep‐foreland system and running parallel to the investigated coastal sector (see Figure 1). Inset shows the location of the mapped marine terraces with respect to the profile below. The tectonic structures responsible of the three different‐scale processes are marked with thick lines in the profile, while the superficial sectors potentially affected by the three tilting processes are indicated on top of the profile. (1) Apulian continental crust; (2) Apulian carbonate sequence; (3) Apennines orogenic wedge including all undifferentiated tectonic units; (4) Quaternary deposits of the Bradanic Foredeep.

Southern Apennines are certainly not an exception. Regarding the evolution of the marine terraces, the major consequence of the tri‐partition and different size of the affected rock volumes is the partial overlap of the effects induced at the surface (viz. uplift and tilting). Indeed, in the southwestern sector of the investigated area, all three geological structures contributed to the progressive uplift and tilt of the terraces, but in the central sector, only the wedge‐rotation and lithospheric duplexing do so, while in the northeastern sector it is suggested that only the lithospheric‐scale structure is responsible for the raising and tilting of the marine terraces (Figures 8, 9 and 10). [55] Based on all presented data, discussed results and proposed tectonic evolution, an innovative and major conclusion of this paper implies that the most external sectors of the Apennines, and likely the whole chain‐foredeep‐ foreland system, remained under a regional prevailingly compressional tectonic regime during the entire Quaternary period. [56] According to several authors [e.g., Cinque et al., 1993; Patacca and Scandone, 2001, 2004], the shortening in the Apennines ceased at the beginning of the Middle

Pleistocene (“Ionian” stage). As a corollary of their statement it is commonly assumed that the frontal blind thrust (viz. basal detachment) was also de‐activated. On the other hand, their interpretation of a mid‐late Quaternary tectonic quiescence in the Bradanic Foredeep is mainly based on, and could be biased by, the lack of a ‘thrust‐related depositional sequence’ which has been defined as “the sedimentary record of a tectonic cycle” [Patacca and Scandone, 2004]. Following the same authors, the beginning of a tectonic cycle is characterized by the activity of a gently dipping sole thrust and the minimum creation of accommodation space within the basin. Due to the fact that southern Italy since at least Middle Pleistocene was affected by a lithospheric‐scale uplift [e.g., Cinque et al., 1993; Westaway, 1993], the Bradanic Trough underwent strong erosional conditions, which probably hindered the local stratigraphic record. In the light of this, the thin Middle‐Upper Pleistocene sedimentary successions associated with the investigated marine terraces [Brückner, 1980a; Tropeano et al., 2002; Cilumbriello et al., 2008] could be interpreted as the “condensed depositional sequence overlying a truncation

19 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

surface,” which is typical of the initial stages of a new tectonic cycle [Patacca and Scandone, 2004]. [57] Following the proposed model, the cumulative shortening along a NE–SW direction across the investigated area during the last 600 ka could be in the range 3–8 km, thus providing a mean regional long‐term convergence‐rate

TC4005

of 5–10 mm/a. Moreover, considering the estimated differential uplift separately for the three processes (Figure 11), it is possible to analyze the most recent activity (latest Pleistocene‐Holocene) of each tectonic structure. In particular, Figure 11a shows that the fold‐related process probably ceased its contribution (or its long‐term slip‐rate has abruptly decreased) well before the Holocene, because the straight interpolation of the best fit curve intersects the abscissa (i.e., 0 m/km) at about 30 ka BP. In contrast, for the other two processes the tilt versus time diagrams (Figures 11b and 11c) suggest that the associated structures still contribute to the general uplift and tilting (viz. differential uplift) of the region hence suggesting their persisting activity. Three further arguments support the conclusion of a recent (latest Quaternary) compressional activity. First, it is the Holocene age of the youngest involved inner edge (Table 2). Second, GPS velocities presented by Ferranti et al. [2008] for the region are in perfect agreement with our estimated values contributed by the two deeper tectonic structures. Third, just few kilometers offshore the investigated coastal sector, seismic profiles clearly document that the frontal thrust of the Apennines accretionary wedge is active and deforms the sea bottom [Pieri et al., 1997; Doglioni et al., 1999a; Scrocca et al., 2003; Finetti, 2005]. Further seismic profiles for hydrocarbon exploration carried out within the Gulf of Taranto have been recently declassified and made available to the scientific community (Ministero dello Sviluppo Economico UNMIG ‐ Società Geologica Italiana ‐ Assomineraria, Italia, Progetto ViDEPI Visibilità Dati Esplorazione Petrolifera in Italia, 2009, available online at http://unmig.sviluppoeconomico.gov.it). They provide additional evidence of contractional structures deforming the sea bottom along the frontal sector of the basal detachment thus confirming its very recent activity. Following a preliminary analysis of not‐migrated sections, the estimated amount of shortening is in the range of several hundreds meters. Taking into account that the cumulative slip could naturally taper toward the northeastern thrust tip and the sliding surface, especially along its frontal and shallower sector, is subhorizontal, parallel to bedding and represents a flat‐on‐flat structural setting, this macroscopic observation and inferred value are in good agreement with the independent estimates suggested on the base of our proposed model (0.6–1.7 km). [58] As a final comment, a direct consequence of the above conclusion is probably of primary importance for the seismic hazard of the region. At this regard, it is useful to

Figure 11. The coast‐parallel slope gradient, or degree of tilting (m/km or ‰), versus time (ka) estimated for the three tectonic structures and associated processes: (a) fold‐related, (b) wedge‐rotation and (c) lithospheric duplexing. During the last 30–40 ka, the contribution provided by the fold‐ related structure seems to have ceased (or strongly slowed down). The regression lines represent the mean tilt‐rate (m/km/ka). The high values of the R2 test on the regression lines suggest relatively uniform rates and hence a likely steady state regime of the causative geological processes. 20 of 24

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

consider separately the activity along the basal detachment and the lithospheric duplex. [59] The relevant information and inferences for the basal detachment could be summarized as follows: (1) possible fault dimensions are tens of kilometers‐long by tens of kilometers‐wide (for example, 70 × 40 km2; Figures 1 and 8), (2) the long‐term slip‐rate ranges between 1.0 and 2.8 mm/a, (3) the corresponding maximum magnitude could be around 7.5 [e.g., Wells and Coppersmith, 1994; Stirling et al., 2002], (4) associated co‐seismic displacements about 2–4 m [e.g., Chinnery, 1969; Scholz, 1982] and hence (5) mean repeat time for major characteristic earthquakes between ca. 700 and 4000 years. Notwithstanding the completeness of the Italian seismic catalogs for such strong events for the last thousand years, no record exists for the investigated area. Taking also into account that the focal depth would be relatively shallow (Figure 10), the potential seismic hazard associated with the basal detachment (wedge‐rotation process) is possibly high. [60] As concerns this lithospheric duplexing process and though the amount of cumulative shortening is possibly larger, the internal strain partitioning typical of these contractional structures characterized by multiple subparallel low‐angle faults generally reduces the slip‐rate on each sliding plane. However, a major uncertainty in assessing the seismic hazard of this shear zone follows the difficulty to define the current mechanical behavior of the affected rock volume, which strongly depends on the real depth of the duplex structure. Indeed, if it occurs in the lower crust below the brittle‐ductile transition (Figure 10), a viscous rheology prevails and deformation likely takes place aseismically. In contrast, if the uppermost lithospheric mantle is involved, the rheological profiles of the area [e.g., Viti et al., 1997] suggest a brittle behavior characterized by a typical stick‐ slip mechanism. In the former case, the corresponding seismogenic potential is low, while it would be much larger if the latter hypothesis holds. Further investigations and

TC4005

geological constraints will probably help in unravelling the situation.

Appendix A [61] In June 2009, the Executive Committee of the International Union of Geological Sciences (IUGS) ratified a proposal by the International Commission on Stratigraphy [see Gibbard et al., 2009] to formally subdivide the Pleistocene Epoch into “Lower,” “Middle” and “Late” Subepochs. The first one includes the stages “Gelasian” (formerly corresponding to the late Pliocene) and “Calabrian,” the second corresponds to the “Ionian” and the third stage is also referred to as “Tarantian” [Cita et al., 2006, Cita, 2008]. A Global Stratotype Section and Point (GSSP) has been defined for both Pliocene‐Gelasian (2.588 Ma) and Gelasian‐Calabrian boundaries (1.806 Ma), while the base of the Ionian stage is posed at the Matuyama‐Brunhes polarity reversal (0.781 Ma) and the proposed criterion guide for the beginning of Late Pleistocene (Tarantian) is the base of MIS 5 (0.127 Ma) [Litt and Gibbard, 2008]. In the present paper, in order to avoid possible confusion between the geographical and chronological meanings of “Calabrian,” “Ionian” and “Tarantian” (three regions of southern Italy and three Pleistocene stages, respectively), when referring to ages we will preferentially use terms like Lower/ Early, Middle and Upper/Late Pleistocene instead of the corresponding stage names. Terms like lower/early, mid/ middle and upper/late remain unformal subdivisions for the Quaternary Period, though widely used in a quasi‐formal way in the European literature. [62] Acknowledgments. Thanks to M. Tropeano, A. Grippa and G. Mastronuzzi for the fruitful discussions and for sharing their accurate stratigraphic and morphological information. Thanks to C.‐D. Reuther, L. Ferranti and an anonymous reviewer for carefully revising the manuscript and to the editors C. Doglioni, C. Faccenna and O. Oncken for their useful comments.

References Abbate, E., et al. (2002), Quaternari chronostratigraphy and the establishment of related standards, Episodes, 25(4), 264–267. Agenzia per la Protezione dell’Ambiente e per i servizi Tecnici (2004), Geological Map of Italy, 1:1,2500,000 scale, edited by B. Compagnoni and F. Galluzzo, Serv. Geol. D’Italia, Soc. Elaborazioni Cartogr., Firenze, Italy. Amato, A., and P. Montone (1997), Present‐day stress field and active tectonics in southern peninsular Italy, Geophys. J. Int., 130(2), 519–534, doi:10.1111/ j.1365-246X.1997.tb05666.x. Amato, A., G. Belluomini, A. Cinque, M. Manolio, and F. Ravera (1997), Terrazzi marini e sollevamenti tettonici quaternari lungo il margine ionico dell’Appennino lucano, Il Quaternario, 10(2), 329–336. Anderson, R. S., A. L. Densmore, and M. A. Ellis (1999), The generation and degradation of marine terraces, Basin Res., 11, 7–19, doi:10.1046/j.13652117.1999.00085.x. Argnani, A., F. Frugoni, R. Cosi, M. Ligi, and P. Favali (2001), Tectonics and seismicity of the Apulian Ridge south of Salento peninsula (Southern Italy), Ann. Geofis., 44(3), 527–540. Basili, R., and S. Barba (2007), Migration and shortening rates in the northern Apennines, Italy: Implications for seimic hazard, Terra Nova, 19(0), 1–7.

Belluomini, G., M. Caldara, C. Casini, M. Cerasoli, L. Manfra, G. Mastronuzzi, G. Palmentola, P. Sansò, P. Tuccimei, and P. L. Vesica (2002), The age of Late Pleistocene shorelines and tectonic activity of Taranto area, Southern Italy, Quat. Sci. Rev., 21, 525–547, doi:10.1016/S0277-3791(01)00097-X. Bennett, R., S. Hreinsdòttir, G. Buble, T. Bašić, Ž. Bačić, M. Marjanović, G. Casale, A. Gendaszek, and D. Cowan (2008), Eocene to present subduction of southern Adria mantle lithosphere beneath the Dinarides, Geology, 36(1), 3–6, doi:10.1130/ G24136A.1. Bentivenga, M., M. Coltorti, G. Prosser, and E. Tavarnelli (2004), Deformazioni distensive recenti nell’entroterra del Golfo di Taranto: Implicazioni per la realizzazione di un deposito geologico per scorie nucleari nei pressi di Scanzano Ionico (Basilicata), Boll. Soc. Geol. Ital., 123, 391–404. Bertotti, G., V. Picotti, C. Chilovi, R. Fantoni, S. Merlini, and A. Mosconi (2001), Neogene to Quaternary basins in the south Adriatic (Central Mediterranean): Foredeeps and lithospheric buckling, Tectonics, 20(5), 771–787, doi:10.1029/2001TC900012. Bianca, M., and R. Caputo (2003), Analisi morfotettonica ed evoluzione quaternaria della Val d’Agric., Appennino meridionale, Il Quaternario, 16(2), 159–170.

21 of 24

Bianca, M., C. Monaco, L. Tortorici, and L. Cernobori (1999), Quaternary normal faulting in southeastern Sicily (Italy): A seismic source for the 1693 large earthquake, Geophys. J. Int., 139, 370–394, doi:10.1046/j.1365-246x.1999.00942.x. Bigi, G., G. Bonardini, R. Catalano, D. Cosentino, F. Lentini, M. Parlotto, R. Sartori, P. Scandone, and E. Turco (1992), Structural Model of Italy, 1:500,000, Cons. Naz. delle Ric., Rome. Boenzi, F., G. Palmentola, and A. Valduga (1976), Caratteri geomorfologici dell’area del Foglio “Matera”, Boll. Soc. Geol. Ital., 95, 527–566. Boenzi, F., M. Caldara, and L. Pennetta (1985), La trasgressione tirreniana nei dintorni di Castellaneta (Taranto), Geol. Appl. Idrogeol., 20(1), 163–175. Bonadonna, F. P., D. Brocchini, M. A. Laurenzi, C. Principe, and G. Ferrara (1998), Stratigraphical and chronological correlations between Monte Vulture volcanics and sedimentary deposits of the Venose Basin, Quat. Int., 47–48, 87–96, doi:10.1016/S1040-6182(97)00074-8. Bonardi, G., et al. (1988), Carta Geologica Dell’appennino Meridionale, 1:250,000, Cons. Naz. delle Ric., Rome. Bordoni, P., and G. Valensise (1998), Deformation of the 125 ka marine terrace in Italy: Tectonic implications, in Coastal Tectonics, edited by I. Stewart and C. Vita‐Finzi, Geol. Soc. Spec. Pub., 146, 71–110.

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

Bosi, C., L. Carobene, and A. Sposato (1996), Il ruolo dell’eustatismo nella evoluzione geologica nell’area mediterranea, Mem. Soc. Geol. Ital., 51, 363–382. Boyer, S. E., and D. Elliot (1982), Thrust systems, Am Assoc. Pet. Geol. Bull., 66, 1196–1230. Brückner, H. (1980a), Marine Terrassen in Süditalien. Eine quartärmorphologische Studie über das Küstentiefland von Metapont, Dusseld. Geogr. Schr., 14, 1–235. Brückner, H. (1980b), Flussterrassen und Flusstäler im Küstentiefland von Metapont (Süditalien) und ihre Beziehung zu Meeresterrassen, Dusseld. Geogr. Schr., 15, 5–32. Brückner, H. (1982), Ausmass von Erosion und Akkumulation im Verlauf des Quartärs in der Basilicata (Süditalien), Z. Geomorph. N. F., 43, suppl., 121– 137. Caldara, M. (1987), La sezione tirreniana di Ponte del Re (Castellaneta Marina, Taranto): Analisi paleoecologica, Atti Soc. Toscana Sc. Nat., XCIII(1986), 129–163. Caldara, M., G. Mastronuzzi, G. Palmentola, P. Sanso, P. Tuccimei, and P. L. Vesica (2003), Reply to the comment by P.J. Hearty and G. Dai Pra, Quat. Sci. Rev., 22, 2369–2371, doi:10.1016/S0277-3791(03) 00150-1. Caputo, R. (2007), Sea level curves: Perplexities of an end‐user in morphotectonic applications, Global Planet. Change, 57(3–4), 417–423, doi:10.1016/j. gloplacha.2007.03.003. Caputo, R., and M. Bianca (2009), Comment on “Late Cenozoic uplift of southern Italy deduced from fluvial and marine sediments: Coupling between surface processes and lower‐crustal flow” by Westaway R. and Bridgland D. (Quaternary International 175, 86–124), Quat. Int., 204, 98–102, doi:10.1016/j.quaint.2008.09.011. Casnedi, R., U. Crescenti, and M. Tonna (1982), Evoluzione dell’avanfossa adriatica meridionale nel Plio‐ Pleistocene, sulla base di dati di sottosuolo, Mem. Soc. Geol. Ital., 24, 243–260. Castello, B., G. Selvaggi, C. Chiarabba, and A. Amato (2005), CSI Catalogo Della Sismicità Italiana 1981–2002, version 1.0, Inst. Naz. Di Geofis. E Vulcanol.–Cent. Naz. Terremoti, Roma. Catalano, S., C. Monaco, L. Tortorici, W. Paltrinieri, and N. Steel (2004), Neogene‐Quaternary tectonic evolution of the southern Apennines, Tectonics, 23, TC2003, doi:10.1029/2003TC001512. Chappell, J., A. Omura, T. Esat, M. McCulloch, J. Pandolfi, Y. Ota, and B. Pillans (1996), Reconciliation of late Quaternary sea levels derived from coral terraces at Huon Peninsula with deep sea oxygen isotope records, Earth Planet. Sci. Lett., 141, 227–236, doi:10.1016/0012-821X(96)00062-3. Chiarabba, C., L. Jovane, and R. Di Stefano (2005), A new view of Italian seismicity using 20 years of instrumental recordings, Tectonophysics, 395, 251–268, doi:10.1016/j.tecto.2004.09.013. Chiarabba, C., P. De Gori, and F. Speranza (2008), The southern Tyrrhenian subduction zone: Deep geometry, magmatism and Plio‐Pleistocene evolution, Earth Planet. Sci. Lett., 268, 408–423, doi:10.1016/j. epsl.2008.01.036. Chinnery, M. A. (1969), Earthquake magnitude and source parameters, Bull. Seismol. Soc. Am., 59(5), 1969–1982. Ciaranfi, N., and A. D’Alessandro (2005), Overview of the Montalbano Jonico area and section: A proposal for a boundary stratotype for the lower‐middle Pleistocene, Southern Italy Foredeep, Quat. Int., 131, 5–10, doi:10.1016/j.quaint.2004.07.003. Ciaranfi, N., M. Marino, L. Sabato, A. D’Alessandro, and R. De Rosa (1996), Studio geologico stratigrafico di una successione infra e mesopleistocenica nella parte sudoccidentale della Fossa Bradanica (Montalbano Ionico, Basilicata), Boll. Soc. Geol. Ital., 115, 379–391. Ciaranfi, N., A. D’Alessandro, A. Girone, P. Maiorano, M. Marino, D. Soldani, and S. Stefanelli (2001), Pleistocene sections in the Montalbano Jonico area and the potential GSSP for Early Middle Pleisto-

cene in the Lucania Basin (Southern Italy), Mem. Sci. Geol., 53, 67–83. Cilumbriello, A., L. Sabato, M. Tropeano, S. Gallicchio, A. Grippa, and P. Pieri (2007), Terraced coarse‐ grained coastal deposits in the southern part of the Bradanic Trough (Middle‐Late Pleistocene, Southern Italy), Epitome, 2, 400. Cilumbriello, A., M. Tropeano, and L. Sabato (2008), The Quaternary terraced marine‐deposits of the Metaponto area (Southern Italy) in a sequence‐ stratigraphic perspective, in Advances in Application of Sequence Stratigraphy in Italy, edited by A. Amorosi et al., GeoActa Spec. Publ., 1, 29–54. Cinque, A., E. Patacca, P. Scandone, and M. Tozzi (1993), Quaternary kinematic evolution of the southern Apennines. Relationship between surface geological features and deep lithospheric structures, Ann. Geofis., 36(2), 249–260. Cita, M. B. (2008), Summary of Italian marine stages of the Quaternary, Episodes, 31(2), 251–254. Cita, M. B., L. Capraro, N. Ciaranfi, E. Di Stefano, M. Marino, D. Rio, R. Sprovieri, and G. B. Vai (2006), Calabrian and Ionian: A proposal for the definition of Mediterranean stages for the Lower and Middle Pleistocene, Episodes, 29(2), 107–114. Cita, M. B., et al. (2008), The Calabrian stage redefined, Episodes, 31(4), 418–429. Consiglio Nazionale Delle Ricerche–Progetto Finalizzato Geodinamica (1991), Synthetic Structural‐ Kinematic Map of Italy. Structural Model of Italy, Sheet n. 5, Soc. Elaborazioni Cartograf., Florence, Italy. Console, R., R. Di Giovambattista, P. Favali, and G. Smriglio (1989), Lower Adriatic Sea seismic sequence (January 1986), Spatial definition of the seismogenic structure, Tectonophysics, 166, 235– 246, doi:10.1016/0040-1951(89)90216-3. Cortini, M. (1975), Età K‐Ar del Monte Vulture (Lucania), Rivista Ital. Geofis., 2, 45–46. Dai Pra, G., and P. J. Hearty (1992), I livelli marini pleistocenici del Golfo di Taranto, sintesi geocronostratigrafica e tettonica, Mem. Soc. Geol. Ital., 41(1988), 637–644. D’Alessandro, A., R. La Perna, and N. Ciaranfi (2003), Response of macrobenthos to changes in palaeoenvironments in the Lower–Middle Pleistocene (Lucania Basin, Southern Italy), Il Quaternario, 16, 167–182. Davison, I. (1994), Linked fault systems; extensional, strike‐slip and continental, in Continental Deformation, edited by P. L. Hancock, 121–142, Pergamon, New York. Del Ben, A., C. Barnaba, and A. Taboga (2008), Strike‐ slip systems as the main tectonic features in the Plio‐Quaternary kinematics of the Calabrian Arc, Mar. Geophys. Res., doi:10.1007/s11001-0079041-6. De Marco, A. (1990), Rapporti tra geodinamica e sedimentazione nella Fossa Bradanica durante il Pleistocene: Testimonianze mineralogiche, Boll. Soc. Geol. Ital., 109, 313–324. Di Bucci, D., B. Massa, M. Tornagli, and A. Zuppetta (2006), Structural setting of the Southern Apennine fold‐and‐thrust belt (Italy) at hypocentral depth: The Calore Valley case history, J. Geodyn., 42, 175–193, doi:10.1016/j.jog.2006.07.001. Doglioni, C. (1990), The global tectonic pattern, J. Geodyn., 12(1), 21–38, doi:10.1016/0264-3707 (90)90022-M. Doglioni, C. (1991), A proposal for the kinematic modelling of W‐dipping subductions ‐ possible applications to the Tyrrhenian‐Apennines system, Terra Nova, 3, 423–434, doi:10.1111/j.1365-3121.1991. tb00172.x. Doglioni, C., F. Mongelli, and P. Fieri (1994), The Puglia uplift (SE Italy): An anomaly in the foreland of the Apenninic subduction due to buckling of a thick continental lithosphere, Tectonics, 13(5), 1309– 1321, doi:10.1029/94TC01501. Doglioni, C., M. Tropeano, F. Mongelli, and P. Pieri (1996), Middle‐Late Pleistocene uplift of Puglia:

22 of 24

TC4005

An “anomaly” in the Apennines foreland, Mem. Soc. Geol. Ital., 51, 101–117. Doglioni, C., S. Merlini, and G. Cantarella (1999a), Foredeep geometries at the front of the Apennines in the Ionian Sea (central Mediterranean), Earth Planet. Sci. Lett., 168, 243–254, doi:10.1016/ S0012-821X(99)00059-X. Doglioni, C., P. Harabaglia, S. Merlini, F. Mongelli, A. Peccerillo, and C. Piromallo (1999b), Orogens and slabs vs. their direction of subduction, Earth Sci. Rev., 45, 167–208, doi:10.1016/S0012-8252 (98)00045-2. Doglioni, C., E. Carminati, M. Cuffaro, and D. Scrocca (2007), Subduction kinematics and dynamic constraints, Earth Sci. Rev., 83, 125–175, doi:10.1016/j.earscirev.2007.04.001. Dumas, B., P. Guérémy, and J. Raffy (2005), Evidence for sea‐level oscillations by the “characteristic thickness” of marine deposits from raised terraces of Southern Calabria (Italy), Quat. Sci. Rev., 24, 2120–2136, doi:10.1016/j.quascirev.2004.12.011. Ferranti, L., et al. (2006), Markers of the last interglacial sea‐level high stand along the coast of Italy: Tectonic implications, Quat. Int., 145–146, 30–54, doi:10.1016/j.quaint.2005.07.009. Ferranti, L., et al. (2008), Active deformation in Southern Italy, Sicily and southern Sardinia from GPS velocities of the Peri‐Tyrrhenian Geodetic Array (PTGA), Boll. Soc. Geol. Ital., 127(2), 299–316. Ferranti, L., E. Santoro, M. E. Mazzella, C. Monaco, and D. Morelli (2009), Active transpression in the northern Calabria Apennines, southern Italy, Tectonophysics, 476, 226–251, doi:10.1016/j.tecto.2008. 11.010. Finetti, I. R. (2005), Understanding the Ionides and their geodynamics, in CROP PROJECT: Deep Seismic Exploration of the Central Mediterranean and Italy, edited by I. R. Finetti, chap. 10, pp. 197–207, Elsevier, San Diego. Finetti, I. R., F. Lentini, S. Carbone, A. Del Ben, A. Di Stefano, P. Guarnieri, M. Pipan, and A. Prizzon (2005), Crustal tectono‐stratigraphiy and geodynamics of the Southern Apennines from CROP and other integrated geophysical‐geological data, in CROP PROJECT: Deep Seismic Exploration of the Central Mediterranean and Italy, edited by I. R. Finetti, chap. 12, pp. 225–262, Elsevier, San Diego. Fuchs, T. (1874), Die Tertiärbildungen von Tarent, Sitzungsherichte Dtsch. Akad. Wiss., 70, 193–197. Ganas, A., and T. Parsons (2009), Three‐dimensional model of Hellenic Arc deformation and origin of the Cretan uplift, J. Geophys. Res., 114, B06404, doi:10.1029/2008JB005599. Gibbard, P. L., et al. (2009), Formal ratification of the Quaternary System/Period and the Pleistocene Series/Epoch with a base at 2.58 Ma, J. Quat. Sci., 25(2), 96–102, doi:10.1002/jqs.1338. Grippa, A. (2006), Studio stratigrafico‐sedimentologico di successioni pleistoceniche sabbioso‐conglomeratiche affioranti nell’area di Marconia di Pisticci (entroterra del Golfo di Taranto), M.Sc. thesis, Univ. of Basilicata, Potenza, Italy. Grippa, A. (2007), Comparison between ancient and present gravelly sandy beach systems: An example from Pisticci and Nova Siri areas (Basilicata, Southern Italy), Epitome, 2, 400. Hearty, P. J., and G. Dai Pra (1992), The age and stratigraphy of Middle Pleistocene and younger deposits along the Gulf of Taranto (Southeast Italy), J. Coastal Res., 8(4), 882–905. Hearty, P. J., and G. Dai Pra (2003), Comment on: “The age of Late Pleistocene shorelines and tectonic activity of Taranto area, Southern Italy” by G. Belluomini, M. Caldara, C. Casini, M. Cerasoli, L. Manfra, G. Mastronuzzi, G. Palmentola, P. Sansò, P. Tuccimei, and P.L. Vesica [Quaternary Science Reviews, 21, 525–547], Quat. Sci. Rev., 22, 2363– 2367, doi:10.1016/S0277-3791(03)00149-5. Hearty, P. J., G. H. Miller, C. E. Stearns, and B. J. Szabo (1986), Aminostratigraphy of Quaternary

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

shorelines around the Mediterranean basin, Geol. Soc. Am. Bull., 97, 850–858, doi:10.1130/00167606(1986)972.0.CO;2. Hillaire‐Marcel, C., C. Gariepy, B. Ghaleb, J. L. Goy, C. Zazo, and J. Cuerda (1996), U series in Tyrrhenian deposits from Mallorca. Further evidence for two last‐Interglacial high sea levels in Balearin Islands, Quat. Sci. Rev., 44, 276–282. Hippolyte, J. C., J. Angelier, F. Roure, and P. Casero (1994), Piggyback basin development and thrust belt evolution: Structural and palaeostress analyses of Plio‐Quaternary basins in the Southern Apennines, J. Struct. Geol., 16(2), 159–173, doi:10.1016/0191-8141(94)90102-3. Hollenstein, C., H. ‐G. Kahle, A. Geiger, S. Jenny, S. Goes, and D. Giradini (2003), New GPS constraints on the Africa‐Eurasia plate boundary zone in southern Italy, Geophys. Res. Lett., 30(18), 1935, doi:10.1029/2003GL017554. Hollenstein, C., M. D. Müller, A. Geiger, and H.‐G. Kahle (2008), Crustal motion and deformation in Greece from a decade of GPS measurements, 1993–2003, Tectonophysics, 449, 17–40, doi:10.1016/j. tecto.2007.12.006. Jackson, J. (1994), Active tectonics of the Aegean region, Annu. Rev. Earth Planet. Sci., 22, 239–271, doi:10.1146/annurev.ea.22.050194.001323. Jackson, J., and D. McKenzie (1988), The relationships between plate motions and seismic moment tensors, and the rates of active deformation in the Mediterranean and Middle East, Geophys. J., 93(1), 45–73, doi:10.1111/j.1365-246X.1988.tb01387.x. Joannin, S., N. Ciaranfi, and S. Stefanelli (2008), Vegetation changes during the late Early Pleistocene at Montalbano Jonico (Province of Matera, southern Italy) based on pollen analysis, Palaeogeogr. Palaeoclimatol. Palaeoecol., 270, 92–101, doi:10.1016/j.palaeo.2008.08.017. Lajoie, K. R. (1986), Coastal tectonics, in Active Tectonics, Stud. Geophys., vol. 15, edited by R. E. Wallace, pp. 95–124, National Acad. Press, Washington D. C. Lentini, F., S. Carbone, A. Di Stefano, and P. Guarnieri (2002), Stratigraphical and structural constraints in the Lucanian Apennines (southern Italy): Tools for reconstructing the geological evolution, J. Geodyn., 34, 141–158, doi:10.1016/S0264-3707(02)00031-5. Le Pichon, X., N. Chamot‐Rooke, S. Lallemant, R. Noomen, and G. Veis (1995), Geodetic determination of the kinematics of central Greece with respect to Europe: Implications for eastern Mediterranean tectonics, J. Geophys. Res., 100(B7), 12,675–12,690. Litt, T., and P. Gibbard (2008), A proposed Global Starotype Section and Point (GSSP) for the base of the Upper (Late) Pleistocene Subseries (Quaternary System/Period), Episodes, 31(2), 260–263. Maiorano, P., M. Marino, E. Di Stefano, and N. Ciaranfi (2004), Calcareous nannofossil event in the Lower‐ Middle Pleistocene transitino at the Montalbano Jonico section and ODP site 964: Calibration with isotope and sapropel stratigraphy, Rivista Ital. Paleont. Stratigr., 110(2), 547–557. Maiorano, P., et al. (2008), Paleoenvironmental changes during the sapropel 19(i‐cycle 90) deposition: Evidences from geochemical, mineralogical and micropaleontological proxies in the mid‐Pleistocene Montalbano Jonica land section (southern Italy), Palaeogeogr. Palaeoclimatol. Palaeoecol., 257, 308–334, doi:10.1016/j.palaeo.2007.10.025. Mantovani, E., D. Albarello, D. Babbucci, C. Tamburelli, and M. Viti (2002), Trench‐arc‐back arc systems in the Mediterranean area: Examples of extrusion tectonics, J. Virtual Explor., 8, 127–144. Martinson, D. G., N. G. Pisias, J. D. Hays, J. Imbrie, T. C. Moore Jr., and N. J. Shakleton (1987), Age dating and the orbital theory of ice ages: Development of a high‐resolution 0 to 300,000 years chronastratigraphy, Quat. Res., 27, 1–29, doi:10.1016/ 0033-5894(87)90046-9. Mastronuzzi, G., and P. Sansò (2002), Pleistocene sea‐ level changes, sapping process and development of

valley networks in the Apulia region (southern Italy), Geomorphology, 46, 19–34, doi:10.1016/ S0169-555X(01)00172-6. Mastronuzzi, G., Y. Quinif, P. Sansò, and G. Selleri (2007), Middle‐Late Pleistocene polycyclic evolution of a stable coastal area (southern Apulia, Italy), Geomorphology, 86, 393–408, doi:10.1016/j.geomorph.2006.09.014. Mazzotti, A., E. Stucchi, G. Fradelizio, L. Zanzi, and P. Scandone (2000), Seismic exploration in complex terrains: A processing experience in the southern Apennines, Geophysics, 65, 1402–1417, doi:10.1190/1.1444830. Meletti, C., E. Patacca, and P. Scandone (2000), Construction of a seismotectonic model: The case of Italy, Pure Appl. Phys., 157, 11–35. Menardi Noguera, A., and G. Rea (2000), Deep structure of the Campaniana‐Lucanian Arc (Southern Apennine, Italy), Tectonophysics, 324, 239–265, doi:10.1016/S0040-1951(00)00137-2. Merritts, D., and W. B. Bull (1989), Interpreting Quaternary uplift rates at the Mendocino triple junction, northern California, from uplifted marine terraces, Geology, 17, 1020–1024, doi:10.1130/0091-7613 (1989)0172.3.CO;2. Mitterer, R. M., and N. Kriausakul (1989), Calculation of amino acid racemisation ages based on apparent parabolic kinetics, Quat. Sci. Rev., 8, 353–357, doi:10.1016/0277-3791(89)90035-8. Moncharmont‐Zei, M. (1957), Foraminiferi e molluschi di un livello tirreniano presso Nova Siri Scalo (Matera), Boll. Soc. Nat. Napoli, 66, 53–68. Montone, P., A. Amato, and S. Pondrelli (1999), Active stress map of Italy, J. Geophys. Res., 104(B11), 25,595–25,610, doi:10.1029/1999JB900181. Moretti, I., and L. Royden (1988), Deflection, gravity anomalies and tectonics of doubly subducted continental lithosphere: Adriatic and Ionian Seas, Tectonics, 7(4), 875–893, doi:10.1029/TC007i004p00875. Mucciarelli, M., M. Bianca, M. R. Gallipoli, and R. Caputo (2007), Faglie o terrazzi marini? Contributo di nuovi dati geofisic allo studio della costa ionica della Basilicata, in Riassunti Estesi Delle Comunicazioni 26th Convegno Nazionale Gruppo NGTS, edited by S. Dario, pp. 63–66, Gruppo Naz. Di Geofis. Della Terra Solida, Rome, Italy, 13–15 Nov. Nardo, A. (2007), Monitoring of seismic areas by means of GPS measurements, Ph.D. thesis, pp. 346, Univ. of Padua, Padua, Italy. Noller, S. J., J. M. Sowers, and W. R. Lettis (Eds.) (2000), Quaternary Geochronology. Methods and Applications, AGU Ref. Shelf, vol. 4, pp. 582, AGU, Washington, D. C. Panza, G. F., A. Peccerillo, A. Aoudia, and B. Farina (2007), Geophysical and pretrological modelling of the structure and composition of the crust and upper mantle in complex geodynamic settings: The Tyrrhenian Sea and surroundings, Earth Sci. Rev., 80, 1–46, doi:10.1016/j.earscirev.2006.08.004. Parea, G. C. (1986), I terrazzi marini tardo‐pleistocenici del fronte della catena appenninica in relazione alla geologia dell’avanfossa adriatica, Mem. Soc. Geol. Ital., 35, 913–936. Patacca, E., and P. Scandone (1989), Post‐Tortonian mountain building in the Apennines. The role of the passive sinking of a relict lithospheric slab, in The Lithosphere in Italy: Advances in Earth Science Research, vol. 80, edited by A. Boriani et al., pp. 157–176, Accad. Naz. Lincei, Roma. Patacca, E., and P. Scandone (2001), Late thrust propagation and sedimentary response in the thrust‐ belt‐foredeep system of the southern Apennines (Pliocene‐Pleistocene), in Anatomy of an Orogene: The Apennines and Adjacent Mediterranean Basin, edited by G. B. Vai and P. Martini, pp. 401–440, Kluwer Acad. Publ. Patacca, E., and P. Scandone (2004), The Plio‐Pleistocene thrust belt‐foredeep system in the Southern Apennines and Sicily (Italy), in Geology of Italy, edited by U. Crescenti et al., pp. 93–129, Ital. Geol. Soc., Rome.

23 of 24

TC4005

Patacca, E., R. Sartori, and P. Scandone (1993), Tyrrhenian basin and Apennines. Kinematic evolution and related dinamic constraints, in Recent Evolution and Seismicity of the Mediterranean Region, edited by E. Boschi et al., pp. 161–171, Kluwer Acad. Publ., Boston, Mass. Patella, D., Z. Petrillo, A. Siniscalchi, L. Improta, and B. Di Fiore (2005), Magnetotelluric profiling along the CROP‐04 section in the Southern Apennines, in CROP PROJECT: Deep Seismic Exploration of the CENTRAL Mediterranean and Italy, edited by I. R. Finetti, chap. 13, 263–280 pp., Elsevier, San Diego. Philip, H. (1987), Plio‐Quateranry evolution of the stress field in Mediterranean zones of subduction and collision, Ann. Geophys., 5B(3), 301–319. Pieri, P., L. Sabato, and M. Tropeano (1996), Significato geodinamico dei caratteri deposizionali e strutturali della Fossa bradanica nel Pleistocene, Mem. Soc. Geol. Ital., 51, 501–515. Pieri, P., et al. (1997), Tettonica quaternaria nell’area bradanico‐ionica, Il Quaternario, 10(2), 535–542. Ricchetti, G., and F. Mongelli (1980), Flessione e campo gravimetrico della micropiastra apula, Boll. Soc. Geol. Ital., 99, 431–436. Ricchetti, G., N. Ciaranfi, E. Luperto Sinni, F. Mongelli, and P. Pieri (1988), Geodinamica ed evoluzione sedimentaria e tettonica dell’avampaese apulo, Mem. Soc. Geol. Ital., 41, 57–82. Sabato, L. (1996), Quadro stratigrafico‐deposizionale dei depositi regressivi nell’area di Irsina (Fossa bradanica), Geol. Romana, 32, 219–230. Scholz, C. H. (1982), Scaling laws for large earthquakes:consequences for physical models, Bull. Seismol. Soc. Am., 72(1), 1–14. Scrocca, D., C. Doglioni, F. Innocenti, P. Manetti, A. Mazzotti, L. Bertelli, L. Burbi, and S. D’Offizi (2003), CROP ATLAS. Seismic reflection profiles of the Italian crust, Mem. Descr. Carta Geol. Ital., 62, 194. Scrocca, D., E. Carminati, and C. Doglioni (2005), Deep structure of the southern Apennines, Italy: Thin‐skinned or thick‐skinned?, Tectonics, 24, TC3005, doi:10.1029/2004TC001634. Sella, M., C. Turci, and A. Riva (1988), Sintesi geopetrolifera della Fossa bradanica (avanfossa della catena appenninica meridionale), Mem. Soc. Geol. Ital., 41, 87–107. Senatore, M., W. R. Normark, T. Pescatore, and S. Rossi (1988), Structural framework of the Gulf of Taranto (Ionian Sea), Mem. Soc. Geol. Ital., 41, 533–539. Serpelloni, E., M. Anzidei, P. Baldi, G. Casula, and A. Galvani (2005), Crustal velocity and strain‐rate fileds in Italy and surrounding regions: New results from the analysis of perrmanent and non‐permanent GPS networks, Geophys. J. Int., 161, 861–880, doi:10.1111/j.1365-246X.2005.02618.x. Serpelloni, E., G. Vannucci, S. Pondrelli, A. Argnani, G. Casula, M. Anzidei, P. Baldi, and P. Gasperini (2007), Kinematics of the Western Africa‐Eurasia plate boundary from focal mechanisms and GPS data, Geophys. J. Int., 169, 1180–1200, doi:10.1111/j.1365-246X.2007.03367.x. Shackleton, N. J. (1987), Oxygen isotopes, ice volume and sea level, Quat. Sci. Rev., 6, 183–190, doi:10.1016/0277-3791(87)90003-5. Shackleton, N. J. (2000), The 100,000‐Year ice‐age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity, Science, 289, 1897–1902, doi:10.1126/science.289.5486.1897. Solarino, S., and R. Cassinis (2007), Seismicity of the upper lithosphere and its relationships with the crust in the Italian region, Boll. Geofis. Teor. Appl., 48(2), 99–114. Spakman, W., S. van der Lee, and R. van der Hilst (1993), Travel‐time tomography of the European‐ Mediterranean mantle down to 1400 km, Phys. Earth Planet. Inter., 79, 3–74, doi:10.1016/00319201(93)90142-V. Steckler, M. S., N. Piana Agostinetti, C. K. Wilson, P. Roselli, L. Seeber, A. Amato, and A. Lemer‐ Lam (2008), Crustal structure in the Southern Apen-

TC4005

CAPUTO ET AL.: MARINE TERRACES IN SOUTHERN ITALY

nines from teleseismic receiver functions, Geology, 36(2), 155–158, doi:10.1130/G24065A.1. Stefanelli, S., L. Capotondi, and N. Ciaranfi (2005), Foraminiferal record and environmental changes during the deposition of the Early Middle Pleistocene sapropels in southern Italy, Palaeogeogr. Palaeoclimatol. Palaeoecol., 216, 27–52, doi:10.1016/j. palaeo.2004.10.001. Stirling, M., D. Rhoades, and K. Berryman (2002), Comparison of earthquake scaling relations derived from data of the instrumental and preinstrumental era, Bull. Seismol. Soc. Am., 92(2), 812–830, doi:10.1785/0120000221. Tanner, P. W. G. (1991), The duplex model: Implications from a study of flexural‐slip duplexes, in Thrust Tectonics, edited by K. R. McKlay, pp. 201–208, Chapman and Hall, London. Tortorici, G., M. Bianca, G. De Guidi, C. Monaco, and L. Tortorici (2003), Fault activity and marine terracing in the Capo Vaticano area (southern Calabria) during the Middle‐Late Quaternary, Quat. Int., 101–102, 269–278, doi:10.1016/S1040-6182(02) 00107-6. Trenhaile, A. S. (1987), The Geomorphology of Rock Coasts, 393 pp., Oxford Univ. Press, Oxford, U.K. Tropeano, M., L. Sabato, and P. Pieri (2002), Filling and cannibalization of a foredeep: Bradanic Trough, southern Italy, Geol. Soc. London Spec. Publ., 191, 55–79.

Turco, E., and A. Zuppetta (1998), A kinematic model for the Plio‐Quaternary evolution of the Tyrrhenian‐ Apenninic system: Implications for rifting processes and volcanism, J. Volcanol. Geotherm. Res., 82, 1–18, doi:10.1016/S0377-0273(97)00055-3. Van Dijk, J. P., et al. (2000), A regional structural model for the northern sector of the Calabrian Arc (southern Italy), Tectonophysics, 324, 267–320, doi:10.1016/S0040-1951(00)00139-6. Vezzani, L. (1967), I depositi plio‐pleistocenici del litorale ionico della Lucania, Atti Accad. Gioenia Sci. Nat. Catania, 18, 159–179. Viti, M., D. Albarello, and E. Mantovani (1997), Rheological profiles in the central‐eastern Mediterranean, Ann. Geofis., 40(4), 849–864. Viti, M., E. Mantovani, D. Babbucci, and C. Tamburelli (2006), Quaternary geodynamics and deformation pattern in the Southern Apennines: Implications for seismic activity, Boll. Soc. Geol. Ital., 125, 273–291. Wells, D. L., and J. K. Coppersmith (1994), New empirical relationships among magnitude, rupture length, rupture width, rupture area, and surface displacement, Bull. Seismol. Soc. Am., 84, 974–1002. Westaway, R. (1993), Quaternary uplift of southern Italy, J. Geophys. Res., 98, 21,741–21,772, doi:10.1029/93JB01566. Westaway, R., and D. Bridgland (2007), Late Cenozoic uplift of southern Italy deduced from fluvial and

24 of 24

TC4005

marine sediments: Coupling between surface processes and lower‐crustal flow, Quat. Int., 175, 86– 124, doi:10.1016/j.quaint.2006.11.015. Zander, A.‐M., A. Fülling, H. Brückner, and G. Mastronuzzi (2006), OSL dating of Upper Pleistocene littoral sediments: A contribution to the chronostratigraphy of raised marine terraces bordering the Gulf of Taranto, South Italy, Geogr. Fis. Dinam. Quat., 29, 33–50. Zazo, C. (1999), Interglacial sea levels, Quat. Int, 55, 101–113, doi:10.1016/S1040-6182(98)00031-7. Zazo, C., J. L. Goy, C. J. Dabrio, T. Bardajì, C. Hillaire‐Marcel, B. Ghaleb, J. A. Gonzalez‐ Delgado, and V. Soler (2003), Pleistocene raised marine terraces of the Spanish Mediterranean and Atlantic coasts: Records of coastal uplift, sea‐level highstands and climate changes, Mar. Geol., 194, 103–133, doi:10.1016/S0025-3227(02)00701-6. M. Bianca and R. D’Onofrio, Di.S.G.G., University of Basilicata, Macchia Romana Campus, I‐85100 Potenza, Italy. R. Caputo, Department o f Earth Sciences, University of Ferrara, via Saragat 1, I‐44100 Ferrara, Italy. ([email protected])