Tectonophysics 656 (2015) 131–141

Contents lists available at ScienceDirect

Tectonophysics

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

Rupture process of the 2014 , , earthquake doublet (Mw6) as inferred from regional and local seismic data

E. Sokos a,⁎,A.Kiratzib,F.Gallovič c,J.Zahradníkc, A. Serpetsidaki a, V. Plicka c, J. Janský c, J. Kostelecký d, G.-A. Tselentis a a University of , Department of Geology — Seismological Laboratory, Greece b Aristotle University of Thessaloniki, Department of Geophysics, Greece c Charles University in Prague, Faculty of Mathematics and Physics, Czech Republic d Research Institute of Geodesy, Topography and Cartography, Geodetic Observatory Pecný, Ondřejov, Czech Republic article info abstract

Article history: We study the 26 January and 3 February, 2014 (~Mw6) events in Cephalonia, combining weak and strong motion Received 2 April 2015 waveforms from regional and local stations. The hypocenter of the January 26 event is located at the southern- Received in revised form 30 May 2015 most tip of the Peninsula, at a depth of ~15 km. The centroid moment tensor (CMT) solution indicates rup- Accepted 6 June 2015 ture along a N20°E dextral strike-slip fault, dipping to the east. The hypocenter of the February 3 event is 10 km Available online 26 June 2015 NNE of the first, at shallower depth (~5 km). The CMT solution of this event is highly uncertain. The kinematic slip model for the January 26 event indicates that the rupture was mainly confined to shallow depths, and it propa- Keywords: Cephalonia earthquake gated upwards and towards NE. The major slip patches, when projected to the surface, cover the western part of Centroid moment tensor the Paliki Peninsula and include the areas where surface ruptures were observed. Our preferred slip model for the Waveform inversion event of February 3 is based on a published two-segment fault model. Although this is our preferred slip model, it Slip history is worth noting, that the single segment inversion provided a similar slip pattern. The rupture propagated Fault complexity predominantly southwards along both segments. The main slip episode on both segments occurred almost fi Stress eld simultaneously. Total duration of the rupture propagation did not exceed 9 and 6 s, respectively. The 2014 Cephalonia doublet did not rupture the Cephalonia Transform Fault (CTF). The diffuse pattern of the aftershocks implies the activation of a network of faults on-shore the Paliki Peninsula, in accordance with the local stress field derived from aftershocks. The 2014 sequence has implications for the seismic hazard assessment: active faults in western Cephalonia exist on-shore; some have gentle dip angles; the strike-slip motions can be combined with thrust components; and the segmented ruptures may introduce time delays that increase the duration of strong ground shaking. © 2015 Elsevier B.V. All rights reserved.

1. Introduction reported and the constructions performed remarkably well (GEER/ EERI/ATC report, 2014). The 2014 sequence attracted the attention of Cephalonia (Greece) belongs to the and is best known the scientific community and a number of publications are already for its beautiful landscape and the strong and frequent earthquakes. The available (Benekos et al., 2015; Boncori et al., 2015; Karakostas et al., earthquake activity in 1953 benchmarks the history of Cephalonia, as 2014; Karastathis et al., 2015; Papadopoulos et al., 2014; Sakkas and the island was destroyed and more than 450 people lost their lives. Lagios, 2015; Valkaniotis et al., 2014). Despite the fact that the sequence This event was the startup point for the Hellenic Antiseismic Code was well recorded by the regional networks in Greece, there are still whose provisions for the constructions in the Ionian Islands are the unresolved issues. Accurate location of earthquakes in Cephalonia is a strictest over Greece (reference ground acceleration for ground type A challenge by itself, due to the absence of stations from the west and (rock) equal to 0.36 g). the sparse seismic coverage from the south. For example, in the west The focus of this work is the sequence that burst in Cephalonia the closest seismic station is in . As a result, in the related publica- Island on January 26, 2014 with an Mw6 earthquake and culminated tions (Boncori et al., 2015; Karakostas et al., 2014; Papadopoulos et al., on February 3 to another Mw6 event (Fig. 1). The reported epicenters 2014), the location, the fault orientation and its dip polarity of the two of the two strong events and all aftershocks are spread along the strong events show considerable variability, reflecting the difficulties western coast (Paliki Peninsula) of Cephalonia. No loss of life was in the data analysis. Another issue that further intrigued the scientists is the unclear, if ⁎ Corresponding author. Tel.: +30 2610 969369; fax: +30 2610 990639. any, connection of the sequence to the well-known dextral Cephalonia E-mail address: [email protected] (E. Sokos). Transform Fault (CTF) that dominates along the western coast of the

http://dx.doi.org/10.1016/j.tecto.2015.06.013 0040-1951/© 2015 Elsevier B.V. All rights reserved. 132 E. Sokos et al. / Tectonophysics 656 (2015) 131–141

Fig. 1. a) Map of Cephalonia Island showing focal mechanisms of post-1966 earthquakes with Mw N 6 (beach-balls) and mapped faults on-shore (Lekkas et al., 2001); the Cephalonia Transform Fault (CTF) zone is highlighted, and black contours denote the sea bathymetry. b) Relocated epicenters of the two major events: the black star denotes the epicenter of the 1st event on Jan 26, 2014 and the black diamond denotes the epicenter of the 2nd event on Feb 3, 2014. For comparison we have included the relocated epicenters for the 1st (gray star) and 2nd (gray diamond) events reported in Karastathis et al. (2015). The CMT solutions for the two events, as calculated here and as reported in GCMT catalog are also shown (beach-balls marked accordingly). The relocated aftershocks (yellow circles) are scaled proportionally to their magnitude.

Ionian Islands (Louvari et al., 1999; Scordilis et al., 1985) and is rather fault surface ruptures were most probably not produced. The abundant described as a ramp in the bathymetry (Shaw and Jackson, 2010). Two surface cracks with cm-size offsets in the northern part of the Paliki branches were identified along CTF — the Cephalonia segment in the Peninsula, ~38.29°N, were interpreted as due to close proximity of the south, where the typical focal mechanisms have parameters: strike ruptured fault to the earth surface (with unclear relation to either of 38°, dip 63° and rake 172°, and the segment in the north the events). Boncori et al. (2015) inferred static ground displacements with: strike 14°, dip 65° and rake 167° (Louvari et al., 1999 and refer- related to the event of 3 February from InSAR images. Their results ences therein). The strike-slip motions, often combined with a thrust were better modeled by a two-segment fault for this event. component, are not confined along the CTF only. On the contrary, a Here we use weak and strong motion waveforms from local and broad zone, ~100 km wide, up to the western is character- regional stations to constrain kinematic rupture models of the two ized by strike-slip motions (Kiratzi, 2014; Louvari et al., 1999; Shaw and major events. For brevity, hereafter we refer to the January 26 and Jackson, 2010). The available fault databases include a few strike-slip February 3 earthquakes as the 1st and 2nd event, respectively. In partic- fault segments offshore Cephalonia (Caputo et al., 2012), and a network ular, we pay careful attention to the determination and consistency of of mapped faults onshore (Lekkas et al., 2001). their hypocenters, centroid moment tensors (location and faulting pa- Below we briefly review the knowledge that has been accumulated rameters), and fault plane geometries. We discuss the two events not so far. Prior to the 2014 sequence, a change in the long-term deforma- only in terms of their source properties, but also in terms of the tion of the Cephalonia Island was geodetically detected. It started in difficulties encountered during their source inversions. In this context, ~2003 and until 2010 the western peninsula of the Cephalonia Island we discuss why the moment tensor solutions, as reported by different (Paliki Peninsula, Fig. 1a) was uplifting at a rate of 1 cm/yr in an abrupt agencies, are quite similar for the 1st event, whereas those reported contrast with the subsidence of the rest of the Cephalonia Island (Lagios for the 2nd event vary significantly and include large non-double- et al., 2012). Karakostas et al. (2014) relocated the sequence and con- couple components. cluded that the two major shocks were related to two adjacent fault seg- The paper is structured as follows: the data and methods used are ments, striking almost N–S and dipping to the east. Karastathis et al. briefly described; observed data of 1st and 2nd event are investigated, (2015) relocated the sequence using the equal differential time and each following a hierarchic scheme (starting from the hypocenter loca- probabilistic non-linear approaches, accounting for the effects of the lat- tion and calculation of the centroid moment tensor, continuing with erally varying crustal structure. They showed that the January 26 and specification of fault plane, ending with slip inversion). Finally, the February 3 events could be related with fault planes dipping to east two events are compared to each other, and to aftershocks, and they and west, respectively. Papadopoulos et al. (2014) made a preliminary are collectively discussed in terms of the local stress field and seismic comprehensive analysis of the 2014 sequence. Their results support hazard. predominantly downward and upward rupture propagation for the January 26 and February 3 events, respectively. They concluded that 2. Data and methods the 2014 sequence ruptured a fault segment which is the SSW-wards continuation of the Lefkada segment as this was defined in Louvari Broad-band waveforms were retrieved from the Hellenic Unified et al. (1999). Seismic Network (HUSN). We adopted the manual P and S phase picks Valkaniotis et al. (2014) analyzed geological effects, such as liquefac- from the Geodynamics Institute of the National Observatory of Athens tion, rock falls, and landslides, concluding that primary (co-seismic) and added manual picks from available local strong motion stations E. Sokos et al. / Tectonophysics 656 (2015) 131–141 133 for aftershock relocation. Stations in Italy were also picked manually for resolution results from synthetic tests (Gallovič and Zahradník, 2011; the two major shocks only. The strong motion records from the local Gallovič et al., 2015). All slip models were obtained using displacement stations were provided to us by the Institute of Engineering Seismology waveforms from local strong motion stations, permanently or tempo- and Earthquake Engineering (ITSAK) and by the Geodynamics Institute rarily deployed on Cephalonia Island. For the strong motion stations of National Observatory of Athens (GI-NOA). DMLN strong motion sta- CHV1, ARG2, and VSK1 accurate timing was not available, thus we tion belongs to the seismic network of Patras University. The locations corrected the timing using the relocated hypocenter of January 26 of the stations used are shown in Fig. S1. event and synthetic travel times calculated by ray-tracing.

2.1. Hypocenter locations 2.4. Apparent source time functions

To locate the hypocenters of the two major events we primarily used We were able to find a suitable aftershock for the 26 January the NonLinLoc code (Lomax et al., 2000). This approach is very suitable (1st event) only, in order to calculate apparent source-time func- because it permits an uncertainty analysis considering the non-linearity tions (ASTFs) using the method of Empirical Green's Functions of the problem. Hence, the error volume, as it is expressed by the (e.g. Courboulex et al., 1997; Roumelioti et al., 2009). The deconvolution probability density function, has a more general shape compared to process was stabilized following the approaches of Bertero et al. (1997) the ellipsoids provided by the standard (linearized) approaches. Parallel and Vallée (2004). to that, many trial locations were performed by four other codes: a) FASTHYPO (Herrmann, 1979), b) HYPOINVERSE (Klein, 2002), 2.5. Stress inversion c) HYPO71PC (Lee and Valdes, 1985) and d) HYPODD (Waldhauser, 2001). We used aftershock mechanisms to examine the local stress field To locate the aftershocks we used HYPOINVERSE and HYPODD. The using the STRESSINVERSE code of Vavryčuk (2014).Thisisamodifica- most suitable velocity model for the geometry of the problem was tion of the method of Michael (1987), based on iterative joint inversion that of Haslinger et al. (1999). Other models were also tested, but the for direction of principal stress axes and fault orientations. The method location uncertainties due to selection of station subsets were always resolves ambiguity of nodal planes for each studied event by applying greater than those due to the choice of the velocity model. the fault instability constraint, which leads to stabilization of the estimated shape ratio. The uncertainty estimates of the stress axes and 2.2. Centroid moment tensor (CMT) solutions shape ratio are provided by randomly perturbing the input mecha- nisms; here we use 200 variations, assuming ±20° uncertainty in the CMT solutions were inferred using the ISOLA code (Sokos and orientation of the P–T axes. Using the Mohr's diagram the code allows Zahradník, 2008, 2013) as further extended elsewhere (Fojtíková us to verify whether the stress inversion is reasonable by means of and Zahradník, 2014; Zahradník and Sokos, 2014). This approach is checking whether the faults identified by the inversion are concentrated based on full waveform least-squares inversion, assuming a point- or in the area of validity of the Mohr–Coulomb failure criterion. multiple-point source representation. Green's functions were calculated by the discrete wavenumber and matrix methods (Bouchon, 1981; 3. Results Coutant, 1989; Kennett and Kerry, 1979) for the 1D velocity model adopted for location (Haslinger et al., 1999). Here and subsequently In the following sections we discuss the source properties of each for the slip inversions, we evaluate the seismogram fit in terms of the event. Before doing so, in Fig. 1a we show surface fault traces on land variance reduction percentage (VR) and/or correlation (corr), which Cephalonia, while in Fig. 1b we show our relative relocation of the are related simply by VR/100 = corr2. The centroid position and sequence (630 events of M N 3). The aftershock distribution is diffused centroid time are calculated by a grid search. A 3D spatial grid search along the Paliki Peninsula, not allowing for any initial constraints on is made for the two main events in order to infer not only the best- the fault planes. fitting centroid location and faulting parameters, but also to assess the uncertainty of the solution. The centroids of the aftershocks are 3.1. The 26 January 2014 event grid-searched in vertical direction below their epicenters. We exhaustively tested the variability in location by jack-knifing the 2.3. Finite fault slip inversions stations, in terms of their proximity to the sequence i.e. local and distant. The epicenter location varied significantly, and this variability is well To obtain the model of rupture propagation along the fault plane we illustrated by the final NonLinLoc results, displayed in Fig. 2(a, b) and adopted the linear slip inversion method of Gallovič et al. (2015).Ithas summarized in Table 1. Fig. 2a shows the location using only the stations been extended, for the purpose of the present work, to perform the deployed on Cephalonia Island. Fig. 2b shows the location using distant inversion simultaneously on multiple fault segments. In this approach, stations in Greece and Italy (shown in Fig. 2d), in an attempt to reduce the rupture process is discretized in space and time along a given fault the azimuthal gap inherent to the near stations. The first striking obser- plane, and the model parameters are the spatial–temporal samples of vation is the significantly different epicenters for this event. the slip rate, spanning the whole rupture duration. Green's functions Our preferred solution is the one obtained by the distant stations are calculated as previously. To stabilize the inverse problem in the and not the local. This choice is based on the fact that this epicenter: least-square sense, we apply: (i) spatial smoothing by means of a a) is supported by the first-motion polarities at near station DMLN prior covariance function with k-2 decay at large wavenumbers k and (Z+, N+, E+), which imply that the epicenter lies south-west of the (ii) a positivity constraint on the slip rate by means of the non- station and b) is in agreement with the provisions imposed by the collo- negative-least squares (NNLS) approach (Lawson and Hanson, 1974). cation of hypocenter and fault plane position (Zahradník et al., 2008). As such, the source description is very general with no a-priori Our preferred solution is in disagreement with the relocations proposed constraints, as for example on the position of the nucleation point, the by other authors (Karakostas et al., 2014; Karastathis et al., 2015), which rupture velocity, and the shape of the slip-rate functions. Nevertheless, are closer to our epicenter obtained from the local data. as in any multi-parameter inversion, the slip model is vulnerable to The CMT solution was calculated by the full waveform inversion of artifacts and biases due to the imperfect station distribution and relatively low-frequency regional data, assuming a single point-source smoothing, requiring careful appraisal. In view of the above, we inter- model. Several frequency ranges and subsets of the stations were tested pret the inversion results with caution taking into account the previous and the obtained source parameters varied as: strike (15°–30°), dip 134 E. Sokos et al. / Tectonophysics 656 (2015) 131–141

Fig. 2. Location of the two major shocks of the 2014 sequence by the NonLinLoc method (see also Table 1). The best-fitting hypocenter position is shown by a blue star, while scattered red dots demonstrate the probability density function. A map-view is supplemented by two vertical cross-sections, W–E (bottom panel) and N–S (right panel). a) 1st event location using only local stations (green triangles), b) our preferred 1st event location using distant stations (green triangles in panel d), c) 2nd event location using local stations (green triangles), d) stations used in panel b.

(70°–85°), rake (140°–180°), scalar moment (1.2–1.4 × 1018 Nm), and very good (variance reduction 86%, Fig. S2a). However, the double- centroid depth (8–12 km). couple (DC) percentage is relatively low, 61%. Nevertheless, some near- The horizontal position of the centroid is also uncertain. A typical by trial source positions (e.g., 7 km north-east of the centroid) provide example of the spatial variability of the MT solution, for the depth of similar strike/dip/rake angles and considerably higher DC ~90%, with 10 km and for the frequency range 0.05–0.08 Hz, is shown in Fig. 3a. just a slightly lower waveform correlation. In particular, the centroid position is less certain in the SW–NE direction In order to perform the slip inversion, we used a single fault due to the distribution of the available stations. The best-fitting solution segment, a valid approach based on the large DC percentage near the is found at the center of the trial source grid with 5 × 5 km increments. centroid position. We chose the nodal plane striking at ~20° and dipping This solution was adopted as our preferred CMT solution (Table 2)and ~75° to the east as the fault plane, in accordance also with previous favors an eastward dipping nodal plane with strike/dip/rake equal to knowledge on the orientation of the main tectonic structures in the 20°/74°/162°. The global waveform match at this centroid position is area (Louvari et al., 1999) and the directivity effects observed on the ASTFs as discussed later. However, the exact position of the fault in space is neither constrained by the (diffuse) aftershocks, nor by surface Table 1 ruptures or other data. Therefore, we try to further constrain its position Hypocenter locations of the two events studied. and the focal mechanism using finite-fault full waveform inversion of 1st event: January 26, 2014 local seismic data. Arrivals used Origin time Latitude (°N) Longitude (°E) Depth (km) To this goal we increase the maximum frequency to 0.2 Hz Local stationsa 13:55:42.91 38.2142 20.4804 18.0 (e.g. frequency range 0.05–0.2 Hz). We assume a planar fault of Distant 13:55:43.27 38.1522 20.3912 15.0 30km×30kmsize(alongstrikeanddip)andperformagridsearch b stations varying the position and orientation of the fault plane. In particular, 2nd Event: February 3, 2014 the so-called fault origin (defined as a point situated 10 km along Arrivals used Origin time Latitude (°N) Longitude (°E) Depth (km) strike and 8 km down-dip from the top-left corner of the fault) is Local stations 03:08:45.23 38.2712 20.4295 5.5 varied around the centroid within the range: 38.1945°–38.2445°N

a – fi Solution not supported by all data. and 20.3519° 20.4119°E. The top of the fault is xed at 0.5 km b Preferred solution. depth. We also vary the pure-shear mechanism in the vicinity of E. Sokos et al. / Tectonophysics 656 (2015) 131–141 135

Fig. 3. Spatial variability of the MT solution on a horizontal grid with step size of 5 km for a) the 1st event and b) the 2nd event. The background color and isolines show the correlation between observed and synthetic waveforms, according to the scale. Beach balls at the trial source positions are color-coded according to their DC percentage. The largest beach ball is for the best-fitting solution (Table 2, Fig. 1b). the preferred CMT parameters: strike 15–25°, dip 70–80° and rake prescribed in the slip inversion). The spatial distribution of the slip 150–180°. For each trial fault location and mechanism (729 in rate denotes that the rupture initially propagated up-dip from the total) we perform the NNLS linear slip inversion with a fixed hypocenter. It reached shallow depths in ~4 s (suggesting a relative- smoothing weight. ly large rupture velocity of ~4 km/s) and continued to propagate The optimal solution was found for the fault origin 38.2195°N, along-strike (NE-wards) for another ~5 s at a rather slower velocity 20.3819°E, depth 8 km, and for mechanism with strike 20°, dip 80°, of ~2 to 3 km/s, perhaps reflecting the decrease of shear modulus and rake 180°. The fault plane agrees with our preferred epicenter with decreasing depth. It is worth noting that these absolute values position as calculated from the distant stations (Fig. 2b). Indeed, when might be obscured by the smoothing constraint imposed during the we allowed the hypocenter depth to increase by less than 10 km, inversion, as well as affected by the imperfect station coverage (see reflecting the location uncertainty, the orthogonal distance of the hypo- also Gallovič et al., 2015). center to the fault plane became smaller than 1 km. This test also Fig. S4a shows the comparison between the synthetic and ob- showed that the epicenter defined by the use of the local stations served displacement waveforms for the strong motion sites. The syn- (Fig. 2a), close to that of Karastathis et al. (2015), is much less favored. thetics reproduce the observed records relatively well (VR 83%), This stems from the fact that even a large increase in the depth still with the exception of the fittotheE–WandN–S components of leaves the hypocenter distance from the fault too large (~5 km) and, stations DMLN and SMHA, respectively. In summary, taking into moreover, such a hypocenter would project closer to the conjugate account that the event might have contained more complex fault WNW–ESE trending nodal plane. geometry and variable focal mechanisms, our slip model reproduces Our calculated slip model, for the 1st event, is presented in Fig. 4 satisfactorily the observed strong motion displacement records. (and Fig. S3a). The slip patch covers almost the entire length of the Attempts to account for the mentioned complexities would go fault (~25 km), being approximately 10 km wide. Slip distributed beyond the resolution of the frequency range used and the crustal over such a relatively large area suggests rather small mean static model available. It is also important to note that in our model the stress drop of the order of 1 MPa. The locus of maximum slip rupture propagated nearly to the surface, which might be in accor- (~25cm)isdepicted~10kmawayfromtheepicenterand~6km dance with the surface ruptures, as marked in Fig. 4a, reported by from the centroid position, towards NNE. The DMLN station is locat- Valkaniotis et al. (2014). ed close to this largest slip, and it should be noted that this close As previously mentioned, we were able to identify an event which station might have controlled the location of the maximum slip could serve as an Empirical Green's Function, in order to obtain appar- (see for example the synthetic tests in Gallovič et al., 2015). Howev- ent source time functions (ASTFs) at regional stations. Thus, the wave- er, still the maximum slip is well within the area of the largest corre- forms of an Mw4.3 aftershock (2014/01/26 at 19:12 UTC, no. 4 in lations in the CMT solutions (Fig. 3) and in these solutions the data Table S1) were deconvolved from the mainshock waveforms. The from DMLN station were not used. inferred ASTFs (Fig. S5) have two distinct lobes. The duration of the Fig. 4b shows the snapshots of the slip rate on the fault plotted first pulse is ~6 s and of the second is ~4 s. The second lobe is more pro- every 1 s. The largest values of the slip rate are attained ~15 km nounced at the stations east and south of the hypocenter. The narrower up-dip the hypocenter (recall that the hypocenter position was not pulse at one station (e.g. IGT) which lies directly north of the epicenter

Table 2 Centroid moment tensor solutions of the two events studied.

Centroid time Lat. (°N) Lon. (°E) Depth (km) Mo (Nm) Mw Strike (°) Dip (°) Rake (°) DC (%) VR (%)

1st event: January 26, 2014 13:55:47.06 38.1944 20.3519 10.5 1.4E + 18 6.1 20° 74° 163° 61 85

2nd event: February 3, 2014 03:08:48.67 38.2845 20.4089 2.5 1.0E + 18 6.0 199° 49° 167° 63 82 136 E. Sokos et al. / Tectonophysics 656 (2015) 131–141

Fig. 4. Slip model for the January 26, 2014 event obtained using strong motion displacement waveforms from the stations depicted on the map. a) Surface projection of the slip distribution along the fault, according to the scale shown below. The epicenter (star) and the centroid location (black circle) together with the CMT solution are also depicted. The relocated aftershocks prior to the occurrence of the 2nd event are also plotted, showing that they are mainly confined off the major slip patches. Blue symbols denote the surface ruptures of Valkaniotis et al. (2014). The fault-top depth is 0.5 km. Coordinates of the fault corners as in Table S2. b) Snapshots every 1 s of the space–time evolution of the slip velocity on the fault. The blue asterisk denotes the vertical projection of the hypocenter on the fault. Note the initially up-dip rupture propagation, which continued to propagate along strike and towards NE.

indicates predominantly along-strike propagation of the rupture. The also systematically shifted northward with respect to the 1st second pulse, which is of shorter duration and amplitude, and well event, analogously to the epicenter. However, the best-fitting lagged in time from the first pulse, probably reflects the shallower centroid position and focal mechanism varied with station sub- slip patch of the slip model (Figs. 4 and S3). A secondary subevent sim- sets more dramatically than for the 1st event. In particular, solu- ilar to the northern slip patch of Fig. 4 has been also indicated by tions with both east- and west-ward dipping nodal plane were multiple-point source models; these tests are not reported here because obtained. Many solutions were characterized by strike ~200°, ac- the secondary subevent had an unstable position and focal mechanism. companied with dip and rake angles ranging between 60°–85° The fact that the northern asperity is not clearly separated from the slip and 130°–180°, respectively. closer to the hypocenter can be attributed to the smoothing constraint Fig. 3b shows the correlation map at 2.5 depth, as an indicative ex- applied in the slip inversion. ample. Similarly to the pattern of the 1st event (Fig. 3a) the correlation isolines are elongated in the SW–NE direction due to the station 3.2. The 3 February 2014 event azimuthal coverage. The best-fitting CMT solution (Table 2)hasawest dipping nodal plane (strike/dip/rake equal to 199°/49°/167°) and After the occurrence of the 1st event temporary stations were DC = 63%. The waveform fit (variance reduction 82%) is shown in deployed in Cephalonia Island by various agencies (e.g. Geodynamics Fig. S2b. The spatial variability of the source mechanism (beach balls Institute of the National Observatory of Athens, GINOA, Institute of in Fig. 3b) is more pronounced compared to that of the 1st event Engineering Seismology and Earthquake Engineering, ITSAK, University (Fig. 3b). We emphasize that although the data are fitted almost equally of Patras Seismology Laboratory, UPSL) which greatly facilitated the well as in the case of the 1st event, the absence of the high-DC solutions location procedure. After jackknifing the stations, to test the location near the best-fitting trial source points and the spatial variability of the stability, we obtained similar epicenter positions for most of the station moment tensor do not allow a straightforward interpretation. One pos- subsets tested. Thus, the NonLinLoc solution for this event, shown in sibility how to explain the MT variability is simply due to the shallow Fig. 2candTable 1, can be considered as a typical result. For this event depth. As shown in Fig. 5, based on the analysis of theoretical MT we are in agreement, within error limits, with the relocation by resolvability (Zahradník and Custodio, 2012), from two hypothetical Karastathis et al. (2015). This second event was considerably shallower sources inverted in the station network, velocity model and (~5 km) than the 1st event, and its epicenter is shifted ~10 km towards low-frequency range, the MT solution of the shallow source is much NNE. more uncertain; in particular, the east and west dipping nodal To find the point-source CMT approximation of the 2nd event, we planes are hard to distinguish, and the DC percentage of the shallow used the same frequency range (i.e. 0.05–0.08 Hz) and initial tests as source is poorly determined. The same shallow-source effect most for the 1st event. The 3D grid search systematically indicated a relatively likely caused the large variation of the MT agency reports of this shallow centroid depth (~2.5–5.0 km). The centroid position was event. Nevertheless, when interpreting the CMT variability, we E. Sokos et al. / Tectonophysics 656 (2015) 131–141 137

Fig. 5. Theoretical resolvability analysis of the moment tensor for a fixed configuration of the stations and two centroid depths, 8.5 km (left panels) and 2.5 km (right panels). Panels a) and b) show the uncertainty of the nodal lines. Panels c) and d) are the uncertainties of the DC percentage. should count also with a possible source complexity, discussed in with less slip near the surface, compared to segment 1 (Fig. S3b). The the following. two segments experienced the main slip episode almost simultaneously Definition of a fault for the slip inversion of this event was a chal- (see snapshots at 4–6sinFig. 5b–c). In view of the above, the 2nd event lenge taking into account the spatial variability of our CMT solutions. should be regarded as an earthquake that ruptured almost simulta- Initial slip inversions considering various fault planes and mechanisms neously the two fault segments, with a joint hypocenter, both involving within the MT uncertainty limits did not provide robust space–time minor thrust components. However, it should be noted that we have slip distributions. Thus, we adopted a fault geometry constrained by tested also slip inversion considering just segment 1. The rupture prop- DInSAR data (Boncori et al., 2015). The authors showed in their analysis agation is very similar to that obtained above (along segment 1), while (in their Fig. 3c) that the uplift and subsidence regions, observed at the fit deteriorates only slightly (Fig. S4c, variance reduction 70%). the northern part of the Paliki Peninsula, are separated by a sharp Therefore, we cannot confirm the existence of segment 2 based on the north–south trending boundary. Along this line, the orientation of the available seismic data. horizontal motions changes abruptly, too. Thus, for our finite fault slip The seismically derived slip distributions of the two-segment and inversion, we adopted their two-segment fault model: segments 1 and single-segment fault models were employed in forward modeling to 2 with strike/dip/rake equal to 180°/86°/147° and 33°/76°/164°, respec- simulate the horizontal and vertical static displacements using the tively (Boncori et al., 2015). This is also justified by the fact, that this closed-form solution of Okada (1992) for homogeneous half space. fault model is consistent with our calculated hypocenter (Table 2)asit Both slip models are able to reproduce very well (see Fig. S6) the is situated close to the junction of the two segments (the orthogonal dis- main features of the DInSAR data (Boncori et al., 2015), as previously tance of the hypocenter from segments 1 and 2 is only 600 and 1400 m, mentioned. respectively). Both segments are simultaneously included in the slip in- Unfortunately, even though the sequence was rich in aftershocks, we version. The data are the displacement waveforms at five strong motion were not able to find a suitable event as Empirical Green's Function to stations (Fig. 6a), and both data and synthetics were filtered between retrieve apparent source time functions (ASTFs) for the 2nd event. 0.05 and 0.2 Hz. There are some interesting features recovered in the space–time 3.3. Local stress field slip-rate history for the two segments (Figs. 6b and S4b for waveform fit with VR = 87%). Segment 1 ruptured mainly at shallow depths, We inverted the local stress field using the focal mechanisms of and the rupture front propagated predominantly southwards, towards the aftershocks (Table S1) and the results are summarized in Table S3, the town of (shown in Fig. 1). After 4 s the rupture activated Figs. 7 and S7. We have found that the best resolved parameter is the deeper part of the fault. In this nearly vertical segment, the northward maximum horizontal compression axis σ1 with azimuth 248° and rupture propagation is less pronounced, as indicated by the relatively plunge 7°. Almost identical orientation of the strain–tensor compression small slip rate amplitudes north of the hypocenter (Fig. S3b). The slip axis was reported by Ganas et al. (2013) for a broader central patch covers approximately 15 km × 10 km large area, corresponding region, based on continuous GPS data. The other two principal stress to static stress drop of ~1 MPa, i.e. the same order as for Event 1. The axes σ2 and σ3 are less stable when randomly perturbing the input less steep eastward dipping segment 2 ruptured rather unilaterally, focal mechanism within their expected uncertainties and/or when 138 E. Sokos et al. / Tectonophysics 656 (2015) 131–141

Fig. 6. Slip model for the February 3, 2014 event, obtained using displacement waveforms from the strong motion stations depicted in the map, and the two-segment fault geometry from Boncori et al. (2015). a) Surface projection of the slip of the two fault segments. The star denotes the epicenter (Table 1, Fig. 2b), the black dot and the beach-ball correspond to the centroid (Table 2, Fig. 3b). Relocated aftershocks after the occurrence of the 2nd event up to March 28 are shown. The fault-top depth is 0.5 km. Coordinates of the fault corners as in Table S2. Blue symbols denote the surface ruptures of Valkaniotis et al. (2014). b) Snapshots every 1 s of the space–time evolution of the slip velocity on segment 1 (left panels) and segment 2 (right panels). The blue asterisk in all panels denotes the horizontal projection of the hypocenter perpendicularly to the strike of the respective fault segment. Note that in segment 1 the rupture initiated at shallow depth, propagated mainly southwards, towards the town of Lixouri, activating a deeper asperity at 4 s.

constraining the friction coefficient at various values. In particular, earthquake ruptured an on-shore strike-slip fault, steeply dipping to according to the scatter plot in Fig. S7c the uncertainty pattern of axis the east, roughly parallel to the major structure that dominates off-

σ2 indicates that this axis is likely not vertical. The stress regime is not shore the Cephalonia Island, the so called Cephalonia Transform Fault strictly strike-slip, which agrees with the observed co-existence of the (CTF). From the two segments of CTF, the January 26 earthquake had strike-slip and reverse faults. roughly the orientation and the dip polarity of the Lefkada segment Interestingly, the shape ratio remains stable (0.8–0.9) when ran- (Louvari et al., 1999). domly perturbing the input mechanisms. The events are well distributed We were able to retrieve the slip model for the January 26 event in the Mohr's diagram, covering only the left part of the upper and lower using data from the local strong motion stations. Although ruptures nor- semicircles, which is the area of validity of the Mohr–Coulomb failure mally propagate in all directions (Gudmundsson, 2011), here, initially, criterion. This indicates that the stress inversion results are reasonable the up-dip rupture propagation from the hypocenter prevailed, (p. 75 of Vavryčuk, 2014). reaching shallow depths in ~4 s and continuing to propagate predomi- nantly along-strike towards the north for another ~5 s. This along-strike 4. Conclusions rupture propagation scenario is in agreement with the shape of the apparent source time functions obtained from the Empirical Green's We studied the January–February 2014 Cephalonia Island earth- Function method. We cannot exclude the possibility that the rupture quake sequence, confined along the Paliki Peninsula. Two major events process along the entire fault had a more complex fault geometry of M ~ 6 occurred on January 26 and February 3. We have exhaustively and/or variable focal mechanism because the resolution capacity of exploited the available local, regional and distant data in order to: locate our data does not permit the retrieval of such details. Our kinematic the two events, calculate their centroid moment tensor solutions, deter- slip model for the 1st event differs from the preliminary model by mine their kinematic rupture models, relocate 569 aftershocks with Papadopoulos et al. (2014) due to the considerable difference in the M N 3 and calculate focal mechanisms for 31 of them. The main results assumed fault planes and since we use strong motion waveforms at of our work summarize as follows. local distances and different frequencies.

4.1. The 26 January 2014 earthquake 4.2. The 3 February 2014 earthquake

The hypocenter location of the January 26 event was very difficult. The epicenter location of the second major shock (February 3) was The local data (stations onshore Cephalonia) had a considerable gap in better constrained thanks to the temporary stations installed during azimuth coverage. Our preferred hypocenter (38.1522°N, 20.3912°E; the sequence. Its hypocenter (38.2712°N, 20.4295°E) is 10 km to the h ~ 15 km) was calculated using stations in Italy and regional stations NNE of the 1st earthquake and considerably shallower (~5 km). The in Greece. Our solution is consistent with the 3-component first- centroid position and MT of the 2nd event featured large uncertainties, motion polarities at the near DMLN strong motion station. This epicen- accompanied by low double-couple (DC) percentage, similar to the ter is located at the southernmost tip of the Paliki Peninsula. The variability among agency reports. E. Sokos et al. / Tectonophysics 656 (2015) 131–141 139

Fig. 7. Moment tensor solutions of 31 aftershocks with M N 4 (listed in Table S1). Black and red beach-balls denote the strike-slip and reverse faulting, respectively. Contours denote the sea bathymetry. The inset shows the principal axes of the stress field and their uncertainty derived from the aftershocks (see also Table S3 and Fig. S7).

As such, the MT solution did not provide enough clues for robust Our seismically derived slip model was able, using forward model- slip models with reasonable fit of the strong motion data. Therefore, ing, to simulate the surface static displacements and dominant features we adopted the two-segment fault geometry proposed by Boncori of the DInSAR data (the sharp north–south trending discontinuity et al. (2015) based on modeling the synthetic aperture radar (SAR) ac- between the uplift and subsidence, as well as a dramatic change in the quisitions. The two-segment fault model provided a kinematic slip horizontal-motion direction). In this sense, our work extends the model with satisfactory fit to the strong motion records. The rupture study by Boncori et al. (2015), because we not only confirmed their of the N–S trending and nearly vertical segment 1 (strike/dip/rake fault interpretation by seismic data, but also provided a kinematic 180°/86°/147°) was confined at shallow depths. The rupture propagat- space–time slip model consistent with the seismic and space-geodesy ed basically along strike with a pronounced directivity southwards, to- data. wards the town of Lixouri, and with less pronounced northward propagation. The second, less steep, eastward dipping segment 2 4.3. Connection to regional tectonics (strike/dip/rake 33°/76°/164°) ruptured unilaterally, with less slip near the surface compared to the N–S segment. The two segments expe- One question that arose after the occurrence of the 2014 Cephalonia rienced the main slip episode almost simultaneously (~4–6 s after initi- sequence was its connection to the regional tectonics, and more specif- ation). Moreover, even though our inversion technique does not require ically, to the Cephalonia Transform Fault that dominates the bathymetry any prescribed hypocenter position, the results indicate that the two along the western coasts of the Ionian Islands. We start by the fact that segments could have started their rupturing from a common hypocen- the 2014 sequence occurred on-land and that the aftershocks do not ter, situated at the junction of the two segments. It is worth mentioning delineate a single fault structure (Fig. 1b). Instead, the rather diffuse pat- that we have tested also slip inversion considering only segment 1, tern is more consistent with the activation of a network of minor faults which provides by itself a reasonable model with just minor deteriora- that have been mapped on land (Lekkas et al., 2001). The focal mecha- tion of the waveform fit. nisms of the two major events clearly show dextral strike-slip motions 140 E. Sokos et al. / Tectonophysics 656 (2015) 131–141 combined with a thrust component along faults that are approximately Bouchon, M., 1981. A simple method to calculate Green's functions for elastic layered media. Bull. Seismol. Soc. Am. 71-4, 959–971. parallel to the strike of the Cephalonia Transform Fault, and more specif- Caputo, R., Chatzipetros, A., Pavlides, S., Sboras, S., 2012. The Greek database of seismogenic ically to the strike of the Lefkada segment (Louvari et al., 1999). This se- sources (GreDaSS): state-of-the-art for northern Greece. Ann. Geophys.-Italy 55-5, quence did not rupture CTF itself, but fault segments that are more or 859–894. http://dx.doi.org/10.4401/ag-5168. Courboulex, F., Santoyo, M.A., Pacheco, J.F., Singh, S.K., 1997. The 14 September 1995 less parallel to it, on-shore Cephalonia. The shear motions in Cephalonia (M = 7.3) Copala, Mexico, earthquake: a source study using teleseismic, regional, and the other Ionian Islands are not confined along the CTF only, but on and local data. Bull. Seismol. Soc. Am. 87-4, 999–1010. the contrary there is a wide zone (~100 km) which extends in the Coutant, O., 1989. Program of Numerical Simulation AXITRA. Research Reports LGIT, east up to western Peloponnese (Kiratzi, 2014 and references therein). University of Joseph Fourier, Grenoble, France. Fojtíková, L., Zahradník, J., 2014. A new strategy for weak events in sparse networks: the We estimated the local stress field and the best resolved stress axis first-motion polarity solutions constrained by single-station waveform inversion. is the maximum compression (azimuth 248°, plunge 7°), while the Seismol. Res. Lett. 85-6, 1265–1274. http://dx.doi.org/10.1785/0220140072. č intermediate axis (less well resolved) is not strictly vertical, hence Gallovi , F., Zahradník, J., 2011. Toward understanding slip inversion uncertainty and artifacts: 2. Singular value analysis. J. Geophys. Res. Solid Earth 116 (B2), B02309. corresponding to a combination of the strike-slip and thrust regime. http://dx.doi.org/10.1029/2010JB007814. In view of the above, the 2014 Cephalonia sequence reflects the general Gallovič, F., Imperatori, W., Mai, P.M., 2015. Effects of three-dimensional crustal structure stress field in the broader region and is in accordance with that. and smoothing constraint on earthquake slip inversions: case study of the Mw6.3 2009 L'Aquila earthquake. J. Geophys. Res. Solid Earth 120 (1), 2169–9356. The conclusions which, to our view, have important implications to http://dx.doi.org/10.1002/2014JB011650. the seismic hazard, can be summarized as follows: a) the active faults Ganas, A., Marinou, A., Anastasiou, D., Paradissis, D., Papazissi, K., Tzavaras, P., Drakatos, G., are not only confined to the CTF, but there are many on-shore faults 2013. GPS-derived estimates of crustal deformation in the central and north Ionian Sea, Greece: 3-yr results from NOANET continuous network data. J. Geodyn. 67-0, the Cephalonia Island; b) the focal mechanisms are not pure dextral 62–71. http://dx.doi.org/10.1016/j.jog.2012.05.010. strike-slip but the thrust component could be significant; c) the dip an- GEER/EERI/ATC, report, 2014. Earthquake Reconnaissance January 26th/February 2nd 2014 gles of these faults are not always nearly vertical, and less steep faults af- Cephalonia, Greece Events, Version 1: June 6, 2014. http://www.geerassociation.org/ GEER_Post%20EQ%20Reports/Cephalonia_Greece_2014/GEER-EERI-ATC_Cephalonia- fect broader regions on land; d) the seismogenic thickness seems to GREECE_v1.pdf. extend from the surface up to ~20 km in the region; and f) the fault rup- Gudmundsson, A., 2011. Rock Fractures in Geological Processes. Cambridge University tures seem to have a tendency to be segmented, which might introduce Press 978-0-521-86392-6 http://dx.doi.org/10.1017/CBO9780511975684. time delays between major patches; this effect may possibly prolong Haslinger, F., Kissling, E., Ansorge, J., Hatzfeld, D., Papadimitriou, E., Karakostas, V., Makropoulos, K., Kahle, H.G., Peter, Y., 1999. 3D crustal structure from local earthquake the duration of strong ground shaking (in time scale of seconds or mi- tomography around the Gulf of Arta (Ionian region, NW Greece). Tectonophysics 304 nutes), and also increase the duration of the elevated seismic hazard (3), 201–218. http://dx.doi.org/10.1016/S0040-1951(98)00298-4. — – (in a time scale of days or weeks). Herrmann, R.B., 1979. Fasthypo ahypocenterlocationprogram.Seismol.Res.Lett.50-2,25 38. Karakostas, V., Papadimitriou, E., Mesimeri, M., Gkarlaouni, C., Paradisopoulou, P., 2014. The 2014 Kefalonia doublet (M w 6.1 and M w 6.0), central Ionian islands, Greece: Acknowledgments seismotectonic implications along the Kefalonia transform fault zone. Acta Geotech. 1–16 http://dx.doi.org/10.2478/s11600-014-0227-4. Karastathis, V.K., Mouzakiotis, E., Ganas, A., Papadopoulos, G.A., 2015. High-precision relo- Constructive comments of E. Burov (Editor), A. Gudmundsson cation of seismic sequences above a dipping Moho: the case of the January–February (Reviewer) and an anonymous reviewer helped to improve the presenta- 2014 seismic sequence on Cephalonia Island (Greece). Solid Earth 6-1, 173–184. tion of the results. We acknowledge the waveform data exchange within http://dx.doi.org/10.5194/se-6-173-2015. Kennett, B.L.N., Kerry, N.J., 1979. Seismic waves in a stratified half space. Geophys. J. Int. the Hellenic Unified Seismic Network (HUSN) for the broad band stations 57-3, 557–583. used in this work. Strong motion data (stations ARG2, VSK1, KRI1, ZAK2 Kiratzi, A., 2014. In: Beer, M., Kougioumtzoglou, I., Patelli, E., Siu-Kui Au, Ivan (Eds.), and CHV1) were provided by the Institute of Engineering Seismology Mechanisms of Earthquakes in Aegean, In: Encyclopedia of Earthquake Engineering. ISBN: 978-3-642-36197-5. and Earthquake Engineering (ITSAK) and by the Geodynamics Institute Klein, F.W., 2002. User's guide to HYPOINVERSE-2000, a Fortran program to solve of the National Observatory of Athens (station SMHA). Phase data for sta- for earthquake locations and magnitudes. U.S. Geol. Surv. Open File Rep. 02–171 tions KEF1, KEF2, KEF3, and KEF4 were downloaded from GI-NOA web (Version 1.0 (2002), 123 pp.). Lagios, E., Papadimitriou, P., Novali, F., Sakkas, V., Fumagalli, A., Vlachou, K., Del Conte, S., site (http://bbnet.gein.noa.gr/HL/). A. Kiratzi acknowledges support by 2012. Combined seismicity pattern analysis, DGPS and PSInSAR studies in the broader the European Union (European Social Fund — ESF MIS377335), Greek na- area of Cephalonia (Greece). Tectonophysics 524–525, 43–58. http://dx.doi.org/10. tional funds — Research Funding Program: Thales — Investing in knowl- 1016/j.tecto.2011.12.015. Lawson, C.L., Hanson, R.J., 1974. Solving Least Square Problems. Prentice-Hall, Inc., New edge society through the European Social Fund (ESF) (MIS377335) and Jersey (340 pp). the Earthquake Planning and Protection Organization (91138) for the Lee, W.H.K., Valdes, C.M., 1985. A personal computer version of the HYPO71 earthquake network operation. Financial support in the Czech Republic: GACR 14- location program. U.S. Geol. Surv. Open File Rep. 85-749, 43. 04372S and CzechGeo/EPOS LM2010008. A number of figures were con- Lekkas, E., Danamos, G., Mavrikas, G., 2001. Geological structure and evolution of Cephalonia andIthakiIslands.Bull.Geol.Soc.GreeceXXXIV(1),11–17. structed using Generic Mapping Tools (Wessel and Smith, 2013). Czech– Lomax, A., Virieux, P., Berge, C., 2000. Probabilistic earthquake location in 3D and layered Greek cooperation under Erasmus+ Program (CZ Praha07 and G Patra01) models: introduction of a Metropolis-Gibbs method and comparison with linear loca- is acknowledged. tions. In: Thurber, C.H., Rabinowitz, N. (Eds.), Advances in Seismic Event Location. Kluwer Academic Publishers. Louvari, E., Kiratzi, A.A., Papazachos, B.C., 1999. The Cephalonia Transform Fault and its Appendix A. Supplementary data extension to western Lefkada Island (Greece). Tectonophysics 308 (1–2), 223–236. http://dx.doi.org/10.1016/S0040-1951(99)00078-5. Michael, A.J., 1987. Use of focal mechanisms to determine stress: a control study. Supplementary data to this article can be found online at http://dx. J. Geophys. Res. 92 (B1), 357–368. http://dx.doi.org/10.1029/JB092iB01p00357. doi.org/10.1016/j.tecto.2015.06.013. Okada, Y., 1992. Internal deformation due to shear and tensile faults in a half-space. Bull. Seismol. Soc. Am. 82-2, 1018–1040. Papadopoulos, G.A., Karastathis, V.K., Koukouvelas, I., Sachpazi, M., Baskoutas, I., References Chouliaras, G., Agalos, A., Daskalaki, E., Minadakis, G., Moshou, A., Mouzakiotis, A., Orfanogiannaki, K., Papageorgiou, A., Spanos, D., Triantafyllou, I., 2014. The Cephalonia, Benekos,G.,Derdelakos,K.,Bountzouklis,C.,Kourkouli,P.,Parcharidis,I.,2015.Sur- Ionian Sea (Greece), sequence of strong earthquakes of January–February 2014: a first face displacements of the 2014 Cephalonia (Greece) earthquake using high report. Res. Geophys. 4-1. http://dx.doi.org/10.4081/rg.2014.5441. resolution SAR interferometry. Earth Sci. Inf. 1–7. http://dx.doi.org/10.1007/ Roumelioti, Z., Benetatos, C., Kiratzi, A., 2009. The 14 February 2008 earthquake (M6.7) se- s12145-015-0205-7. quence offshore south Peloponnese (Greece): source models of the three strongest Bertero, M., Bindi, D., Boccacci, P., Cattaneo, M., Eva, C., Lanza, V., 1997. Application of events. Tectonophysics 471-3 (4), 272–284. http://dx.doi.org/10.1016/j.tecto.2009. the projected Landweber method to the estimation of the source time function in 02.028. seismology. Inverse Prob. 13-2, 465–486. Sakkas, V., Lagios, E., 2015. Fault modelling of the early-2014 ~M6 Earthquakes in Boncori, J.P.M., Papoutsis, I., Pezzo, G., Tolomei, C., Atzori, S., Ganas, A., Karastathis, V., Cephalonia Island (W. Greece) based on GPS measurements. Tectonophysics Salvi, S., Kontoes, C., Antonioli, A., 2015. The February 2014 Cephalonia earthquake 644–645, 184–196. http://dx.doi.org/10.1016/j.tecto.2015.01.010. (Greece): 3D deformation field and source modeling from multiple SAR techniques. Scordilis, E.M., Karakaisis, G.F., Karacostas, B.G., Panagiotopoulos, D.G., Comninakis, Seismol. Res. Lett. 86-1, 124–137. http://dx.doi.org/10.1785/0220140126. P.E., Papazachos, B.C., 1985. Evidence for transform faulting in the Ionian Sea: E. Sokos et al. / Tectonophysics 656 (2015) 131–141 141

the Cephalonia Island earthquake sequence of 1983. Pure Appl. Geophys. 123, Vavryčuk, V., 2014. Iterative joint inversion for stress and fault orientations from focal 388–397. mechanisms. Geophys. J. Int. 199-1, 69–77. http://dx.doi.org/10.1093/gji/ggu224. Shaw, B., Jackson, J., 2010. Earthquake mechanisms and active tectonics of the Hellenic Waldhauser, F., 2001. hypoDD: a computer program to compute double-difference subduction zone. Geophys. J. Int. 181-2, 966–984. http://dx.doi.org/10.1111/j.1365- earthquake locations. U.S. Geol. Surv. Open File Rep. 01–113. 246X.2010.04551.x. Wessel, P., Smith, W.H.F., Scharroo, R., Luis, J., Wobbe, F., 2013. Generic Mapping Tools: Sokos, E.N., Zahradník, J., 2008. ISOLA a Fortran code and a Matlab GUI to perform improved version released. Eos. Trans. AGU 94-45, 409–410. http://dx.doi.org/10. multiple-point source inversion of seismic data. Compt. Rendus Geosci. 34-8, 1002/2013EO450001. 967–977. http://dx.doi.org/10.1016/j.cageo.2007.07.005. Zahradník, J., Custodio, S., 2012. Moment tensor resolvability: application to southwest Sokos, E., Zahradník, J., 2013. Evaluating centroid moment tensor uncertainty in the new Iberia. Bull. Seismol. Soc. Am. 102-3, 1235–1254. http://dx.doi.org/10.1785/ version of ISOLA software. Seismol. Res. Lett. 84-4, 656–665. http://dx.doi.org/10. 0120110216. 1785/0220130002. Zahradník, J., Sokos, E., 2014. The Mw 7.1 Van, Eastern Turkey, earthquake 2011: two- Valkaniotis, S., Ganas, A., Papathanassiou, G., Papanikolaou, M., 2014. Field observations of point source modelling by iterative deconvolution and non-negative least squares. geological effects triggered by the January–February 2014 Cephalonia (Ionian Sea, Geophys. J. Int. 196-1, 522–538. http://dx.doi.org/10.1093/gji/ggt386. Greece) earthquakes. Tectonophysics 630, 150–157. http://dx.doi.org/10.1016/j. Zahradník, J., Gallovic, F., Sokos, E., Serpetsidaki, A., Tselentis, A., 2008. Quick fault-plane tecto.2014.05.012. identification by a geometrical method: application to the Mw 6.2 Leonidio earth- Vallée, M., 2004. Stabilizing the empirical Green function analysis: development of the quake, 6 January 2008, Greece. Seismol. Res. Lett. 79 (5), 653–662. http://dx.doi. projected Landweber method. Bull. Seismol. Soc. Am. 94-2, 394–409. http://dx.doi. org/10.1785/gssrl.79.5.653. org/10.1785/0120030017.