You are on page 1of 11

Geophysical Journal International

Geophys. J. Int. (2015) 203, 18211831

doi: 10.1093/gji/ggv397

GJI Seismology

High-precision differential earthquake location in 3-D models:


evidence for a rheological barrier controlling the microseismicity at
the Irpinia fault zone in southern Apennines
Grazia De Landro,1 Ortensia Amoroso,1,2 Tony Alfredo Stabile,3 Emanuela Matrullo,4
Antony Lomax5 and Aldo Zollo1
1 Department

of Physics, University of Naples Federico II, Naples, Italy. E-mail: delandro@fisica.unina.it


S.c. a r.l., Analysis and Monitoring of Environmental Risk Via Nuova Agnano 11, I-80125 Naples, Italy
3 Consiglio Nazionale delle Ricerche, Istituto di Metodologie per lAnalisi Ambientale (CNR-IMAA), Tito (PZ), Italy
4 INERISInstitut National de lEnvironnement Industriel et des Risques, Nancy, France
5 Alomax Scientific, Mouans-Sartoux, France
2 AMRA

SUMMARY
A non-linear, global-search, probabilistic, double-difference earthquake location technique is
illustrated. The main advantages of this method are the determination of comprehensive and
complete solutions through the probability density function (PDF), the use of differential arrival
times as data and the possibility to use a 3-D velocity model both for absolute and doubledifference locations, all of which help to obtain accurate differential locations in structurally
complex geological media. The joint use of this methodology and an accurate differential
time data set allowed us to carry out a high-resolution, earthquake location analysis, which
helps to characterize the active fault geometries in the studied region. We investigated the
recent microseismicity occurring at the Campanian-Lucanian Apennines in the crustal volume
embedding the fault system that generated the 1980 MS 6.9 earthquake in Irpinia. In order
to obtain highly accurate seismicity locations, we applied the method to the P and S arrival
time data set from 1312 events (ML < 3.1) that occurred from August 2005 to April 2011
and used the 3-D P- and S-wave velocity models optimized for the area under study. Both
manually refined and cross-correlation refined absolute arrival times have been used. The
refined seismicity locations show that the events occur in a volume delimited by the faults
activated during the 1980 MS 6.9 Irpinia earthquake on subparallel, predominantly normal
faults. We find an abrupt interruption of the seismicity across an SWNE oriented structural
discontinuity corresponding to a contact zone between different rheology rock formations
(carbonate platform and basin residuals). This barrier appears to be located in the area
bounded by the fault segments activated during the first (0 s) and the second (18 s) rupture
episodes of the 1980s Irpinia earthquake. We hypothesize that this geometrical barrier could
have played a key role during the 1980 Irpinia event, and possibly controlled the delayed times
of activation of the two rupture segments.
Key words: Seismicity and tectonics; Computational seismology; Continental tectonics:
extensional; Fractures and faults.

1 I N T RO D U C T I O N
The knowledge of the spatial distribution of microseismicity allows
the identification and geometrical characterization of active fault
structures (e.g. Got et al. 1994; Waldhauser & Schaff 2008), the
study of the regional stress field (e.g. De Matteis et al. 2012),
the characterization of the small-scale variability of faulting style,
stress and strength (e.g. Hardebeck 2006; Stabile et al. 2012), the

C

evaluation of the source parameter scaling and radiation efficiency


(e.g. Zollo et al. 2014) and the assessment of the seismic hazard
(e.g. Emolo et al. 2011) in active seismic regions.
The uncertainty on seismicity location parameters is strongly
influenced by the network geometry, the knowledge of the crustal
structures and the number and quality of phase readings (Pavlis
1986; Gomberg et al. 1990). On the other hand, the effects due to
the inaccuracy of the crustal velocity model can be minimized using

The Authors 2015. Published by Oxford University Press on behalf of The Royal Astronomical Society.

1821

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Accepted 2015 September 15. Received 2015 September 14; in original form 2014 December 17

1822

G. De Landro et al.

2 METHODOLOGY
For absolute, non-linear, global-search location in 3-D models we use the methodology proposed by Lomax et al.
(2000), implemented in the NonLinLoc (non-linear location;

http://www.alomax.net/nlloc) software, where the complete probabilistic solution of the earthquake location problem is represented
by a posterior probability density function (PDF).
In order to perform dd earthquake locations, we propose a
methodology that optimizes the relative spatial x, y, z, and t (origin
time) coordinates for a set of hypocentres given a set of differential
phase arrival times measured at each station for multiple hypocentres. The global optimization process is based on the annealing Metropolis algorithm search (Mosegaard & Tarantola 1995;
Lomax et al. 2000; Lomax et al. 2009). It has a temperature parameter which allows searching using a large step-size in x, y, z, and t
for event perturbation initially, with a cooling, but not freezing
to an adaptive, nearly constant step size for later sampling. Thus
the later sampling becomes Metropolis sampling to define the PDF
for each hypocentre in x, y, z, and t. This Metropolis sampling uses
an adaptive step size for each event, which is increased slowly if
the acceptance rate for new, perturbed hypocentres for the event is
high (e.g. > 14 ) and decreased more rapidly if the acceptance rate is
low (e.g. < 14 ). This global optimization is much faster than a pure
Monte Carlo approach and likely faster, simpler and more thorough
than a cascading grid-search.
An important reason for using a non-linear, directed, Metropolistype stochastic search instead of a linearized approach, in addition to
the simple use of 3-D models (3-D traveltime fields) with arbitrary
complexity and parameterization, is the chance of better exploring
complicated cost functions with multiple minima, curved, irregular,
and non-parabolic valleys (see Supporting Information Fig. S11).
Though the extreme high dimensionality of the multievent inverse
problem (curse of dimensionality) does make this extremely difficult for all methods (including and sometimes especially linearized
methods). The disadvantage of this global search method relative
to a linearized one is that the computation time is certainly much
longer and the maximum number of differential times and events
that can be processed is much less.
In the model parameters space exploration, the algorithm seeks
for the parameter solution that maximizes the likelihood function
(based upon the misfit between measured and calculated differential
phase arrival times) while perturbing the hypocentre coordinates.
In particular, the misfit and the solution likelihood are determined
by evaluating the dd equation proposed by Waldhauser & Ellsworth
(2000).
The equation describing the likelihood for a set of arrival-time
difference measures that concerns an event i with coordinates x, y,
z, and origin time t is
L i (x, y, z, t)



obs
cal 2

j
j

i
i

tk tk
tk tk

1 

= exp
wk

2 k
2T 2

(1)

where (. . . )obs is the difference between arrival-times of an event i


and another event j at the same station k, (. . . )cal is the calculated
arrival-time difference of these two events i and j at the station k,
is the uncertainty for a differential time measure, and T is the
current Metropolis (simulated annealing) temperature parameter
(T 1.0) (Lomax et al. 2000).
During the application of the Metropolis random walk, one of the
hypocentres, i, with current likelihood L icurrent (eq. 1), is perturbed
in x, y, z, and t. Next, the likelihood for the perturbed hypocentre i,

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

relative location methods (Poupinet et al. 1984; Got et al. 1994).


When the hypocentral distance between two earthquakes is, in fact,
small as compared to the event-station distance and to the scale
length of velocity heterogeneity, the ray paths between the source
region and a common station are similar. Then, when ray geometry
is favourable, the difference in travel times for two events observed
at one station can be attributed to the spatial offset between the
events, which can be estimated with high accuracy (Got et al. 1994;
Waldhauser & Ellsworth 2000).
The double-difference (dd), relative location method has been
widely used to determine fine-scale fault geometry and produces
high-precision earthquake locations with linearized inversion in the
assumption of a 1-D model of seismic wave propagation (Got et al.
1994; Rubin et al. 1999; Waldhauser & Schaff 2008; Hauksson &
Shearer 2005; Valoroso et al. 2013).
However, it has been demonstrated that the use of an incorrect
velocity model can produce artefacts in the location of hypocentres
even applying the dd method (Michelini & Lomax 2004). In crustal
volumes with strong lateral variation in velocity and irregular topographic surface, the use of a simple 1-D velocity model may be not
appropriate and a 3-D velocity model may be needed to properly
account for such crustal heterogeneities.
3-D crustal seismic velocity models are increasingly available
from seismic traveltime inversion and eventually from geologic
and geophysical interpretations. Probabilistic, absolute earthquake
location using non-linear, global-search methods allows the use of
complex 3-D models and produces comprehensive uncertainty and
resolution assessment (Lomax et al. 2000; Husen et al. 2003; Lomax
et al. 2009). On the other hand, the use of differential traveltimes
with 3-D velocity models is restricted to linearized, dd tomography
(Zhang & Thurber 2003; Thurber et al. 2004; Monteiller et al.
2005; Bannister et al. 2011; Allam & Ben-Zion 2012); in this case
the relative event locations can be determined jointly with the 3-D
velocity model.
In this article, we present a non-linear, global-search, probabilistic dd earthquake location method, implemented in the NLDiffLoc
(non-linear differential location) code, which uses both P- and Swave differential arrival-time data and related uncertainties to determine refined location solutions in complex 3-D media. We show an
application of the methodology to the microseismicity recorded in
the Campania-Lucania region (Southern Italy, see Fig. 1) using both
P- and S-wave manually refined and cross-correlation refined data
set in the high-resolution 3-D Vp and Vs velocity model of the area
(Amoroso et al. 2014a). Synthetic tests are also carried out in order
to assess the performance of the methodology and the parameter
resolution achievable from the available data. In the present study,
we qualitatively compared the quality of cross-correlation refined
data with the manually revised one verifying that the two data sets
provide similar accuracy in earthquake locations. Due to a higher
number of event locations obtained with the manually revised times,
we finally prefer this data set for the interpretation of results and
namely the identification of the discontinuity in the seismicity
spatial distribution, providing new insights on the geometry of the
active faults and their relation to the geological structure and the
present tectonic stress regime.

Differential earthquake location in 3-D models

1823

perturbed

Li
, is evaluated to check if the perturbation improves the fit
to the data, in the sense of the Metropolis rule:
perturbed

(1) always accept the perturbed solution if L i


has higher
likelihood than L icurrent ;
perturbed
(2) otherwise, accept it with probability L i
/L icurrent (see
Mosegaard & Tarantola 1995).
Note that this application of the Metropolis rule acts on the partial
likelihood Li for event i, and not on the total likelihood L for all
hypocentres and over arrival-time difference measures.
Because a main goal of the algorithm is to define the full PDF for
each hypocentre, the main cut-off criterion is a fixed total number
of accepted samples. However, the location of an event can be

automatically rejected (aborted) if either its acceptance rate is very


low or its PDF value remains very low. One or more events can
also be fixed at their initial hypocentres, useful for examining the
relative constraint between differential times for subsets of events.
The maximum likelihood point of the complete, non-linear location PDF is selected as an optimal hypocentre. The significance
and uncertainty of this maximum likelihood hypocentre cannot be
assessed independently of the complete solution PDF. Indeed, the
Gaussian or normal estimators, such as the expectation E(x) and
covariance matrix C may be obtained from the gridded values of
the normalized location PDF or from samples of this function (e.g.
Tarantola & Valette 1982; Sen & Stoffa 1995) in the same way as
in NonLinLoc (Lomax et al. 2000).

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Figure 1. Geological sketch map of the Campania-Lucania region integrated with seismogenic sources (DISS Working group 2010), location of seismic ISNet
stations (blue triangles) and INGV stations (turquoise triangles), main historical earthquakes with their focal mechanisms and aftershock zones (modified after
De Matteis et al. 2012).

1824

G. De Landro et al.

Thus re-weighting of data, for example, due to sensitivity of L2


to outliers, is not necessary, but could be included. For example,
in NLDiffLoc, an adaptive Metropolis step-size algorithm and rejection of events with low acceptance rate or PDF values partly
compensate for outlier data.
We made several synthetic tests in order to show how the two
methods (absolute and dd) works depending on the choice of the
velocity model, of the maximum possible inter-event distance and
on the arrival-time data set accuracy (see Supplementary Material).

3 3 - D M I C R O E A RT H Q UA K E L O C AT I O N
I N T H E I R P I N I A A C T I V E FAU LT Z O N E
3.1 Geological setting

(1) the Lagonegro Basin (LB, Middle Triassic-Miocene):


shallow-water, shelf-margin, and basinal facies successions located
between the ACP and the Western Carbonate Platforms (WCPs);
(2) the WCP succession (or ACP) overthrusts on the LB units.
It consists of a Mesozoic and Paleogene carbonate deposits succession;
(3) internal basin domains, which are related to the Sannio and
Sicilide complex.
Syntectonic terrigenous sequences do not uniformly cover the
thrust sheets stack and represent the infill of Upper Tortonian to
Lower Pleistocene satellite basins (Patacca & Scandone 2001).
The tectonic of this area is controlled by the collision between
the Adriatic microplate and the Apenninic belt, derived by the convergence between the EuroAsian and African plates. The eastward
migration of the thrust-beltforedeepforeland system derived by
the west-dipping subduction process of the Adriatic microplate is
related to the opening of the Tyrrhenian basin (Patacca et al. 1990).
The front of the orogenic wedge reached the present-day location
and stopped at the beginning of the middle Pleistocene. Indeed, a
geodynamic change occurred around 800 km, when an SWNE extension became dominant over the core of the Apennines, as shown
by geological and geomorphological analyses (Cinque et al. 1993;
Galadini 1999; DAgostino et al. 2001). This tectonic regime is still
active, as demonstrated by breakout and seismicity data (Montone
et al. 2004; Pasquale et al. 2009; De Matteis et al. 2012).

3.2 Seismicity data


Since 2005, the capability of detecting and locating smallmagnitude events in this area has been greatly improved (Stabile
et al. 2013) with the deployment of a dense, wide dynamic range,
Irpinia Seismic Network (ISNet; Iannaccone et al. 2010). Presently,
the area exhibits low-magnitude seismicity (ML < 3.5) primarily
concentrated in very limited regions along the 1980 Irpinia and
19901991 Potenza earthquake fault zones (De Matteis et al. 2012).
These weakness zones produce repeated earthquakes and swarmtype microearthquakes sequences, which are concentrated in a few
specific zones of the fault system (Stabile et al. 2012). Qs and Qp
exhibit similar values whereas Vp/Vs is high and a peak in seismicity distribution is observed, is the evidence for a highly fractured,
partially fluid-saturated medium embedding the Irpinia fault zone
(Amoroso et al. 2014b; Zollo et al. 2014). The significant concentration of background seismicity is likely controlled by high pore
fluid pressure. Its increase in fluid-filled cracks around major faults
can lead to the episodic nucleation of moderate to large earthquakes
(Amoroso et al. 2014a).
We analysed a seismic data set consisting of 17 202 traces
recorded by 42 ISNet and INGV stations from 1312 microearthquakes with moment magnitude ranging between 0.9 and
3.1, occurred from August 2005 to April 2011 (De Matteis et al.
2012).
We used the manually picked first P- and S-wave arrival times of
earthquakes recorded by at least four stations. A weighting factor
inversely related to the uncertainty on arrival time picking has been
assigned. A selection based on the location quality was preliminarily
performed: 704 events with at least 5 P and 2 S picked arrival times,
an azimuthal gap smaller than 180 , and root mean square (rms) of
location lower than 0.5 s are taken into account for this study. A total
amount of 10 875 absolute arrival times have been collected and
combined to construct the differential times (see Section 3.3). The
differential times obtained by this procedure constitute the manually
refined picking (MRP) data set.
We can improve the location precision using arrival times refined
by waveform cross-correlation techniques. This technique was used
to construct a cross-correlation refined picking (CRP) data set used
separately from the MRP data set in order to validate the locations
results obtained from the previous data set.
To enhance and improve the accuracy of S-wave picking, we
have applied a phase picking technique based on polarization filtering and waveform coherence analysis (Amoroso et al. 2012). In

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

The Campania-Lucania region (Fig. 1) is located in the axial portion of Southern Apennines, an Adriatic-verging duplex system,
orogenically transported over the flexured southwestern margin of
the Apulia foreland (Patacca et al. 1990). This duplex system consists of a complex architecture of carbonate horsts deriving from
the Apulia Carbonate Platform (ACP), which is overthrusted by
rootless nappes. The belt is located between the Tyrrhenian backarc
basin to the west and with the Bradano foredeep to the east.
The ACP consists of 78 km thick Mesocenozoic carbonate sequence, which overlies Permotriassic clastic deposits (e.g. Verrucano Fm.; Roure et al. 1991). Plio-pleistocene terrigenous deposits
stratigraphically cover the flexed ACP in the eastern margin of the
Bradano Trough (Casnadei 1988). Moving westwards to the external zone of the belt, the ACP progressively dips below the rootless
nappes and is in turn involved in the folds and thrusts of the thrust
belt.
The orogenic stack overlying the ACP is formed by thrust sheets
coming from the deformation of the stratigraphic successions associated with the following main paleogeographic domains (Patacca
et al. 1992):

The Campania-Lucania Apenninic belt is one of the regions in


Italy with the highest seismogenic potential, having experienced
historical and recent destructive earthquakes. The most recent destructive earthquake was the 1980 November 23 MS 6.9 Irpinia
earthquake, generated by the rupture of at least three distinct normal fault segments (Fig. 1; Bernard & Zollo 1989). Since the 1980
earthquake and within its epicentral area, the largest recorded event
(the 1996 April 3 ML 4.9 earthquake) was also characterized by a
normal-faulting mechanism (Fig. 1; Cocco et al. 1999).
In the neighbour of the town of Potenza, about 60 km southeast
of the 1980 Irpinia main shock location, a strike-slip fault zone is
located where two moderate magnitude seismic sequences occurred
between 1990 and 1991 (Ekstrom 1994). The two ML 5 main
shocks were characterized by strike-slip faulting mechanisms with
the preferred fault planes having an eastwest orientation (Fig. 1;
Di Luccio et al. 2005).

Differential earthquake location in 3-D models

3.3 Absolute and double-difference location


of microearthquakes
For both data sets, we have used the same location procedure, based
on the maximum interevent distance, to properly compare the results
of the earthquake locations. For this reason, and also in order to use
a more inclusive criterion of event selection than the minimum
cross-correlation coefficient one, it was chosen an absolute CRP
data set.
The software NLLoc and the tool Loc2ddct, which allow us to
calculate the differential times for a specified maximum distance
between event couples, have been used to obtain the two inputs of
the dd location code, initial absolute location of events, and the
corresponding differential traveltimes.
In order to obtain more accurate and reliable locations by using
the MRP data set (see Supporting Information Fig. S12), we have
made a subdivision in subsets of events located in overlapping rectangular subareas of the region under study. We consider rectangular
boxes with size 10 10 30 km3 and overlapping over an area
of 10 5 km2 . For each box, a maximum of 100 events has been
considered.
Instead, for the CRP data set we used a maximum interevent
distance of 5 km (see Supporting Information Figs S13 and S14),
with no need of a previous section subdivision.
We used 3-D P- and S-wave velocity models, obtained with an
iterative, linearized, tomographic approach in which the P and S arrival times are simultaneously inverted for the earthquake locations
and velocity parameters (Amoroso et al. 2014a). The 3-D P-wave
velocity model is characterized by the presence of a strong lateral
velocity variation at 48 km depth, along the direction orthogonal
to the Apenninic chain, so defining two domains characterized by
relatively low (3.54.8 km s1 ) and high (5.26.5 km s1 ) P-wave
velocities. The S-wave velocity model also shows a change along
the SWNE direction but smaller than the one observed for P-wave
velocity.
In order to demonstrate the improvement of dd application, we
show in Fig. 2 the comparison between the distribution of ISNet
bulletin (catalogue) events obtained from a manual picking data set
by using Hypo2000 and the 1-D velocity model (Matrullo et al.
2013, Figs 2(a) and (b), and the dd events locations obtained with

NLDiffLoc and the 3-D velocity model (Amoroso et al. 2014a;


Figs 2c and d).
The locations of bulletin events exhibit a widespread distribution
throughout the survey area without any specific pattern. Instead, the
locations obtained with the dd method exhibit a pattern of seismicity
with interesting features that will be discussed in the next section.
The mean rms of final dd location is 0.13 s for the MRP data set,
and 0.15 s for the CRP data set. The two dd epicentral distributions
of the events, Figs 2(c) and (d), exhibit very similar characteristics,
both in terms of distribution centroids and relative interevent distances. In particular, according to the previous works (De Matteis
et al. 2012; Matrullo et al. 2013; Amoroso et al. 2014a), the background seismicity shows a diffuse distribution along the NWSE
direction of the Apenninic chain for the Irpinia area, and in the EW
direction for the Potenza area.
The map view in Fig. 2(c), together with the histogram of the
events number in the NWSE direction in Fig. 2(e), shows that
there is a continuous broad zone of seismicity but a rapid decrease
in the rate of events in the SE area is observed. These two zones to
the NW and to the SE, characterized by different rate of seismicity,
are well spatially correlated to the fault segments activated during
the 1980 earthquake and separated by a zone of low seismicity.
The zone to the NW is associated with the 1980 fault segment that
first ruptured and has a higher density of events than the SE zone,
associated with the second fault segment that ruptures 18 s after the
first one.
The depth of events ranges from a few km to about 20 km,
Fig. 2(d), with a higher density of events at around 5 and 12 km
depths, respectively (Fig. 2g). The errors histograms of absolute
(light grey) and dd locations (grey and turquoise) are shown in
Fig. 2(f). The horizontal location error (ErrHo, defined as the length
of the largest projection of the three principal errors on a horizontal
plane) and the depth error (ErrZ, defined as the largest projection of
the three principal errors on a vertical line) decrease significantly
with the use of dd technique. In particular, both dd location errors
are smaller than 0.5 km for the most of the events.
The cross-section in Fig. 2(h) indicates that the seismicity along
the Apenninic chain does not occur on a single, isolated fault but
instead within a volume possibly delimited by the faults activated
during the 1980 earthquake. This is consistent with the extensional
tectonic stress regime characterizing the South Apennines region,
where the seismicity rather occurs along multiple, subparallel, and
en-echelon normal faults associated with the present-day tectonic
deformation acting along the Apenninic belt (Evans et al. 1985;
Suarez et al. 1990; Rigo et al. 1996; De Matteis et al. 2012).
The location errors in Fig. 2(f) show that the two data sets provide
similar accuracy in dd earthquake locations. Moreover, the two
dd location distributions show the same features. Due to a higher
number of events in the MRP data set, we have chosen this data set
for the analysis of the results.

4 DISCUSSION
An innovative dd location method has been applied to the seismicity (1312 microearthquakes) recorded by 42 ISNet and INGV stations from August 2005 to April 2011 along the Campania-Lucania
Apennine chain (Southern Italy). The proposed methodology solves
the dd equations (Waldhauser & Ellsworth 2000) using the nonlinear, global-search, probabilistic location approach developed by
Lomax et al. (2000) instead of linearized approaches. Moreover,
the methodology gives the advantage of relocating the seismicity

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

order to obtain highly accurate, arrival-time readings, the automatic


refined re-picking technique proposed by Rowe et al. (2002) has
been used. The waveforms of all analysed seismic events have been
preliminarily organized in common receiver gathers. For each pair
of traces recorded at the same station, the similarity was evaluated
using a cross-correlation function in a window bracketing a reference picking time. The cross-correlation values are used to identify
clusters of highly similar events, and to estimate relative lags within
the same cluster. The pick adjustment and associated uncertainty
is therefore evaluated through an iterative conjugate gradient technique (see fig. S1 in the Auxiliary Material of Amoroso et al. 2014a).
The uncertainty on refined picking measurements is assigned as the
standard deviation, after estimation via a Monte Carlo sampling
technique (Tarantola 2005). The quality check of this selected data
set provided an estimation of the uncertainty modal value of 2.6 ms
for P-waves and 3.5 ms for S-waves, corresponding to an extremely
high quality data set.
With this technique we have finally obtained 6756 absolute arrival times, corresponding to 513 events, combined to obtain the
differential times that compose the CRP data set.

1825

1826
(a)

G. De Landro et al.
1500'

1530'

1600'

(c)

1500'

1530'

4100'

4100'

1600'

(f)
300

200
100

0
0

1
2
MRP ErrHo (km)

1
2
MRP ErrZ (km)
200

100

4030'

A
0

20

(b)

0
km
10

0.0
20

(d)

0.5 1.0 1.5 0.0 0.5 1.0 1.5


CRP ErrHo (km)
CRP ErrZ (km)

(g)
10
Depth (km)

10

20

Depth (km)

Depth (km)

10

20

20
15.5
Longitude

16.0

15.0

15.5
Longitude

(e) A

16.0

10

20

(h) B

30 40 50 60
Number of events

70

80

20

C
15
60
40
20
0

20
25
10

10

10

10

20

30
40
Distance (km)

50

60

70

Num. of events

Bullettin data set


Crosscorrelation RP data set
Manual RP data set

Depth (km)

10
20
20

10

0
10
Distance (km)

20

Figure 2. (a) Events of the ISNet bulletin located with hypo2000 and the 1-D velocity model. (b) Eastwest vertical section of the ISNet bulletin events. (c)
Double-difference earthquake location of seismicity from 2005 August to 2011 April, from MRP data (in grey) and from CRP data (in turquoise). Red lines are
the surface projection of the three fault segments that ruptured during the 1980 Irpinia earthquake (Pantosti & Valensise 1990). (d) Eastwest vertical section
of the seismicity. (e) Anti-Apenninic AA vertical section of seismicity and histogram of events number moving in the NWSE direction along the same profile.
(f) Comparison between horizontal (left) and vertical (right) errors of absolute location (light grey) and dd locations for the MRP data set (top, dark grey) and
CRP data set (bottom, turquoise). (g) Histogram of the events as function of depth. (h) Earthquake cross-section along the profiles BB and CC indicated in
the map in panel (c).

by using the same 3-D P- and S-wave velocity model (and by considering station elevations) adopted for the absolute location of events,
which is a crucial condition for more reliable and comprehensive
dd locations in complex media such as the area investigated in this
study. Indeed, even if the dd technique allows for the minimization of errors due to un-modelled velocity structures, it has been
shown (e.g. Michelini & Lomax 2004) that the use of an inappropriate 1-D velocity model will, in most cases, lead to bias and error
in the dd locations. Therefore, the combination of the proposed
methodology with cross-correlation differential times allowed us
to perform a robust, high-resolution study for exploring new insights concerning the distribution and the geometry of the analysed
microseismicity.
Actually, the investigated crustal volume is cut by an NW
SE striking normal fault system along the axis of the Apenninic
chain, and to an approximately EW striking, strike-slip fault system, oblique with respect to the main trend of the belt. It has to be
noted that these fault systems correspond with those responsible for
the 1980 MS 6.9, and the 19901991 ML 5.2 and ML 4.7 earthquakes
(De Matteis et al. 2012).
The recent background microseismicity along the chain axis identifies the hanging wall volume delimited by the main fault and its
antithetic one responsible for the 1980 MS 6.9 Irpinia earthquake.
The present-day low-magnitude seismicity follows the same pattern
of the 1980 event aftershocks, with a main elongation parallel to the

strike of the fault segments activated during the Irpinia earthquake.


The earthquakes occurring at the core of the Apennine chain affect
the uppermost 20 km of the crust. Most the earthquakes occur in
carbonate rocks and the underlying crystalline basement (Picentini
and Marzano Mounts). In general, the depth distribution is nearly
uniform except in the Sele River Valley area, where a seismic gap in
the upper crust above about 8 km depth is observed, which is likely
due to the presence of a thick sedimentary cover probably infilling
a paleovalley (Amato & Selvaggi 1993).
We ascribed to the background seismicity the microseismicity
along the 19901991 Potenza seismogenic fault system, suggesting that this EW striking, subvertical, right-lateral structure is still
active at the low-magnitude level. The Potenza earthquakes were
generated within the most internal buried foreland, where it tends to
deepen below the outer front of the Apulia anti-formal stack (i.e. the
deepest part of the Apennine chain; Boncio et al. 2007). This means
that, where present, right-lateral EW striking shear zones could be
active at least as far as the buried Adriatic foreland is not involved in
thrusting. With respect to the Apennine chain, the foreland buried
below the outer front of the Apulia anti-formal stack is the most
internal structural domain where active tectonics and seismicity are
known to occur along EW striking shear zones.
There is a general consensus that EW striking structures, cutting
the foreland crust, are older, pre-existing faults inherited at least
since Mesozoic times, and that the widespread seismicity associated

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

15.0

Depth (km)

km
10

Depth (km)

4030'

Differential earthquake location in 3-D models

1827

is due to their reactivation under the present-day stress field although


not always perfectly oriented with respect to it (e.g. Di Bucci et al.
2010; Latorre et al. 2010). These structures are large regional fault
zones, which dissect the foreland crust and have experienced longlasting activity under different tectonic regimes, that is to say, under
different kinematics at different times (Di Bucci et al. 2010).
The most important result of this paper is the evidence, from
both the distribution of seismicity and tomography, of the presence
of a rheological discontinuity (Fig. 3). Indeed, in the southeastern
part of the Marzano massif is clearly visible the alignment of the
epicentres SWNE and the abrupt limit of seismicity in the orthogonal direction of Apenninic chain. We hypothesize that this barrier
can be due to the contact between units with different rheological
behaviour in response to the NESW stress regime acting in the
chain: carbonate units (ACP) at NW and the most recent basinal
facies (LB) at SE (e.g. Pescatore et al. 1999; Scrocca et al. 2005).
It appears indeed that the seismicity distribution is controlled by
the geometry of the ACP (Improta et al. 2003): the barrier in the
seismicity highlighted by these data is actually located between a
high and a low in the Apulian carbonates (Fig. 3a). This is more
evident in correspondence of a low depth of ACP (Fig. 3a) where a
seismic gap is observed until 6 km in depth. The presence of surface
deposits of the plio-Pleistocene is a consequence of the presence of
a depression of the Apulian carbonates.
This is also confirmed by the 3-D velocity model retrieved in
the area which shows, by moving in the SWNE direction, an indepth of the Apulian Platform and a decrease in the rate of events
in the area corresponding to the basinal deposits (Fig. 3b). The
different rheology of contact geological formations leads, thus,
to a substantial decrease in the rate of seismicity moving in SW
direction.

A schematic geological interpretation of the BB section (Fig. 3b)


is shown in Fig 4. To construct this figure we jointly used the
distribution of relocated seismicity, the tomographic images and
the topography of the Apulian Platform. The various geological
formations are shown at different depths. The shallowest area of
basinal deposits is between 0 and 2 km; the Lagonegro unit area
is between 2 and 6 km, where P-waves have velocity between 4.4
and 6.2 km s1 ; the Apulian platform, whose top is between 5 and
6 km in depth, in which the P-wave velocity ranges between 5.6 and
6.3 km s1 ; finally the basement between 9 and 12 km depth.
We suggest that this discontinuity is a transverse fault which
displaces a portion of the Apulian carbonate platform. This fault is
oriented in such a way as not to be easily activated in the current
stress field (the strike of this structure is approximately parallel to
the axis of minimum compression, NESW) and could constitute a
barrier to the propagation of seismic ruptures towards SE.
This hypothesis is in agreement with previous work (Bernard &
Zollo 1989; Cocco et al. 1999), according to which the cause of
the delay time between the dislocation of the first two segments
of 1980s earthquake was associated to the possible presence of a
geometrical barrier. Indeed, it is observed that the lower limit
of the DISS fault (DISS Working Group 2010) relative to the first
segment corresponds to the discontinuity of the seismicity (Fig. 2a).
During the 1980 Irpinia earthquake, the Monte Marzano segment
ruptured first, and its nucleation occurred near its northeastern edge
with a normal fault mechanism on a plane dipping 60 towards the
northeast. The rupture propagated bilaterally for 15 km towards NE
and for a few km in SW direction.
The southern segment rupture started 18 s after nucleation of
the first event and 15 km southeast of its hypocentre, on a lowangle normal fault (30 ) dipping towards the northeast (Bernard &

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Figure 3. (a) Comparison of the dd location of seismicity with the topography of the Apulian Carbonate Platform top (modified after Improta et al. 2003). (b)
Comparison of tomographic cross-sections reported in map in panel (a) and the earthquake occurred in this area.

1828

G. De Landro et al.

Zollo 1989). This discontinuity has, thus, formed a barrier that has
resisted and slowed down the evolution of the rupture SE, forcing
it to change its geometry and restarted with a delay of about 20
s along a low-angle fault deeper, probably where the fracture has
found again the carbonate platform (Bernard & Zollo 1989).

(5) The hypothesis of the presence of a barrier can be confirmed


by the delay time in the rupture of the first two segment of the same
1980s Irpinia earthquake (Bernard & Zollo 1989; Cocco et al. 1999)
and can be hypothesized that the individuated discontinuity had
an important role in this highly destructive event.

5 C O N C LU S I O N S

D ATA A N D R E S O U R C E S

In this work we presented a non-linear, global-search, probabilistic


dd earthquake location method that uses as data P- and S-wave
differential arrival-times:

Data can be obtained from the ISNet Bulletin at http://isnet.na.


infn.it/cgibin/isnetevents/isnet.cgi (last accessed month year) and
from the INGV data management centre http://iside.rm.ingv.
it/iside/standard/index.jsp.
Some plots were made using the Generic Mapping Tools
(www.soest.hawaii.edu/gmt; Wessel & Smith 1998). The Non
Linear Difference Location code (NLDiffLoc) is available at
http://www.alomax.net.

(1) The method, implemented in the NLDiffLoc code, has the advantage to obtain from one hand comprehensive solutions through
the PDF using as data the differential traveltimes and, on the other
hand, is implemented taking into account of complex media requiring a 3-D velocity models.
(2) This method was applied to relocate the microseismicity in
the Campania-Lucania region inverting MRP and CRP P- and Swave arrival-times, and using the high-resolution, 3-D Vp and Vs
velocity model of the area recently determined by Amoroso et al.
(2014a).
(3) The relocated seismicity shows a diffuse distribution in the
NWSE direction along the Appeninic chain for the Irpinia area
and the seismicity along the chain does not occur on a single fault
but in a volume delimited by the faults activated during the 1980
Irpinia earthquake. This is consistent with previous observations
and seismicity analyses (De Matteis et al. 2012; Matrullo et al.
2013; Amoroso et al. 2014a).
(4) The distribution of relocated seismicity and the tomographic
image show the presence of a rheological discontinuity due to the
contact between the carbonate units (APC) and the basinal facies
(LB) that have different rheological behaviour in to the chain stress
regime. This discontinuity is underlined by a significant reduction
in the rate of events in the zone corresponding to a low of the
Apulian Platform filled by the deposits of the LB.

AC K N OW L E D G E M E N T S
This work has been partially funded by Project SHEERSHale
gas Exploration and Exploitation induced Risks (GA N.640896
H2020-LCE-2014-2015/H2020-LCE-2014-1), and Project EPOS
IP (GA N.676564H2020-INFRADEV-2014-2015). The authors
wish to thank the editor Yehuda Ben-Zion and the reviewers Amir
Allam and Jean-Luc Got for their valuable comments which have
been contributed to improve the quality of the manuscript.

REFERENCES
Allam, A.A. & Ben-Zion, Y., 2012. Seismic velocity structures in the Southern California plate-boundary environment from double-difference tomography, Geophys. J. Int., 190, 11811196.
Amato, A. & Selvaggi, G., 1993. Aftershock location and P-wave velocity
structure in the epicentral region of the 1980 Irpinia earthquake, Ann.
Geofis., 36(1), 315.

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Figure 4. Schematic geological interpretation. The positions of the interfaces are deduced from both tomographic model and hypocentre-relocated events. The
name of the lithologies and the position of main topographic references are also indicated and supposed from Improta et al. (2003). The crosses indicate the
rough position of the retrieved hypocentres. A dashed line is used to represent the hypothetical barrier/ fault scarp.

Differential earthquake location in 3-D models

Hardebeck, J.L., 2006. Homogeneity of small-scale earthquake faulting,


stress, and fault strength, Bull. seism. Soc. Am., 96(5), 16751688.
Hauksson, E. & Shearer, P., 2005. Southern California hypocenter relocation with waveform cross-correlation, Part 1: results using the doubledifference method, Bull. seism. Soc. Am., 95(3), 896903.
Husen, S., Kissling, E., Deichmann, N., Wiemer, S., Giardini, D. & Baer,
M., 2003. Probabilistic earthquake location in complex three-dimensional
velocity models: application to Switzerland, J. geophys. Res., 108, 2077
2102
Iannaccone, G. et al., 2010. A prototype system for earthquake early-warning
and alert management in southern Italy, Bull. Earthq. Eng., 8, 1105
1129.
Improta, L., Bonagura, M., Capuano, P. & Iannaccone, G., 2003. An integrated geophysical investigation of the upper crust in the epicentral area of
the 1980, Ms = 6.9, Irpinia earthquake (Southern Italy), Tectonophysics,
361(12), 139169.
Latorre, D., Amato, A. & Chiarabba, C., 2010. High-resolution seismic imaging of the Mw 5.7, 2002 Molise, southern Italy, earthquake area: evidence of deep fault reactivation, Tectonics, 29, TC4014,
doi:10.1029/2009TC002595.
Lomax, A., Virieux, J., Volant, P. & Thierry, B.C., 2000. Probabilistic earthquake location in 3D and layered models: introduction of a Metropolis
Gibbs method and comparison with linear locations, in Advances in Seismic Event Location, pp. 101134, eds Thurber, C.H. & Rabinowitz, N.,
Kluwer Academic.
Lomax, A., Michelini, A. & Curtis, A., 2009. Earthquake location, direct,
Global-SearchMethods, in Encyclopedia of Complexity and Systems Science, Part 5, pp. 24492473, ed. Meyers, A., Springer.
Matrullo, E., De Matteis, R., Satriano, C., Amoroso, O. & Zollo, A.,
2013. An improved 1-D seismic velocity model for seismological studies in Campania-Lucania region (Southern Italy), Geophys. J. Int., 195,
460473.
Michelini, A. & Lomax, A., 2004. The effect of velocity structure errors
on double-difference earthquake 590 location, Geophys. Res. Lett., 31,
L09602, doi:10.1029/2004GL019682.591.
Monteiller, V., Got, J.-L., Virieux, J. & Okubo, P., 2005. An efficient algorithm for double difference tomography and location in heterogeneous
media, with an application to the Kilauea volcano, J. geophys. Res.,
110(B12), 122.
Montone, P., Mariucci, M.T., Pondrelli, S. & Amato, A., 2004. An improved
stress map for Italy and surrounding regions (central Mediterranean), J.
geophys. Res., 88, 64156429.
Mosegaard, K. & Tarantola, A., 1995. Monte Carlo sampling of solutions
to inverse problems, J. geophys. Res., 100, 12 43112 447.
Pantosti, D. & Valensise, G., 1990. Faulting mechanism and complexity of
the November 23, 1980, Campania Lucania earthquake, inferred from
surface observations, J. geophys. Res., 95(15), 319341.
Pasquale, G., De Matteis, R., Romeo, A. & Maresca, R., 2009. Earthquake
focal mechanisms and stress inversion in the Irpinia Region (southern
Italy), J. Seism., 13, 107124.
Patacca, E., Sartori, R. & Scandone, P., 1990. Tyrrhenian basin and Apenninic arcs: kinematic relations since late Tortonian times, Memorie della
Societ`a, Geologica Italiana, 45, 425451.
Patacca, E., Scandone, P., Ballatalla, M., Perilli, N. & Santini, U., 1992.
The Numidian-sand event in the Southern Apennines, Mem. Soc. Geol.
Padova, 43, 297337.
Patacca, E. & Scandone, P., 2001. Late thrust propagation and sedimentary
response in the thrust belt foredeep system of the Southern Apennines
(Pliocene Pleistocene), in Anatomy of an Orogen: The Apennines and
Adjacent Mediterranean Basins, pp. 401440, eds Vai, G.B. & Martini,
I.P., Kluwer Academic.
Pavlis, G.L., 1986. Appraising earthquake hypocenter location errors: a
complete, practical approach for 623 single event locations, Bull. seism.
Soc. Am., 76, 16991717.
Pescatore, T.S., Renda, P., Schiattarella, M. & Tramutoli, V., 1999. Stratigraphic and structural relationships between Meso-Cenozoic Lagonegro basin and coeval carbonate platforms in southern Apennines, Italy,
Tectonophysics, 315, 269286.

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Amoroso, O., Maercklin, N. & Zollo, A., 2012. S-wave identification by


polarization filtering and waveform coherence analyses, Bull. seism. Soc.
Am., 102(2), 854861.
Amoroso, O., Ascione, A., Mazzoli, S., Virieux, J. & Zollo,
A., 2014a. Seismic imaging of a fluid storage in the actively
extending Apennine mountain belt, southern Italy, Geophys. Res. Lett.,
41, doi:10.1002/2014GL060070.
Amoroso, O., Russo, G., Orefice, A., Virieux, J. & Zollo, A., 2014b. 3D
seismic imaging of the Irpinia fault system (Southern Italy) from multiscale velocity and attenuation tomography, in 2ECEES, Istanbul, August
2529, doi:10.13140/2.1.1529.7922.
Bannister, S., Fry, B., Reyners, M., Ristau, J. & Zhang, H., 2011. Fine-scale
relocation of aftershocks of the 22 February Mw 6.2 Christchurch earthquake using double-difference tomography, Seismol. Res. Lett., 82(6),
839845.
Bernard, P. & Zollo, A., 1989. The Irpinia (Italy) 1990 earthquake: detailed
analysis of a complex normal fault, J. geophys. Res., 94, 16311648.
Boncio, P., Mancini, T., Lavecchia, G. & Selvaggi, G., 2007. Seismotectonics
of strikeslip earthquakes within the deep crust of southern Italy: Geometry, kinematics, stress field and crustal rheology of the Potenza 19901991
seismic sequences (Mmax 5.7), Tectonophysics, 445, 281300.
Casnadei, R., 1988. Subsurface basin analysis of fault-controlled turbidite
system in Bradano Trough, Southern Adriatic foredeep, Italy, Bull. Am.
Assoc. Pet. Geol., 72(11), 13701380.
Cinque, A., Patacca, E., Scandone, P. & Tozzi, M., 1993. Quaternary kinematic evolution of Southern Apennines. Relationships between surface
geological features and deep lithospheric structures, Ann. Geofis., 37,
249260.
Cocco, M. et al., 1999. The April 1996 Irpinia seismic sequence: evidence
for fault interaction, J. Seism., 3, 105117.
DAgostino, N., Jackson, J.A., Dramis, F. & Funiciello, R., 2001. Interactions
between mantle upwelling, drainage evolution and active normal faulting:
an example from the central Apennines (Italy), Geophys. J. Int., 147, 475
497.
De Matteis, R., Matrullo, E., Rivera, L., Stabile, T.A., Pasquale, G. & Zollo,
A., 2012. Fault delineation and regional stress direction from the analysis
of background microseismicity in Southern Apennines, Italy, Bull. seism.
Soc. Am., 102(4), 18991907.
Di Bucci, D., Burrato, P., Vannoli, P. & Valensise, G., 2010. Tectonic evidence for the ongoing Africa-Eurasia convergence in central Mediterranean foreland areas: a journey among long-lived shear zones, large
earthquakes, and elusive fault motions, J. geophys. Res., 115, B12404,
doi:10.1029/2009JB006480.
Di Luccio, F., Fukuyama, E. & Pino, N.A., 2005. The 2002 Molise earthquake sequence: what can we learn about the tectonics of Southern Italy?,
Tectonophysics, 405, 141154.
DISS Working Group, 2010. Database of Individual Seismogenic Sources
(DISS), Version 3.1.1: a compilation of potential sources for earthquakes larger than M 5.5 in Italy and surrounding areas. Available
from the Istituto Nazionale di Geofisica e Vulcanologia (INGV) at:
http://diss.rm.ingv.it/dissNet/, last accessed July 2011.
Ekstrom, G., 1994. Teleseismic analysis of the 1990 and 1991 earthquakes
near Potenza, Ann. Geofis., XXXVII, 15911599.
Emolo, A., Convertito, V. & Cantore, L., 2011. Ground-motion predictive
equations for low-magnitude earthquakes in the CampaniaLucania area,
Southern Italy, J. geophys. Eng., 8(1), 4660.
er, S.B., 1985. Tectonics of the
Evans, R., Asudeh, I., Crampin, S. & Uc
Marmara Sea region of Turkey: new evidence from microearthquake fault
plane solutions, Geophys. J. R. astr. Soc., 83, 4760.
Galadini, F., 1999. Pleistocene changes in the central Apennine fault kinematics: a key to decipher active tectonics in central Italy, Tectonics, 18,
877894.
Gomberg, J.S., Shedlock, K.M. & Roecke, S.W., 1990. The effect of S-wave
arrival times on the accuracy of 557 hypocenter estimation, Bull. seism.
Soc. Am., 80, 16051628.
Got, J.-L., Frechet, J. & Klein, F.W., 1994. Deep fault plane geometry inferred
from multiplet relative relocation beneath the south flank of Kilauea, J.
geophys. Res., 99, 15 37515 386.

1829

1830

G. De Landro et al.

S U P P O RT I N G I N F O R M AT I O N
Additional Supporting Information may be found in the online version of this paper:
Figure S1. Test with the 3-D velocity model: maximum distance
5 km versus 10 km. The synthetic arrival times were inverted to

calculate the initial absolute locations with NLLoc. For the dd location, we built the differentials arrival times considering all possible
combinations of events with maximum distances of 5 and 10 km.
The figure shows comparison between the dd location results of the
synthetic events using the 3-D velocity model and considering 5 km
(blue) and 10 km (green) of maximum distance to calculate the differential times. Red stars indicate the reference positions of events.
Grey triangles are the ISNet seismic stations. (a) Seismicity distribution in plane. (b) Eastwest vertical section of the seismicity. (c)
Distance in plane between the calculated and the true position (left)
and distance between the calculated and the true depth of the events
(right). (d) Horizontal (left) and vertical (right) locations errors. (e)
Root mean square (rms) distribution.
Figure S2. Spatial location error distribution for synthetic events
considering the maximum distance of 5 km. (a) Confidence ellipsoid
at 68 per cent projections in the XY plane. (b) Confidence ellipsoid
at 68 per cent projections in the XZ plane. (c) Confidence ellipsoid at
68 per cent projections in the YZ plane. (d) Zoom of the confidence
ellipsoids in the zone with higher events density. (e) Zoom of the
confidence ellipsoids in the zone with lower events density.
Figure S3. Plot of events interconnection. (a) Considering 5 km
maximum distance, every event has a maximum of eight connections and at least one single connection. With this condition the obtained locations have mean distance from the true position of about
200 m in plane and 200 m in depth (Fig. S1c, blue). (b) Considering 10 km maximum distance, every event has at least seven
connections and a maximum of 16 connections. With this condition
the obtained locations are significantly improved, in fact, the mean
distance is about 30 m in plane and 25 m in depth (Fig. S1c, green).
Thus the minimum number of connections to constrain the position
of a single event is about 10, equal to the number of connections of
the central events of the finer grid. (c) In the case of real locations, for
the MRP data set with the use of the sections every event has on
average 40 connections, whereas for the CRP data set with the use
of 5 km as the maximum distance every event has on average 15
connections (see Fig. S12b).
Figure S4. Test with the 1-D velocity model: absolute versus dd
locations. The 3-D model synthetic arrival times were inverted with
the 1-D velocity model to calculate the initial absolute locations
with NLLoc. For the dd location, we built the differential arrival
times considering all possible combinations of events with maximum distance of 5 and 10 km, and then locate considering the
1-D velocity model (Matrullo et al. 2013). (a) Comparison between
the absolute initial location (blue) and the dd final location of the
synthetic events (green) with the 1-D velocity model. Red star indicate the reference positions of events. Grey triangles are the ISNet
seismic stations. (a) Seismicity distribution in plane. (b) Eastwest
vertical section of the seismicity. (c) Distance in plane between the
calculated and the true position. (d) Distance between the calculated
and the true depth of the events. (e) rms distribution.
Figure S5. (a) Comparison between absolute initial location (blue)
and dd final location (green) of the synthetic events with the 3-D
velocity model. Red stars indicate the events reference positions.
Grey triangles are the ISNet seismic stations and red stars are the
reference positions of events. (a) Seismicity distribution in plane.
(b) Eastwest vertical section of the seismicity. (c) In plane (XY)
and vertical (Z) distance between the locations and the true position
of events. (d) Horizontal (XY) and vertical (Z) location errors. (e)
rms distributions.
Figure S6. Spatial location error distribution of synthetic
events considering data with noise. (a) Confidence ellipsoid at
68 per cent projections in the XY plane. (b) Confidence ellipsoid

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

Poupinet, G., Ellsworth, W.L. & Frechet, J., 1984. Monitoring velocity variations in the crust using earthquake doublets: an application to the Calaveras Fault, California, J. geophys. Res., 89(B7), 57195731.
Rigo, A., Lyon-Caen, H., Armijo, R., Deschamps, A., Hatzfeld, D., Makropoulos, K., Papadimitriou, P. & Kassaras, I., 1996. A microseismic
study in the western part of the Gulf of Corinth (Greece): implications for large-scale normal faulting mechanisms, Geophys. J. Int., 126,
663688.
Roure, F., Casero, P. & Vially, R., 1991. Growth processes and melange
formation in the Southern Apennines accretionary wedge, Earth planet.
Sci. Lett., 102, 395412.
Rowe, C., Aster, R., Brochers, B. & Young, C., 2002. An automatic, adaptive
algorithm for refining phase picks in large seismic data sets, Bull. seism.
Soc. Am., 92, 16601674.
Rubin, A.M., Gillard, D. & Got, J.L., 1999. Streaks of microearthquakes
along creeping faults, Nature, 400, 635641.
Scrocca, D., Carminati, E. & Doglioni, C., 2005. Deep structure of the
southern Apennines, Italy: thin-skinned or thick-skinned?, Tectonics, 24,
TC3005, doi:10.1029/2004TC001634.
Sen, M.K. & Stoffa, P.L., 1995. Global Optimization Methods in Geophysical
Inversion, Elsevier, 281 pp.
Stabile, T.A., Satriano, C., Orefice, A., Festa, G. & Zollo, A., 2012. Anatomy
of a microearthquake sequence on an active normal fault, Sci. Rep., 2,
doi:10.1038/srep00410.
Stabile, T.A., Iannaccone, G., Zollo, A., Lomax, A., Ferulano, M.F., Vetri,
M.L.V. & Barzaghi, L.P., 2013. A comprehensive approach for evaluating network performance in surface and borehole seismic monitoring,
Geophys. J. Int., 192(2), 793806.
Suarez, G., Monfret, T., Wittlinger, G. & David, C., 1990. Geometry of subduction and depth of the seismogenic zone in the Guerrero gap, Mexico,
Nature, 345, 336338.
Tarantola, A., 2005. Inverse Problem Theory and Methods for Model Parameter Estimation, SIAM.
Tarantola, A. & Valette, B., 1982. Inverse problems = quest for information,
J. geophys. Res., 50, 159170.
Thurber, C., Roecker, S., Zhang, H., Baher, S. & Ellsworth, W.,
2004. Fine-scale structure of the San Andreas fault zone and location of the SAFOD target earthquakes, Geophys. Res. Lett., 31(12),
doi:10.1029/2003GL019398.
Valoroso, L., Chiaraluce, L., Piccinini, D., Di Stefano, R., Schaff, D. &
Waldhauser, F., 2013. Radiography of a normal fault system by 64 000
high-precision earthquake locations: the 2009 LAquila (central Italy) case
study, J. geophys. Res., 118, 11561176.
Waldhauser, F. & Ellsworth, W.L., 2000. A double-difference earthquake
location algorithm: method and application to the northern Hayward Fault,
California, Bull. seism. Soc. Am., 90, 13531368.
Waldhauser, F. & Schaff, D.P., 2008. Large-scale relocation of two
decades of Northern California seismicity using cross-correlation
and double-difference methods, J. geophys. Res., 113, B08311, doi:
10.1029/2007jb005479.
Wessel, P. & Smith, W.H.F., 1998. New, improved version of the Generic
Mapping Tools Released, EOS, Trans. Am. geophys. Un., 79, 579.
Zhang, H. & Thurber, C.H., 2003. Double-difference tomography: the
method and its application to the Hayward fault, California, Bull. seism.
Soc. Am., 93, 18751889.
Zollo, A., Orefice, A. & Convertito, V., 2014. Source parameter scaling and radiation efficiency of microearthquakes along the Irpinia
fault zone in southern Apennines, Italy, J. geophys. Res., 119,
doi:10.1002/2013JB010116.

Differential earthquake location in 3-D models

Figure S11. Plot of a normalized irregular likelihood function of an


event of the MRP data set located in the 3-D velocity model considering all the event with maximum distance of 5 km to calculate the
differential times. (a) Normalized likelihood function in 3-D. (b)
Likelihood function section in the latitudelongitude plane around
the maximum likelihood depth. (c) Likelihood function section in
the latitudedepth plane around the maximum likelihood longitude.
Figure S12. dd location of the MRP data set with the 3-D velocity model considering all events with maximum distance of 5 km
(turquoise dots) and subdivision in section of 10 10 30 km3
(grey dots). Grey triangles are the seismic stations. (a) Seismicity
distribution in plane. (b) East-west vertical section of the seismicity. (c) Anti-Apenninic section according the dotted red lines in
panel (a). (d) Horizontal (right) and vertical (left) location error
distributions. (e) rms distribution.
Figure S13. dd location of the CRP data set with the 3-D velocity model considering all events with maximum distance of 5 km
(turquoise dots) and subdivision in section of 10 10 30 km3
(grey dots). Grey triangles are the seismic stations. (a) Seismicity distribution in plane. (b) East-west vertical section of the
seismicity. (c) Anti-Appeninic section according the dotted red lines
in (a). (d) Horizontal (right) and vertical (left) location error distributions. (e) rms distribution.
Figure S14. Plot of events interconnections for the CRP (a) and for
MRP (b) data set considering 5 km as maximum distance (http://gji.
oxfordjournals.org/lookup/suppl/doi:10.1093/gji/ggv397/-/DC1).
Please note: Oxford University Press is not responsible for the content or functionality of any supporting materials supplied by the
authors. Any queries (other than missing material) should be directed to the corresponding author for the paper.

Downloaded from http://gji.oxfordjournals.org/ at Seoul National University on May 22, 2016

at 68 per cent projections in the XZ plane. (c) Confidence ellipsoid at 68 per cent projections in the YZ plane. (d) Zoom of the
confidence ellipsoids in the zone with higher events density. (e)
Zoom of the confidence ellipsoids in the zone with lower events
density.
Figure S7. (a) Comparison between the dd traveltime residuals
of the initial absolute location (green) and the final dd location
(blue) for the MRP data set. (b) Comparison between the horizontal
(right) and vertical (left) location errors. (c) Comparison between
rms. Residual histograms of final dd location are improved in terms
of narrower traveltimes residual distribution, in terms of rms that
decreases on average of 0.05 s, and in terms of location errors that
decrease on average of about 0.5 km.
Figure S8. (a) Comparison between the dd traveltime residuals of
the initial absolute location (green) and the final dd location (blue)
for the CRP data set. (b) Comparison between the horizontal (right)
and vertical (left) location errors. (c) Comparison between rms. The
residual histograms of final dd location are improved both in terms
of narrower traveltime residual distribution, in terms of rms that
decreases on average of 0.1 s, and in terms of location errors, that
decrease on average of about 0.5 km.
Figure S9. Histograms of normalized likelihood function for four
different events of the MRP data set located with the 3-D velocity
model.
Figure S10. Plot of the normalized near-ellipsoidal likelihood
function of an event of the MRP data set located in the 3D velocity model. (a) Normalized likelihood function in 3-D.
(b) Likelihood function section in the latitudelongitude plane
around the maximum likelihood depth. (c) Likelihood function section in the latitudedepth plane around the maximum likelihood
longitude.

1831

You might also like