You are on page 1of 15

Review Article

Earthquake nests as natural laboratories for the study of intermediate-depth


earthquake mechanics
Germn A. Prieto
a,
, Gregory C. Beroza
b
, Sarah A. Barrett
b
, Gabriel A. Lpez
c
, Manuel Florez
a
a
Departamento de Fsica, Universidad de los Andes, Colombia
b
Department of Geophysics, Stanford University, Palo Alto, CA, USA
c
Department of Imaging Science and Technology, TU Delft, The Netherlands
a b s t r a c t a r t i c l e i n f o
Article history:
Received 18 January 2012
Received in revised form 9 July 2012
Accepted 21 July 2012
Available online 1 August 2012
Keywords:
Intermediate-depth earthquakes
Earthquake nests
Earthquake clustering
Rupture mechanism
Earthquake location
Stress drops
The physical mechanism of intermediate-depth earthquakes is still under debate. In contrast to conditions in the
crust and shallow lithosphere, at temperatures and pressures corresponding to depths >50 km, rocks ought to
yield by creep or ow rather than brittle failure. Some physical process has to enable brittle or brittle-like failure
for intermediate-depth earthquakes. The two leading candidates for that are dehydration embrittlement and ther-
mal shear runaway. Given their great depth, intermediate-depth earthquake processes can't be observed direct-
ly. Instead we must rely on a combination of seismology and the study of laboratory analogs to understand them.
Earthquake nests are regions of highly concentrated seismicity that are isolated from nearby activity. In this
paper we focus on three intermediate-depth earthquake nests Vrancea, Hindu Kush and Bucaramanga, and
what they reveal about the mechanics of intermediate-depth earthquakes. We review published studies of
tectonic setting, focal mechanisms, precise earthquake locations and earthquake source physics at these
locations, with an emphasis on the Bucaramanga nest. All three nests are associated with subducting litho-
sphere and at least two of the nests have consistently larger stress drops compared to shallow seismicity.
In contrast, the Bucaramanga nest has a larger b-value, larger variability of focal mechanisms and shows
no evidence of aftershock sequences unlike the other two. We also report for the rst time nding a signi-
cant number of repeating earthquakes, some with reverse polarity.
Given the nature and characteristics of earthquake nests, they can be thought as natural laboratories. Future
seismological studies of intermediate-depth earthquakes in nests will likely enlighten our understanding of
their physical mechanisms.
2012 Elsevier B.V. All rights reserved.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2. Proposed mechanisms for intermediate depth earthquakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.1. Dehydration embrittlement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.2. Thermal shear runaway instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3. Earthquake nests in the Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4. Tectonic setting of earthquake nests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.1. Vrancea nest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2. Hindu Kush nest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3. Bucaramanga nest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5. Location of earthquakes within the Bucaramanga nest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.1. Repeating earthquakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6. Temporal behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.1. Aftershock sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7. Earthquake source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7.1. Size distribution (b-values) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7.2. Focal mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
7.3. Stress drops and source physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Tectonophysics 570571 (2012) 4256
Corresponding author.
E-mail address: gprieto@uniandes.edu.co (G.A. Prieto).
0040-1951/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tecto.2012.07.019
Contents lists available at SciVerse ScienceDirect
Tectonophysics
j our nal homepage: www. el sevi er . com/ l ocat e/ t ect o
8. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
1. Introduction
Just what constitutes an intermediate-depth earthquake? Early de-
nitions seemto come fromthe expectation that improved understanding
of earthstructure wouldprovide clear denitions for distinct populations.
For example, Wadati (1929) denes shallowearthquakes as b60 kmand
intermediate earthquakes as 100200 km. While Gutenberg and Richter
(1949), use 70300 km in most (but not all) of their publications.
Presently, without fully understanding the mechanism by which inter-
mediate and deep earthquakes occur, it is difcult to dene a depth
cut-off. Another denition might use the distribution of global seismicity,
dening the start of intermediate-depth seismicity at the onset of the
approximately exponential decay in seismic frequency, extending this
range to ~300 km. We choose a denition based on the maximum
depth of inter-plate subduction zone earthquakes of ~50 km (Bilek
et al., 2004). This is also the depth at which well-developed aftershock
sequences become less frequent (Frohlich, 2006).
Earthquakes deeper than 50 km represent approximately 25% of
global seismicity (Frohlich, 2006). Intermediate-depth earthquakes,
the focus of this contribution, are those earthquakes with depths
ranging from about 50 to 300 km (Frohlich, 2006; Houston, 2007).
They occur exclusively at convergent plate boundaries within
subducting lithosphere, although in some cases the subduction that
gave rise to the lithosphere at depth may no longer be active at the
surface (Chung and Kanamori, 1976).
Intermediate-depth earthquakes occur at temperatures and pres-
sures above the point where ordinary fractures ought to occur, and
despite their abundance, the physical mechanismbehind themremains
uncertain (Frohlich, 1989; Green and Houston, 1995; Houston, 2007).
Perhaps the leading hypothesis is dehydration embrittlement, in
which hydrated minerals release uids at particular pressures and
temperatures allowing brittle failure to occur (Frohlich, 1989; Green
and Houston, 1995; Hacker et al., 2003; Kirby et al., 1996a; Wiens,
2001). Another plausible mechanism is known as thermal shear run-
away instability (Ogawa, 1987; Hobbs and Ord, 1988; Frohlich, 2006;
Houston, 2007; Keleman and Hirth, 2007; John et al., 2009), which
would occur through a positive, rapid feedback between shear strain
localization and thermal heating.
A particular kind of intermediate-depth earthquake concentration
or clustering known as an earthquake nest is characterized with high
activity rate that is isolated from nearby activity. An earthquake nest
is different from an aftershock sequence or an earthquake swarm
because the activity rate at nests is both high and persistent over
time.
The three most famous and well-studied earthquake nests are
the Vrancea, Romania, Hindu-Kush, Afghanistan; and Bucaramanga,
Colombia nests. In this review, we discuss all three of these nests
but devote most of our attention to the Bucaramanga nest, where
we have started a thorough analysis of regional seismic data from
the Colombian seismic network. Although there may be other nests
(Pulpan and Frohlich, 1985; Sacks et al., 1967; Zari and Havskov,
2003) we will concentrate on these three in this paper.
Earthquake nests provide perhaps the best setting for understanding
the physical mechanism responsible for intermediate-depth earth-
quakes. They represent high temporal and spatial concentrations of
earthquakes that have been recorded at teleseismic and in some cases,
regional and local distances. The magnitudes range from very small
(Mb2.0) to large (M>7.0), and they occur within a limited volume,
so that attenuation and other path corrections will have similar effects
on their waveforms. Complex tectonic and geodynamic models have
been proposed to explain the high earthquake concentration at each
earthquake nest, but less has been done on the physical mechanism
that enables brittle-like failure.
Intermediate-depth earthquake nests also pose a signicant seismic
hazard. Signicant earthquakes with ~1000 fatalities have occurred in
Romania (Mw 7.5, 4 March, 1977; Frohlich, 2006), and Hindu Kush
(Mw 6.9, 20 August, 1988; Frohlich, 2006) and signicant damage in
Bucaramanga and nearby cities (M 6.3, 27 July, 1967; Ramirez, 2004).
We do not attempt to cover the entire literature on earthquake
nests, but rather we wish to highlight some key aspects and seismolog-
ical observations of earthquake nest seismicity and how these may be
useful for constraining the physical mechanism of intermediate-depth
earthquakes. The reader may nd useful reviews on deep earthquakes
(Frohlich, 1989; Green and Houston, 1995; Green and Marone, 2002;
Houston, 2007; Kirby et al., 1996a), on earthquake nests (Zari and
Havskov, 2003), and of course an entire book dedicated to deep earth-
quakes (Frohlich, 2006).
This paper is organized as follows: First we describe candidate physi-
cal mechanisms that may explain intermediate-depth earthquake occur-
rence, then we describe earthquake nests froma general perspective and
discuss the tectonic setting of the Vrancea, Hindu Kush and Bucaramanga
nests. Next we discuss in more detail precise earthquake locations and
repeating events, aftershock productivity, GutenbergRichter statistics,
focal mechanisms, and source parameters with a special attention to
recent results from the Bucaramanga nest and how it compares to the
other nests. Finally, we discuss howall of these seismological observables
may contribute to the main goal of better dening the mechanism of
intermediate-depth earthquakes and how this may point towards future
research on earthquake nests.
2. Proposed mechanisms for intermediate depth earthquakes
In the following we outline the two primary and most widely accept-
ed mechanisms proposed for the generation of intermediate-depth
earthquakes. A third mechanism known as metastable phase transition
(Frohlich, 2006; Green and Houston, 1995; Kirby et al., 1996a) but it
is unlikely to be relevant for intermediate-depths (50300 km) and we
will not discuss it further.
2.1. Dehydration embrittlement
In this mechanism, hydrated minerals in the subducting slab
undergo phase changes to anhydrous forms, releasing uids in the
process that counteract the high normal stresses expected at depth.
The reduced effective normal stress that results is hypothesized to
enable brittle failure (Frohlich, 1989; Green and Houston, 1995; Hacker
et al., 2003; Kirby et al., 1996a,b). Rayleigh and Paterson (1965)
suggested that partial dehydration of serpentinite caused the embrittle-
ment and weakening of the samples in experiments performed at
700 C and 0.5 GPa. Multiple experiments at different pressures and
temperatures conrm that dehydration embrittlement can operate
given the availability of uids (Chollet et al., 2009; Jung et al., 2004;
Meade and Jeanloz, 1991).
43 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
The transformation of basaltic oceanic crust to eclogite provides a
likely source of uids (Kirby et al., 1996b). The distribution of earth-
quakes in Japan supports the dehydration embrittlement hypothesis,
in that it is consistent with the predicted depthof dehydration reactions
(Peacock and Wang, 1999). Probably the most compelling argument in
favor of dehydration embrittlement comes from thermo-petrological
models that are able to explain double seismic zones in a wide variety
of slab conditions (Hacker et al., 2003). Seno and Yamanaka (1996)
and Jiao et al. (2000) found that intermediate-depth earthquake fault
orientations could be explained by re-activation of faults that were
formed, and hydrated, in the outer-rise. Kiser et al. (2011) imaged
large M>7.0 intermediate-depth earthquakes using backprojection
and found that many of these large events are comprised of large
sub-events that are separated in depth along sub-horizontal faults.
They suggested a combination of dehydration that preferentially
weakens pre-existing sub-horizontal faults in the subducting crust
and dynamic triggering to explain their observations.
2.2. Thermal shear runaway instability
The shear runaway hypothesis holds that a localized shear or plastic
instability forms and leads to a runaway process due to increased weak-
ening with increasing temperature (Griggs and Handin, 1960; Ogawa,
1987; Hobbs and Ord, 1988; Frohlich, 2006; Houston, 2007; Keleman
and Hirth, 2007; John et al., 2009). In this model, localized shear pro-
duces a positive feedback between temperature dependent rheology
and shear deformation that generates viscous heating (Houston,
2007) and leads to the exponential acceleration and extreme localiza-
tion of shear strain.
Geological observations support this possibility through brittle
and viscous deformation and the formation of shear zones and
pseudotachylite (e.g., Andersen et al., 2008; John and Schenk, 2006;
John et al., 2009). Numerical experiments also provide support for
this mechanism (Kelemen and Hirth, 2007; John et al., 2009). A
hybrid model invokes a twist in which local perturbations to material
properties, perhaps due to the presence of a dehydrating phase, initi-
ate a process that leads to a thermally induced shear runaway (John
et al., 2009).
Seismological evidence for a possible thermal shear runaway asso-
ciated with partial melting was presented by Kanamori et al. (1998).
Estimates of temperature rise for the 1994 M8.2 deep Bolivia earth-
quake suggest that melting would occur if fault zone thickness is
less than 0.3 m. Based on estimates of radiation efciency (see also
Venkataraman and Kanamori, 2004) the authors conclude that a large
amount of the available energy was dissipated, perhaps by melting. As
pointed out by Andersen et al. (2008) the geological evidence of
pseudotachylite formation even for small faults at intermediate-depths
suggests that very high stress drops (200500 MPa) are observed and a
signicant amount of energy is converted to heat during faulting.
3. Earthquake nests in the Earth
An earthquake nest is a volume of intense seismic activity that is
isolated from nearby activity (Richter, 1958). Nests are distinct from
aftershock sequences because they have a persistent high activity
rate while an aftershock sequence has high activity rate after the
main-shock but that dies down over a period of time (Pavlis and
Das, 2000). Nests are also distinct from earthquake swarms, because
even though earthquake swarms are concentrated areas of intense
seismicity, they are localized in time. Earthquake nests, on the other
hand, persist for (at least) decades.
The denition of what represents an earthquake nest is somewhat
arbitrary. There are many regions where clustering of seismic activity
is observed, many times associated with subducting slabs, that may be
categorized as earthquake nests. Zari and Havskov (2003) suggest
that a seismic nest is dened by stationary seismicity within a volume
that is substantially more active than its surroundings.
There are strong similarities and strong differences between the
seismic nests discussed in this contribution so that a single set of
criteria to dene seismic nests is not available. In fact, other seismic
nests have been suggested in the past (see for example an attempt
to evaluate some of them in Zari and Havskov, 2003) but very little
work has been published on these clusters.
So, in conclusion even though we will use the term earthquake
nest throughout the paper, this term represents a particular kind of
intermediate-depth clustering of seismicity that has been observed
for a long time. Other possible nests may be found or conrmed as
global and regional networks improve. Also, the use of the term
nest here does not preclude other clustering as being relevant and
useful for understanding the mechanism of intermediate-depth
earthquakes.
It was probably Santo (1969a, 1969b) who rst used the term
seismic nest to refer to the intermediate depth earthquake cluster in
Bucaramanga (Schneider et al., 1987). The term nest was used to
distinguish it from earthquake swarms or aftershock sequences. At
that time, it was evident that at least for the Bucaramanga nest, the
seismicity was isolated in space, surrounded by areas of much lower
seismic activity rates (Frohlich, 1989).
Several other intermediate-depth clusters have been called earth-
quake nests and have been compared to the Bucaramanga nest. The
two most well-known of these earthquake nests are the Hindu
Kush, Afghanistan (Chatelain et al., 1980) and the Vrancea, Romania
(Oncescu, 1984; Oncescu and Trifu, 1987). Although other earth-
quake nests have been suggested in the literature (Socampa Nest in
Peru, Sacks et al., 1967; the Iliamma Cluster beneath Cook Inlet, Alaska,
Pulpan and Frohlich, 1985; Fiji, Argentina-Chile border region and
Ecuador, Zari and Havskov, 2003), the Vrancea, Hindu Kush, and
Bucaramanga nests will be used throughout this paper for discussion
because a signicant literature by multiple authors exists for each. In
this paper much attention is centered in the Bucaramanga nest due to
local seismic data provided by the Colombian Seismic network (RSNC).
Fig. 1 shows the location of the three prominent seismic nests and
lists some of the most signicant features for comparison. The numbers
shown about their seismic activity correspond to locations from the
ISC catalog (International Seismological Centre, 2001) and additional
references (see gure caption).
The Vrancea seismic nest is located around 45.7N and 26.5E,
with an areal extend of around 2050 km and depths between 70
and 180 km (Sperner et al., 2001). Signicant earthquakes have been
reported (e.g., Bse et al., 2009; Oncescu et al., 1999; Oth et al., 2008,
2009) with magnitudes of Mw 7.4 (1940), Mw 7.7 (1977), and Mw7.1
(1986), in some cases with more than 1000 fatalities (Oncescu et al.,
1999). It is probably one of the best-studied cases due to its proximity
to a large urban area.
The Hindu Kush nest is broadly associated with the collision of the
Indian and Eurasian plates and shows an S-shaped form (Pegler and
Das, 1998). It is the most active of the three nests discussed here.
Santo (1969b) rst suggested that the nest had dimensions of about
30 km at a depth of around 215 km and multiple researchers have
proposed widely different tectonic interpretations (see Section 4.2).
This nest is also capable of signicant earthquakes M>7.0, with at
least 15 earthquakes in the last century, some of them with over
1000 fatalities (Frohlich, 2006).
The Bucaramanga nest is a unique concentration of seismic
activity with a depth concentration around 160 km below the Earth's
surface and with approximately one M4.7 earthquake per month,
near the city of Bucaramanga, Colombia (Trygvasson and Lawson,
1970; Pennington et al., 1979; Schneider et al., 1987, and many
others). Compared with other concentrations of intermediate-depth
earthquakes, the Bucaramanga nest has a higher activity rate, a smaller
source volume, and a clearer isolation from nearby activity (Schneider
44 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
et al., 1987). Zari and Havskov (2003) showed that the Bucaramanga
nest has at least ve times more events per unit volume compared to
other nests. It is believed (Ramirez, 2004) that the Bucaramanga
nest has experienced at least a M6.3 earthquake according to the ISC,
although it was Mb 6.0 according to Cortes and Angelier (2005), in
the last century, leading to signicant damage in Bogota and nearby
cities.
As suggested in Fig. 1, the Bucaramanga nest represents the
highest concentration of intermediate-depth seismicity in the world.
Even though the Hindu Kush area presents a larger number of events,
the Bucaramanga nest has the largest number of earthquakes per unit
volume of the three as can be seen from the histogram inset. As we
will show later, using local earthquake data, that the concentration
and isolation from nearby activity becomes even more apparent.
4. Tectonic setting of earthquake nests
The tectonic setting of each of the nest is complex and in all cases
there is debate on important aspects of the tectonic setting, specically
whether subduction or delamination is involved, or if plate collision
at depth is responsible for the high activity rates. Below we briey
describe the general tectonic setting and discuss some of the proposed
models for each of the three regions. This is not intended to be compre-
hensive and the reader is encouraged to consult the literature in each
case for further information.
4.1. Vrancea nest
The Vrancea region is located in the SE-Carpathian mountain
system and is part of the greater Alpine fold-and-thrust system
(Koulakov et al., 2010). The tectonic history of the Carpathians is
associated with the retreating subduction of the Tethys Ocean towards
the SWW (Csontos, 1995; Sperner et al., 2002; Stampi and Borel,
2002). Subduction ceased in the northern part about 1214 Ma ago
with the arrival of the buoyant continental East European lithosphere
into the subduction zone and later continued towards the SE (Jiricek,
1979). The progression of subduction from NW to SE is corroborated
by foredeep depocenter ages (Meulenkamp et al., 1997) and systematic
decrease in age of volcanic activity (Linzer, 1996; Pcskay et al., 1995;
Szakacs and Seghedi, 1995). Plate convergence in the northern and
eastern Carpathians seems to be currently inactive, while in the SE
Carpathians, where the Vrancea nest is located, active processes are
still taking place. The Vrancea region thus marks the youngest part of
the subduction/collision along the Carpathians.
Fig. 2 shows the location of seismicity in the Vrancea region based
on the ISC catalog (20002010). The map view shows the highly
concentrated seismicity of the Vrancea nest over an area approxi-
mately 3070 km. The NW-SE cross-section shows almost vertical,
nger-shaped intermediate-depth seismicity that is spatially separated
from the shallow activity.
Sperner et al. (2001) proposed one of the most well-accepted inter-
pretations of the tectonics inthe SE Carpathianregion (e.g., Martinet al.,
2006; Wenzel et al., 2002). Based on seismic tomography, earthquake
locations, and stress patterns Sperner et al. suggested that there is a
clear separation between the shallow crust and the subducted slab,
either by slab break-off or delamination. This decoupling is complete
in the northern and eastern Carpathians where subduction ceased
while in the Vrancea region (Fig. 2) it is in the process of detachment
and some coupling is still present. Fig. 3 illustrates the slab break-off
model of Sperner et al. (2001) where slab segments are detached in
the northern Carpathians while in the southern part, slab segments
may still be mechanically coupled to the shallower lithosphere. This
model explains the absence of intermediate-depth seismicity in the
northern and eastern Carpathians, due to diminished slab-pull once
detachment was completed.
Previous tomographic studies (Fan et al., 1998; Fuchs et al., 1979;
Koch, 1985; Lorenz et al., 1997; Oncescu, 1982, 1984; Popa et al., 2001;
Wortel and Spakman, 1992) had been performed based on regional
and local seismicity. More recent tomographic results (Koulakov et al.,
2010; Martinet al., 2006) based onteleseismic andlocal data respectively
have conrmed the presence of a nearly vertical high-velocity region
extending from about 60 to more than 200 km depth (Martin et al.,
2006 suggest that the high-velocity region extends to 350 km depth).
The geodynamic model of Martin et al. (2006) based on
teleseismic tomography suggests a high-velocity volume that they
interpret as a subducted, but not fully detached, slab. The slab hosts
the Vrancea nest where it is coupled with the continental block and
is aseismic where it is not.
Koulakov et al. (2010) do not interpret the presence of this
high-velocity body as a slab break-off, but rather as delamination of
continental material. Their interpretation is as follows: Due to conti-
nental collision, lithospheric thickening leads to higher P-T conditions
and the formation of eclogite from the mac crust and upper mantle.
Once a critical mass of eclogite is concentrated, the denser layer
begins to drop and may produce the high stress conditions that drive
seismic deformation in the Vrancea nest. Koulakov et al. (2010) suggest
that slab break-off may not be able to explain the presence of seismicity
in the central part (highest velocity anomaly) of the interpreted slab.
Fig. 1. Location map of the Vrancea, Hindu Kush and Bucaramanga nests. Insets list relevant features for each based on the ISC catalog from 2000 to 2010. Extension and depth range
of each nest are taken from Sperner et al., 2001 (Vrancea), Nowroozi, 1971 (Hindu Kush) and Schneider et al., 1987 (Bucaramanga). Number of earthquakes with M>4 is listed
based on ISC data from 2000 to 2010. Histogram shows seismicity depth concentration for each nest.
45 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
4.2. Hindu Kush nest
The Hindu Kush and Pamir areas are located in the western
syntaxis of the IndiaEurasia collision zone. The tectonic history
(Windley, 1988) is mainly associated with the closure of the Tethys
Ocean, the northward accretion of multiple plates and the nal colli-
sion with the Indian plate along the Indus suture zone (Dewey and
Bird, 1970; Gansser, 1966, 1977). After collision, signicant crustal
shortening has taken place as the Indian plate indents into Eurasia
(DeMets et al., 1990; Pegler and Das, 1998).
Since the 1970s, the tectonic arrangement that would explain the
distribution of earthquakes in the Pamir Hindu Kush has been under
debate. The main feature clearly observed since Nowroozi (1971)
has been an S-shaped conguration of the intermediate-depth
seismicity. Billington et al. (1977) used earthquake locations and
focal mechanisms to suggest that a preexisting oceanic lithosphere
subducted due to converging Indian and Eurasian plates. They identi-
ed a contorted (S-shaped) Benoiff zone with a north-dipping zone in
the Hindu Kush and a south-east dipping zone in the Pamir area. They
interpreted the contorted volume as a single slab but did not discount
the possibility of two opposing subduction zones.
In contrast, Chatelain et al. (1980) and Roecker et al. (1980) using
data from microseismic studies concluded that two subducted slabs
with opposing directions were involved in the region. Later studies
have reached similar conclusions (Burrman and Molnar, 1993; Fan
et al., 1994; Hamburger et al., 1992), although it must be noted that
in some cases it has been the presence of seismic gaps that has lead
to the idea of a two slab model. More recently, using teleseismic
tomography, seismicity and thermo-kinematic numerical modeling
Negredo et al. (2007) also suggested a two-slab model. Lou et al.
(2009) relocated seismicity using hypoDD (Waldhauser and Ellsworth,
2000) and proposed a two-slab model, but with a collision of the slabs
at about 130 km depth.
A thorough relocation of more than 6000 earthquakes in the area
in combination with focal mechanisms led Pegler and Das (1998) and
later Pavlis and Das (2000) to support the initial model of Billington
et al. (1977). Pegler and Das (1998) suggest that the seismicity and
the alignment of the T-axes along the Pamir area with the contorted
subduction can be explained with a single S-shaped seismic zone.
Similar to what is observed in Vrancea, the lack of seismicity above
70 km depth in the Pamir region suggests that the slab has become
decoupled from the surface.
Fig. 4 illustrates the seismicity features in the Hindu Kush, using
well-located earthquakes from the ISC catalog between 2000 and
2010. Both overall seismicity and depth sections clearly show the
S-shaped Benioff zone (or zones). As previously suggested in the
Fig. 2. Vrancea area seismicity based on the 20002010 ISC catalog. Left panel shows the regional location of the seismicity (color coded by depth) and the areal extent of the
Vrancea nest. Right panel shows a cross section between A-A showing the vertical extend of the Vrancea nest and its apparent separation from shallow seismicity. Only well
located events, based on low rms arrival time residuals and number of reporting stations are shown.
Fig. 3. Model for slab break-off around the Carpathian arc. The subducted slab is already detached from the continental lithosphere while the SE segment is still mechanically
coupled and hosts the Vrancea earthquake Nest (From Sperner et al., 2001 with permission by Wiley).
46 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
literature, the cross sections across the Hindu Kush and Pamir regions
show opposing dips. It is also evident that the maximum depth of
seismicity is different in the two regions.
A recent work by Lister et al. (2008) suggests that the Hindu Kush
nest seismicity is associated with slab break-off, similar to some sugges-
tions for the Vrancea nest. Using the global CMT solutions, the authors
suggest that the intermediate-depth seismicity in the area is associated
with slab beak-off, ductile faults and shear zones, similar to what is
observed at a much smaller scale in gneisses. Their model is focused on
the Hindu Kush area and more specically on the reduced and thinned
seismicity at around 150 km (see cross-section A-A) and does not
discuss in detail whether one or two slabs are present.
4.3. Bucaramanga nest
The tectonic setting inthe regionsurrounding the Bucaramanga nest
is complicated and there are a variety of competing models to explain
the active tectonics of the region (Christeson et al., 2008; Cortes and
Angelier, 2005; Higgs, 2009; Pennington et al., 1979; Pindell and
Kennan, 2009; Suter et al., 2008; Taboada et al., 2000; van der Hilst
and Mann, 1994; Zari et al., 2007). The models differ to such a degree
that it is difcult to say with certainty which slab, or slabs, the
Bucaramanga nest seismicity is associated with. The seismicity could
occur within a subducted portion of the Caribbean plate, within the
Nazca plate, or perhaps at the interaction/collision of two plates
(Zari et al., 2007).
Northern Colombia is located at the intersection of three plates:
the Nazca to the west, the Caribbean to the north and the South
American. Several models (Cortes and Angelier, 2005; Pennington,
1983; Taboada et al., 2000) suggest that a portion of the Caribbean
plate is subducting southeastward, and the Bucaramanga nest is located
within it. In these models, the Nazca plate also subducts eastward but to
the south of the Bucaramanga nest. Another model proposed by van der
Hilst and Mann (1994) suggests that the Bucaramanga nest is located in
the Nazca plate in a segment they call the redened Bucaramanga
slab. A third model suggested recently by Zari et al. (2007) based on
locations and focal mechanisms of Bucaramanga nest earthquakes sug-
gests that the collisionbetweenthe Nazca andCaribbeanplates at depth
is responsible for the Bucaramanga nest seismicity.
Fig. 5 shows the general seismicity in northwestern South America.
From seismicity located by the RSNC (Colombian seismic network)
one can easily dene two distinct regions, one on the south (south of
Fig. 4. Seismicity in the Hindu Kush Nest area. Middle panel shows overall seismicity color coded by depth. Red box shows area plotted in the right panel rotated 30 clearly showing
the S-shaped nature of subduction in the area as well as the different maximum depths of the Hindu Kush and Pamir areas. Red lines (A-A and B-B) correspond to the
cross-sections shown in the left panel. Note how in the western cross-section (Hindu Kush) subduction dips towards the north, while in the Pamir region (B-B) subduction
dips towards the SSE.
Fig. 5. Seismicity in northern South America based on the RSNC catalog color-coded by depth. Map view shows major plate boundaries. Seismicity shows distinct behavior north
and south of about 5N, which some authors associate with a continental tear (C. Vargas, pers. comm., 2011) Right panels show two cross-sections along the Nazca subduction zone
(B) and the interpreted Caribbean subducting slab (A), where the Bucaramanga nest is located.
47 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
about 5N) and one to the north. Both regions show a dipping Benioff
zone, but it is not clear whether these two regions are connected
(similar to the S-shaped zone in Hindu Kush) or if they represent two
different subducting plates.
It may seem surprising, but the Bucaramanga nest is not easily
discerned from the seismicity shown in map view or in cross section
(Fig. 5, cross section A). If the Bucaramanga nest is the highest
concentration of intermediate-depth seismicity, shouldn't it be obvious
from the seismicity? Fig. 6 shows the location of the Benioff zone based
on the seismicity of Fig. 5. It also shows (bottom panel) a 3D isosurface
map representing the number of earthquakes per unit volume (number
of earthquakes per 555 kmvolume). The Bucaramanga nest is now
clearly marked by the volume with a density of over 2000 earthquakes,
while the density in other parts of the country is well below 20 earth-
quakes. The Bucaramanga nest is interpreted in many cases as occurring
on the subducting Caribbean plate (Cortes and Angelier, 2005;
Pennington et al., 1979; Schneider et al., 1987; Taboada et al., 2000),
but there are few events above 70 km depth (see Fig. 5, cross-sections
A) that may suggest that, similar to what is observed in Vrancea or
Hindu Kush (Lister et al., 2008), the slab is mechanically detached or
is in the process of break-off.
In contrast to Zari et al. (2007) the earthquake locations and
Benoiff zones (Figs. 5 and 6) do not support a collision between two
subducting plates. Instead it supports the model of Taboada et al.
(2000) and Cortes and Angelier (2005) in that the nest is associated
with a subducting Caribbean plate, although the data does not provide
information on whether the plate is in the process of slab break-off.
5. Location of earthquakes within the Bucaramanga nest
Trygvasson and Lawson (1970) used seismicity located in the
1960s and concluded that in principle the distribution of hypocenters
was co-located, and suggested a volume of about 10 km in extent.
Similarly Santo (1969a, 1969b) were not able to distinguish the
Bucaramanga nest froma point source. Later, portable seismic networks
were deployed in 1976 and 1979, which recorded 27 nest earthquakes
(Penington, 1981; Pennington et al., 1979; Schneider et al., 1987). Using
89 well-located events (Schneider et al., 1987) suggested that the
volume of the Bucaramanga nest is 844 km. Both Pennington et
al. (1979) and Schneider et al. (1987) using these local deployments
were able to show that the seismicity actually came from a volume
and not from a point source.
Frohlich et al. (1995) relocated Bucaramanga nest earthquakes
using teleseismically determined phases (including depth phases)
and estimated a 131812 km volume. Zari et al. (2007) using
local earthquake data suggest an elongated Bucaramanga nest with
major axes of 24 km and 15 km width. But neither of the relocation
studies so far has found planar features in the Bucaramanga nest.
Recent results of intermediate-depth earthquakes elsewhere suggest
that seismicity may align on sub-horizontal (or sub-vertical) planes
(Kiser et al., 2011; Warren et al., 2007, 2008).
Using the data from the Colombian national seismic network
(RSNC) we perform earthquake relocations with a double-difference
algorithm (Waldhauser and Ellsworth, 2000). We started from catalog
phase picks from the national calatog and were reviewed manually.
Waveform cross-correlations were also used to improve the selected
picks. Most seismic stations have both P and S wave picks and in some
cases additional arrival times were picked during processing. Fig. 7
shows relocated Bucaramanga nest earthquakes M>3.5 using 15
regional seismic stations from the RSNC. The error ellipses shown in
Fig. 7 are estimated using a bootstrap resampling technique (Prieto
et al., 2007; Shearer, 1997; Waldhauser and Ellsworth, 2000).
The high-quality locations have average errors of about 2 km
horizontally and 1 km vertically, as the long axis is generally on a
horizontal plane. Fig. 7 suggests that the Bucaramanga nest is not a
volume of seismic activity, but that it shows planar features, some
sub-horizontal and some sub-vertical. This observation, if corroborated
with waveform cross-correlations and more dedicated deployments
could prove fundamental for better understanding the physical mecha-
nismand structural features associated with seismic nests. These results
also suggest that the volume of the nest is likely larger than what was
proposed by Schneider et al. (1987) but it has internal structures and
planar features that were not evident until now.
5.1. Repeating earthquakes
Earthquake nests have seismicity, which is highly localized in
space and has the tendency to have earthquakes that occur repeatedly
at the same or almost the same location. The question that arises is
whether a signicant number of these earthquakes represent repeated
rupture of the same fault plane. This feature has certainly beenobserved
at shallowdepths (Marone et al., 1995; Poupinet et al., 1984; Schaff and
Beroza, 2004; Schaff and Richards, 2004; Schaff et al., 1998; Uchida et
al., 2007) and may be present for deep moonquakes as well (Frohlich
and Nakamura, 2009).
Typically, repeating earthquakes can be identied because they
exhibit nearly identical seismograms at common stations, suggesting
nearby locations, similar focal mechanisms and rupture of the same as-
perity or patch (see for example Uchida et al., 2007). As far as we know,
there have not been any systematic searches for repeating earthquakes
in intermediate-depth earthquake nests or even for intermediate-depth
earthquakes. An example of repeating deep earthquakes was presented
by Wiens and Snider (2001), where they found clusters of M4-5 earth-
quakes with identical waveforms in the Tonga slab (560600 km
depth). Zhang et al. (2008) searched for doublet or multiplet clusters
of similar waveforms at the Bucaramanga nest for studying the rotation
of the inner core and found a signicant number of similar waveforms
that they actually designated as a 9-event multiplet.
As pointed out by Houston (2007), the detection of similar wave-
forms does not directly mean that the events are rupturing the same
Fig. 6. Benioff zone and earthquake density in northern South America. Top panel shows
the estimatedtop of the Benioff zone based on the seismicity inFig. 5. Bottompanel shows
the Benioff zone (transparency of 40%) and isosurface volumes representing different
number of earthquakes (20, 200, 2000) on a small volume (555 km). A clear yellow
iso-surface represents a region with over 2.000 earthquakes inside the small volume
and where the Bucaramanga nest is located, clearly isolated from any signicant nearby
seismicity.
48 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
fault patch. Evenusing precise locations, uncertainties are at least 12 km
(see Fig. 7) meaning earthquakes may be occurring on sub-parallel fault
planes. Nevertheless, the observation of possible repeating events
and their precise location is key for constraining the physical mechanism
involved in intermediate-depth and deep earthquakes. Wiens and Snider
(2001) interpret the repeating earthquakes they found as rupturing
the same fault patch, and suggest that these earthquakes are due to
a thermal shear runaway process. Mechanisms involving dehydration
embrittlement or phase transformations might not be expected to foster
repeating earthquakes.
We have recently started a systematic search for repeating earth-
quakes in the Bucaramanga nest using the local seismic network
(RSNC). We have found a signicant number of very similar waveforms
(correlation coefcient CC>0.9) at multiple stations. Waveform
Fig. 7. Relocated Bucaramanga Nest earthquakes. E-W and N-S cross sections of relocated seismicity and error ellipses using hypoDD (Waldhauser and Ellsworth, 2000).
Sub-horizontal features of concentrated activity are evident from the relocations and within errors at least three distinct regions (160, 151 and 148 depths) can be identied.
Fig. 8. Similar and reverse polarity waveform records for repeating Bucaramanga nest earthquakes (yellow star). Records at multiple stations of similar waveforms (CC>0.9 in at
least 5 stations) with respect to a master event (ID 63214) are shown. For each panel green waveforms represent waveforms with reverse polarity (w.r.t. the master event, large
negative CC), while red waveforms have equal polarity. Bottom signals (thick waveforms) represent the stack of the reverse and normal polarity waveforms (note that green stack
has been ipped for comparison purposes). Extremely precise locations are required to determine whether the events are occurring on the same faults, or whether the reversed
polarity events occur on faults sub-parallel to the normal polarity earthquakes.
49 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
similarity holds for at least 15 s after P and S wave arrivals, suggesting
that these earthquakes occur at very short separation distances from
each other and that focal mechanism and source complexity are similar.
Perhaps more interesting, however, we found also reversed polarity
waveforms, that show the same waveform complexity, but with the
polarity of ground motion reversed (Fig. 8). The presence of reversed
polarity earthquakes has been previously observed in volcanic regions
as well as in deep moonquakes (Frohlich and Nakamura, 2009). The
presence of reversed polarity earthquakes is also in agreement with
the observed variation of focal mechanisms in the Bucaramanga nest
(Cortes and Angelier, 2005; Frohlich and Nakamura, 2009; Frohlich
et al., 1995). We have made a careful quality control and in Fig. 8 pres-
ent those events where abs(CC)>0.9 is observed in at least 5 stations
(negative CC for reverse polarity events). The DD relocations of the
repeating events show that reversed polarity earthquakes are located
on sub-parallel faults with respect to normal polarity ones, and are
separated by 12 km. Unfortunately this distance is similar to the
uncertainties in relocating these events, so it is not possible to conclude
whether these are repeating and anti-repeating earthquakes, but this
is clearly a topic for future research.
6. Temporal behavior
6.1. Aftershock sequences
Trygvasson and Lawson (1970) analyzed seismicity in the
Bucaramanga nest between 1962 and 1968 and noted that there didn't
seem to be any evidence of swarm, aftershock or foreshock sequences.
The behavior of aftershock sequences for intermediate-depth and
deep earthquakes is still under debate, mainly because the number of
aftershocks is small at these depths, and in just a few cases have 20 or
more aftershocks been reported (Frohlich, 2006). There are some
examples of very large deep earthquakes with no aftershocks at all
(Frohlich, 2006). There are other examples with well-developed after-
shock sequences (Wiens et al., 1997) such as the Bolivian 1994 M8.3
earthquake where a signicant number of aftershocks were detected
(Myers et al., 1995).
In the three earthquake nests variable aftershock productivity is
observed. Pavlis and Hamburger (1991) presented strong evidence
for aftershock sequences in the Hindu Kush area. They studied 40
earthquakes (M>5.5) and found that only 3 of these earthquakes
presented clear aftershock productivity. The three events that showed
aftershock sequences had very similar locations, focal mechanisms
and magnitudes and seem to follow an Omori law with p ~1.0.
In the Vrancea nest, very clear aftershock sequences have been
detected after the major M ~7.0 earthquakes in 1977, 1986 and 1990
(Enescu et al., 2008; Fuchs et al., 1979). Even though the number of
aftershocks is far less than for shallow sequences, Enescu et al. were
able to detect more than 300 aftershocks M>2.8 for all three examples
and showed that at least for the 1990 sequence, aftershock productivity
followed an Omori law with p=0.83.
In the Bucaramanga nest aftershock productivity has not been
reliably detected. Trygvasson and Lawson (1970) found that the
seismicity in the Bucaramanga nest followed a Poisson distribution
in time. Frohlich et al. (1995) used the ratio test (e.g., Van der Elst
and Brodsky, 2010), where one takes the ratio of the time of a preced-
ing event and the time of the event following the mainshock, and
found that the ratios obtained were undistinguishable from a Poisson
distribution. No evidence of aftershock sequences was observed.
Fig. 9 shows aftershock productivity for the three earthquake
nests, as discussed above, taken from Pavlis and Hamburger (1991)
and Enescu et al. (2008). For the Bucaramanga nest we take the cata-
log from the RSNC and use the method proposed by Davis and
Frohlich (1991) by aligning all earthquakes with M
L
>4.5 at time
zero and counting the number of events after the mainshock time.
As can be observed and compared to the results in Hindu Kush and
Vrancea, the Bucaramanga nest has little evidence of aftershock
productivity. It is possible that this productivity is masked by the
large background seismicity and the fact that there haven't been
any large (M>6.5) earthquakes in the Bucaramanga nest.
7. Earthquake source
7.1. Size distribution (b-values)
Both shallow and deep earthquakes follow GutenbergRichter
statistics, i.e., that over a given area or volume, the number of small
earthquakes compared to the large ones ts a power low distribution
of the form:
logN abM;
Fig. 9. Aftershock decay properties for the Hindu Kush, Vrancea and Bucaramanga
nests. Data for Hindu Kush (top panel) taken from Pavlis and Hamburger (1991),
each color symbol represents one of the three earthquakes that showed clear
aftershock-like behavior. Vrancea data (middle panel) from Enescu et al. (2008). The
RSNC catalog is used for the Bucaramanga nest data (bottom panel). Each panel
shows the minimum magnitude of events based on estimated catalog completeness.
The Bucaramanga nest shows no clear Omori-like behavior, while the other two
nests show aftershock sequences. Note the large value for background seismicity for
the Bucaramanga nest.
50 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
where N is the number of earthquakes of magnitude M, a is a constant
which describes the productivity of earthquakes, and b (sometimes
known in the literature as b-value) gives the fall-off rate in the
number of large earthquakes compared to the smaller ones. The
frequencymagnitude relation was originally dened in this manner,
but it is commonly applied with M representing the cumulative
number of earthquakes of that magnitude or greater.
It is clear that a values for the Hindu Kush nest are higher than
those in the Vrancea and Bucaramanga nests (Fig. 1); there is a larger
total number of earthquakes (more productivity) in Hindu Kush. The
b-value is a different story, and based on global catalogs of both deep
and shallow earthquakes the b-value is close to 1.0 (Frohlich and
Davis, 1993; Kagan, 1999). A b-value of 1.0 indicates that there are
10 times more earthquakes for every unit magnitude decrease. If we
assume b=1 in the Hindu Kush nest and we have 549 M>4 earth-
quakes in the ISC catalog, we would expect about 5500 M>3 and
only 55 M>5 earthquakes (see Fig. 10 below).
Spatial and temporal variations of b-values for shallow and deep
earthquakes have been reported (Frohlich and Davis, 1993; Trifu
and Radulian, 1991), but it is more evident for spatial variation with
deep earthquakes (e.g., Frohlich, 1989; Frohlich and Davis, 1993;
Giardini, 1988). A very signicant and robust difference is observed
between the slopes of deep earthquakes at Tonga-Karmadec and
South America (e.g., Frohlich, 2006; Houston, 2007). Some authors
have interpreted the spatial variations of b-values as a consequence
of the thermal parameters of slabs (Wiens, 2001; Wiens and Gilbert,
1996) or shear stresses near the source region (Amitrano, 2003;
Scholz, 1968; Schorlemmer et al., 2005; Wiemer and Wyss, 2002;
Wyss, 1973) although there doesn't seem to be any such relationship
with stress drops of small earthquakes in southern California (Shearer
et al., 2006).
Fig. 10 shows the frequencymagnitude relationship for the three
earthquake nests. For the Vrancea nest the b-values is 1.15, slightly
larger than previous estimates between 0.5 and 1.0 (Mantysniemi
et al., 2003; Mrza et al., 1991; Trifu and Radulian, 1991; Zari and
Havskov, 2003) and for the Hindu Kush nest it is 0.95. Drakopoulos
and Srivastava (1972) and Zari and Havskov (2003) report values
of 1.4, while Gutenberg and Richter (1954) and Chouiian and
Srivastava (1970) report 0.6 for the b-values for this nest.
What is more striking in Fig. 10 is the distinct behavior of the
frequency-magnitude statistics for the Bucaramanga nest. Using ISC
data the b-value of 1.35 is larger compared to the other two nests
and is slightly smaller than previous studies. Frohlich et al. (1995)
reported a value of 2.0, while more recently Frohlich and Nakamura
(2009) obtain 1.6.
The ISC data for the Bucaramanga nest (Fig. 10, right panel) shows
a dip in the number of earthquakes of large magnitude (data does not
follow a straight slope), but the local RSNC data show a consistent
estimated b-value of 1.05. It is evident that there is a discrepancy
between local magnitude and ISC magnitudes, and this is evident as
the largest ISC earthquake in the period of interest (20002010) is a
magnitude 5.2, while for the local RSNC data it is a M6.0. We correct
local magnitudes using earthquake magnitudes of the same event
from both catalogs and nding a linear relationship between both
magnitude estimates (e.g., Scordilis, 2006). The corrected M
L
shows a
magnitude frequency relationship with a constant slope and a b-value
of 1.6, similar to recent results presented by Frohlich and Nakamura
(2009) for the Bucaramanga nest. It is important to notice that the
much higher b-value of the Bucaramanga nest is both observed using
the ISC and the corrected local magnitudes.
7.2. Focal mechanisms
Focal mechanisms in the Vrancea nest have been studied extensively
(Bala et al., 2003; ConstantinescuandEnescu, 1964; OncescuandBonjer,
1997; OncescuandTrifu, 1987; Radulian et al., 2000; Telesca et al., 2011;
Trifu, 1991). Most studies agree that the shallow crustal seismicity
shows a wide variety of plane orientations, but when only the
intermediate-depth earthquakes are compared, a clear pattern is
observed, indicating a reverse-faulting environment with vertical
extension and horizontal compression. For the larger earthquakes
(M>6.0) the prevalent fault strike is NE-SW with a few cases of
NW-SE striking fault planes (e.g., Radulian et al., 2000). Recent results
by Bala et al. (2003) and Telesca et al. (2011) agree in nding mostly
vertical T axes, and horizontal P axes. For the more than 700 combined
earthquakes studied in these two papers, there is a slight clustering of
NW-SE P axes orientation.
Chatelain et al. (1980) studied fault plane solutions in the Hindu
Kush area, and although signicant variations were observed, the
solutions showed predominantly reverse faulting with nearly vertical
T axes (e.g., Billington, et al., 1977). Pavlis and Das (2000) agree that
in the Hindu Kush region vertical T axes suggest that focal mechanisms
were aligned with a dipping slab, but found that in the Pamir area, the
T axes were horizontal and followed the trend of the contorted
S-shaped Benioff zone. This was interpreted as not being a reection
of subduction but rather of bending of a single subducting plate. Khan
(2003) has a different interpretation of focal mechanisms, and nd
mostly vertical T axes, but with a signicant scatter of compressional
stresses. Lister et al. (2008) found no obvious preferred orientation
of Hindu Kush earthquake focal mechanisms at shallower depths
Fig. 10. Frequencymagnitude relation for the Hindu Kush, Vrancea and Bucaramanga nests. Left panel shows the relationship for the three nests based on the 20002010 ISC cat-
alog, inset shows the b-value t for each region. The gray symbols are not used in tting the b-values. Right panel shows results for the Bucaramanga nest based on the ISC catalog
(Red triangles) and the local RSNC catalog. Note that the ISC uses Mb while the RSNC uses M
L
(green triangles) with a signicant discrepancy between magnitudes for events in both
catalogs. The blue symbols show corrected M
L
for the RSNC data with a resulting change in the b-value.
51 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
(b180 km) but found that between 180 and 280 km reverse-faulting
geometry dominated with sub-parallel and conjugate vertically dipping
slip directions.
The Bucaramanga nest shows highly variable focal mechanisms
(Frohlichand Nakamura, 2009; Schneider et al., 1987) and ina signicant
number of focal mechanisms, a large percentage of CLVD (compensated
linear vector dipole) components is observed (Frohlich, 2006; Zari et
al., 2007). Schneider et al. (1987) based on fault plane solutions of 59
micro-earthquakes found no discernable trend and suggested that this
may be due to ascent of uids. In contrast, using teleseismically deter-
mined mechanisms, Frohlich et al. (1995) found some variation, but a
clear tendency for P axes dipping towards the W and T axes dipping to-
wards the east. Several of the largest earthquakes in the Bucaramanga
nest have substantial CLVD components (Frohlich, 2006; see Fig. 11). It
is possible that these may be composite ruptures of multiple sub-events
as observed in the Hindu Kush nest by Kiser et al. (2011). Additionally,
previous studies observe shorter durations for large intermediate-depth
earthquakes than expected using a relationship for shallow events
(M
o
0.33
). The simultaneous (or partially simultaneous) rupture of
subevents might explain this discrepancy.
This last conclusion (Frohlich et al., 1995) is in agreement with re-
cent results by Cortes and Angelier (2005) who found a similar trend
and interpret this as due to a sinking slab that is being torn or is break-
ing off (similar to what is observed in Vrancea). Zari et al. (2007)
performed stress inversion of Bucaramanga nest earthquakes and
agree with these results in that there may be dominant down-dip ten-
sion, but suggest that the observed variability of micro-earthquake
focal mechanisms can be explained by an interaction between two
colliding slabs.
Careful review of the estimated focal mechanisms by Cortes and
Angelier (2005) still nd a signicant scatter with strikeslip, normal
and reverse mechanisms. Fromthe global CMT catalog we nd19 earth-
quakes with focal mechanism solutions, as shown in Fig. 11. Even for
the larger magnitude earthquakes, focal mechanisms inthe Bucaraman-
ga nest are highly variable. A number of focal mechanisms showstrike
slip mechanisms and some have large CLVD components. We plot the P
and T axes of each of the CMT solutions and nd that on average focal
mechanisms show a SE dipping T axis and a W dipping P axis, in agree-
ment with previous work, but signicant scatter is evident.
7.3. Stress drops and source physics
Understanding the nature and the characteristics of earthquake
rupture based on seismological observables (seismic waveforms) is
fundamental to understanding the physical mechanism responsible
for intermediate-depth earthquakes. Earthquake nests are particularly
useful for this purpose, because they provide a unique view of earth-
quake rupture at a wide range of magnitudes, but over a small source
region such that source effects can be more effectively separated from
propagation effects than would otherwise be possible.
There are many important earthquake source parameters that can
provide insight into the static and dynamic character of earthquake
rupture, including: stress drop, rupture size, rupture velocity, radiated
seismic energy and seismic efciency. Based on some of these source
parameters, Kanamori et al. (1998) argued that frictional melting
could be invoked during the rupture of the 1994 Bolivian deep focus
earthquake, which would lead to very lowestimates of seismic efcien-
cy and correspondingly high stress drops. The high stress drops and low
seismic efciencies suggest some energy dissipation during seismic
rupture, and points toward a thermal shear runaway process (John
et al., 2009). Does this behavior stand for different earthquake sizes, or
do small earthquakes at depth behave differently than the larger ones?
Do all intermediate-depth earthquakes have particularly high stress
drops and low seismic efciencies? How do these earthquakes compare
to their shallow counterparts? Earthquake nests may be the key for
answering these questions.
Few studies have investigated the scaling characteristics of
intermediate-depth earthquakes (Radulian and Popa, 1996; Gusev et
al., 2002; Oth et al., 2007, 2009, mainly for the Vrancea nest). Except
for the larger earthquakes, stress drops, radiated seismic energy or
even seismic efciency is not routinely estimated. Gusev et al. (2002)
found that 16 intermediate to large magnitude, intermediate-depth
Vrancea earthquakes followed a constant stress-drop model, except
for the larger ones, which had relatively high-stress drops (>10 MPa),
similar to previous results in the region (Radulian and Popa, 1996).
For the larger Vrancea nest earthquakes: the 1977 M7.5 earthquake
stress drops of 4.4 MPa and 90 MPa, and for the 1986 M7.1 values of
5 MPa and 3085 MPa were obtained by Radulian and Popa (1996)
and Oth et al. (2007, 2009) respectively. It is important to note that
stress-drop estimates can be subject to large uncertainties, and signi-
cant differences between estimates can be obtained for the same data
under different measurement techniques or model assumptions (see
for example uncertainty estimates of source parameters in Prieto
et al., 2007; Kane et al., 2011).
Other than in the Vrancea nest, we are unaware of any systematic
study of earthquake stress drops in earthquake nests. Fig. 12 shows
stress drops of intermediate-depth and deep earthquakes as reported
by Frohlich (2006) compared to estimated stress drops in the Vrancea
Fig. 11. Focal mechanisms of earthquakes in the CMT catalog for the Bucaramanga nest. Left panel shows the P and T axes of the focal mechanisms shown in the map as well as the
estimated principal directions of stress as estimated fromCortes and Angelier (2005) with down dip extension. Note the widely variable focal mechanisms present in the 19 earthquakes
shown in the map, with reverse, normal and strikeslip faulting on a very tight volume.
52 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
nest and our recent work on the Bucaramanga nest based on RSNC
seismic data (Lopez and Prieto, 2010; Prieto et al., 2011). The results
are similar to other earthquakes deeper than 50 km, in that the nest
earthquakes in Vrancea and Bucaramanga have higher stress drops than
their shallow counterparts, with an average stress drop >10 MPa.
Further work is needed to estimate radiation efciency, radiated seismic
energies and other dynamic source parameters that may constrain the
mechanism of these nest earthquakes.
8. Discussion
Dehydration embrittlement provides a good explanation of the loca-
tion of intermediate-depth earthquakes, especially as it is able to predict
the presence of double-seismic zones (Hacker et al., 2003; Yamasaki and
Seno, 2003). It is also a good candidate for earthquake nests because a
common constituent of subducting lithosphere, antigorite serpentinite,
is expected to dehydrate up to 250 kmdepth. Nevertheless, dehydration
embrittlement may not explain the presence of repeating ruptures along
the same fault planes as suggested here (Figs. 8 and 9).
Once a section of the subducted slab has been dehydrated, one
would not expect it to trigger a repeating earthquake. It is possible
that locations of these repeating events, or similar locations along
sub-horizontal faults as suggested in Fig. 7 are due to location accura-
cy and that in reality earthquakes occur on close but yet different fault
planes. Pre-existing fault planes along bend-related faults at the outer
rise (Ranero et al., 2003, 2005) may provide the explanation for
both sub-horizontal and repeating earthquakes. Sub-horizontal fault
planes have been observed in various locations (Warren et al., 2007,
2008) and recently have been imaged by Kiser et al. (2011) in the
Hindu Kush area for example. These sub-parallel fault planes may
concentrate volatiles from dehydration reactions elsewhere and
allow for repeating rupture along the same faults.
Thermal shear instability is another viable mechanism. This mech-
anism does allow for repeating earthquakes (see for example Wiens
and Snider, 2001) and is also consistent with high stress drops as
shown in Fig. 12 for the Bucaramanga nest. High stress drops are
expected based on geologic observations of shear zones and
pseudotachalyte formation (Andersen et al., 2008; John et al., 2009).
Partial melting and high stress drops have been inferred in particular
cases (Kanamori et al., 1998) pointing towards a thermal instability
mechanism. The high stress drops shown in Fig. 12 agree with the ob-
servations above, but are insufcient to conrm whether thermal
shear instability is responsible for those earthquakes. Estimates of
seismic efciency (not shown) for the data shown in Fig. 12 are sim-
ilar to those of the deep Bolivia earthquake (Kanamori, et al., 1998)
and agree with expected large energy dissipation during rupture for
small and large intermediate-depth earthquakes. It may be difcult
to estimate other source parameters like rupture velocity in the Buca-
ramanga nest because the largest events are around M5.0, but this is
possible in the Hindu Kush area with much higher productivity of
large M>6.5 earthquakes.
Other possible seismic observables that have not been document-
ed carefully include high-resolution tomographic Vp/Vs in the source
area. We expect to observe a Vp/Vs signature if dehydration embrit-
tlement is responsible, although care must be taken since antigorite
may have strong anisotropy and high Vp/Vs (Reynard et al., 2010).
Some results of anisotropy in earthquake nests can be found in the lit-
erature but have not been discussed here (e.g., Shih et al., 1991a,
1991b). Earthquake nests are ideal candidates for high-resolution to-
mography at the source region due to their compact nature and con-
centration of large number of events.
9. Conclusions
Some of the key observations discussed above for earthquake
nests are consistent with one or both of the proposed mechanisms
found in the literature. It is this kind of seismological observation
that may help resolve this issue in the future. Table 1 lists some of
the relevant observations in the literature and as presented in this
paper.
In particular, for the Bucaramanga nest we show for the rst time
that precise locations suggest that the nest has linear structures, per-
haps sub-horizontal planes of seismicity. By searching the recorded
seismograms we have found a signicant number of repeating earth-
quakes and more interestingly reversed polarity waveforms. Re-
versed polarity may suggest repeating rupture on the same or
sub-parallel faults but with reversed slip directions. A simple model
that may explain these observations is an extruding block model,
where a block moves faster with respect to the surrounding material,
leading to sub-parallel faults with contrary slip directions.
Previous studies and our own results showhigh stress drops in the
Vrancea and Bucaramanga nests. This may be due to small rupture
areas and/or slow rupture velocities compared to shallow earth-
quakes. To evaluate shear instability as a mechanism, a complete en-
ergy budget is needed and estimates of radiated seismic energy and
seismic efciency are required. We are working on a more complete
analysis of Bucaramanga nest earthquakes to address this issue.
As discussed in this review, seismological characterization of earth-
quake nests may provide a unique tool for constraining the physics of
rupture mechanism. More precise earthquake locations and focal mech-
anisms will be valuable, but additionally high-resolution source-region
tomography and earthquake physics studies are all needed for improv-
ing our understanding of this behavior. The results presented here
cannot conclusively favor a particular mechanism but are in the right
direction to help in deciphering this key seismological question.
Acknowledgments
We would like to thank two anonymous reviewers for thoughtful
comments on an earlier version of this paper. We thank Adrien Oth
Fig. 12. Estimated static stress drop for intermediate-depth and deep earthquakes.
Vrancea (green) and Bucaramanga nest earthquake (red) stress drop estimates have
average between 4070 and 2040 MPa respectively. Compilation of results for
intermediate-depth and deep earthquakes by Frohlich (2006) is shown for comparison
purposes (gray bars). Vrancea results (M3.57.1) are taken from Oth et al. (2009).
Magnitude range for Bucaramanga nest earthquakes is M4.05.7.
53 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
for sharing his stress drop estimates from Vrancea. We also would
like to thank the Editor for inviting us to contribute this article. This
work was partially supported by NSF Grant EAR-1045684.
References
Amitrano, D., 2003. Brittle-ductile transition and associated seismicity: experimental
and numerical studies and relationship with the b value. Journal of Geophysical
Research 108 (B1), 2044. http://dx.doi.org/10.1029/2001JB000680.
Andersen, T.B., Mair, K., Austrheim, H., Podladchikov, Y.Y., Vrijmoed, J.C., 2008. Stress-
release in exhumed intermediate-deep earthquakes determined from ultramac
pseudotachylyte. Geology 36, 995998.
Bala, A., Radulian, M., Popescu, E., 2003. Earthquakes distribution and their focal mecha-
nismincorrelation with the active tectonic zones of Romania. Journal of Geodynamics
36, 129145.
Bilek, S.L., Lay, T., Ruff, L.J., 2004. Radiated seismic energy and earthquake source dura-
tion variations from teleseismic source time functions for shallow subduction zone
earthquakes. Journal of Geophysical Research 109. http://dx.doi.org/10.1029/
2004JB003039, B00308.
Billington, S., Isacksa, B., Barazangi, M., 1977. Spatial distribution and focal mechanisms
of mantle earthquakes in the Hindu Kush-Pamir region: a contorted Benioff zone.
Geology 5, 699704.
Bse, M., Sokolov, V., Wenzel, F., 2009. Shake map methodology for intermediate-depth
Vrancea (Romania) earthquakes. Earthquake Spectra 25, 497514. http://dx.doi.org/
10.1193/1.3148882.
Burrman, V.S., Molnar, P., 1993. Geological and geophysical evidence for deep subduction
of continental crust beneath the Pamir, Spec. Pap. Geological Society of America 281,
175.
Chatelain, J.L., Roecker, S.W., Hatzfeld, D., Molnar, P.m, 1980. Microearthquake seismicity
and fault plane solutions inHindu Kush region and their tectonic implications. Journal
of Geophysical Research 85, 13651387.
Chollet, M., Daniel, I., Koga, K.T., Petitgirard, S., Morard, G., 2009. Dehydration kinetics
of talc and 10 phase: consequences for subduction zone seismicity. Earth and
Planetary Science Letters 284, 5764.
Chouiian, R., Srivastava, V., 1970. Global variation of b in the Gutenberg Richter's
relation log N=abM with depth. Pure and Applied Geophysics 124132.
Christeson, G.L., Mann, P., Escalona, A., Aitken, T.J., 2008. Crustal structure of the
Caribbeannortheastern South America arc-continent collision zone. Journal of
Geophysical Research 113, B08104. http://dx.doi.org/10.1029/2007JB005373.
Chung, W.-Y., Kanamori, H., 1976. Source process and tectonic implications of the
Spanish deep focus earthquake of March 29, 1954. Physics of the Earth and Planetary
Interiors 13, 8596.
Constantinescu, L., Enescu, D., 1964. Fault-plane solutions for some Romanian earth-
quakes and their seismotectonic implication. Journal of Geophysical Research 69,
667674.
Cortes, M., Angelier, J., 2005. Current states of stress in the northern Andes as indicated
by focal mechanisms of earthquakes. Tectonophysics 403, 2958.
Csontos, L., 1995. Tertiary tectonic evolution of the Intra-Carpathian area: a review.
Acta Vulcanologica 7 (2), 113.
Davis, S.D., Frohlich, C., 1991. Single-link cluster analysis of earthquake aftershocks: decay
laws and regional variations. Journal of Geophysical Research 96 (B4), 63356350.
DeMets, C., Gordan, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophysical
Journal International 101, 425478.
Dewey, J.F., Bird, J.M., 1970. Mountain belts and new global tectonics. Journal of
Geophysical Research 75 (14), 26252647.
Drakopoulos, J.C., Srivastava, H.N., 1972. The dependence of earthquake frequency
magnitude relationship and strain energy release upon the focal depth in Hindu
Kush region. Annali di Geosica 25 (1972), 593606.
Enescu, B., Struzik, Z., Kiyoto, K., 2008. On the recurrence time of earthquakes: insight
from Vrancea (Romania) intermediate-depth events. Geophysical Journal International
172, 395404.
Fan, G., Ni, J.F., Wallace, T.C., 1994. Active tectonics of the Pamirs and Karakoram.
Journal of Geophysical Research 99, 71317160.
Fan, G., Wallace, T.C., Zhao, D., 1998. Tomographic imaging of deep velocity structure
beneath the eastern and southern Carpathians, Romania: implications for continental
collision. Journal of Geophysical Research 103, 27052724. http://dx.doi.org/10.1029/
97JB01511.
Frohlich, C., 1989. The nature of deep-focus earthquakes. Annual Review of Earth and
Planetary Sciences 17, 227254.
Frohlich, C., 2006. Deep Earthquakes. Cambridge University Press, Cambridge.
Frohlich, C., Davis, S.D., 1993. Teleseismic b-values or, much ado about 1.0. Journal of
Geophysical Research 98, 631644.
Frohlich, C., Nakamura, Y., 2009. The physical mechanisms of deep moonquakes and
intermediate-depth earthquakes: how similar and how different? Physics of the
Earth and Planetary Interiors 173, 365374.
Frohlich, C., Kadinsky-Cade, K., Davis, S.D., 1995. A reexamination of the Bucaramanga
Nest. Bulletin of the Seismological Society of America 85 (6), 16221634.
Fuchs, K., Bonjer, K.-P., Bock, G., Cornea, I., Radu, C., Enescu, D., Jianu, D., Nourescu, A.,
Merkler, G., Moldoveanu, T., Tudorache, G., 1979. The Romanian earthquake of
March 4, 1977: II. Aftershocks and migration of seismic activity. Tectonophysics 53,
225247.
Gansser, A., 1966. The Indian Ocean and the Himalayas a geological interpretation.
Eclogae Geologicae Helvetiae 59 (831848), 1966.
Gansser, A., 1977. The great suture zone between Himalaya and Tibet a preliminary
account. Himalaya: Sciences de la Terre. Centre National de la Recherche Scientique,
Paris, pp. 209212.
Giardini, D., 1988. Frequency distribution and quantication of deep earthquakes.
Journal of Geophysical Research 93, 20952105.
Green, H.W., Houston, H., 1995. The mechanics of deep earthquakes. Annual Review of
Earth and Planetary Sciences 23, 169213.
Green, H.W., Marone, C., 2002. Instability of deformation. In: Korato, S., Wenk, H. (Eds.),
Reviews in Mineralogy and Geochemistry: Plasticity of minerals and Rocks, 51.
Blackwell, Oxford, pp. 181199.
Griggs, D.T., Handin, J.H., 1960. Observations on fracture and a hypothesis of earth-
quakes. In: Griggs, D.T., Handin, J.H. (Eds.), Rock Deformation: Mem. Geol. Soc.
Am., 79, pp. 347373.
Gusev, A., Radulian, M., Rizescu, M., Panza, G.F., 2002. Source scaling of intermediate-
depth Vrancea earthquakes. Geophysical Journal International 151, 879889.
Gutenberg, B., Richter, C.F., 1949. Seismicity of the Earth and Associated Phenomena,
2nd edn. Princeton University Press, Princeton, NJ. 310 pp.
Gutenberg, B., Richter, C., 1954. Seismicity of the Earth and Associated Phenomena.
Princeton University Press, Princeton.
Hacker, B.R., Peacock, S.M., Abers, G.A., Holloway, S.D., 2003. Subduction factory 2. Are
intermediate-depth earthquakes in subducting slabs linked to metamorphic
dehydration reactions? Journal of Geophysical Research 108, 2462724637.
http://dx.doi.org/10.1029/ 2001JB001129.
Hamburger, M.W., Sarewitz, D.R., Pavlis, T.L., Popandopulo, G.A., 1992. Structural and
seismic evidence for intracontinental subduction in the Peter the First range,
central Asia. Geological Society of America Bulletin 104 (397408), 1992.
Higgs, R., 2009. Caribbean-South America oblique collision model revised. Geological
Society of London Special Publication 328, 613657.
Hobbs, B.E., Ord, A., 1988. Plastic instabilities: implications for the origin of intermediate
and deep focus earthquakes. Journal of Geophysical Research 93, 10,52110,540.
Houston, H., 2007. Deep earthquakes. In: Schubert, G. (Ed.), Treatise on Geophysics.
Elsevier, Amsterdam, pp. 321350. http://dx.doi.org/10.1016/B978-044452748-
6.00071-7.
International Seismological Centre, 2001. On-line Bulletin. Internatl. Seis. Cent.
Thatcham, United Kingdom. http://www.isc.ac.uk.
Jiao, W., Silver, P.G., Fei, Y., Prewitt, C.T., 2000. Do intermediate- and deep-focus earth-
quakes occur on preexisting weak zones? An examination of the Tonga subduction
zone. Journal of Geophysical Research 105, 28,12528,138.
Jiricek, R., 1979. Tectogenetic development of the Carpathian arc in the Oligocene and
Neogene. In: Mahel, M. (Ed.), Tectonic Proles Through the West Carpathians.
Geol. Inst. D. Stur, Bratislava, pp. 20032214.
John, T., Schenk, V., 2006. Interrelations between intermediate-depth earthquakes and
uid ow within subducting oceanic plates: constraints from eclogite-facies
pseudotachylytes. Geology 34, 557560.
John, T., Medvedev, S., Rupke, L.H., Andersen, T.B., Podladchikov, Y.Y., Austrheim, H.,
2009. Generation of intermediate-depth earthquakes by self localizing thermal
runaway. Nature Geoscience 2, 137140. http://dx.doi.org/10.1038/NGEO419.
Jung, H., Green, H.W., Dobrzhinetskaya, L.F., 2004. Intermediate-depth earthquake
faulting by dehydration embrittlement with negative volume change. Nature 428,
545549.
Kagan, Y.Y., 1999. Universality of the seismic momentfrequency relation. Pure and
Applied Geophysics 155, 537573.
Kanamori, H., Anderson, D.L., Heaton, T.H., 1998. Frictional melting during the rupture
of the 1994 Bolivian earthquake. Science 279, 839842.
Kane, D.L., Prieto, G.A., Vernon, F.L., Shearer, P.M., 2011. Quantifying seismic source
parameter uncertainties. Bulletin of the Seismological Society of America 101 (2),
535543. http://dx.doi.org/10.1785/ 0120100166.
Keleman, P.B., Hirth, G., 2007. A periodic shear-heating mechanism for intermediate-
depth earthquakes in the mantle. Nature 446, 787790.
Khan, P.K., 2003. Stress state, seismicity and subduction geometries of the descending
lithosphere below the Hindukush and Pamir. Gondwana Research 6, 867877.
Kirby, S., Stein, S., Okal, E.A., Rubie, D.C., 1996a. Metastable mantle phase transformations
and deep earthquakes in subducting oceanic lithosphere. Reviews of Geophysics 34
(2), 261306.
Kirby, S.H., Engdahl, E.R., Denlinger, R., 1996b. Intraslab earthquakes and arc volcanism:
dual physical expressions of crustal and uppermost mantle metamorphism in
Table 1
Comparison of some observed features of the three nests studied in this study. Depth
ranges and number of earthquakes are estimated from the ISC locations. b-values
from the ISC catalog and local catalog for the Bucaramanga case. Focal mechanism as
reported in the references, Global CMT for the Bucaramanga nest. Average stress
drops from Oth et al. (2007, 2009) for the Vrancea nest and from local data for the
Bucaramanga nest.
Nest Hindu Kush Vrancea Bucaramanga
Location (lat, lon) 36.5, 71.0 45.7, 26.5 6.8, 73.1
Depth range (km) 175250 70180 145165
# M>4 (20002010) 549 50 151
b-value 0.95 1.15 1.35 (Local) 1.60
Focal mechanism Vertical T-axes
Variable P-axes
Vertical T-axes
Horiz. P-axes
Highly variable
East Dipping T-axes
Some cases CLVD
Avg. stress drop N.A. 45 MPa 30 MPa
54 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
subducting slabs, in Bebout, G.E., et al., eds., Subduction: top to bottom: American
Geophysical Union. Geophysical Monograph 96, 195214.
Kiser, E., Ishii, M., Langmuir, C.H., Shearer, P.M., Hirose, H., 2011. Insights into the
mechanism of intermediate-depth earthquakes from source properties as imaged
by back-projection of multiple seismic phases. Journal of Geophysical Research.
http://dx.doi.org/10.1029/2010JB007831.
Koch, M., 1985. Nonlinear inversion of local seismic travel times for the simultaneous
determination of the 3D-velocity structure and hypocentersapplication to the
seismic zone Vrancea. Journal of Geophysics 56, 160173.
Koulakov, I., Zaharia, B., Enescu, B., Radulian, M., Popa, M., Parolai, S., Zschau, J., 2010.
Delamination or slab detachment beneath Vrancea? New arguments from local
earthquake tomography. Geochemistry Geophysics Geosystems 10, Q03002.
http://dx.doi.org/10.1029/2009GC002811.
Linzer, H.G., 1996. Kinematics of retreating subduction along the Carpathian arc,
Romania. Geology 24, 167170.
Lister, G., Kennett, B., Richards, S., Forster, M., 2008. Boudinage of a stretching slablet
implicated in earthquakes beneath the Hindu Kush. Nature Geoscience 1, 196201.
http://dx.doi.org/10.1038/ngeo132.
Lopez, G., Prieto, G.A., 2010. Earthquake scaling of intermediate-depth earthquakes in
the Bucaramanga Nest. Abstract S21C-2065 presented at 2010 Fall Meeting. AGU,
San Francisco, Calif. 13-17 Dec.
Lorenz, F.P., Martin, M., Sperner, B., Wenzel, F., Popa, M., 1997. Teleseismic travel-
time tomography of the compressional-wave velocity structure in the Vrancea
zone, Romania. EOS Transactions American Geophysical Union 78 (46) (Fall
Meet. Suppl., F497).
Lou, X.T., Cai, C., Yu, C.Q., Ning, J.Y., 2009. Intermediate-depth earthquakes beneath the
Pamir-Hindu Kush Region: evidence for collision between two opposite subduction
zones. Earthquake Science 22, 659665.
Mantysniemi, P., Marza, V.I., Kijko, A., Retief, P., 2003. A new probabilistic seismic
hazard analysis for the Vrancea (Romania) seismogenic zone. Natural Hazards
29, 371385.
Marone, C., Vidale, J.E., Ellsworth, W.L., 1995. Fault healing inferred from time depen-
dent variations in source properties of repeating earthquakes. Geophysical Re-
search Letters 22, 30953098.
Martin, M., Wenzel, F., the CALIXTO Working Group, 2006. High-resolution teleseismic
body wave tomography beneath SE-RomaniaII. Imaging of a slab detachment
scenario. Geophysical Journal International 164, 579595. http://dx.doi.org/10.1111/
j.1365- 246X.2006.02884.x.
Mrza, V.I., Kijko, A., Mntyniemi, P., 1991. Estimate of earthquake hazard in the
Vrancea (Romania) region. Pure and Applied Geophysics 136, 143154.
Meade, C., Jeanloz, R., 1991. Deep-focus earthquakes and recycling of water into the
Earth's mantle. Science 252, 6872.
Meulenkamp, J.E., Kova c, M., Cicha, I., 1997. On Late Oligocene to Pliocene depocentre
migrations and the evolution of the CarpathianPannonian system. Tectonophysics
266 (14), 301318.
Myers, S.C., Wallace, T.C., Beck, S.L., Silver, P.G., Zandt, G., Vandecar, J., Minaya, E., 1995.
Implications of spatial and temporal development of the aftershock sequence for
the Mw 8.3 June 9, 1994 deep Bolivian earthquake. Geophysical Research Letters
22, 22692272.
Negredo, A.M., Replumaz, A., Villasenor, A., Guillot, S., 2007. Modeling the evolution
of continental subduction processes in the PamirHindu Kush region. Earth and
Planetary Science Letters 259, 212225.
Nowroozi, A.A., 1971. Seismo-tectonics of the Persian Plateau, Eastern Turkey, Caucasus,
and Hindu Kush regions. Bulletinof the Seismological Society of America 61, 317341.
Ogawa, M., 1987. Shear instability in a viscoelastic material as the cause of deep focus
earthquakes. Journal of Geophysical Research 92, 1380113810.
Oncescu, M.C., 1982. Velocity structure of the Vrancea region, Romania. Tectonophysics
90, 117122. http://dx.doi.org/10.1016/0040-1951(82)90256-6.
Oncescu, M.C., 1984. Deep structure of the Vrancea region, Roumania, inferred from
simultaneous inversion for hypocenters and 3-D velocity structure. Annals of
Geophysics 2 (1), 2328.
Oncescu, M.C., Bonjer, K.-P., 1997. A note on the depth recurrence and strain release of
large Vrancea earthquakes. Tectonophysics 272, 291302.
Oncescu, M.C., Trifu, C.I., 1987. Depth variation of moment tensor principal axes in
Vrancea (Romania) seismic region. Annals of Geophysics 5B, 149154.
Oncescu, M.C., Bonjer, K.-P., Rizescu, M., 1999. Weak and strong ground motion of
intermediate-depth earthquakes from the Vrancea region. In: Wenzel, F., Lungu, D.,
Novak, O. (Eds.), Vrancea Earthquakes: Tectonics, Hazard and Risk Mitigation. Springer,
New York, pp. 2742.
Oth, A., Wenzel, F., Radulian, M., 2007. Source parameters of intermediate-depth
Vrancea (Romania) earthquakes from empirical Green's functions modeling.
Tectonophysics 438, 3356.
Oth, A., Bindi, D., Parolai, S., Wenzel, F., 2008. S-wave attenuation characteristics beneath
the Vrancea region in Romania: new insights from the inversion of ground motion
spectra. Bulletin of the Seismological Society of America 98 (5), 24822497.
Oth, A., Parolai, S., Bindi, D., Wenzel, F., 2009. Source spectra and site response from
S waves of intermediate-depth Vrancea, Romania, earthquakes. Bulletin of the
Seismological Society of America 99 (1), 235254.
Pavlis, G.L., Das, S., 2000. The PamirHindu Kush seismic zone as a strain marker for
ow in the upper mantle. Tectonics 19, 103115.
Pavlis, G.L., Hamburger, M.W., 1991. Aftershock sequences of intermediate-depth
earthquakes in the Pamir-Hindu Kush seismic zone. Journal of Geophysical Research
96 (1810718117), 1991.
Peacock, S.M., Wang, K., 1999. Seismic consequences of warm versus cool subduction
metamorphism: examples from southwest and northeast Japan. Science 286,
937939.
Pcskay, Z., Lexa, J., Szaka cs, A., et al., 1995. Space and time distribution of Neogene-
Quaternary volcanism in the Carpatho-Pannonian region. Acta Vulcanologica 7 (2),
1528.
Pegler, G., Das, S., 1998. An enhanced image of the Pamir-Hindu Kush seismic zone
from relocated earthquake hypocenters. Geophysical Journal International 134,
573595.
Penington, W.D., 1981. Subduction of the eastern Panama basin and seismotectonics of
northwestern South America. Journal of Geophysical Research 86, 10,57910,597
(1981).
Pennington, W.D., 1983. The role of shallow phase changes in the subduction of oceanic
crust. Science 220, 10451047.
Pennington, W.D., Mooney, W.D., van Hissenhoven, R., Meyer, H., Ramirez, J.E., Meyer,
R.P., 1979. Results of a reconnaissance microearthquake survey of Bucaramanga,
Colombia. Geophysical Research Letters 6, 6568.
Pindell, J.L., Kennan, L., 2009. South America in the mantle reference frame: an update
Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America
in the mantle reference frame: an update. Geological Society of London Special
Publication 328, 135.
Popa, M., Kissling, E., Radulian, M., Bonjer, K.P., Enescu, D., Dragan, S., the CALIXTO
Working Group, 2001. Local source tomography using body waves to deduce a
minimum 1D velocity model for the Vrancea (Romania) zone. Romanian Reports
in Physics 53, 519536.
Poupinet, G., Ellsworth, W.L., Frchet, J., 1984. Monitoring velocity variations in the
crust using earthquake doublets: an application to the Calaveras Fault, California.
Journal of Geophysical Research 89, 57195731.
Prieto, G.A., Thomson, D.J., Vernon, F.L., Shearer, P.M., Parker, R.L., 2007. Condence
intervals of earthquake source parameters. Geophysical Journal International
168, 12271234. http://dx.doi.org/10.1111/j.1365- 246X.2006.03257.x.
Prieto, G.A., Lopez, G., Barrett, S., Beroza, G., 2011. Earthquake source scaling, stress
drops and radiated seismic energies of intermediate depth earthquakes. Abstract
T22A-06 Presented at 2011 Fall Meeting. AGU, San Francisco, Calif. 5- 9 Dec.
Pulpan, H., Frohlich, C., 1985. Geometry of the subducted plate near Kodiak Island and
Lower Cook Inlet, Alaska, determined from relocated earthquake hypocenters.
Bulletin of the Seismological Society of America 75, 791810.
Radulian, M., Popa, M., 1996. Scaling of the source parameters for the Vrancea interme-
diate depth earthquakes. Tectonophysics 261, 6781.
Radulian, M., Mandrescu, M.N., Panza, G.F., Popescu, E., Utale, A., 2000. Characterization
of seismogenic zones of Romania. Pure and Applied Geophysics 157 (2000), 5777.
Ramirez, J.E., Instituto Geofsico Universidad Javeriana, 2004. Actualizacin de la
historia de los terremotos en Colombia. Ponticia Universidad Javeriana, Bogot,
D. C.
Ranero, C.R., Phipps-Morgan, J., McIntosh, K., Reichert, C., 2003. Bending-related
faulting and mantle serpentinization at the Middle America trench. Nature 425,
367373.
Ranero, C.R., Villasenor, A., Morgan, J.P., Weinrebe, W., 2005. Relationship between
bend-faulting at trenches and intermediate-depth seismicity. Geochemistry Geophysics
Geosystems 6, Q12002.
Rayleigh, C.B., Paterson, M.S., 1965. Experimental deformation of serpentinite and its
tectonic implications. Journal of Geophysical Research 70, 39653985.
Reynard, B., Nakajima, J., Kawakatsu, H., 2010. Earthquakes and plastic deformation of
anhydrous slab mantle in double WadatiBenioff zones. Geophysical Research
Letters 37, L24309. http://dx.doi.org/10.1029/2010GL045494.
Richter, C.F., 1958. Elementary Seismology. W.H. Freeman, San Francisco, Calif, p. 342.
Roecker, S.W., Soboleva, V., Nersesov, I.L., Lukk, A.A., Hatzfeld, D., Chatelain, J.L., Molnar,
P., 1980. Seismicity and fault plane solutions of intermediate depth earthquakes in
the Pamir-Hindu Kush region. Journal of Geophysical Research 85, 13581364.
Sacks, I.S., Suyehiro, S., Kamitsuki, A., Tuve, M.A., Otsuka, M., et al., 1967. A tentative
value of Poisson's coefcient from the seismic nest of Socampa. Annual Report
of the Director. Carnegie Inst. Dep. Terr. Magn, pp. 4345. 1965-1966.
Santo, T., 1969a. Characteristics of seismicity in South America. Bulletin of the Earth-
quake Research Institute University of Tokyo 47, 635672.
Santo, T., 1969b. Regional study on the characteristic seismicity of the world, I, Hindu
Kush region. Bulletin of the Earthquake Research Institute University of Tokyo
47, 10351049.
Schaff, D.P., Beroza, G.C., 2004. Coseismic and postseismic velocity changes measured
by repeating earthquakes. Journal of Geophysical Research 109, B10302. http://
dx.doi.org/10.1029/2004JB003011.
Schaff, D.P., Richards, P.G., 2004. Repeating seismic events in China. Science 303, 11761178.
Schaff, D.P., Beroza, G.C., Shaw, B.E., 1998. Postseismic response of repeating after-
shocks. Geophysical Research Letters 25, 45494552.
Schneider, J.F., Pennington, W.D., Meyer, R.P., 1987. Microseismicity and focal mechanisms
of the intermediate-depth Bucaramanga Nest, Colombia. Journal of Geophysical
Research 92, 1391313926.
Scholz, C.H., 1968. The frequencymagnitude relation of microfracturing in rock and
its relation to earthquakes. Bulletin of the Seismological Society of America 58,
399415.
Scordilis, E.M., 2006. Empirical global relations converting M S and m b to moment
magnitude. Journal of Seismology 10, 225236.
Seno, T., Yamanaka, Y., 1996. Double seismic zones, compressional deep trench-outer
rise events, and superplumes. In: Bebout, G.E., et al. (Ed.), Subduction: Top to
Bottom: Geophys. Monogr. Ser., 96. AGU, Washington, D. C., pp. 347355.
Schorlemmer, D., Wiemer, S., Wyss, M., 2005. Variations in earthquake-size distribution
across different stress regimes. Nature 437, 539542.
Shearer, P., 1997. Improving local earthquake locations using the L1 norm and wave-
form cross correlation: application to the Whittier Narrows, California, aftershock
sequence. Journal of Geophysical Research 102, 82698283.
55 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256
Shearer, P.M., Prieto, G.A., Hauksson, E., 2006. Comprehensive analysis of earthquake
source spectra in southern California. Journal of Geophysical Research 111, B06303.
http://dx.doi.org/10.1029/2005JB003979.
Shih, X.R., Meyer, R.P., Schneider, J.F., 1991a. Seismic anisotropy above a subducting
plate. Geology 19, 807810.
Shih, X.R., Schneider, J.F., Meyer, R.P., 1991b. Polarities of P and S waves, and shear
wave splitting observed from the Bucaramanga Nest, Colombia. Journal of Geophysical
Research 96, 1206912082.
Sperner, B., Lorenz, F., Bonjer, K., Hettel, S., Muller, B., Wenzel, F., 2001. Slab breakoff
abrupt cut or gradual detachment? Newinsights fromVrancea Region (SE Carpathians,
Romania). Terra Nova 13, 172179.
Sperner, B., Ratschbacher, L., Nemcok, M., 2002. Interplay between subduction retreat
and lateral extrusion: tectonics of the Western Carpathians. Tectonics 21 (6),
1051. http://dx.doi.org/10.1029/2001TC901028.
Stampi, G.M., Borel, G.D., 2002. A plate tectonic model for the Paleozoic and Mesozoic
constrained by dynamic plate boundaries and restored synthetic oceanic isochrons.
Earth and Planetary Science Letters 196 (12), 1733.
Suter, F., Sartori, M., Neuwerth, R., Gorin, G., 2008. Structural imprints at the front of
the Choco-Panama indenter. Tectonophysics 460, 134157.
Szakacs, A., Seghedi, I., 1995. Time-space evolution of Neogene-Quarternary volcanism
in the Calimani-Gurghiu-Harghita volcanic chain. Romanian Journal of Strategy 76
(4), 24.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
Rivera, C., 2000. Geodynamics of the northern Andes: subductions and intracontinental
deformation (Colombia). Tectonics 19 (5), 787813.
Telesca, L., Alcaz, V., Sandu, I., 2011. The stress eld of Vrancea region from fault plane
solution (FPS). Natural Hazards and Earth System Sciences 11, 28172820.
Trifu, C.-I., 1991. The study of Vrancea earthquake mechanisms by a composite
technique. Revue Roumaine de Physique 36, 457469.
Trifu, C.-I., Radulian, M., 1991. Frequencymagnitude distribution of earthquakes in Vracea:
Relevance for a discrete model. Journal of Geophysical Research 96 (B3), 43014311.
Trygvasson, E., Lawson, J.E., 1970. The intermediate earthquake source near Bucaramanga,
Colombia. Bulletin of the Seismological Society of America 60, 269276.
Uchida, N., Matsuzawa, T., Ellsworth, W.L., Kazutoshi, I., Okada, T., Hasegawa, A., 2007.
Source parameters of aM4.8 and its accompanying repeating earthquakes off
Kamaishi, NE Japan: implications for the hierarchical structure of asperities and earth-
quake cycle. Geophysical Research Letters 34, L20313 (doi:1029/2007GL031263).
Van Der Elst, N., Brodsky, E., 2010. Connecting near-eld and far-eld earthquake
triggering to dynamic strain. Journal of Geophysical Research 115, B07311.
van der Hilst, R., Mann, P., 1994. Tectonic implications of tomographic images of
subducted lithosphere beneath northwestern South America. Geology 22, 451454.
Venkataraman, A., Kanamori, H., 2004. Observational constraints on the fracture ener-
gy of subduction zone earthquakes. Journal of Geophysical Research 109, B05302.
http://dx.doi.org/10.1029/2003JB002549.
Wadati, K., 1929. Shallow and deep earthquakes (2nd paper). Geophysical Magazine 2,
136.
Waldhauser, F., Ellsworth, W.L., 2000. A double-difference earthquake location
algorithm: method and application to the Northern Hayward Fault, California.
Bulletin of the Seismological Society of America 90, 13531368. http://dx.doi.org/
10.1785/0120000006.
Warren, L.M., Hughes, A.N., Silver, P.G., 2007. Earthquake mechanics and deforma-
tion in the Tonga-Kermadec subduction zones from fault-plane orientations of
intermediate- and deep-focus earthquakes. Journal of Geophysical Research 112
(B05314). http://dx.doi.org/10.1029/2006JB004677.
Warren, L.M., Langstaff, M.A., Silver, P.G., 2008. Fault-plane orientations of intermediate-
depth earthquakes in the middle America trench. Journal of Geophysical Research
113 (B01304). http://dx.doi.org/10.1029/2007JB005028.
Wenzel, F., Sperner, B., Lorenz, F., Mocanu, V., 2002. Geodynamics, tomographic images
and seismicity of the Vrancea region (SE-Carpathians, Romania): Stephan Mueller
Spec. Publ. Ser., 3, pp. 95104.
Wiemer, S., Wyss, M., 2002. Mapping spatial variability of the frequencymagnitude
distribution of earthquakes. Advances in Geophysics 45, 259302.
Wiens, D.A., 2001. Seismological constraints on the mechanisms of deep earthquakes:
temperature dependence of deep earthquake source properties. Physics of the
Earth and Planetary Interiors 127, 145163.
Wiens, D.A., Gilbert, H.J., 1996. Effect of slab temperature on deep-earthquakes
aftershock productivity and magnitudefrequency relations. Nature 384, 153156.
Wiens, D.A., Snider, N.O., 2001. Repeating deep earthquakes: evidence for fault
reactivation at great depth. Science 293, 14631466.
Wiens, D.A., Gilbert, H.J., Hicks, B., Wysession, M.E., Shore, P.J., 1997. Aftershock
sequences of moderate-sized intermediate and deep earthquakes in the Tonga
subduction zone. Geophysical Research Letters 24, 20592062.
Windley, B.F., 1988. Tectonic framework of the Himalaya, Karakoram and Tibet, and the
problem of their evolution. Philosophical Transactions of the Royal Society of London
A326, 316.
Wortel, M.J.R., Spakman, W., 1992. Structure and dynamics of subducted lithosphere in
the Mediterranean region. Proceedings of the Koninklijke Nederlandse Akademie
van Wetenschappen 95, 325347.
Wyss, M., 1973. Towards a physical understanding of the earthquake frequency distri-
bution. Geophysical Journal of the Royal Astronomical Society 31, 341359.
Yamasaki, T., Seno, T., 2003. Double seismic zone and dehydration embrittlement of the
subducting slab. Journal of Geophysical Research 108. http://dx.doi.org/10.1029/
2002JB001918.
Zari, Z., Havskov, J., 2003. Characteristics of dense nests of deep and intermediate-
depth seismicity. Advances in Geophysics 46, 237278.
Zari, Z., Havskov, J., Hanyga, A., 2007. An insight into the Bucaramanga nest.
Tectonophysics 443, 93105.
Zhang, J., Richards, P.G., Schaff, D.P., 2008. Wide scale detection of earthquake wave-
form doublets and further evidence for inner core super-rotation. Geophysical
Journal International 174, 9931006.
56 G.A. Prieto et al. / Tectonophysics 570571 (2012) 4256

You might also like