Download 3-D crustal structure of the extensional Granada Basin in the

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Geophysics wikipedia , lookup

Deep sea community wikipedia , lookup

Geology wikipedia , lookup

Geochemistry wikipedia , lookup

Seismic anisotropy wikipedia , lookup

Large igneous province wikipedia , lookup

Provenance (geology) wikipedia , lookup

Earthquake engineering wikipedia , lookup

Seismic inversion wikipedia , lookup

Surface wave inversion wikipedia , lookup

Earthscope wikipedia , lookup

Transcript
Tectonophysics 344 (2002) 61 – 79
www.elsevier.com/locate/tecto
3-D crustal structure of the extensional Granada Basin in the
convergent boundary between the Eurasian and African plates
Inmaculada Serrano a,b,*, Dapeng Zhao a, José Morales b,c
a
Department of Earth Sciences, Ehime University, Bunkyo-cho 2-5, Matsuyama 790-8577, Japan
b
Instituto Andaluz de Geofı́sica, Universidad de Granada, 18071-Granada, Spain
c
Departamento de Fisica Teórica y del Cosmos, Facultad de Ciencias, Universidad de Granada, Granada, Spain
Received 27 March 2001; accepted 10 September 2001
Abstract
Three-dimensional P and S wave velocity models of the crust under the Granada Basin in Southern Spain are obtained with a
spatial resolution of 5 km in the horizontal direction and 2 to 4 km in depth. We used a total of 15407 P and 13704 S wave highquality arrival times from 2889 local earthquakes recorded by both permanent seismic networks and portable stations deployed
in the area. The computed P and S wave velocities were used to obtain three-dimensional distributions of Poisson’s ratio (r) and
the porosity parameter (Vp Vs). The 3-D velocity images show strong lateral heterogeneities in the region. Significant velocity
variations up to ± 7% in P and S velocities are revealed in the crust below the Granada Basin. At shallow depth, high-velocity
anomalies are generally associated with Mesozoic basement, while the low-velocity anomalies are related to the neogene
sedimentary rocks. The south – southeastern part of the Granada Basin exhibits high r values in the shallowest layers, which
may be associated with saturated and unconsolidated sediments. In the same area, Vp Vs is high outside the basin, indicating
low porosity of the mesozoic basement. A low-velocity zone at 18-km depth is found and interpreted as a weak – ductile crust
transition that is related to the cut-off depth of the seismic activity. In the lower crust, at 34-km depth, a clear slow Vp and Vs
anomalous zone may indicate variations in lithology and/or with the rigidity of the lower crust rocks. D 2002 Elsevier Science
B.V. All rights reserved.
Keywords: Crustal earthquakes; Low-velocity layer; P waves; S waves; Seismic tomography; Granada Basin
1. Introduction
The western Mediterranean experiences seriously
associated with the interaction between the Eurasian
and African plates. The spatial distribution of the
earthquakes is in agreement with a well-defined plate
*
Corresponding author. Department of Earth Sciences, Ehime
University, Bunkyo-cho 2-5, Matsuyama 790-8577, Japan. Tel.:
+81-89-927-9652; fax: +81-89-927-9640.
E-mail address: [email protected] (I. Serrano).
boundary on the Atlantic side (Menard, 1965), where
in its eastern segments, an ocean – ocean convergence
between the two plates is taking place (Minster and
Jordan, 1978; Argus et al., 1989, Sartori et al., 1994);
earthquakes are scattered and focal mechanisms show
a combination of thrust faulting and strike – slip
motion (McKenzie, 1972; Fukao, 1973; Udias et al.,
1976; Grimison and Chen, 1986). The plate boundary
is less clear between the Iberian Peninsula and
Morocco-Algeria, its seismicity is diffuse and the
zone delimited by the seismicity reaches a maximum
0040-1951/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 4 0 - 1 9 5 1 ( 0 1 ) 0 0 2 0 1 - 3
62
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
width of 300 km (Anderson and Jackson, 1987,
Buforn et al., 1988). This region is characterized by
a high rate of seismicity with many earthquakes of
magnitude 5.0 (Fig. 1).
The Betic Cordillera in Southern Spain and the Rif
in Morocco (Fig. 2) are among the most important
Alpine orogene of the western Mediterranean, which
were built up as the responses to the convergence
between the Eurasian and African plates (Dewey,
1988). The Granada Basin is one of the most distinct
intramountaineous basins of the Betic Cordillera,
which is filled with sedimentary rocks lying over the
External (EZ) and Internal Zones (IZ), which constitute the basement of Granada Basin (Fig. 2). The first
features of the formation of the Granada Basin started
in the Middle Miocene whose rocks crop out in the SE
area; nevertheless, it is from the Late Miocene (Early
Tortonian) to the present that it acquired its principal
characteristics. The Tortonian rocks are marine and
their outcrops reach up to 1500 m above sea level
(Rodriguez-Fernandez, 1989), while the Turolian and
subsequent rocks are continental. The sedimentary
infill, like the metamorphic basement, is affected by
upright open folds trending NE – SW and by normalfault set with NW –SE strikes (Galindo-Zaldivar et al.,
1999). The External Zones, which formed most of the
southern and part of the eastern margin of the Iberian
Massif, are mainly composed by limestones, dolostones, marls, sandstones and clays of Mesozoic and
Cenozoic ages. Moreover, igneous rocks (ophiolites)
and pillow-lava levels are intercalated into the Triassic
rocks and the Jurassic and Cretaceous sedimentary
series. The Internal Zones, situated farther to the east,
consist of several tectonic units mostly composed by
schists and marbles. The types of deformation occurring in the Internal and External Zones and their
history are different, and it is only the great displacement (primarily during the Early Mioceno) that created the present-day juxtaposition (Sanz de Galdeano,
1990).
The main aim of this study is to analyze in detail the
features of the crust below the Granada Basin and
surrounding regions by using high-resolution seismic
tomography. The availability of a dense seismic net-
Fig. 1. Seismicity of the South Spain, Morocco and West Argelia for the period 1983 – 2001 (magnitude 2.8) registered by Andalusian Seismic
Network (Andalusian Institute of Geophysics, Granada University, Spain).
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
63
Fig. 2. Simplified geological map of Central Betic Cordilleras. Crosses, triangles, squares and dark circles denote the earthquakes, permanent seismic stations belongs to RSA,
temporary seismic stations belongs to RSA and seismic stations belongs to National Geographic Institute, Ministry of Fomento and Royal Institute and Navy’s Observatory at San
Fernando, respectively. The insert map on the upper right side shows the location of the present study area.
64
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
work has allowed us to collect high-quality data from
many earthquakes of small magnitudes in the area. The
Granada Basin and the neighboring areas have the
highest rate of microseismicity (mb 5.5) in the Iberian Peninsula (De Miguel et al., 1989). The seismic
activity, on average, does not exceed 20 km in depth;
the lower crust is essentially aseismic and the seismic
activity is concentrated in the upper crust between 5
and 17 km. This lower cut-off in the seismic activity in
the crust is interpreted as the limit brittle – ductile,
limiting the thickness of the seismogeneic layer
(Morales et al., 1997). These authors pointed out the
coincidence of the lower cut-off in the seismic activity
with the presence of an intracrust reflector detected by
Banda et al. (1993) and Galindo-Zaldivar et al. (1997)
and which would be layering the crust in two: a seismic
part (brittle) and another aseismic (ductile), according
to the differences observed in the style of the deformation between the upper and lower crust (GalindoZaldivar et al., 1997).
Fig. 3. (A and B) Trade-off curves between the variance of the solutions and arrival time residuals for inversions with irregular discontinuities
(3D) and flat discontinuities (1D). The selected values (damping) for the final results are marked within the circle. (A) For P wave velocity. (B)
For S wave velocity.
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
The present state of stress determined by focal
mechanism solutions shows an extensional regime.
Both the surface geological data and the focal mechanisms indicate a present-day regional NE –SW extension, with triaxial to prolate stress ellipsoids. However,
the stress field is heterogeneous with both spatial and
temporal variations, sometimes even acting simultaneously in adjacent areas. The most frequent changes
consist of radial or NW –SE extension favored by the
low axial ratio of the stress ellipsoids and NW – SE
subhorizontal compression favored by the regional
65
tectonic (Galindo-Zaldivar et al., 1999). The extension
is normal to the regional convergence direction (NW –
SE) (DeMets et al., 1990). Beneath this region, there are
also deep earthquakes with focal depths down to 640
km although there is a lack of intermediate earthquakes
(Buforn et al., 1997).
Numerous seismic tomographic studies have already
been performed in this region: Blanco and Spakman
(1993), Plomerová et al. (1993), Sallarés (1996), Gurrı́a
et al. (1997), Serrano et al. (1998) and Calvert et al.
(2000) have found important velocity anomalies in this
Fig. 4. Fractional P wave velocity perturbations (in percentage) at three depth layers for the case when the Moho and the mid-crust discontinuity
are flat.
66
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
Fig. 5. (a) Fractional P wave velocity perturbations (in percentage) at the first six depth layers for the case when the Moho and the mid-crust discontinuity have lateral depth
variations. The velocity perturbation is from the mean value of the inverted velocity at each layer. The depth of the layer is shown at the lower-left corner of each map. Blue and red
colors denote fast and slow velocities, respectively. The velocity perturbation scale is shown on the right. Black lines show active faults. Red lines show the boundary between the
Granada Basin (sedimentary rocks) and External – Internal Zones. (b) The same to the second six depth layers.
67
Fig. 5 (continued).
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
68
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
region. However, the low resolution of these results do
not allow a detailed correlation between the tomographic results and shallow geological structures.
In local earthquake tomography, body-wave arrival
times are used to estimate P and S wave velocities and
from these, we deduce variations in lithology and
physical properties of rocks. According to Christensen
(1989), velocity usually increases with depth; however, in some regions, velocity reversals have been
recognized. Comparisons of laboratory velocity data
with seismically measured velocities are subject to
many considerations because the influences of temperatures and pressures must be taken into account to
infer mineralogy from seismic velocities. For crystalline rocks, the initial application of pressure affects
the seismic velocity by reducing microporosity and
within the upper part of the crust, large-scale fracturing may have an analogous effect. In deeper portions
of the crust and in the upper mantle, however, it is
likely that fractures and microcracks are no longer
present because of the high confining pressures and
metamorphic recrystallization. In sedimentary rocks,
especially sandstone, porosity is much higher and
does not completely close at elevated pressures and
so will have a dominant influence on seismic velocities. For most common rock types, velocity decreases
with increasing temperature. Velocities have been
found to correlate well with mineralogy at pressure
high enough to eliminate the influence of cracks,. The
lower density minerals, such as quartz, have lower
velocities than the high-density pyroxenes, olivine
and garnet. In general, metamorphic rocks of mineral
composition similar to igneous rocks have identical
velocities although this may be complicated by anisotropy (Christensen, 1989). A brief review of how rock
properties relate to seismic velocity and attenuation
can be found in Sander et al. (1995) and Lees and Wu
(2000). Poisson’s ratio is directly related to Vp/Vs, the
ratio of compressional and shear-wave velocities and
can be a useful indicator of lithology and pore fluid
pressure. On average, Poisson’s ratio is 0.25 for
Earth’s crust and upper mantle (Holbrook et al.,
1988). Any variation from this value may indicate a
change in the properties of the material. On the other
hand, the product of compressional and shear-wave
velocities, Vp Vs has been used to delineate porosity
in sedimentary rocks (Iverson et al., 1989). It has been
observed that lower Vp Vs indicates an increase of
porosity, whereas Vp/Vs, constant for specific lithology, does not change with porosity (Pickett, 1963;
Tatham, 1992). Vp/Vs, or r, is commonly used to
delineate lithology, while the product Vp Vs can be
used to identify variations in porosity for the shallow
crustal rocks (Iverson et al., 1989; Tatham, 1992). In
this work, we try to relate the three-dimensional
variations in r and Vp Vs to fluctuations in fluid
content, porosity and changes in the lithologies of
different zones of the Granada Basin.
2. Data selection
We used arrival time of P and S waves from local
earthquakes recorded by the permanent and portable
seismic stations of the Andalusian Seismic Network
(RSA) that is operated by the Andalusian Institute of
Geophysics (Granada University). Although the station coverage of this network is dense and adequate
for locating local earthquakes, we have also used the
data from other stations that belong to other institutions (i.e., National Geographic Institute, Ministry of
Fomento and Royal Institute and Navy’s Observatory
at San Fernando) in order to achieve a better ray
coverage for the tomographic inversion. The data used
were gathered from the events that occurred between
1983 and 1999. Since 1983 to 1988, the most of the
data were recorded in analog formats, from 1988
onwards all the recordings are digital (except three
stations belong to Ministry of Fomento). The seismic
stations used are densely and uniformly distributed in
the studied area (Fig. 2).
We selected a set of events which are located
between 36°480N to 37°120N and from 3°240W to
4°030W. These earthquakes were selected on the basis
of minimum number of arrival times, at least 10, as
well as the fact that they have a uniform spatial
distribution in the study area. The 96% of the earthquakes selected have rms smaller than 0.2 s and the
rms maximum for the hypocenter locations is 0.9 s.
All the events are located by more than five stations.
Finally, we selected a total of 2889 earthquakes, with
hypocenters located using the method of Lienert et al.
(1986). The mean error is 2 km for the latitude, 0.8
km for the longitude and 1 km for the depth provided
by the location program. The depth of the selected
earthquakes ranges from 0 to 40 km. The majority of
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
the events have focal depths shallower than 20 km
below the Granada Basin; only the events located
beneath the southwestern part of the basin are deeper.
Fig. 2 shows the epicenter distribution of the selected
earthquakes. A total of 15407 P arrivals and 13704 S
arrivals were collected from these earthquakes in the
study area of 97 52 km. For stations of the Andalusian Seismic Network, the accuracy of time picking
of the P and S digital arrivals may be estimated in the
most favorable cases as ± 0.01 s. In the case of less
impulsive arrivals and/or poor signal-to-noise ratio,
the accuracy is degraded, but not more than 0.1 s. For
the analog recordings (less than 4% of the total
69
number of data used in this study), the mechanical
characteristics of the drums and the velocity of development of the traces (2 mm/s) set a limit to the
accuracy of about ± 0.2 s. With regard to the picking
up of the S arrivals, all seismic stations, except one,
have vertical short period sensors.
3. Method and analysis
In this study, we have used the tomography method
of Zhao et al. (1992). Although the conceptual
approach of this method is derived from Aki and Lee
Fig. 6. Distribution of the number of P wave rays passing through each grid node (hit counts). The hit count scale is shown at the bottom.
70
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
(1976), it has some additional features. The technique
can deal with the complex geometry of seismic velocity discontinuities, such as the Moho or the subducting
slab boundary, and it uses a 3-D ray tracing scheme to
compute travel times and ray paths. For details of the
method, see Zhao et al. (1992, 1994).
Our velocity model contains three layers: the
upper crust, the lower crust and the upper mantle,
which are bounded by the Moho and the mid-crust
discontinuity. We set three-dimensional grid nets
independently for every layers to express the threedimensional velocity structure for layers, which are
bounded by two adjacent discontinuities. Velocities at
grid points are taken to be unknown parameters
except those at the outermost grid. Velocities at the
outermost grid are used just to interpolate velocities
outside of the modeling space. A velocity at any point
in the model is calculated by linearly interpolating the
velocities at the grids surrounding that point (Zhao et
al., 1992).
The P wave velocity (Vp) for the upper crust, lower
crust and the uppermost mantle is 6.0, 6.8 and 8.0 km/
s, respectively. Vp in the upper mantle has a vertical
gradient of 0.003 km/s/km. Vp/Vs is set to be 1.7 in the
initial model. We have constructed this initial velocity
model for the shallow layers taking into consideration
the results of Wadati Diagrams of 335 earthquakes
occurred in the region (Morales et al., 1997) and
Fig. 7. Results of checkerboard resolution test for P (a) and S (b) waves. Black and white symbols denote slow and fast velocities, respectively.
The depth of each layer is shown at the upper part of the map. The grid spacing is 5 km in the horizontal direction and 2 – 4 km in depth. The
perturbation scale is shown on the right.
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
71
Fig. 7 (continued).
considering for all the layers the seismic profiles and
gravimetric data from Galindo-Zaldivar et al. (1997,
1998) and the results for the Betic Cordillera from
Banda et al. (1993).
In a preliminary inversion, we assumed that the
mid-crust and the Moho discontinuities had constant
depths at 17 and 38 km, respectively. Then, we tried to
use a more realistic model with lateral depth changes in
the Moho and mid-crust discontinuity. The Moho and
mid-crust discontinuity geometries were constructed
by referring to the seismic explosion profiles and
gravimetric data (Galindo-Zaldivar et al., 1997,
1998). The mid-crust discontinuity depth ranges from
17 to 15 km; the Moho depth ranges from 38 to 36 km.
After we compared the results for both of the cases, the
most meaningful outcome is that the inversion with the
curved discontinuities results in a final root-meansquare residual of 18% for Vp and 12% for Vs smaller
than those with flat discontinuities (Fig. 3). This
suggests that it is important to take into account the
lateral depth variations of the seismic discontinuities in
order to determine a detailed crustal structure. When
the geometry of the discontinuities is taken into
account, ray paths and travel times can be calculated
more accurately (Zhao et al., 1992, 1994). Also, we
found that the general patterns of velocity distributions
in both cases are almost the same although there are
some differences in amplitudes of the velocity anomalies for the upper and lower crust (Figs. 4 and 5).
We solved the inverse problem for 3996 (2015 for
P and 1981 for S waves) velocity parameters at the
grid nodes with hit counts (number of rays sampling a
72
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
cell) greater than 10. Fig. 6 shows the distribution of
ray paths at several layers. The errors for the velocity
perturbations obtained are estimated to be < 1%. The
damping parameter was selected based on an empirical approach (Eberhart-Phillips, 1986). A number of
inversions were run with different damping values.
Then the reduction in travel time residual is compared
to the variance of the solutions and we draw a tradeoff curve between them. The selected value of the
damping parameter is the one which gives the optimal
residual reduction and the solution variance (Fig. 3).
4. Hit counts and resolution test
Before describing the features of the obtained
models, we first show the distribution of ray paths
in the study area and evaluate the resolution of the
tomographic image. Fig. 6 shows the distribution of
hit counts at several layers. The density of ray path
coverage varies throughout the study area. The coverage is very good down to a depth of 18 km in almost
the whole area, with the highest density in the central
part of the Granada Basin. From 22 to 38 km depth,
the coverage is good in the western part of the study
area and it is not good in the eastern part. Hence,
reliable results are expected for the upper crust layers
of the study area (from 2 to 14 km).
By the concept of resolution, we wish to know how
the true structure is reconstructed in the calculated
image. The most direct means of testing the resolution
of the inverted solution is to first calculate the sets of
travel time delays that resulted from tracing the actual
ray set through a synthetic test structure, then invert
Fig. 8. The same as Fig. 5 but for S wave velocity structure. (a) Fractional S wave velocity perturbations (in percentage) at the first six depth
layers for the case when the Moho and the mid-crust discontinuity have lateral depth variations. (b) The same to the second six depth layers.
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
73
Fig. 8 (continued).
those delays as though they are data and, finally,
compare the synthetic inversion with the initial structure (Zhao et al., 1992). We used the checkerboard
resolution test (see Humphreys and Clayton, 1988).
Positive and negative perturbations are assigned to a
3-D grid nodes, which are arranged in the modeling
space, the image of which is straightforward and easy
to remember. We conducted a number of inversions
by changing the grid spacing between grid nodes,
which are set up in the study area. Through resolution
analyses for the different grid spacing, we found that a
grid spacing of 5 km in the horizontal direction and
2– 4 km in depth is the minimum space for that the
data give reasonable results (Fig. 7). The results of the
checkerboard test are good down to a depth of 18 km;
however, the results for depths from 22 to 38 km
under eastern part of the basin are not very good. As
expected, the resolution is good in areas where most
of the earthquake occurred. The total number of grid
nodes is 13 23 14.
5. Results and discussion
The velocity at a grid point represents the best
estimate for the volume surrounding the grid point.
Similarly, the size and shape of velocity anomalies can
provide reasonable estimates, but may not correspond
to the exact boundaries of true velocity features
(Eberhart-Phillips, 1986). Considering the good resolution in the shallow layers, we will mainly focus our
discussion on features obtained in those layers. There
is no volcanism and geothermal activity in the Granada Basin and surrounding areas, then anomalies in Vp
and Vs may be related to changes in lithology, cracks
and fluid distribution in the crust.
74
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
5.1. Layer: 2 km
One of the distinct features of this layer is the high P
and S wave velocity anomaly under the eastern boundary of the Granada Basin (Figs. 5 and 8). This
anomaly coincides with the outcrop of dolomies and
marbles belonging to the Mesozoic basement (IZ). The
anomaly also exhibits high Vp Vs anomaly in a small
area ( + 15%), which may indicate low porosity, compacted and weakly fractured material here (Fig. 9).
In the south –central and south –southeast parts of
Granada Basin, there is a high Vp zone that goes
through neogene – quaternary sedimentary rocks
(inside the basin) and marble – dolomies of the mesozoic basement (outside the basin, IZ). Its trend may be
parallel to the local normal-fault set with NW –SE
strikes. In this same zone, the results for Vs are
different: inside the basin (sedimentary layers) shows
low Vs ( 6%) and outside the basin shows high Vs
( + 4%), which may correspond to low-crack density
(mesozoic basement). The value of r is high inside the
basin (Fig. 10), which may be related to saturated and
unconsolidated sediments. Vp Vs is high outside
basin, indicating low porosity of the mesozoic basement.
The W Granada Basin (EZ) exhibits high Vp
( + 6%) and Vs ( + 2%), generally consistent with the
magnetic data (Ardizone et al., 1989). The positive
magnetic anomaly and fast Vp anomaly can indicate
the existence of a body formed by basic igneous rocks
or metabasites, which belong to the Iberian Massif.
The Poisson ratio shows higher values than those
obtained at S –SE and E Granada Basin, indicating
different lithologies on the two sides of the basin.
However, the Vp Vs values are similar, indicating a
low fracture density or highly compacted material on
the two sides of the outside basin.
A low Vp anomaly is observed in the central and
north Granada Basin, which is correlated with the
neogene and quaternary rocks. Vs shows, in general,
low values except in a small area, where we can see a
high-velocity anomaly (Fig. 8). Also, in this small
area, r is very low ( 25%), indicating rocks with a
higher rigidity than the surroundings. This high Vs and
low r in this small area may be related to a thin
sedimentary layer inside the basin.
Under the southwestern part of the Granada Basin,
there are low Vp and Vs anomalies and we obtain the
smallest values of the porosity parameter ( 14.8%)
of the whole study area, indicating a high-porosity
Fig. 9. Percent perturbation of Vp Vs (in percentage) at the first three depth layers. Black areas are high Vp Vs ; white areas are low Vp Vs.
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
material. This velocity anomaly shows a general N –S
trend. This anomaly (about 150 km2) corresponds to
the southern part of a great extension of Jurassic
carbonated rocks (Sierra Gorda) and alluvial materials
belonging to ‘‘Polje de Zafarraya’’, formed by sand
levels. This aquifer is one of the most important
features in this region (MAGNA, 1979).
A P and S low-velocity zone is imaged outside the
NW Granada Basin and the porosity parameter is very
low. The major part of this anomaly corresponds to
one of the thick continental areas filled of the Granada
Basin (MAGNA, 1988). It is formed by marls, silts,
limestones, conglomerates, sands and silt clays.
5.2. Layers from 4 to 14 km
The eastern part of the Granada Basin continues
showing high Vp ( + 5%) and high Vs ( + 5%) at 4-km
75
depth, which disappear at 8-km depth. The porosity
parameter also shows high values ( + 11%) down to a
depth of 6 km, indicating low porosity in this area.
At a depth of 4 km, the high Vp and Vs anomaly in
the southeastern part of the basin decreases to smaller
values. This anomaly again appears at a depth of 10
km, showing high values of Vp ( + 5%) and Vs ( + 5%).
Under the southeastern part of the basin, Poisson ratio
is high ( + 22%) down to a depth of 4 km, indicating
low rigidity.
The western zone of the Granada Basin exhibits
high values of Vp and Vs down to a depth of 10 km,
while the porosity parameter is high down to a depth of
4 km. From a depth of 6 to 10 km, the area west of the
basin shows high Vp ( + 5%) and lower values of Vs.
It is important to point out that from 6 to 10 km
depth outside the N –NW Granada Basin, the Vp model
shows a high-velocity body with a NS trend. At the
Fig. 10. Percent perturbation of Poisson’s ratio at (a) the first six depth layers and (b) the second six layers.
76
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
Fig. 10 (continued).
surface, this area shows outcrops of igneous rocks
(Subbetic units), such as ophiolites, which are normally found located together with the Triassic rocks of
keuper facies and pillow-lava levels, intercalated into
the Jurassic and Cretaceous sedimentary series. One of
the most intense positive aeromagnetic anomalies of
the Betic Cordillera is located in this area (Ardizone et
al., 1989). The maximum in the total field surface
anomaly map reaches up to 120 nT and its place is
coincident with the positive Vp anomaly obtained in
this study. According to Bohoyo et al. (2000), this
large extent of the magnetic anomaly maximum is
probably the consequence of a large body of basic
igneous rocks. They investigated this area with new
magnetic and gravity data and their results indicate that
the high aeromagnetic anomaly values and the small
difference between the land and aeromagnetic anomaly values indicate that it has a deep origin.
Under the central Granada Basin, we can observe a
NE – SW preferential orientation of the boundary
between high and low velocity, the southeastern part
of the basin shows higher velocity, in general, than
that of the northwest part. This boundary can be in
relation to the NE – SW-trending contact between the
EZ and IZ.
5.3. Layer: 18 km
One interesting feature is the presence of a lowvelocity zone at mid-crustal depths beneath the Granada Basin. This slow Vp and Vs appear at a depth of 18
km and strong changes of seismic velocity are visible
from a depth of 14 to 18 km (Fig. 11). The low-velocity
zone is in agreement with the depth range of crustal
seismicity (Morales et al., 1997), which might imply
that it is a crustal decoupling zone at the base of the
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
77
Fig. 11. E – W vertical cross sections of P wave velocity perturbations (in percent) along lat 37°N. Blue and red colors denote fast and slow
velocities, respectively. The velocity perturbation scale is shown at the bottom. Crosses denote earthquakes that occurred within a width of 25
km along the profile.
seismogenetic zone. A strong high Vp and Vs anomaly
is also detected in the southeastern boundary of the
basin. This anomaly (Fig. 10) coincides with the earthquake swarm activity located in the zone (Saccorotti et
al., in press) and could be associated with an asperity.
These results together with the highest values of r
in two small areas, NW Loja and N Sierra Alhama,
may be indicating the weak sections of the seismogenic crust due to overpressurized, fluid-filled, fractured rock matrices, similar to those detected in the
source area of the 1995 Kobe earthquake in Japan
(Zhao et al., 1996).
Vp ( 5%), slow Vs ( 3%) and low r are obtained in
the NW of the study area at a depth of 34 km. The low
resolution in the lower crust is considered to be due to
the lack of earthquakes in these layers and the resolution for S wave velocity structure is inferior to that
for P wave. For this reason, we prefer not to do
interpretation of the results although these values
may indicate variations of the lithology or a high
rigidity of the crustal material as compared with the
surrounding areas.
6. Conclusions
5.4. Layers from 22 to 38 km
The low Vp and Vs values continue under the
Granada Basin down to a depth of 26 km. Very slow
High-resolution P and S wave tomographic images
are determined for the Granada Basin. The results
show strong lateral heterogeneities in the crust.
78
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
In the depth range 2 to 4 km, the SE part of the
Granada Basin shows a strong high Vp zone (+ 6%).
Its trend is parallel to the local normal-fault set with
NW – SE strikes. We obtained low Vs ( 6%) inside
the basin (sedimentary layers) and high Vs (+ 4%)
outside the basin (IZ, mesozoic basement), which may
correspond to the difference in crack density. The
value of r is high inside the basin, which may show
saturated and unconsolidated sediments. Vp Vs is
high outside the basin, indicating a low porosity for
the mesozoic basement. The eastern part of the
Granada Basin shows high Vp (+ 6%) and Vs (+ 7%),
low r, in accordance with the presence of Mesozoic
basement (IZ). A strong high anomaly of Vp Vs in a
small area (+ 15%) indicates a material of very low
porosity, compacted and weakly cracked. The western
part of the Granada Basin (Sierra Gorda, EZ) shows
high Vp (+ 6%) and Vs (+ 2%), generally consistent
with magnetic data. The positive magnetic anomaly
and fast Vp anomaly can indicate the existence of a
body probably comprising basic igneous rocks or
metabasites, which belong to the Iberian Massif.
r shows higher values than those obtained in the
southeast and eastern, indicating different lithologies
in the two sides of the basin. However, the Vp Vs
values are similar, indicating a low fracture density or
highly compacted material.
In the depth range of 6 to 14 km, the southeastern
part of the basin shows high Vp (+ 5%) and high Vs
(+ 5%). In the depth range of 18 to 38 km, slow Vp
and Vs appear at a depth of 18 km under the south part
of the basin. The maximum depth of seismicity is 14–
17-km depth. This low-velocity layer may be related
to a highly fractured and fluid-filled zone. r is highest
in two small areas: NW Loja and N Sierra Alhama. It
may indicate the weak sections of the seismogenic
crust due to overpressurized, fluid-filled, fractured
rock matrices.
Acknowledgements
This work has been supported by the Comision
Interministerial de Ciencia y Tecnologia project
AMB99-0795-C02-01 and REN2001-2418-C04-04
(Spain). The first author (I. Serrano) thanks the
University of Granada (Spain) and Ministerio de
Ciencia y Tecnologı́a for two postdoctoral fellowships
at Ehime University (Japan). The authors would also
like to thank Claudio Chiarabba and a anonymous
reviewer for their very constructive reviews and
suggestions.
References
Aki, K., Lee, W.H.K., 1976. Determination of the three-dimensional
velocity anomalies under a seismic array using first P arrival
times from local earthquakes: Part 1. A homogeneus initial
model. J. Geophys. Res. 81, 4381 – 4399.
Anderson, H., Jackson, J., 1987. Active tectonics of the Adriatic
region. Geophys. J. R. Astron. Soc. 91, 937 – 983.
Ardizone, J., Mezcua, J., Socias, I., 1989. Mapa Aeromagnético de
España Peninsular, IGN. 1:1000000.
Argus, D.F., Gordon, R.G., Demets, C., Stein, S., 1989. Closure of
the Africa – Eurasia – North America plate motion circuit and
tectonics of the Gloria fault. J. Geophys. Res. 94, 5585 – 5602.
Banda, E., Gallart, J., Garcı́a-Dueñas, V., Dañobeitia, J.J., Makris,
J., 1993. Lateral variation of the crust in the Iberian peninsula:
new evidences from the Betic Cordillera. Tectonophysics 221,
53 – 66.
Blanco, M.J., Spakman, W., 1993. The P-wave velocity structure of
the mantle below the Iberian Peninsula: evidence for subducted
lithosphere below southern Spain. Tectonophysics 221, 13 – 34.
Bohoyo, F., Galindo-Zaldivar, J., Serrano, I., 2000. Main features of
the basic rock bodies of the Archidona region derived from
geophysical data (External Zones, Betic Cordillera). C. R. Acad.
Sci. Paris 330, 667 – 674.
Buforn, E., Udias, A., Mezcua, J., 1988. Seismicity and focal mechanisms in south Spain. Bull. Soc. Am. 78, 2008 – 2024.
Buforn, E., Coca, P., Udias, A., Lasa, C., 1997. Source mechanism
of intermediate and deep earthquakes in southern Spain. J. Seismol. 1, 113 – 130.
Calvert, A., Sandvol, E., Seber, D., Barazangi, M., Roecker, S.,
Mourabit, T., Vidal, F., Alguacil, G., Jabour, N., 2000. Geodynamic evolution of the lithosphere and upper mantle beneath the
Alboran region of the western Mediterranean: constraints from
travel time tomography. J. Geophys. Res. 105 (B5), 10871 –
10898.
Christensen, N.I., 1989. In: James, D.E. (Ed.), The Encyclopedia of
Solid Earth Geophysics ISBN 0-442-24366-9.
DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current
plate motions. Geophys. J. Int. 101, 425 – 478.
De Miguel, F., Vidal, F., Alguacil, G., Guirao, J.M., 1989. Spatial
and energic trends of the microearthquakes activity in the Central Betics. Geodin. Acta 3, 87 – 94.
Dewey, J.F., 1988. Extensional collapse of orogens. Tectonics 7,
1123 – 1139.
Eberhart-Phillips, D., 1986. Three dimensional velocity structure in
Northern California Coast Ranges from inversion of local earthquake arrival times. Bull. Seismol. Soc. Am. 76, 1025 – 1052.
Fukao, Y., 1973. Thrust faulting at a lithospheric plate boundary: the
Portugal earthquake of 1969. Earth Planet. Sci. Lett. 18, 205 –
216.
I. Serrano et al. / Tectonophysics 344 (2002) 61–79
Galindo-Zaldivar, J., Jabaloy, A., Gonzalez-Lodeiro, F., Aldaya, F.,
1997. Crustal structure of the central sector of the Betic Cordillera (SE Spain). Tectonics 16, 18 – 37.
Galindo-Zaldivar, J., Gonzalez-Lodeiro, F., Jabaloy, A., Maldonado,
A.A., Scheider, A.A., 1998. Models of magnetic and Bouguer
gravity anomalies for the deep structure of the central Alboran
Sea basin. Geo-Mar. Lett. 18 (1), 10 – 18.
Galindo-Zaldivar, J., Jabaloy, A., Serrano, I., Morales, J., GonzalezLodeiro, F., Torcal, F., 1999. Recent and present-day stresses in
the Granada Basin (Betic Cordilleras): example of a late Miocene – present-day extensional basin in a convergent plate boundary. Tectonics 18 (4), 686 – 702.
Grimison, N.L., Chen, W.P., 1986. The Azores – Gibraltar plate
boundary: focal mechanisms, depth of earthquake, and their
tectonic implications. J. Geophys. Res. 91, 2029 – 2047.
Gurrı́a, E., Mezcua, J., Blanco, M.J., 1997. Crustal and upper mantle velocity structure of Southern Iberia, the sea of Alborán, and
the Gibraltar arc determined by local earthquake tomography.
Ann. Geofis., 1.
Holbrook, W.S., Gajewski, P., Kramer, A., Prodehl, C., 1988. An
interpretation of wide-angle compressional and shear wave data
in southwest Germany: Poisson’s ratio and petrological implications. J. Geophys. Res. 93, 12081 – 12106.
Humphreys, E., Clayton, R.W., 1988. Adaptation of back projection
tomography to seismic travel time problems. J. Geophys. Res.
93, 1073 – 1085.
Iverson, W.P., Fahmy, B.A., Smithson, S.B., 1989. VpVs from
mode converted P – SV reflections. Geophysics 54, 843 – 852.
Lees, J.M., Wu, H., 2000. Poissons ratio and porosity at Coso geothermal area, California. J. Volcanol. Geotherm. Res. 95, 157 –
173.
Lienert, B.R., Berg, E., Frazer, L.N., 1986. AHypocenter@: an
earthquake location method using centered, scaled and adaptively lamped least squares. Bull. Seismol. Soc. Am. 76, 771 –
783.
MAGNA, 1979. Mapa Geologico de Espana. 1:50.000. Insituto
Geologico y Minero de España. Servicio de Publicaciones del
Ministerio de Industria y Energia, 1040, pp. 18 – 43.
MAGNA, 1988. Mapa Geologico de Espana. 1:50.000. Insituto
Geologico y Minero de España. Servicio de Publicaciones del
Ministerio de Industria y Energia, 1008, pp. 18 – 41.
McKenzie, D.P., 1972. Active tectonics of the Mediterranean region. Geophys. J. R. Astron. Soc. 30, 109 – 185.
Menard, H.W., 1965. Sea floor relief and mantle convection. Phys.
Chem. Earth 6, 315 – 364.
79
Minster, J.B., Jordan, T.H., 1978. Present-day plate motions. J. Geophys. Res. 83, 5231 – 5254.
Morales, J., Serrano, I., Vidal, F., Torcal, F., 1997. The depth of the
earthquake activity in the Central Betic (Southern Spain). Geophys. Res. Lett. 24, 3289 – 3292.
Pickett, G.R., 1963. Acoustic character logs and their applications in
formation evaluation. J. Petrol. Tech. 15, 650 – 667.
Plomerová, J., Payo, G., Babuska, V., 1993. Teleseismic P-residual
study in the Iberian Peninsula. Tectonophysics 221, 1 – 12.
Rodriguez-Fernandez, J., 1989. El Mioceno en el sector central de
las Cordilleras Béticas. Tesis doctoral, Universidad de Granada,
Granada, Spain, 224 pp.
Saccorotti, G., Carmona, E., Ibáñez, J.M., Del Pezzo, E., 2000.
Phys. Earth Planet. Int. (in press).
Sallarés, V., 1996. Tomografı́a Sı́smica a partir de terremotos locales. Aplicación en la Cordillera Bética y el Mar de Alborán.
Master thesis, Univ. Barcelona, 96 pp.
Sander, C.O., Ponko, S.C., Nixon, L.D., Schwart, E.A., 1995. Seismological evidences for magmatic and hydrothermal structure in
Long Valley caldera from local earthquake attenuation and velocity tomography. J. Geophys. Res. 100, 8311 – 8326.
Sanz de Galdeano, C., 1990. Geological evolution of the Betic
Cordillera, in the western Mediterranean, Miocene to the
present. Tectonophysics 172, 107 – 119.
Sartori, R., Torelli, L., Zitellini, N., Peis, D., Lodolo, E., 1994.
Eastern segment of the Azores – Gibraltar line (central – eastern
Atlantic): an oceanic plate boundary with diffuse compressional
deformation. Geology 22, 555 – 558.
Serrano, I., Morales, J., Zhao, D., Torcal, F., Vidal, F., 1998. PWave tomographic images of the Central Betics Alboran sea
(South Spain) using local earthquakes: contribution for a continental collision. Geophys. Res. Lett. 20, 4031 – 4034.
Tatham, R.H., 1992. Vp/Vs and lithology. Geophysics 47, 333 –
344.
Udias, A., Lopez Arroyo, A., Mezcua, J., 1976. Seismotectonics of
the Azores – Alboran region. Tectonophysics 31, 259 – 289.
Zhao, D., Hasewaga, A., Horiuchi, S., 1992. Tomographic imaging
of P and S wave velocity structure beneath northeastern Japan. J.
Geophys. Res. 97, 19909 – 19928.
Zhao, D., Hasewaga, A., Kanamori, H., 1994. Deep structure of
Japan subduction zones as derived from local, regional and teleseismic events. J. Geophys. Res. 99, 22313 – 22329.
Zhao, D., Kanamori, H., Negishi, H., Wiens, D., 1996. Tomography
of the source area of the 1995 Kobe earthquake: evidence for
fluids at the hypocenter? Science 274, 1891 – 1894.