Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of Hydrology 553 (2017) 71–87

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Technical Note

Multiband PSInSAR and long-period monitoring of land subsidence in a


strategic detrital aquifer (Vega de Granada, SE Spain): An approach to
support management decisions
Rosa María Mateos a,b,⇑, Pablo Ezquerro b, Juan Antonio Luque-Espinar a, Marta Béjar-Pizarro b,
Davide Notti c, Jose Miguel Azañón c, Oriol Montserrat d, Gerardo Herrera b, Francisca Fernández-Chacón a,
Tomás Peinado a, Jorge Pedro Galve c, Vicente Pérez-Peña c, Jose A. Fernández-Merodo b, Jorge Jiménez a
a
Geological Survey of Spain (IGME), Urb. Alcázar del Genil, 4-Edif. Bajo, 18006 Granada, Spain
b
Geohazards InSAR Laboratory and Modelling Group, Geohazards Unit, Geological Survey of Spain (IGME), Spain
c
Department of Geodynamics, University of Granada, Avd. Fuentenueva s/n, 18071 Granada, Spain
d
Centre Tecnològic de Telecomunicacions de Catalunya, Parc Mediterrani de la Tecnologia (PMT)-Building B4, Av. Carl Friedrich Gauss 7, 08860-Castelldefels, Spain

a r t i c l e i n f o a b s t r a c t

Article history: This work integrates detailed geological and hydrogeological information with PSI data to obtain a better
Received 23 February 2017 understanding of subsidence processes detected in the detrital aquifer of the Vega de Granada (SE Spain)
Received in revised form 14 June 2017 during the past 13 years. Ground motion was monitored by exploiting SAR images from the ENVISAT
Accepted 26 July 2017
(2003–2009), Cosmo-SkyMed (2011–2014) and Sentinel-1A (2015–2016) satellites. PSInSAR results show
Available online 28 July 2017
an inelastic deformation in the aquifer and small land surface displacements (up to 55 mm). The most
widespread land subsidence is detected during the ENVISAT period (2003–2009), which coincided with a
Keywords:
long, dry period in the region. The highest displacement rates recorded during this period (up to 10 mm/
PSI monitoring
Land subsidence
yr) were detected in the central part of the aquifer, where many villages are located. For this period, there
Detrital aquifer is a good correlation between groundwater level depletion and the augmentation of the average subsi-
Clay content dence velocity and slight hydraulic head changes (<2 m) have a rapid ground motion response. The
Management Cosmo-SkyMed period (2011–2014) coincided with a rainy period, and the land subsidence is only con-
Sentinel-1 centrated in some points. Rates of average subsidence up to 11.5 mm/yr are obtained for this period and
Spain are anthropogenic in origin, being related to earthmoving works. During the Sentinel-1A monitoring per-
iod (2015–2016) most of the region showed no deformation, except for some points of unknown origin in
the NE sector. A general conclusion is that there is a clear lithological control in the spatial distribution of
ground subsidence; all the subsiding areas detected are located where a higher clay content was identi-
fied. Although the SE sector of the aquifer had more intense groundwater exploitation, no land subsi-
dence processes were detected, as coarse-grained sediments predominate in the substratum. This
research will contribute to the drawing-up of a management plan for the sustainable use of this strategic
aquifer, taking into account critical levels of groundwater depletion to avoid land subsidence in the areas
identified as vulnerable. The European Space Agency satellite Sentinel-1A could be an effective decision-
making tool in the near future.
Ó 2017 Elsevier B.V. All rights reserved.

1. Introduction

⇑ Corresponding author at: Geological Survey of Spain, Research Centre in


In many agricultural regions worldwide, prolonged groundwa-
Granada, Urb. Alcázar del Genil, 4-Edif Bajo, 18006 Granada, Spain. ter exploitation has caused land subsidence related to falling
E-mail addresses: rm.mateos@igme.es (R.M. Mateos), p.ezquerro@igme.es (P. groundwater levels. During recent decades, numerous cases of land
Ezquerro), ja.luque@igme.es (J.A. Luque-Espinar), m.bejar@igme.es (M. Béjar- subsidence related to intensive agricultural practices have been
Pizarro), davidenotti@gmail.com (D. Notti), jazanon@ugr.es (J.M. Azañón), oriol. reported in many developed aquifer systems (Motagh et al.,
monserrat@cttc.cat (O. Montserrat), g.herrera@igme.es (G. Herrera), paquifcha-
2008; Amelung et al., 1999; Calderhead et al., 2011; Galloway
con@gmail.com (F. Fernández-Chacón), t.peinado@igme.es (T. Peinado), jpgal-
ve@ugr.es (J.P. Galve), geolovic@gmail.com (V. Pérez-Peña), jose.fernandez@igme. and Burbey, 2011; Papadaki, 2014; Zhu et al., 2015; Farr and Liu,
es (J.A. Fernández-Merodo), j.jimenez@igme.es (J. Jiménez). 2015; Faunt et al., 2016). In Spain, the most arid country in Europe,

http://dx.doi.org/10.1016/j.jhydrol.2017.07.056
0022-1694/Ó 2017 Elsevier B.V. All rights reserved.
72 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

groundwater irrigates around one million hectares (Hernßndez- satellite, Sentinel-1, represents a significant improvement for land
Mora et al., 2015), most of which are concentrated in some inten- subsidence monitoring due to the shorter revisit time (6 days in
sively exploited aquifers in the Mediterranean basin, and particu- the near future), which provides good coherence compared to
larly on the SE fringe of the Iberian Peninsula. Similarly, other SAR sensors (Barra et al., 2016). In this sense, Sentinel-1
literature also reports many cases of land subsidence related to could be a very effective decision-making tool for the aquifer in
groundwater withdrawal in susceptible aquifer systems located coming years.
in SE Spain; they are usually unconsolidated alluvial or basin-fill
aquifers (Herrera et al., 2009; Herrera et al., 2010; Pulido-Bosch
2. The study area
et al., 2012; Rigo et al., 2013; Boni et al., 2015; Notti et al., 2016;
Béjar et al., 2016). This kind of unconsolidated aquifer system usu-
2.1. Geographical setting and climate
ally has great heterogeneity in the distribution of facies and a sig-
nificant proportion of compressible fine-grained materials. When
The Vega the Granada is located on the western flank of the
subjected to water head declines that exceed critical levels, much
Sierra Nevada and is very close to the metropolitan area of the city
of the compaction is related to an inelastic deformation and, con-
of Granada (Fig. 1). It is an almost flat region with an extension of
sequently, the subsidence is permanent (Galloway et al., 2016).
200 km2, with a maximum length of 20 km and an average width
DInSAR and more specifically, Persistent Scatterer Interferome-
of 10 km, the average altitude being 630 m a.s.l. The River Genil
try (PSI) approaches allow monitoring ground subsidence over
flows through the centre of the basin, from SE to NW, and is joined
large areas and for a long period of time. These techniques have
by smaller tributaries (Darro, Beiro, Dílar etc.). Traditionally, the
been successfully applied in numerous regions affected by land
Vega de Granada has been one of the most important agricultural
subsidence where intense exploitation of aquifers occurs (Bell
regions in SE Spain with very fertile soils and both, surficial and
et al., 2008; Herrera et al., 2009; Herrera et al., 2010; González
groundwater resources. Numerous dams regulate the rivers of
and Fernández, 2011; Ezquerro et al., 2014; Tomás et al., 2014;
Granada; the Cubillas and Colomera reservoirs (Fig. 1) meet part
Farr and Liu, 2015; Tessitore et al., 2016; Boni et al., 2016; Chen
of the demand for irrigation water from the fields of the Vega de
et al., 2016).
Granada (Chacon et al., 2012). Additionally, its proximity to the
The present work focuses on the Vega de Granada aquifer,
city of Granada has determined important urban growth during
located in the postorogenic intermontane Basin of Granada (SE
recent decades (since the 80’s) and the population has increased
Spain) (Fig. 1). The Vega de Granada aquifer (with an extension
exponentially in most of the villages located there: Armilla, Chur-
of 200 km2) is one of the largest groundwater reservoirs in Andalu-
riana, Santa Fe, Chauchina, Atarfe, Maracena, and Vegas del Genil
sia and is considered of strategic importance for the economy of
(Fig. 1) etc., which exerts substantial pressure on the aquifer
this semi-arid region. Ground subsidence had already been
(Chica-Olmo et al., 2014).
detected using DInSAR based techniques in the Vega de Granada.
The climate in the region is dry-Mediterranean. The average
Fernandez et al. (2009) exploited ERS images covering a period of
annual precipitation is 450 mm and the average temperature lies
7 years (from June 1993 to December 2000), and obtained defor-
around 15 °C. The warm season lasts from June to September with
mation rates of up to 8 mm/yr in the village of Santa Fe, located
an average daily temperature above 30 °C and with very scarce
in the central part of the aquifer (Fig. 1). They concluded that
precipitation. Rainfall is mainly concentrated during the autumn
‘‘the only process which could be related to the subsidence is an inten-
and winter months and rainfall data series reflect highly variable
sive aquifer exploitation due to crop irrigation and urban water sup-
alternating periods of rain and drought (AEMET, State Meteorolog-
ply”, but the origin was not investigated in depth. Sousa et al.
ical Agency of Spain). Prolonged droughts have occurred in the
(2008), Sousa et al. (2010) and Notti et al., (2016) analysed ground
region during recent decades. During the period spanning 1992–
subsidence processes related to the exploitation of a small-scale
1995, a devastating drought (the most severe of the 20th century)
detrital aquifer on the southeastern edge of the Granada Basin, in
took place and numerous wells had to be drilled in the eastern
the municipality of Otura (Fig. 1). Notti et al. (2016) integrated
fringe of the aquifer to solve urban supply problems. Similar con-
detailed geological and hydrogeological data with differential
ditions (although not so prolonged) occurred again in 2004–
SAR interferometry monitoring from ENVISAT (2003–2009) and
2005, and new wells were drilled to supply the municipalities of
Cosmo-SkyMed (2011–2014) images. They concluded that a clear
Santa Fe, Vegas del Genil and Armilla. However, the period span-
lithological control exists in the spatial distribution of the ground
ning from 2009 to 2013 was extremely rainy in the region, with
subsidence; the sector with highest rates of subsidence (up to 15
accumulated rainfall at almost twice the average value (Mateos
mm/yr) does not correspond to the area with the most intense
et al., 2016).
groundwater exploitation but to the area with the greatest pres-
The aquifer of the Vega de Granada is a water reservoir of
ence of clays.
strategic importance with a very irregular exploitation pattern in
In the present study, multiband PSInSAR measurements have
time: (1) during summer, the water demand rises to very high val-
been used to obtain a detailed spatial and temporal distribution
ues compared to the rest of the year; (2) during periods of drought,
of the ground surface deformation affecting the aquifer of the Vega
the aquifer plays a vital role in supplying water.
de Granada, covering a large temporal span of 13 years (from 2003
to 2016). Synthetic Aperture Radar (SAR) images from the ENVISAT
(2003–2009, C band), Cosmo-SkyMed (2011–2014, X band) and 2.2. Geology
Sentinel-1A (May 2015–April 2016, C band) satellites have been
exploited. Additionally, a thorough analysis of the geology and The Vega de Granada is located in the central sector of the Betic
hydrogeology of the region has been carried out, and also a litho- Cordillera and within the Granada Basin domain. The term ‘‘Gran-
logical description of the aquifer system based on the interpreta- ada Basin” is given to an outcrop of Neogene to Quaternary sedi-
tion of borehole data. The integration of all these ground data ments lying over the NE-SW trending contact between the
with the results obtained from PSInSAR monitoring aims to give External and Internal zones of the Betic Cordillera. The sedimen-
a reliable view of the processes involved in the land subsidence tary sequence is over 2 km thick in some areas, containing lower
as well as to analyse the deformational behaviour induced by Miocene to Holocene sediments. The Tortonian marine sediments
groundwater exploitation which could contribute to a future man- identified in the Basin indicate that the Alborán Sea covered this
agement plan for the aquifer. The European Space Agency (ESA) depression during the late Miocene (Galindo et al., 1999; Rodrí
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 73

Fig. 1. Location of the Vega de Granada aquifer in the intramontane Basin of Granada (SE Spain) and on the western flank of the Sierra Nevada (3479 m). The River Genil flows
through the centre of the flat basin, which is joined by numerous tributaries. Many localities exploit the aquifer not only for irrigation but also for their urban water supply.
The population trend in Vegas del Genil (VG) since 1997 is shown, as an example of the substantial urban growth which has occurred in the region during recent decades.
Additionally, time distribution of SAR imagery from different satellites exploited in the present work is also represented.

guez-Fernández and Sanz de Galdeano, 2006; Rodríguez-Fernán fans. By contrast, in the eastern area of the aquifer, closer to the
dez et al., 2012). From the late Miocene to the present, a rapid Sierra Nevada and the city of Granada, there is a predominance
uplift of the cordilleras has occurred, and a system of coalescent of coarse-grained material (conglomerates) from the alluvial fan
alluvial and delta fans has been deposited along all the range facies (Alhambra Formation, upper Pliocene–lower Pleistocene in
edges; they are predominately coarse-grained deposits which pro- age). To summarise, the Vega de Granada aquifer is a multilayer
grade to the centre of the Basin (Rodríguez- Fernández, 1982; Rodrí detrital deposit with small-scale heterogeneity and frequent
guez-Fernández et al., 2012). Additionally, the differential uplift of lateral-spatial changes of lithofacies.
each area has determined the development of normal faults which For the present research, data from numerous boreholes (38)
delineate the boundaries of the Granada Basin (Azanón et al., 2004; drilled in the study area (FAO/IGME, 1972) have been interpreted.
Rodríguez-Fernández and Sanz de Galdeano, 2006). In this geolog- Results are shown in Fig. 2B where 3 cross-sections have been
ical context, the Vega de Granada aquifer corresponds to detrital drawn up. This study confirms the sedimentary pattern described
sediments filling the tectonic depression, which presents a sedi- above: (1) in the central part of the aquifer, close to the River Genil,
mentary multilayer structure with Quaternary levels of gravel, the aquifer is unconfined and is mainly composed of fluvial grav-
sand, silt and clay, and with a thickness of up to 250 m in the cen- els; some boreholes (N. 12, 13) register a thickness of up to 150
tral part of the aquifer (Luque-Espinar et al., 2008). The central part m of this coarse sediment. (2) In the northern area of the aquifer,
of the plain is dominated by river sedimentation and consequently, where the village of Atarfe is located, a predominance of clays is
coarse-grained sediments (gravel and sands) related to the fluvial recorded, with values of up to 120 m in clay thickness (N. 23).
dynamics clearly predominate in the central axis of the aquifer (3) In the southern part of the aquifer, and where Chauchina and
(Fig. 2A). Additionally, the prograding coalescent alluvial fans, Santa Fe are located, a predominance of fine sediments is also
coming from the surrounding ranges, enter the basin and leave dis- recorded and clay thickness can exceed 65 m (N. 3, 4, 33). Based
tal fine sediments intercalated among coarse-grained sediments on this information, isolines of clay percentage (in the first 50 m
from the river sedimentation. Fig. 2A shows that numerous villages only, as the borehole depths are variable) have been drawn over
are located on this fine-grained substratum (Chauchina, Santa Fe, the geological map (Fig. 2A) and show a higher clay content in
Atarfe, Maracena), corresponding to the distal facies of the alluvial the northern and southern extremes of the aquifer as well as in
74 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Fig. 2. Geology. 2A) Geological map of the area where the Vega de Granada aquifer is located, in the contact between the External and Internal zones of the Betic Cordillera.
The aquifer corresponds to the detritic sediments filling the tectonic depression of the Granada Basin, which presents a sedimentary multilayer structure with Quaternary
levels of gravel, sand, silt and clay, as a result of the coalescence between river sedimentation and that of the alluvial fan. A schematic geological cross-section derived from
Azañón et al. (2004) is included, based on the interpretation of seismic profiles. Three cross sections parallel to the River Genil are shown which correlate the information
obtained from the interpretation of 38 boreholes drilled in the region between 1970 and 1972 (FAO/IGME, 1972). Based on the borehole information, the isolines of clay
percentage (in the first 50 m) are drawn over the geological map 2A.

its central part, where the urban settlements of Caballo Blanco and ered as the steady state conditions of the aquifer. Isolines show a
Vegas del Genil are located. This is a key contribution to knowledge minor recharge area (where the groundwater flows from NE to
about the aquifer that may later explain its spatial response to SW) in the northeastern part of the aquifer (Maracena) charac-
hydraulic head changes and the subsequent vertical land terised by a steeper hydraulic gradient where low permeable
movements. materials exist (Castillo, 1986). The unsaturated zone thickness
ranges from 1 m to 70 m, with the lesser thickness being in the
3. Hydrogeology extreme western area and close to the river channel. The transmis-
sivity values show a great spatial variability as a result of the sed-
3.1. Aquifer characteristics imentary structure of the aquifer. They range from between 100
and 40,000 m2/day, with an average value of 4000 m2/day. The
The Vega de Granada aquifer is of regional importance and has highest values are situated in the eastern part of the aquifer and
been used for agricultural irrigation and the supply of drinking in zones neighbouring the River Genil. The mean effective porosity
water since the second half of the 20th century. With renewable is estimated at 6%, although most values range from 1% to 10%. The
water resources of about 160 hm3/yr (Castillo, 2005), groundwater permeability of the aquifer is greater beneath the River Genil
exploitation has intensified considerably since the 1970s and now (40 m/day) and in the eastern sector, decreasing towards the limits
reaches about 40 hm3 annually (Pardo-Igúzquiza et al., 2015). Gen- of the aquifer (1 m/day). The main recharge sector is located in the
erally, the groundwater flows from east to west, following the southeastern fringe of the aquifer, where most of the riverbeds
direction of the River Genil which is the aquifer’s main drainage (Genil, Monachil, Dílar etc.) merge (Chica-Olmo et al., 2014).
axis (Kohfahl et al., 2008). Fig. 3 shows the first piezometric map Agriculture is the main land use (66% of the surface area), with
of the aquifer drawn up by Trac (1968) and which can be consid- herbaceous irrigation crops, poplar groves and fruit and vegetable
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 75

Fig. 2 (continued)

products (Chica-Olmo et al., 2014). Urban development and indus- Atarfe, in the northern and central part of the aquifer. As regards
trial zones cover 25% of the surface area of the region. Many vil- depth, the deepest wells are located in the southeastern area of
lages and population centres are located in the study area, with the aquifer (with depths of over 100 m) and in the northeastern
ongoing growth in recent years. The Vega de Granada aquifer con- area (50–100 m in depth). The well depth progressively decreases
tributes to the water supply for the metropolitan area of Granada following the groundwater flow direction (towards NW).
(more than 450,000 inhabitants) and represents one of the most
important aquifers in the south of Spain. Its strategic value was 3.2. Hydraulic head changes in the aquifer
revealed during the severe drought of 1992–1995, when 10 emer-
gency wells had to be drilled in the southeastern sector (Fig. 4) to The Regional Water Agency of Andalusia has facilitated infor-
meet urban demand. Almost 10 hm3 of water was drawn out dur- mation regarding the piezometric network in the area (36
ing the 6-month period spanning from June 1995 to December piezometers), and many piezometric time series are available from
1995. The updated well database from the Geological Survey of 2003 to 2015. Additionally, for the present work, a piezometric
Spain (IGME) is shown in Fig. 4, and their use and depth can be campaign was carried out in June 2016 to update groundwater
observed. 36% of the wells are used for irrigation and 14% for urban levels across the aquifer. The piezometric information exploited
water supply. Additionally, many wells were drilled to supply for this research has been summarised in Table 1, indicating, for
water to the industrial parks of Granada, Maracena, Albolote and each piezometer, the maximum depletion level detected for each
76 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Fig. 3. 1967 Piezometric map of the aquifer modified from Trac (1968) which can be considered as the steady state conditions of the aquifer with no exploitation. The
groundwater flows from southeast to northwest, following the direction of the River Genil. The main recharge area is located in the southeastern sector, at the foot of the
Sierra Nevada. Additionally, a recharge area in the northeastern part of the aquifer (Maracena) can be observed and which is characterised by a higher hydraulic gradient.

monitoring time period. To analyse the piezometric level fluctua- ing the past 13 years. Fig. 6A shows the map corresponding to the
tions over time, nine piezometers have been selected (Numbers: period (2003–2009) and three areas with higher water depletion
3, 6, 8, 9, 10, 13, 29, 30, 38) as they have the longest series and can be observed: one is located in the surroundings of Armilla, in
are distributed throughout the aquifer. Piezometric level evolution the southern part of the aquifer, a second is located in the northern
has been represented against monthly precipitation for the period extreme, in the surroundings of Atarfe, and a third is located in the
from January 2002 to June 2016 (Fig. 5), and includes data from the western-central part of the aquifer, between the localities of
nearest rainfall station, the 5530 AEMET weather station (height Fuente Vaqueros, Chauchina and Santa Fe. The highest water head
567 m a.s.l), which is located at Granada airport (Chauchina). Both variation is detected in the second and northernmost area with
parameters present a clear correlation; the piezometric series values of up to 38 m. Fig. 6B shows the results for the period
show a general decreasing tendency starting in December 2004, (2011–2014), the areas with the highest groundwater depletion
when a long, dry period occurred in the region, and levels started are concentrated in the northeastern part of the aquifer (Atarfe
to recover at the end of 2008, and continued to do so from 2009 and Marcena) and in the southern part of Armilla, with values of
to 2014 when a rainy period took place. Since 2013, groundwater up to – 22 m in both cases. Fig. 6C shows the map for the period
levels have shown a stable trend. Piezometers N.13 and N.30 show (2015–2016), it can be observed that the most exploited sector is
hardly any groundwater level variations, as they are located in the concentrated in the southern part of the aquifer, with water head
discharge sector of the aquifer. descents of up to 10 m. An area with lesser water head variations
Taking the information shown in Table 1 into consideration, (values of up to 5 m) is detected between the localities of Atarfe
maximum groundwater depletion maps of the aquifer have been and Maracena for this time-period.
drawn for each radar-monitoring time period. We have applied
geostatistical analysis, and specifically the ordinary kriging method
(Isaaks and Srivastava, 1989), which assumes that the dataset has a 4. PSI analysis and results
stationary variance but also a non-stationary mean value within
the search area (500 m  500 m in the present case). Taking into 4.1. SAR data and processing methodologies
account the available piezometric information, it is considered that
cluster-based ordinary kriging better estimates the spatial variabil- In the present study, three independent sets of images acquired
ity of groundwater decline and global errors are lower than when by three different satellites have been used: the ENVISAT (C-band)
applying other geostatistical methods (Abedini et al., 2008; Pardo- and the Sentinel-1A satellites (C-band) belonging to the European
Igúzquiza et al., 2009; Pardo-Igúzquiza et al., 2015). Maps for the Space Agency (ESA), and the Cosmo-skyMed constellation (X-band)
three periods considered are represented in Fig. 6 and it can be owned by the Italian Space Agency. Table 2. summarises their main
observed that the aquifer has been heterogeneously exploited dur- features and characteristics. The ENVISAT dataset consists of a set
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 77

Fig. 4. Well database from the Geological Survey of Spain in the Vega de Granada aquifer (updated in 2016) indicating their use and depth. The deepest wells are located in
the southeastern part of the aquifer, where the 10 emergency wells drilled during the severe drought in 1992–1995 are located. The well depth decreases towards the NW,
following the aquifer drainage direction. Wells are used for irrigation (36% of the total number) and urban water supply (14% of the total). In the northern and central part of
the aquifer, many wells were drilled to supply water to the industrial parks of Granada, Maracena, Albolote and Atarfe.

of 25 SAR ascending images acquired from May 2003 to December Two different advanced-DInSAR methodologies have been used
2009. From these images, 138 differential interferograms were cal- for processing the data: (1) PSIG Cousins analysis, described in
culated. The Cosmo-SkyMed dataset consists of a set of 20 Devanthéry et al. (2014) applied to both the ENVISAT and
descending images, obtained from May 2011 to February 2014, Cosmo- SkyMed datasets. The key element of this method is given
where 85 interferograms were calculated. The Sentinel-1A dataset by the so-called Cousin PSs (CPSs), which are PSs characterised by a
consists of 25 Interferometric Wide Swath images acquired from moderate spatial phase variation that ensures a correct phase
descending relative orbit 81 between May 15th 2015 and April unwrapping and takes into account the atmospheric phase screen
3rd 2016. The combined use of these datasets has allowed a better to derive the PS deformation velocity and time series. The numbers
defining of the area affected by ground deformation as well as its of interferograms, as well as the selected pairs, were selected by
temporal evolution. considering the temporal coverage of the network (trying to make
Differential Interferometric Synthetic Aperture Radar (DInSAR) the number of interferometric pairs uniform throughout the period
techniques exploit the information contained in the radar phase covered), the temporal baselines (less than 2 years when possible)
of two or more SAR images acquired at different times over the and the geometric baseline of the satellite (below 500 m). (2) the
same area to estimate the phase difference or interferogram. If Sentinel-1A dataset was processed using a simplified PSI approach
the scattering characteristics of the ground remain approximately based on temporally consecutive interferograms (Barra et al., 2016;
the same, the interferogram contained in the interferogram is Crosetto et al., 2016a,b). First the 25 images were coregistered and
equivalent to the change in the instrument-to-ground distance then 24 consecutive interferograms, covering a region of  250 km
(Massonnet and Feigl, 1998). The conventional DInSAR approach x 180 km (27 bursts of three swaths), were produced. The topo-
extracts the deformation signal from individual interferograms graphic phase contribution was corrected using the Shuttle Radar
and allow the detection of fast deformation phenomena or dis- Topography Mission data (Farr et al., 2007). The perpendicular
placements concentrated in time (e.g. cosesimic deformation, baseline has a maximum length of 144 m and the temporal base-
Béjar-Pizarro et al., 2010). PSI methods, on the other hand, exploit line is 12 days in all cases except for two interferograms spanning
multiple SAR images acquired over the same area and allow the 24 and 36 days due to the lack of data covering the region. Using
detection of more subtle deformation processes through the reduc- this short temporal baseline, the coherence of the interferograms
tion of error sources (Ezquerro et al., 2014; Notti et al., 2016). PSI is fully exploited. Taking advantage of the reduced Sentinel-1 orbi-
methods are therefore ideally suited to measure the temporal evo- tal tube, this approach disregards the residual topographic compo-
lution of ground deformation associated with aquifer-system nent. The procedure includes the phase unwrapping of the
compaction. interferograms, a direct integration of the interferometric phases
78 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Table 1 and, the estimation and removal of the atmospheric signals to


Piezometric information exploited for the present research. For each piezometer (see obtain the deformation time series.
location in Fig. 6), the maximum depletion level detected for each time series has
been indicated.
The accuracy of the technique depends on multiple factors,
including the sensor, the number of SAR images, the time over
Piezometer Max. depletion (m) Max. depletion (m) Max. depletion (m) which the images are acquired, the distance from the reference
Number 2003–2009 2011–2014 2014–2016
point, the quality of the reference point, the coherence of the PS,
1 9.55 – – the PS density and quality, etc (Hooper et al., 2012, Crosetto
3 11.55 4.6 1.1
4 3.73 2.1 0.7
et al., 2016a,b). Under ideal conditions (large datasets, high PS den-
5 2.58 1 0.7 sity, minimum atmospheric effects, etc) sub-millimetre accuracy
6 5.79 4.3 1.5 can be obtained (Ferretti et al., 2007) although millimetre accuracy
7 2.4 3 2.3 is usually obtained under more realistic conditions (e.g. Marinkovic
8 11.62 3.8 1.3
et al., 2008, Crosetto et al., 2016a,b). Regarding our specific dataset,
9 16.62 6.7 0.9
10 6.34 8.7 3 the standard deviation values estimated for all the PSs in the urban
11 11.6 11 7.3 area of Granada (stable area) are 0.73, 0.80 and 6.63 for the ENVI-
12 4.16 2.4 2 SAT, CSK and Sentinel-1 data, respectively. The worst performances
13 2.9 3.3 1 on Sentinel-1 are due both to the short monitored period and to
14 5.51 6.5 2.6
the number of available images.
16 10.9 – –
18 16.9 4.4 1 The main results of this combined procedure are (for each net-
19 12.4 6.1 1 work): the interpolated deformation map at each acquisition time
20 3.26 2.5 0.9 and the deformation velocity map derived from the time series
21 8.5 – –
(Fig. 7). The joint analysis of the maps obtained from the different
24 3.82 4.5 2.9
27 – 3.5 0.6 networks allows the delineating of the deformation field, and to
29 10.5 1.3 1.2 locate the maximum deformation areas during the whole period
30 3.14 1.4 1 monitored. Additionally, the spatial and temporal evolution of
31 13.04 6.2 4.7 ground surface displacements can characterize the deformational
35 12.7 7.7 1.2
behaviour of the aquifer.
36 16 7.4 1.1
37 12 7.8 –
38 15 7.3 1.3
4.2. Results
39 16 –
40 9.7 8.3 1.1
41 9.3 – – Fernandez et al. (2009) carried out a DInSAR analysis in the
42 9.3 – – metropolitan area of Granada based on ERS1 and ERS2 images, cov-
43 9.4 7.6 1.2 ering the period from 1993 to 2000. Within the Vega de Granada
44 8.5 1.4 1
aquifer, land subsidence was only detected in the village of Santa

Fig. 5. Piezometric level fluctuations from nine piezometers located in the aquifer and their correlation to monthly rainfall recorded in the study area (from January 2002 to
June 2016). The annual moving average of rainfall has also been represented.
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 79

Fig. 6. Maximum groundwater level depletion maps of the aquifer for each monitoring time period considered in the present work. (A) The map corresponding to the period
2003–2009; (B) the map for the period 2011–2014 and (C) the map for the period 2015–2016. Cluster-based ordinary kriging has been used to draw up the maps. It can be
observed that the aquifer has been heterogeneously exploited during the past 13 years, with higher groundwater extractions in the eastern part of the aquifer (the main
recharge area).
80 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Fig. 6 (continued)

Table 2
Main characteristics of ENVISAT, Cosmo-SkyMed and Sentinel-1A data as well as the processing methodology applied for each dataset.

Parameters/Satellite ENVISAT Cosmo-SkyMED Sentinel-1A


Wavelength C-BAND/5.6 cm X-BAND/3.1 cm C-BAND/5.6 cm
Temporal span May 2003–Dec 2009 May 2011–Feb 2014 May 2015-April 2016
Geometries of acquisitions Ascending Descending Descending
Incidence angle 23° 42° 38°
N° Images 25 20 25
N° of Interferograms used (PSI) 138 85 24
Spatial resolution (approx.) 4  20 m 33m 30  30 m
Methodology of processing PSIG Cousins PSIG Cousins Direct integration approach (Barra et al., 2016)
(Devanthéry et al., 2014) (Devanthéry et al., 2014)

Fe, where deformation rates of up to 8 mm/yr were observed with – A central area of 6 km2, where Vegas del Genil and Caballo
maximum values in the central area, while values declined Blanco are located. Most of the PS show displacement rates
towards the outskirts of the town in an almost concentric pattern between 4 and 6 mm/yr and some also attained values of above
reaching null values outside the urban area. It was indicated in that 10 mm/yr.
study that intensive aquifer exploitation was the only process – A southern fringe, 11 km in length, which includes the localities
which could have been related to the subsidence. The monitoring of Chauchina, Santa Fé and Vegas del Genil, where the subsi-
period covers part of the severe drought which occurred during dence values increase westward and are higher in the localities
the years 1992–1995. of Santa Fe and north-Chauchina (values of 10 mm/yr). It is here
where the highest values of accumulated displacements have
4.2.1. ENVISAT (2003–2009) been recorded (up to 55 mm).
Fig. 7A shows the results of the surface displacement rates
(mm/yr) in the study area for the ENVISAT monitoring period 4.2.2. Cosmo-SkyMed (2011–2014)
(May 2003–Dec 2009). Results of PSI processing show a high den- During this period (May 2011-Feb 2014), most of the previously
sity of PS in urban areas and linear infrastructures. Three subsiding detected areas of subsidence disappeared. Some points of subsi-
areas have been detected within the aquifer: dence were detected in the localities of Chauchina and Cijuela as
well as in some parts surrounding Maracena (Fig. 7B). A small sub-
– A northern fringe, 4 km in length, between Atarfe and Granada, siding fringe of 3 km in length was detected in the northern part of
with higher surface displacement rates concentrated in Atarfe, Cijuela and Chauchina. It is clearly concentrated in the newer
with values of up to 10 mm/yr. urban areas of both localities where the most extended new
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 81

Fig. 7. PSI Vertical ground motion velocities (mm/yr) recorded for the Vega de Granada aquifer for the different monitoring time periods. (A) ENVISAT (May 2003–Dec 2009);
(B) Cosmo-SkyMed (May 2011- Feb 2014); and (C) Sentinel-1A (May 2015- Apr 2016).
82 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Fig. 7 (continued)

construction type is the single family home with private swimming iod considered in the present work. The aim of this combination is
pool. The most relevant area (with displacement rates of up to 11.5 to understand the aquifer ground motion response to hydraulic
mm/yr and accumulated displacements of -14.5 mm) is located in head changes and to determine the lithological control in the spa-
the southern part of Chauchina, where Granada airport is located. tial distribution of the ground subsidence.
It has been verified that numerous earthmoving and fill-in works
were carried out to level the runway and car park area during 5.1. Ground motion versus clay thickness (spatial relationship)
the monitoring period and could be the origin of the deformation
detected. In the area of Maracena, most of the PS show very low The presence of compressible layers (fine-grained sediments) in
rates of surface displacement (between 2 and 5 mm/yr). the aquifer and their thickness is a very relevant factor which
On the other hand, in Fig. 7B we can observe local uplift areas in determines the location of land subsidence processes related to
Santa Fe and Vegas del Genil, with small positive surface displace- groundwater withdrawal (Notti et al., 2016; Calderhead et al.,
ment between +2 and +3 mm/yr. In an area located at the south of 2011; Bru et al., 2013; Zhu et al., 2015). To analyse the spatial rela-
Armilla, there is a great concentration of PS with positive displace- tionship between clay thickness and subsidence, Fig. 8A focuses on
ments (up to +6.5 mm/yr). the ENVISAT period (2003–2009), when the most widespread
ground subsidence was detected. A very clear lithological control
4.2.3. Sentinel-1A (2015–2016) in the spatial distribution of land subsidence is observed. The three
For the period spanning from May 2015 to April 2016, no rele- subsiding areas identified and described in Fig. 7A are located
vant land subsidence was detected in the aquifer area, even the where a higher clay content was identified from boreholes.
points of subsidence detected previously in Chauchina and Cijuela Fig. 8B and C also show occasional points of subsidence in the sur-
have disappeared. Only some small and isolated areas in the east- roundings of Maracena; although borehole information for this
ern sector of Maracena were observed but with the highest dis- sector is not available, geological maps show that the substratum
placement rates ever detected, having values of up to 45 mm/yr. of this area corresponds to distal alluvial facies. The presence of
We suspect that these anomalous values are not related to ground- these fine sediments is corroborated by the original piezometric
water pumping. map shown in Fig. 3 and where the Maracena sector is charac-
terised by a steep hydraulic gradient.
5. Discussion
5.2. Ground motion versus hydraulic head changes (spatial
The information set out and analysed previously is integrated to relationship)
interrelate land subsidence results with geological and hydrogeo-
logical data. Fig. 8 shows the results of this combination, overlap- Fig. 8A, B and C show two main areas of higher groundwater
ping the following information: PSI data, percentage of ground clay depletion for the past 13 years: the northern part of Maracena
content in the first 50 m (isolines) and maximum groundwater and Atarfe, and the southeastern part of the aquifer, in the sur-
depletion maps of the aquifer for each radar-monitoring time per- roundings of Armilla and Granada. Widespread subsidence has
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 83

Fig. 8. Maps with the combination of: PSI data, the isolines of clay percentage (in the first 50 m) and the maximum groundwater level depletion values in the aquifer for each
monitoring time period. (A) ENVISAT (May 2003–Dec 2009); (B) Cosmo-SkyMed (May 2011- Feb 2014); and (C) Sentinel-1A (May 2015-Apr 2016).
84 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

Fig. 8 (continued)

not been detected in any of these areas for any period, only occa- lated displacement curve is steeper during the period 2004–2005,
sional spots in the surroundings of Maracena. Nevertheless, wide- which coincided with the driest period for the last 13 years (Fig. 5).
spread subsidence was detected in the central part of the aquifer Piezometers 1 and 6 are located in areas with higher clay content
during the ENVISAT period (2003–2009), Fig. 8A, where smaller and abrupt changes in both curves can be observed. The cumulated
changes in groundwater level took place (maximum depletions displacement changes even under small variations in groundwater
below 12 m). It has been confirmed that the ground subsidence level (<2 m).
detected in Cijuela and Chauchina for the Cosmo-SkyMed period For the Cosmo-SkyMed analysis, data from piezometer N.5 has
(2011–2014), Fig. 8B, when minor groundwater level depletion been exploited; it being the only one located in the area with land
occurred (below 1 m in this sector, piezometer N.5), is related to subsidence detected during this period (Fig. 10). Although there
engineering works carried out at that time (earthmoving and fill- are many gaps in the piezometric evolution, the water head level
in works). Sentinel 1A results, Fig. 8C, indicate points of minor changes are not significant (around 1 m) and correlation between
ground subsidence to the east of Maracena, where groundwater groundwater level descents and the SAR measured displacements
level depletion values lower than 5 m were recorded. They are can be observed. This may support the anthropogenic origin of
located in an industrial area where additional information was the land subsidence. In fact, there is a subsidence peak (up to 3
not available. mm) during the first year of the monitoring period (2011–2012),
which later returns to very low values. However, these values are
5.3. PSI deformation time series versus hydraulic head changes close to the detection limits of the technique.
Looking at some plots in Figs. 9 and 10 we can observe some
The average vertical displacement based on the time series was uplift related to a rise in the water table (starting in 2008). It can
estimated by using buffer areas with a radius of 1000 m for each be considered as part of the elastic behaviour of the aquifer, in
piezometer (Ezquerro et al., 2014; Béjar et al., 2016), and consider- agreement with the uplift areas detected during the Cosmo-
ing all PS time series included in the area. Groundwater level data SkyMed monitoring period (Fig. 7B).
of a group of selected piezometers (those with a longer time series)
were later compared with the average vertical displacement of the 5.4. Contributions to management of the aquifer
time series within each buffer area. Results are shown in Fig. 9, for
the ENVISAT data (2003–2009), and in Fig. 10 for the Cosmo- The present work aims to contribute to the drawing up of a
SkyMed data (2011–2014). Sentinel 1A results have not been eval- management plan for the sustainable exploitation of the detrital
uated as only a piezometric campaign is available for this short aquifer of the Vega de Granada, taking into account land subsi-
time period (May 2015–April 2016). dence processes for the first time. The results have been discussed
For the ENVISAT analysis, the data from 6 piezometers (N. 3, 5, and the following points should be considered:
8, 29, 1, 6) located within the areas where subsidence occurred has
been used. A general, good correlation is observed between  During periods of drought, there are greater groundwater with-
groundwater level descents and the increase in the average of ver- drawals in the areas of the aquifer which are vulnerable to sub-
tical displacements (Fig. 9). For the six cases analysed, the accumu- sidence, that is, where the localities of Cijuela, Chauchina, Santa
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 85

Fig. 9. Comparison of water table changes with ground motion changes at six piezometers located in the areas with land subsidence identified for the ENVISAT period (May
2003- Dec 2009). We have considered all the PS time series included in a buffer area of 1000 m for each piezometer and the average displacement has been calculated. A good
general correlation is observed between the two parameters. For the drought period 2004–2005 the average displacement curve is steeper.

shorter revisit time. In this sense, the high spatial and temporal
resolution of Sentinel-1 A may contribute to a better defining of
the deformation pattern of the aquifer and it could be an effec-
tive decision-making tool in the near future.

6. Conclusions

Fig. 10. Comparison of water table changes with ground motion changes at In the present work multiband PSInSAR results for the past
piezometer N.5, located in the SW area of the aquifer and near the localities of
13 years and geological and hydrogeological data have been com-
Cijuela and Chauchina, where points of land subsidence were detected during the
Cosmo-SkyMed period (May 2011-Feb 2014). No correlation is observed between bined to obtain a better understanding of the ground subsidence
the two parameters, which supports the anthropogenic origin of the subsidence. affecting the aquifer of the Vega de Granada, one of the largest
groundwater reservoirs in SE Spain and is considered strategic
for the economy of this semi-arid region. The aquifer has been used
Fe, Vegas del Genil, Caballo Blanco, Atarfe and Maracena are for agricultural irrigation and currently, 36% of the total number of
located. The compaction detected in these areas is attributed wells are used for agricultural supply. Additionally, the aquifer is
to the inelastic deformation of the fine-grained sediments, part of the Granada metropolitan area, where rapid urban growth
and much of the accompanying subsidence is permanent. This has occurred over recent decades, which has caused a substantial
effect could produce legal and economic problems in the near increase in urban water supply requirements. New pumping wells
future and must be addressed. were drilled for this purpose in some towns located on the aquifer
 The need to develop a plan for the conjunctive use of surface (Vegas del Genil, Santa Fe, Chauchina, etc.); they represent 14% of
and groundwater during periods of drought and which takes the total number of wells. The Vega de Granada aquifer corre-
into account the operating thresholds (critical levels) of ground- sponds to detrital sediments filling a tectonic depression, which
water level depletion to avoid land subsidence in these vulner- presents a multilayer deposit with Quaternary levels of gravel,
able areas should be considered. sand, silt and clay. The aquifer has small-scale heterogeneity and
 A short-term management plan is required for the progressive frequent lateral-spatial changes of lithofacies as a result of the coa-
replacement of the urban supply wells in many vulnerable lescence between the sedimentation of the river and the alluvial
localities. The groundwater pumping pattern in the aquifer fan.
needs to be reorganised and, in this sense, the Water Adminis- The main outcomes of this research are:
tration may reinforce groundwater exploitation in the SE
recharge area of the aquifer where the drought emergency wells 1. Interpretation of data from numerous boreholes drilled in
are located. This area is not vulnerable to land subsidence, as the study area reveals a higher clay content in the central
coarse sediments predominate in the substratum. part of the aquifer where numerous villages are located:
 The need to carry out a specific study in the areas surrounding Cijuela, Chauchina, Santa Fe, Atarfe, Caballo Blanco, Vegas
Maracena, where new points of subsidence have lately been del Genil etc. In these sectors, the aquifer is partially con-
identified. A detailed hydrogeological and geological study has fined. Additionally, in the northeastern part of the aquifer
to be carried out to determine the origin of this subsidence. (Maracena), fine-grained sediments corresponding to distal
 The piezometric control network in the aquifer has to be alluvial facies also outcrop and a steep hydraulic gradient
improved. Some strategic areas are poorly covered and there is detected.
are many gaps in the time series. 2. There is a good correlation between rainfall pattern and
 Sentinel-1 imagery provides an accurate tool for the rapid esti- piezometric trend in the aquifer. Piezometric time series
mation of even small ground movements in the region due to its show a general decreasing tendency starting in December
86 R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87

2004, when a long, dry period occurred in the region, and related to earthmoving and fill-in works. There is no correla-
levels begin to recover at the end of 2008, and continue to tion between groundwater level descents and the increase in
do so from 2009 to 2014 when a rainy period took place. ground surface displacements.
The dry episode coincides with the ENVISAT monitoring per- 10. The causes of the ground subsidence identified in the NE
iod (2003–2009) and the highest values of water table des- part of Maracena during the Sentinel period (2015–2016)
cent were recorded during this time. are undocumented. They are located in new industrial zones
3. The aquifer has been heterogeneously exploited during the and no specific information regarding them is available.
past 13 years. Two main areas of groundwater depletion 11. A management plan for the sustainable exploitation of the
have been identified: the northern part of Maracena and detrital aquifer of the Vega de Granada has to be drawn up
Atarfe, and the southeastern part of the aquifer, in the sur- and must address the existence of land subsidence pro-
roundings of Armilla and Granada. Both sites are located in cesses. Sentinel-1 could be an effective decision-making tool
the recharge areas of the aquifer and the deepest wells are in the near future.
there located.
4. The subsidence monitoring has been carried out by applying
two different methodologies for processing the radar data:
PSIG Cousins analysis, applied to ENVISAT and Cosmo- Sky- Acknowledgements
Med datasets, and a simplified PSI approach for the
Sentinel-1A dataset. This combination allows an in-depth The research leading to these results has received funding from
analysis of the ground deformation pattern and both tempo- the Spanish National Research Plan ‘‘RETOS 2013” under Grant
ral and spatial dimensions of the subsidence for the past Agreement N. ESP2013-47780-C2-2-R- AQUARISK Project. Addi-
13 years are obtained. tionally, part of the work was carried out with the support of the
5. SAR results show an inelastic deformation of the aquifer- VIII Spanish National Research Plan under LITHOSURF project
system and small, accumulated land surface displacements Grant Agreement N. CGL2015-67130-C2-1-R.
(up to -55 mm). The most widespread land subsidence was
detected during the ENVISAT period (2003–2009) in three
References
areas of the aquifer, with low values of vertical displacement
velocities (up to 10 mm/yr); the highest values were Abedini, M.J., Nasseri, M., Ansari, A., 2008. Cluster-based ordinary kriging of
recorded in the locality of Santa Fe and north-Chauchina. piezometric head in West Texas/New Mexico – Testing of hypothesis. J. Hydrol.
351, 360–367.
These values are similar to those obtained previously in
Amelung, F., Galloway, D.L., Bell, J.W., Zebker, H.A., Laczniak, R.J., 1999. Sensing the
Santa Fe by Fernandez et al. (2009) for the period spanning ups and downs of Las Vegas: InSAR reveals structural control of land subsidence
from 1993 to 2000 (deformation rates of 8 mm/yr). and aquifer-system deformation. Geology 27, 483–486.
6. Most of the subsiding areas observed during the ENVISAT Azanón, J.M., Azor, A., Booth-Rea, G., Torcal, F., 2004. Small-scale faulting,
topographic steps and seismic ruptures in the Alhambra (Granada, southeast
period (2003–2009) disappeared during the Cosmo- Spain). J. Quaternary Sci. 19, 219–227.
SkyMed period (2011–2014), but the deformation previ- Barra, A., Monserrat, O., Mazzanti, P., Esposito, C., Crosetto, M., Scarascia Mugnozza,
ously detected is permanent and did not recover during this G., 2016. First insights on the potential of Sentinel-1 for landslides detection.
Geomatics Nat. Hazard. Risk. 1 1.
rainy period. Subsidence was only detected in the northern Béjar-Pizarro, M., Carrizo, D., Socquet, A., Armijo, R., Barrientos, S., Bondoux, F.,
part of Cijuela and Chauchina, the most relevant point being Bonvalot, S., Campos, S., Comte, D., Chabalier, J.B., Charade, O., Delorme, A.,
located at the Granada airport site (with deformation rates Gabalda, G., Galetzka, J., Genrich, J., Nercessian, A., Olcay, M., Ortega, F., Ortega,
I., Remy, D., Ruegg, J.C., Simons, M., Valderas, C., Vigny, C., 2010. Asperities,
of up to 11.5 mm/yr and accumulated displacements of barriers and transition zone in the North Chile seismic gap: State of the art after
14.5 mm). For the Sentinel period (2015–2016) no relevant the 2007 Mw 7.7 Tocopilla earthquake inferred by GPS and InSAR data.
land subsidence was detected in the aquifer, only some Geophys. J. Int. 183, 390–406.
Béjar, M., Guardiola-Albert, C., García-Cárdenas, R.P., Herrera, G., Barra, A., López
points in the eastern sector of Maracena but with the highest Molina, A., Tessitore, S., Staller, A., Ortega-Becerril, J.A., García-García, R.P., 2016.
displacement rates ever detected, with values of up to 45 Interpolation of GPS and geological data using InSAR deformation maps:
mm/yr. method and application to land subsidence in the alto guadalentín aquifer (SE
Spain). Remote Sens. 8, 965–982.
7. There is a clear lithological control in the spatial distribution
Bell, J.W., Amelung, F., Ferretti, A., Bianchi, M., Novali, F., 2008. Permanent scatterer
of ground subsidence in the aquifer. For the ENVISAT period InSAR reveals seasonal and long-term aquifer-system response to groundwater
(2003–2009), when the most widespread ground subsidence pumping and artificial recharge. Water Resour. Res. 44 (2), W02407.
was detected, the subsiding areas identified are clearly Boni, R., Herrera, G., Meisina, C., Notti, D., Béjar-Pizarro, M., Zucca, F., González, F.,
Palano, M., Tomás, R., Fernández, J.J., Fernández-Merodo, J.A., Mulas, J., Aragón,
located where a higher clay content was identified. The com- R., Guardiola-Albert, C., Mora, O., 2015. Application of multi-sensor advanced
paction is attributed to the inelastic deformation of the fine- DInSAR analysis to severe land subsidence recognition: Alto Guadalentín Basin
grained sediments. Occasional points of subsidence detected (Spain). Proc. IAHS 2015 (374), 45–48.
Boni, R., Cigna, F., Bricker, S., Meisina, C., Mc Cormarck, H., 2016. Characterization of
later in the surroundings of Maracena are also located in an hydraulic head changes and aquifer properties in the London Basin using
area with a fine-grained substratum. By contrast, the SE sec- Persistent Scatterer Interferometry groundmotion data. J. Hydrol. 540, 835–849.
tor of the aquifer had more intense groundwater exploita- Bru, G., Herrera, G., Tomás, R., Dur, J., De la Vega, R., Mulas, J., 2013. Control of
deformation of buildings affected by subsidence using persistent scatterer
tion but no land subsidence processes were identified as interferometry. Struct. Infrastruct. Eng. 9, 188–200.
coarse-grained sediments predominate in the substratum. Calderhead, A.I., Therrien, R., Rivera, A., Marte, L.R., Garfias, J., 2011. Simulating
8. For the ENVISAT period (2003–2009), which coincided with pumping-induced regional land subsidence with the use of InSAR and field data
in the Toluca Valley, Mexico. Adv. Water Resour. 34, 83–97.
a dry, long episode, there is a good correlation between Castillo, A., 1986. Estudio hidroquímico del acuífero de la Vega de Granada PhD
headwater level depletion and the augmentation in ground manuscript. University of Granada. 658.
surface displacements. In areas identified as having a higher Castillo A, 2005. El acuífero de la Vega de Granada. Ayer y hoy (1966-2004). In:
Agua, Minería y Medio Ambiente, Libro Homenaje al Profesor Rafael Fernández
clay content and where the highest rates of subsidence were
Rubio, López Geta et al. (Eds.), 2005 IGME pp. 161-172.
recorded, small changes in headwater level (<2 m) produce a Chacon, J., Irigaray, C., El Hamdouni, R., Valverde-Palacios, I., Valverde-Espinosa, J.,
rapid ground motion response. Calvo, F., Lamas, F., 2012. Engineering and environmental geology of granada
9. The points of subsidence detected in the localities of Chau- and its metropolitan area (Spain). Environ. Eng. Geosci. 18, 217–260.
Chen, M., Tomás, R., Li, Z., Motagh, M., Li, T., Hu, L., Gong, H., Li, X., Yu, J., Gong, X.,
china and Cijuela during the Cosmo-SkyMed monitoring 2016. Imaging land subsidence induced by groundwater extraction in Beijing
period (2011–2014) had an anthropogenic origin, being (China) using satellite radar interferometry. Remote Sens. 8 (6), 468.
R.M. Mateos et al. / Journal of Hydrology 553 (2017) 71–87 87

Chica-Olmo, M., Luque-Espinar, J.A., Rodríguez-Galiano, V., Pardo-Igúzquiza, E., and karstic environments: a field study in the Granada Basin (Southern Spain).
Chica-Rivas, L., 2014. Categorical Indicator Kriging for assessing the risk of Appl. Geochem. 23, 846–862.
groundwater nitrate pollution: the case of Vega de Granada aquifer (SE Spain). Luque-Espinar, J.A., Chica-Olmo, M., Pardo-Igúzquiza, E., García-Soldado, M.J., 2008.
Sci. Total Environ. 470–471, 229–239. Influence of climatological cycles on hydraulic heads across a Spanish aquifer. J.
Crosetto, M., Monserrat, O., Devanthéry, N., Cuevas-González, M., Barra, A., Crippa, Hydrol. 354, 33–52.
B., 2016. Persistent scatterer interferometry using Ssentinel-1 data. In: The Marinkovic, P., Ketelaar, G., van Leijen, F., Hanssen, R., 2008. InSAR quality control:
International Archives of the Photogrammetry, Remote Sensing and Spatial analysis of five years of corner reflector time series. European Space Agency,
Information Sciences, Volume XLI-B7. 2016 XXIII ISPRS Congress, 12–19 July (Special Publication) ESA SP-649.
2016, Prague, Czech Republic. Doi: 10.5194/isprsarchives-XLI-B7-835-2016. Massonnet, D., Feigl, K.L., 1998. Radar interferometry and its application to changes
Crosetto, M., Monserrat, O., Cuevas-González, M., Devanthéry, N., Crippa, B., 2016. in the earth’s surface. Rev. Geophys. 36, 441–500.
Persistent Scatterer Interferometry: A review, ISPRS Journal of Photogrammetry Mateos, R.M., Azañón, J.M., Roldán, F.J., Notti, D., Pérez-Peña, V., Galve, J.P., Pérez-
and Remote Sensing, (115) 78-89, ISSN 0924-2716, https://doi.org/10.1016/j. García, J., Colomo, C.M., Gómez-López, J.M., Montserrat, O., Devantèry, N.,
isprsjprs.2015.10.011. Lamas-Fernández, F., Fernández-Chacón, F., 2016. The combined use of PSInSAR
Devanthéry, N., Crosetto, M., Monserrat, O., Cuevas-González, M., Crippa, B., 2014. and UAV photogrammetry techniques for the analysis of the kinematics of a
An approach to persistent scatterer interferometry. Remote Sensing (6) 6662- coastal landslide affecting an urban area (SE Spain). Landslides. DOI 10.1007/
6679. s10346-016-0723-5.
Ezquerro, P., Herrera, G., Marchamalo, M., Tomás, R., Béjar-Pizarro, M., Martínez, R., Motagh, M., Walter, R.T., Sharifi, M.A., Fielding, E., Schenk, A., Anderssohn, J., Zschau,
2014. A quasi-elastic aquifer deformational behavior: Madrid aquifer case J., 2008. Land subsidence in Iran caused by widespread water reservoir over
study. J. Hydrol. 519, 1192–1204. exploitation. Geophys. Res. Lett. 35 (16), 403–412.
FAO/IGME, 1972. Proyecto piloto de utilización de aguas subterráneas para el Notti, D., Mateos, R.M., Oriol, M., Devanthéry, N., Peinado, T., Roldán, F.J., Fernández-
desarrollo agrícola de la cuenca del Guadalquivir, España. Utilización de las Chacón, F., Galve, J.P., Lamas, F., Azañón, J.M., 2016. Lithological control of land
aguas subterráneas para la mejora del regadío en la Vega de Granada. Inf. subsidence induced by groundwater withdrawal in new urban areas (Granada
Técnico n° 2. AGL: SF/SPA 16. 218 p. Internal report, unpublished. Basin, SE Spain) Multiband DInSAR monitoring. Hydrol. Processes 30, 2317–
Farr, T.G., Rosen, P., Caro, E., Crippen, R., Duren, R., Hensley, S., Alsdorf, D., 2007. The 2331.
shuttle radar topography mission. Rev. Geophys. 2007, 45. Papadaki, E.S., 2014. Monitoring subsidence at Messara basin using radar
Farr, T.G., Liu, Z., 2015. Monitoring subsidence associated with groundwater interferometry. Environ. Earth Sci. 72, 1965–1977.
dynamics in the Central Valley of California using interferometric radar. In: Pardo-Igúzquiza, E., Chica-Olmo, M., Garcia-Soldado, M.J., Luque Espinar, J.A., 2009.
Lakshmi, V. (Ed.), Remote sensing of the terrestrial water Hydrogeol J cycle. Using semivariogram parameter uncertainty in hydrogeological applications.
Geophysical Monograph 206, American Geophysical Union, Washington, DC, pp Ground Water 47, 25–34.
397–406. Pardo-Igúzquiza, E., Chica-Olmo, M., Luque-Espinar, J.A., Rodríguez-Galiano, V.,
Faunt, C.C., Sneed, M., Traum, J., Brandt, J.T., 2016. Water availability and land 2015. Compositional cokriging for mapping the probability risk of groundwater
subsidence in the Central Valley, California, USA. Hydrogeol. J. 24, 675–684. contamination by nitrates. Sci. Total Environ. 532, 162–175.
Fernandez, P., Irigaray, C., Jimenez, J., El Hamdouni, R., Crosetto, M., Monserrat, O., Pulido-Bosch, A., Delgado, J., Sola, F., Vallejos, A., Vicente, F., López-Sánchez, J.M.,
Chacon, J., 2009. First delimitation of areas affected by ground deformations in Mallorquí, J.J., 2012. Identification of potential subsidence related to pumping in
the Guadalfeo River Valley and Granada metropolitan area (Spain) using the the Almeria Basin (SE Spain). Hydrol. Process. 26, 731–740.
DInSAR technique. Eng. Geol. 105, 84–101. Rigo, A., Béjar-Pizarro, M., Martínez-Díaz, J., 2013. Monitoring of Guadalentín valley
Ferretti, A., Savio, G., Barzaghi, R., Borghi, A., Musazzi, S., Novali, F., Prati, C., Rocca, F., (southern Spain) through a fast SAR Interferometry method. J. Appl. Geophys.
2007. Submillimeter accuracy of InSAR time series: experimental validation. 91, 39–48.
IEEE TGRS 45 (5), 1142–1153. Rodríguez- Fernández, J., 1982. El Mioceno del sector central de las Cordilleras
Galindo, J., Jabaloy, A., Serrano, I., Morales, J., González-Lodeiro, F., Torcal, F., 1999. Béticas. Ph D thesis, University of Granada, 224 p.
Recent and present-day stresses in the Granada Basin (Betic Cordilleras): Rodríguez-Fernández, J., Sanz de Galdeano, C., 2006. Late orogenic intramontane
Example of a late Miocene-present-day extensional basin in a convergent plate basin development: The Granada basin, Betics (Southern Spain). Basin Res. 18,
boundary. Tectonics 18, 686–702. 85–102.
Galloway, D.L., Burbey, T.J., 2011. Review: land subsidence accompanying Rodríguez-Fernández, J., Miguel Azañón, J., Azor, A., 2012. The betic intramontane
groundwater extraction. Hydrogeol. J. 19, 1459–1486. basins (SE Spain): stratigraphy, subsidence, and tectonic history. Tectonics
Galloway, D.L., Erkens, G., Kuniansky, E., Rowland, J., 2016. Preface: land subsidence Sedimen Basins Recent Adv., 461–479
processes. Hydrogeol. J. 24, 547–550. Tessitore, S., Fernández-Merodo, J.A., Hererra, G., Tomás, R., Ramondini, M.,
González, P.J., Fernández, J., 2011. Drought-driven transient aquifer compaction Sanabria, M., Duro, J., Mulas, J., Calcaterra, D., 2016. Comparison of water-
imaged using multitemporal satellite radar interferometry. Geology 39, 551– level, extensometric, DInSAR and simulation data for quantification of
554. subsidence in Murcia (SE Spain). Hydrogeol. J. 24, 727–747.
Hernández-Mora, N., Martínez Cortina, L., Llamas, R., Custodio, E., 2015. Tomás, R., Romero, R., Mulas, J., Marturià, J.J., Mallorquí, J.J., López-Sánchez, J.M.,
Groundwater in the Southern Member States of the European Union: an Blanco, P., 2014. Radar interferometry techniques for the study of ground
assessment of current knowledge and future prospects. Country report for subsidence phenomena: a review of practical issues through cases in Spain.
Spain. http://www.easac.eu/fileadmin/PDF_s/reports_statements/ Environ. Earth Sci. 71, 163–181.
Spain_Groundwater_country_report.pdf. On 14 october 2016. Trac, N., 1968. Mapa hidrogeológico de la Vega de Granada (escala 1:25.000). FAO-
Herrera, G., Fernández, J.A., Tomás, R., Cooksley, G., Mulas, J., 2009. Advanced IGME report. Unpublished.
interpretation of subsidence in Murcia (SE Spain) using A-DInSAR data- Sousa, J.J., Ruiz. A.M., Hanssen, R.F., Perski, Z., Bastos, L., Gil, A.J, Galindo-Zaldívar, J.,
modelling and validation. Nat. Hazard. Earth Syst. Sci. 9, 647–661. 2008. PS-INSAR measurement of ground subsidence in Granada area (Betic
Herrera, G., Tomás, R., Monells, D., Centolanza, G., Mallorquí, J.J., Vicente, F., Navarro, Cordillera, Spain). In: Proceedings of the 13th FIG International Symposium on
V.D., López-Sánchez, J.M., Sanabria, M., Cano, M., Mulas, J., 2010. Analysis of Deformation Measurements and Analysis, Lisbon, Portugal, pp 12-15.
subsidence using TerraSAR-X data: Murcia case study. Eng. Geol. 116, 284–295. Sousa, J.J., Ruiz, A.M., Hanssen, R.F., Bastos, L., Gil, A., Galindo-Zaldívar, J., Sanz de
Hooper, A., Bekaert, D., Spaans, K., Arıkan, M., 2012. Recent advances in SAR Galdeano, C.S., 2010. PS-InSAR processing methodologies in the detection of
interferometry time series analysis for measuring crustal deformation. field surface deformation—Study of the Granada basin (Central Betic Cordilleras,
Tectonophysics 514, 1–13. southern Spain). J. Geodyn. 49, 181–189.
Isaaks, E.H., Srivastava, R.M., 1989. An Introduction to Applied Geostatistics. Oxford Zhu, L., Gong, H., Teatini, P.L., Wang, R., Chen, B., Dai, Z., 2015. Land Subsidence due
University Press, New York, p. 561p. to groundwater withdrawal in the northern Beijing plain, China. Eng. Geol. 193,
Kohfahl, C., Sprenger, C., Herrera, J.B., Meyer, H., Chacón, F.F., Pekdeger, A., 2008. 243–255.
Recharge sources and hydrogeochemical evolution of groundwater in semiarid

You might also like