Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Science of the Total Environment 636 (2018) 670–687

Contents lists available at ScienceDirect

Science of the Total Environment

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

SAR interferometry monitoring of subsidence in a detritic basin related to


water depletion in the underlying confined carbonate aquifer
(Torremolinos, southern Spain)
A. Ruiz-Constán a,⁎, A.M. Ruiz-Armenteros b,c,d, S. Martos-Rosillo a, J. Galindo-Zaldívar e,f, M. Lazecky g,
M. García e, J.J. Sousa h, C. Sanz de Galdeano f, J.M. Delgado-Blasco c,i, P. Jiménez-Gavilán j,
M. Caro-Cuenca k, J.A. Luque-Espinar a
a
Instituto Geológico y Minero de España, Urb. Alcázar del Genil 4, Edf. Zulema bajo, 18006 Granada, Spain
b
Departamento de Ingeniería Cartográfica, Geodésica y Fotogrametría, Universidad de Jaén, Campus Las Lagunillas s/n, 23071 Jaén, Spain
c
Grupo de Investigación Microgeodesia Jaén, Universidad de Jaén, Campus Las Lagunillas s/n, 23071 Jaén, Spain
d
Centro de Estudios Avanzados en Ciencias de la Tierra (CEACTierra), Universidad de Jaén, Campus Las Lagunillas s/n, 23071 Jaén, Spain
e
Departamento de Geodinámica, Universidad de Granada, 18071 Granada, Spain
f
Instituto Andaluz de Ciencias de la Tierra, CSIC-Universidad de Granada, 18071 Granada, Spain
g
IT4Innovations VSB-TU Ostrava, Czech Republic
h
Universidade de Trás-os-Montes e Alto Douro, INESC-TEC (formerly INESC Porto), Vila Real, Portugal
i
Progressive System s.r.l., Rome, Italy
j
Departamento de Ecología y Geología, Facultad de Ciencias, Universidad de Málaga, Málaga, Spain
k
Department of Radar Technology, TNO, The Netherlands

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Influence of intensive exploitation of a


confined carbonate aquifer in ground
deformation of the overlying sedimen-
tary infill
• Accumulated descent of the confined
water level at Sierra de Mijas of 140 m
during the period 1992-2009.
• Subsidence related to decrease in inter-
stitial pressure due to water withdrawal

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

Article history: This research underlines the need to improve water management policies for areas linked to confined karstic
Received 20 November 2017 aquifers subjected to intensive exploitation, and to develop additional efforts towards monitoring their subsi-
Received in revised form 21 March 2018 dence evolution. We analyze subsidence related to intensive use of groundwater in a confined karstic aquifer,
Accepted 20 April 2018
through the use of the InSAR technique, by the southern coast of Spain (Costa del Sol). Carbonates are overlain
Available online 1 May 2018
by an unconfined detritic aquifer with interlayered high transmissivity rocks, in connection with the Mediterra-
Editor: R Ludwig nean Sea, where the water level is rather stable. Despite this, an accumulated deformation in the line-of-sight
(LOS) direction greater than −100 mm was observed by means of the ERS-1/2 (1992–2000) and Envisat

⁎ Corresponding author at: Instituto Geológico y Minero de España, Unidad de Granada, Urb. Alcázar del Genil. Edificio Zulema, bajo., 18006 Granada, Spain.
E-mail address: a.ruiz@igme.es (A. Ruiz-Constán).

https://doi.org/10.1016/j.scitotenv.2018.04.280
0048-9697/© 2018 Elsevier B.V. All rights reserved.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 671

Keywords: (2003–2009) satellite SAR sensors. During this period, the Costa del Sol experienced a major population increase
Confined carbonate aquifer due to the expansion of the tourism industry, with the consequent increase in groundwater exploitation. The
Detritic aquifer maximum LOS displacement rates recorded during both time spans are respectively −6 mm/yr and
Betic Cordillera
−11 mm/yr, respectively. During the entire period, there was an accumulated descent of the confined water
Rapid urbanization
Radar
level of 140 m, and several fluctuations of more than 80 m correlating with the subsidence trend observed for
InSAR the whole area. Main sedimentary depocenters (up to 800 m), revealed by gravity prospecting, partly coincide
with areas of subsidence maxima; yet ground deformation is also influenced by other factors, the main ones
being the fine-grained facies distribution and rapid urbanization due to high touristic pressure.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction Subsidence related to karst aquifer exploitation has usually been re-
ported in relation to areas where surface and subsurface dissolutions
The intensive use of groundwater from carbonate aquifers results in form cavities and natural sinkholes, filled or not by unconsolidated de-
many adverse consequences (Custodio, 2002; Llamas and Custodio, posits, which could collapse (e.g. Beck, 1984; Beck et al., 1999; Salvati
2003). Among them, we can underline the reduction of natural dis- and Sasowsky, 2002; Waltham et al., 2007; García-Moreno and
charge or even the disappearance of springs, the reduction of subsurface Mateos, 2011; Rucker et al., 2013). Areas related to individual sinkholes
discharge towards rivers or wetlands and subsequent environmental ef- are relatively small —diameters ranging from a meter to hundreds of
fects, contamination due to the mixture of polluted or brackish water, meters— but they have been widely documented, considered a signifi-
land subsidence and sinkhole collapses (Rosenthal, 1988; Cardenal cant geological hazard causing significant damage and economic ex-
et al., 1994; Pulido-Bosch et al., 1995; Martos-Rosillo et al., 2009; pense ($300 million/year in USA; Weary, 2015). Nevertheless,
Rodríguez-Estrella, 2012). These effects have a direct influence on soci- subsidence in carbonate aquifers may also be related to water depletion
ety owing to increased groundwater operating costs and user conflicts, if they are confined by a multilayer detritic formation with interlayered
among others (Foster, 1992; Van der Gun and Lipponen, 2010). aquitard levels (Stringfield and Rapp, 1976). This process is not well

Fig. 1. Geographic location of the study area in western Europe (A) and in southern Spain (B).
672 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

described and only few examples have been documented, such as those were afterwards confirmed, for the 2003–2008 period, through the pro-
of the Savannah area (Georgia, USA) with up to 100 mm of subsidence cessing of an Envisat ASAR dataset with the SBAS InSAR Service of
after 33 m of water level decline (Davis et al., 1963; Poland and Davis, European Space Agency's Geohazard Exploitation Platform (GEP), an
1969), or the Montellano aquifer (S Spain) with 33 mm of subsidence unsupervised MT-InSAR analysis (Galve et al., 2017). In this study, we
related to 43 m of groundwater depletion (Ruiz-Constán et al., 2016). apply spaceborne SAR to estimate and analyze ground motion in the
The Costa del Sol in southern Spain (Fig. 1), has undergone great Benalmádena-Torremolinos area using SAR images from the European
population growth in recent decades, manly during 1990–2010, with ERS-1/2 (1992–2000) and Envisat (2003–2009) satellites. We also per-
more than 4% population increase per year (INE, 2015). This phenome- formed a gravity survey that allowed us to locate thickness variations in
non is particularly important in the towns of Torremolinos and the sedimentary infill that could condition the observed subsidence pat-
Benalmádena (Málaga province), the focus of our research. Each of tern. The complete dataset was compared to detect areas of the alluvial
these towns had around 25,000 inhabitants in 1990; in 2015, each of sedimentary infill that are most seriously affected by induced subsi-
their populations had grown to 67,000 inhabitants, which is a 160% in- dence due to intensive use of the karstic aquifer. These results shed
crease in just 25 years. Besides, the summer season sees a massive ar- light on the need to implement management plans for a sustainable
rival of tourists, with an additional population increase of 60% in July use of this important aquifer, given the influence of groundwater deple-
and August. The high population density (around 3000 inhabitants/ tion in the subsidence of neighboring areas.
km2) relies heavily on groundwater pumping to meet water demand
under unbridled urban growth. In this study, we look at the water 2. Geological and hydrogeological setting
level evolution of the aquifers supplying freshwater to the towns of
Torremolinos and Benalmádena, the so-called Sierra de Mijas and Bajo Sierra de Mijas and the Málaga Basin (Fig. 2) are located along the
Guadalhorce aquifers (carbonate and detritic aquifers, respectively) in Betic Cordillera, the main relief of southern Spain. Sierra de Mijas is con-
order to determine their degree of overexploitation during the stituted by metamorphic rocks that are unconformably overlain by
1992–2009 period, when the major population increase took place. Neogene-Quaternary sediments (Piles Mateo et al., 1978; Estévez-
Spaceborne SAR has proven to be an excellent technique for study- González and Chamón Cobos, 1978) that constitute the sedimentary
ing ground motion caused by underground water flow of large areas infilling of the Málaga Basin. From a hydrogeological point of view, the
(Galloway et al., 1998; Amelung et al., 1999; Hoffmann et al., 2001; study area comprises part of the aquifers of Sierra de Mijas (carbonate
Colesanti et al., 2003; Schmidt and Burgmann, 2003; Caro Cuenca, aquifer) and Bajo Guadalhorce (detritic aquifer).
2012). Subsidence in several areas of the Costa del Sol was first identi-
fied (Ruiz et al., 2011) applying Multi-Temporal InSAR techniques 2.1. Sierra de Mijas carbonate aquifer
(MT-InSAR) to an ERS-1/2 SAR dataset (1992–2000) through the
Stanford Method for Persistent Scatterers – Multi-Temporal Interferom- The stratigraphic succession of Sierra de Mijas is formed at the bot-
etry (StaMPS MT-InSAR). It was mainly located in the towns of tom by more than 400 m of Paleozoic rocks (schists and gneisses) su-
Benalmádena, Torremolinos and in the SW of Málaga. These results perposed by more than 500 m of Triassic marbles which form the

Fig. 2. Geological map (A) and lithological column (B) of Sierra de Mijas carbonate aquifer and Bajo Guadalhorce detritic aquifer. Modified from 1066 and 1067 MAGNA 50 geological maps
(Piles Mateo et al., 1978; Estévez-González and Chamón Cobos, 1978). White dotted-lines represent the location of the main towns while white dashed-line indicates the sediment/
basement boundary. Black lines locate the cross-sections represented in Fig. 7.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 673

main reliefs. The total outcropping surface of the calcareous massif is 5–800 m2/day, average permeability is 10−3 m/s and the effective po-
80 km2 and its permeability is mainly related to fractures and rosity is 5–10% (Vadillo et al., 2007). During rainy periods, there are
karstification processes (Andreo, 2007). The S and SW boundaries of wetlands at the river mouth, but the water level significantly descends
the carbonate aquifer are constituted by the cited metapelites, forming in dry periods. Marine intrusion is detected mostly in the right margin
a reverse contact; meanwhile, at the E and N boundaries the marbles of the river, through the high permeable Pliocene sandy levels and the
plunge below the detritic sediments of the Bajo Guadalhorce aquifer, alluvial, where pumping is greater (Carrasco et al., 2012, Urresti-Estala
also cut by faults that contribute to their sinking. et al., 2016). The study area comprises the SW part of the Malaga
The sequence is deformed by ESE-WNW to ENE-WSW recumbent Basin and the E border of Sierra de Mijas (Figs. 1 and 2).
northward-verging folds and NNO-SSE normal faults (Andreo and
Sanz de Galdeano, 1994). These structures and the progressive decline
of the relief, mainly to the E, condition the existence of springs at differ- 3. Hydrogeological and meteorological analysis
ent heights: 400–485 m.a.s.l. in Mijas; 280–300 m.a.s.l. in Alhaurín el
Grande; 210–220 m.a.s.l. in Benalmádena; 125–145 m.a.s.l. in Alhaurín We analyzed piezometric data from 32 piezometers from the
de la Torre; and 55–74 m.a.s.l. in Torremolinos. Most of them are now Instituto Geológico y Minero de España (IGME, 2017) and the
dry as a consequence of excessive pumping. Water resources are esti- Demarcación Hidrográfica de las Cuencas Mediterráneas Andaluzas
mated in ~28 hm3/yr, while pumping is around 35 hm3/yr (BOE, (DHCMA, 2017) control networks. Piezometers with longer time series
2016), surpassing the resources of the system and causing an accumu- and with fewer information gaps were selected (Figs. 2 and 3): six of
lated water table decrease of 80–100 m in the period 1997–2003 them located at the Bajo Guadalhorce detritic aquifer (D1–D6) and 15
(Andreo, 2007). in Sierra de Mijas (C1–C15). In addition, we analyzed rainfall data at
The specific flow rate at Sierra de Mijas is around 10–100 l/s/m, with the Torremolinos meteorological station (TMS; Fig. 2) from the Agencia
transmissivity values of 103 m2/day. Spring flow variations are cush- Estatal de Meteorología network (AEMET, 2017). In order to obtain a
ioned in natural regime by the rainfall, with peak floods registered sev- continuous register for the 1992–2009 period, these data were com-
eral months after precipitation (Andreo et al., 2002a). Recharge of the pleted through simple linear regression with those of the Benalmádena
aquifer is from rainfall infiltration; discharge, at present, is chiefly re- and Alhaurín de la Torre AEMET stations (BMS and AMS; Fig. 2). The
lated to pumping activity in Sierra de Mijas (Andreo et al., 1996), but accumulated deviation of monthly precipitation was obtained, as it
also to lateral flow to the neighboring Bajo Guadalhorce detritic aquifer. provides an estimation of the relative deficit or surplus of precipitation
Lateral flow values range from 10 to 15 hm3/yr (Vadillo et al., 2007) to (Paccum in Fig. 3). This value is computed evaluating the average
25 hm3/yr (IGME, 1983a). The water has calcium magnesium bicarbon- monthly precipitation over a long-term period (this study,
ate facies without important hydrochemical variations, although its 1992–2009) and then, the obtained value is successively subtracted
quality is worsening because of agricultural impact and urban develop- from each monthly rainfall value. Major drought periods could be easily
ment (Andreo et al., 2002a). distinguished as they correspond to downward trends in the rainfall ac-
cumulated deviation curve, for example those comprised between 1992
2.2. Guadalhorce detritic aquifer and 1995 and 1998–2002. Annual fluctuations correspond to the wet
and dry periods of the same hydrological year. This parameter usually
The aquifer is constituted by Miocene to Quaternary sediments evolves in parallel to the temporal evolution of the piezometric level
(IGME, 1983b) cropping out at the depression where the Guadalhorce in aquifers in natural regime.
River flows (Málaga Basin) and continuing offshore. The basement of Control points near the coastline show lower water table fluctua-
the basin is constituted by the Triassic marbles —and the Paleozoic tions (e.g. D5, D6) than those placed far from the coast (e.g. D2 and
rocks under them— of Sierra de Mijas and Sierra de Cártama, in the D3) due to their proximity to the base level of the aquifer and the
southern and central sectors. Northwards, the basement is formed by high transmissivity of the alluvial sediments (Fig. 3). The average
Paleozoic slates and greywackes of Montes de Málaga and, tectonically water table change in coastal piezometers is of 2 m (D1–D6). It should
over them, Cretaceous-Tertiary flysch clays, also including lower Mio- also be pointed out that the cumulative level drops observed at the
cene formations (Figs. 1 and 2). The sedimentary infill (Fig. 2B) is detritic aquifer during the period 1995–1996 (D3; minimum of −4 m.
slightly deformed and it is subdivided into several aquifer formations a.s.l.) correspond to the increase of groundwater extractions at that
(Vadillo et al., 2007; Nieto et al., 2016) which, from bottom to top, are: area during an intense drought period (Vadillo et al., 2007; Carrasco
et al., 2012), the worst in the last 50 years. These declines are rapidly off-
i) Upper Miocene aquifer: calcarenites and conglomerates with set during the rainy season, and the water table remains roughly at the
maximum thickness of 100 m. In the northern border of Sierra same level for the whole period.
de Mijas, these rocks are in contact with the marbles and con- On the other hand, the continuous decrease of the piezometric level
fined by Pliocene marly sediments (Andreo et al., 2002b). How- at the carbonate aquifer is not entirely compensated by recoveries in the
ever, they were eroded in many areas even before the Pliocene. wet season, achieving historic minima in each drought period. During
ii) Lower Pliocene aquifer: Pliocene deposits formed by 10–70 m of the 80s, the water table marked by most piezometers at the carbonate
basal conglomerates generally confined by Upper Pliocene marls, aquifer was at 60–70 m.a.s.l. After a dry period, it decreased to 40 m.a.
although they are locally open laterally. s.l., but then fully recovered in 1989–1990 (Andreo, 2007). The drought
iii) Upper Pliocene aquifer: 400 m of marly series and alternations of period between 1990 and 1995 produced a descent of 80 m in the water
gravels and sandy levels (20 m) partially connected with the al- column (−10 m.a.s.l.) with another important recovery in the following
luvial aquifer through wells. two years (50 m.a.s.l.). Since then, a noteworthy descent is observed
iv) Quaternary alluvial aquifer: alluvial deposits of the Guadalhorce until reaching −70 m.a.s.l. in 2009, representing a rate of 11 m/yr. To
River with a maximum thickness of 50 m (Vadillo et al., 2007). sum up, a descending trend of 8 m/yr is observed for the whole period
(1992–2009). The evolution of every particular piezometer is reflected
in Fig. 3, although for correlation purposes (Figs. 8 and 11) we have
Depths of the pumping wells are generally lower than 50 m; water is used the one with the most complete time series (C10). These data
obtained from the Quaternary alluvial aquifer (26 hm3/yr; BOE, 2016) allow us to affirm that the carbonate aquifer shows an excellent recov-
and, to a lesser extent, from the gravel alternations of the Upper Plio- ery capacity in the rainy seasons. Yet since the 90s the recovery capacity
cene aquifer (6 hm3/yr). Flow rates are around 40 l/s (Carrasco et al., is surpassed by the pumping activity, resulting in continuous water re-
2003), transmissivity values of the alluvial aquifer are in the range serve consumption.
674 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

Fig. 3. Groundwater level evolution (black line) of the carbonate (A) and detritic (B) aquifers together with accumulated deviation of monthly precipitation (dark blue line), for the period
1992–2010. Groundwater level is expressed in meters above sea level (m.a.s.l.) and location of the piezometers is marked in Fig. 2.

4. InSAR analysis recent decades, to model atmosphere delays of the radar signal and pro-
vide for a precise estimation of displacements using a larger number of
4.1. SAR datasets SAR images as input, hence an unprecedented precision of one millime-
ter per year, or even better (Ferretti et al., 2007a). For a general review
We used two independent sets of SAR scenes acquired by satellites of InSAR, the reader is referred to Bamler and Hartl (1998), Rosen et al.
belonging to the European Space Agency (ESA) with sensors operating (2000), Ferretti et al. (2007b), Zhou et al. (2009), Hetland et al. (2012),
in C-band (~5.7 wavelength, 5.3 GHz) to delineate areas of ground mo- Crosetto et al. (2016), and Pepe and Calò (2017), among others.
tion using interferometric synthetic aperture radar (InSAR). The first
dataset consisted of a stack of 30 ERS-1/2 SAR scenes acquired in as-
Table 1
cending mode along the satellite track 230, frame 729, between October ERS-1/2 SAR data for Torremolinos area (Track 230 ascending, Frame 729). Parameters are
1992 and November 2000 (Table 1). The second dataset consisted of 20 relative to the master imagen acquired on 13/08/1997 (text in bold).
Envisat ASAR scenes from the same track and frame as ERS-1/2, ob-
No. Acq. date Orbit Sensor B⊥ (m) Btemp (days) fdc (Hz)
tained from November 2002 to February 2009 (Table 2). Both SAR
datasets, characterized by a 35-day nominal repeat cycle, have an 23° 1 19,921,004 6387 ERS-1 −380,4 −1774 −357
2 19,930,606 9894 ERS-1 −271,8 −1529 −309
incidence angle of the radar signal from the vertical direction and a 5
3 19,930,711 10,395 ERS-1 172 −1494 −298
× 25 m (azimuth × range) nominal pixel dimension. Due to ERS-2 on- 4 19,930,815 10,896 ERS-1 −482,5 −1459 −358
board gyroscope problems in January 2001, only images until the end 5 19,930,919 11,397 ERS-1 −448,7 −1424 −344
of 2000 were selected to avoid high Doppler centroid differences over 6 19,931,024 11,898 ERS-1 −144,4 −1389 −367
the critical value of 700 Hz (Fig. S1 in Supplementary material). 7 19,931,128 12,399 ERS-1 261,8 −1354 −373
8 19,950,425 19,757 ERS-1 −371,6 −841 −369
9 19,950,704 20,759 ERS-1 504,2 −771 −321
4.2. Processing methodology 10 19,950,705 1086 ERS-2 404,8 −770 −564
11 19,951,121 22,763 ERS-1 378,5 −631 −378
By design, SAR acquisitions are processed with so-called focusing al- 12 19,951,122 3090 ERS-2 259,3 −630 −611
13 19,960,305 24,266 ERS-1 −152,3 −526 −301
gorithms, to arrive at images where two radio signal characteristics are
14 19,961,002 7599 ERS-2 252,6 −315 −578
represented within each image pixel; it shows the intensity of reflection 15 19,970,326 10,104 ERS-2 94,2 −140 −538
of the radio wave sent by SAR antenna and its phase, i.e. a portion of 16 19,970,813 12,108 ERS-2 0 0 −605
reflected wavelength. Both parameters are used within SAR interferom- 17 19,971,022 13,110 ERS-2 73,8 70 −603
etry (InSAR), though the phase value is crucial by the interferometric 18 19,971,231 14,112 ERS-2 −568,2 140 −586
19 19,980,311 15,114 ERS-2 −97 210 −600
(phase) combination of two well coregistered SAR images taken from 20 19,980,520 16,116 ERS-2 360,7 280 −573
the same track in a specific time delay (at least 35 days in the case of 21 19,980,729 17,118 ERS-2 −556,6 350 −572
ERS-1/2 and Envisat), an interferogram image is created. The interfero- 22 19,981,007 18,120 ERS-2 −289,6 420 −579
gram contains information about phase differences for single resolution 23 19,981,216 19,122 ERS-2 −296,9 490 −533
24 19,990,713 41,801 ERS-1 −126,7 699 −236
cells, normally linked to surface displacements over the observed area,
25 19,990,714 22,128 ERS-2 118,5 700 −545
together with other physical characteristics affecting the phase signal, 26 19,991,130 43,805 ERS-1 826 839 −354
including topography and atmosphere artefacts (Hanssen, 2001). Un- 27 20,000,419 26,136 ERS-2 −135 980 −722
wanted signals can be removed, for example, using external informa- 28 20,000,628 27,138 ERS-2 −725,4 1050 −464
tion such as a digital elevation model (DEM). To achieve accurate 29 20,000,906 28,140 ERS-2 −196,6 1120 −360
30 20,001,115 29,142 ERS-2 407,9 1190 −1087
displacement maps, MT-InSAR techniques have been developed in
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 675

Table 2 material), 1774 days being the maximum temporal baseline (Table 1,
Envisat ASAR data for Torremolinos area (Track 230 ascending, Frame 729). Parameters and Fig. S1A and B in Supplementary material). For Envisat scenes, per-
are relative to the master image acquired on 02/02/2005 (text in bold).
pendicular baselines range from −654.5 to 597.4 m (Table 2, and
No. Acq. date Orbit Sensor B⊥ (m) Btemp (days) fdc (Hz) Fig. S1C in Supplementary material); Doppler centroid baselines range
1 20,021,120 3791 ASAR −361,5 −805 −553 from −26.6 to 39.6 Hz (Fig. S1D in Supplementary material) with re-
2 20,030,723 7298 ASAR −263 −560 −510 spect to the master scene dated February 2, 2005, in this case
3 20,030,827 7799 ASAR −189,5 −525 −561 1470 days being the maximum temporal baseline (Table 2, and
4 20,031,105 8801 ASAR 84,2 −455 −576
Fig. S1C and D in Supplementary material).
5 20,031,210 9302 ASAR 442,9 −420 −514
6 20,040,324 10,805 ASAR −325,6 −315 −547 Fig. 4A shows the relationship between temporal and spatial base-
7 20,040,428 11,306 ASAR −105,2 −280 −556 lines for the ERS-1/2 dataset used in this study, with both PS (dashed
8 20,040,602 11,807 ASAR −654,5 −245 −537 red lines) and SB (green lines) processing. The same baseline distribu-
9 20,040,707 12,308 ASAR 430,8 −210 −560 tion is shown in Fig. 4B for Envisat dataset. Based on the PS processing,
10 20,040,811 12,809 ASAR 597,4 −175 −561
11 20,050,202 15,314 ASAR 0 0 −550
we constructed 29 and 19 interferograms, respectively for ERS-1/2 and
12 20,050,831 18,320 ASAR 231,5 210 −542 Envisat datasets. The interferograms for the SB processing (80 for ERS-1/
13 20,051,109 19,322 ASAR −294,7 280 −516 2 and 45 for Envisat) were selected in view of the following criteria:
14 20,060,607 22,328 ASAR 278,2 490 −512 maximum coherence = 0.5, maximum temporal baseline =
15 20,060,920 23,831 ASAR −253 595 −520
1500 days, and maximum perpendicular baseline = 1070 m. These
16 20,061,025 24,332 ASAR 166,1 630 −512
17 20,070,207 25,835 ASAR −150,7 735 −516 thresholds are optimized for C-band data and used by default in StaMPS.
18 20,070,801 28,340 ASAR 317,8 910 −513 To estimate accurate deformation maps, highly coherent pixels in
19 20,080,402 31,847 ASAR 340,2 1155 −521 the scene are required. StaMPS pixel selection is based on a converged
20 20,090,211 36,356 ASAR −110,1 1470 −514 probability distribution, computed as presented in Hooper et al.
(2007). The selection strategy consists of finding the selection threshold
of the probability distribution for each PS candidate. If the converged
One major limitation is that conventional estimation methods such probability distribution is greater than the selection threshold, the
as the least squares method cannot be directly applied to InSAR mea-
surements, because the observations —phase differences— are wrapped.
A wrapped phase is the 2π-modulus of the absolute (or unwrapped)
phase. Therefore, the number of wave periods between the satellite
and observed reflector cannot be known. For this reason, techniques A
of phase unwrapping must be introduced in order to achieve proper dis-
placement information, especially if the rate of displacements is rela-
tively high, e.g. Caro Cuenca et al. (2011). Taking into account other
phase-affecting characteristics and the number of available SAR images
used in processing, the phase unwrapping algorithms are considered
successful in cases when displacement does not exceed a quarter of
the SAR wavelength (Hanssen, 2001).
A second limitation stems from the fact that the signal is affected by
changes in the area of interest that are not related to ground displace-
ments. Since the radar wavelength is 5.65 cm, the phase signature of a
ground displacement can be overlain by a disturbance from the motion
of leaves, vegetation growth, snow cover, or similar sources. This signal
appears as noise, and it is referred to as a temporal decorrelation.
A rate of decorrelation is characterized by a parameter called inter-
ferometric coherence. Decorrelated (low coherent) areas are not only
noisy but also very problematic for a proper phase unwrapping, and
stand as the main source of errors. MT-InSAR algorithms developed spe- B
cific approaches to overcome this problem. The Persistent Scatterers
(PS) technique (Ferretti et al., 2001) combines all SAR images with a
common reference image (master image) and processes them for a se-
lected set of PS points that are not decorrelated and show a persistent
stable reflection, such as bare rocks and built structures whose signal
dominates their decorrelating surroundings. Another approach known
as Small Baselines (SB) relies on interferometric combinations of SAR
images with the shortest temporal differences possible (Berardino
et al., 2002); this approach involves a set of pixels that slowly
decorrelate in time (SDFP pixels).
The PS and SB approaches are complementary, and their combina-
tion is able to extract a signal with greater coverage than that of either
technique alone (Hooper, 2008; Ferretti et al., 2011). In this work, we
use the open-source StaMPS MT-InSAR package (Hooper et al., 2012,
2013) to perform MT-InSAR analysis using a combination of both
these basic techniques (PS + SB). For PS processing, the ERS-1/2 scenes
Fig. 4. Temporal vs spatial baseline distribution for the SAR datasets used in this study (A
have perpendicular baselines ranging from −725.4 to 826 m (Table 1,
for ERS-1/2 and B for Envisat). The red lines represent the single master interferograms
and Fig. S1A in Supplementary material) with respect to the central ref- used for PS processing while green lines represent the small baseline ones. Master
erence image (master) dated August 13, 1997. Doppler centroid base- image is symbolized as a black star (August 13, 1997, acquisition 16, Table 1, for ERS-1/
lines range from −482.2 to 368.2 Hz (Fig. S1B in Supplementary 2; February 2, 2005, acquisition 11, Table 2 for Envisat).
676
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687
Fig. 5. Spatio-temporal evolution of the accumulated line-of-sight (LOS) displacement for the combined (PS + SB) processing referred to the first data of the area of Fig. 2. All the single master interferograms are ordered by date from left to right and
top to bottom (A for ERS-1/2 and B for Envisat).
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 677

Fig. 6. Mean SAR line-of-sight (LOS) velocity, assuming a linear deformation rate, superimposed over a Google Earth image (A for ERS-1/2 and C for Envisat) of the area of Fig. 2. (B and
D) Standard deviation for mean LOS velocity. Separate results for PS and SB processing are shown in E and F for ERS-1/2 and in G and H for Envisat.

candidate will be initially selected as a coherent pixel, and otherwise it the processing chain described in greater detail in Hooper (2006,
will be discarded. For the initial topographic estimation, and later on for 2008, and 2010), Hooper et al. (2004, and 2007) and Sousa et al.
terrain geocoding, we used heights from a DEM provided by the Instituto (2010, 2011), spatio-temporally unwrapped phases of single pixels
Geográfico Nacional with 25 m resolution (IGN, 2017a). After applying were used to estimate satellite look-angle errors and atmosphere

Table 3
Short statistics for the processed SAR datasets. Fig. S2 is included in the Supplementary material.

Satellite ERS-1/2 Envisat

Sensor ERS-1/2 SAR Envisat ASAR


Acquisition period 1992/10/04–2000/11/15 2002/11/20–2009/02/11
Number of scenes 20 30
AOI dimension (km2) 56.11
Band C
Wavelength (cm) ~5.7
Incident angle (°) 23
Orbital Track Ascending
Azimuth × slant range resolution (m) 4×8
Multilook factor (Az/Rg) 11
Georeference accuracy (m) 25
Minimum temporal span between two 35
acquisitions (day)
Processing method PS SB Combined (PS + SB) PS SB Combined (PS + SB)
(PS/SDPF/PS + SDPF) density (PS/km2) ~129 ~162 ~252 ~172 ~336 ~422
Mean LOS velocity = [−1.5, 1.5] mm/yr ~81 ~84 ~82 ~62 ~57 ~61
(% population) (Figs. 6E and S2A) (Figs. 6F and S2C) (Figs. 6A and S2E) (Figs. 6G and S2G) (Figs. 6H and S2I) (Figs. 6C and S2K)
Mean LOS velocity = [−2.5, 2.5] mm/yr ~94 ~94 ~95 ~81 ~71 ~76
(% population) (Figs. 6E and S2A) (Figs. 6F and S2C) (Figs. 6A and S2E) (Figs. 6G and S2G) (Figs. 6H and S2I) (Fig. 6C and S2K)
σMean LOS velocity b ±1.5 mm/yr (% population) 100 ~88 ~97 ~99 ~25 ~52
(Fig. S2B) (Fig. S2D) (Figs. 6B and S2F) (Fig. S2H) (Fig. S2J) (Figs. 6D and S2L)
σMean LOS velocity b ±2.5 mm/yr (% population) 100 100 100 100 ~89 ~94
(Fig. S2B) (Fig. S2D) (Figs. 6B and S2F) (Fig. S2H) (Fig. S2J) (Figs. 6D and S2L)
678 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

phase screens (APS). Finally, estimated displacement time series within data, respectively, for the ERS-1/2 and Envisat datasets. The mean LOS
pixels and final maps of an average velocity could be created. velocity patterns together with their standard deviations are shown in
Fig. 6A and B for the ERS-1/2 dataset, and in Fig. 6C and D for Envisat
4.3. InSAR results (both for the combined (PS + SB) processing). Individual results for
PS and SB processing are shown in Fig. 6E and F for ERS-1/2 and in
We applied a MT-InSAR technique spanning from October 1992 Fig. 6G and H for Envisat. A NNE-SSW subsidence bowl is seen over
to November 2000 (ERS-1/2) and from November 2002 to February this area during the ERS-1/2 period, with a maximum deformation
2009 (Envisat), to outline the ground deformation behavior in the rate of −6 mm/yr. The same pattern is detected in the period covered
Benalmádena-Torremolinos area. Fig. 5A and B show the spatio- by the Envisat data, reaching a maximum subsidence of −11 mm/yr,
temporal evolution of the accumulated line-of-sight (LOS) surface dis- which represents an acceleration in the deformation. In both cases the
placement for the whole analyzed period with reference to the first reference area, that is, the area assumed to be stable for phase

Fig. 7. Temporal evolution of line-of-sight (LOS) displacement along the cross-sections indicated in Fig. 6A (EA-EA′, EB-EB′ and ED-ED′) and 6C (AA-AA′, AB-AB′ and AD-AD′).
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 679

Fig. 8. Displacement time series resolved by ERS-1/2 (A) and Envisat (B) ascending observations for PS (PS + SDFP) points located in Fig. 7 compared with the water level of the C10
piezometer (carbonate aquifer) expressed in meters above sea level (m.a.s.l.).

unwrapping, is depicted by a white triangle in the bottom left part of Table 3. The PS distributions for the mean LOS velocity and its standard
Fig. 6A, C, E, F, G, and H. This area was selected manually as it corre- deviation, for the three different processing, are shown in Fig. S2A to F
sponds to a stable urban zone located over basement rocks. We con- for ERS-1/2 (Supplementary material), and in Fig. S2G to L for Envisat
firmed its stable behavior with different processing tests and it (Supplementary material).
corresponds to the average of all the PS inside a radius of 50 m. The al- The mean LOS velocities for most of our ERS-1/2 PS (PS + SDFP)
gorithm selected a total of 14,165 and 23,672 PS (PS + SDFP) pixels in population (82–91%) lie within a ±1.5–2 mm/yr interval, with a mean
the cropped area shown in Fig. 6, resulting in average densities of value of 0.23 mm/yr (Fig. S2E in Supplementary material), and inside
~252 and ~422 PS/km2, respectively for ERS-1/2 and Envisat processing. a ±1–1.5 mm/yr interval (74–97%) in the case of the mean LOS velocity
The main statistics of the processing datasets, including separate values standard deviations (Fig. S2F in Supplementary material). For Envisat
for PS, SB and the combined (PS + SB) processing, are summarized in dataset, we estimated the mean LOS velocities to have a mean value of

Fig. 9. Bouguer (A) and residual (B) anomaly maps. Modeled profiles (in red) and gravity stations located in the sedimentary infill (black triangles) and basement rocks (white triangles)
are marked.
680 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

−1.1 mm/yr, and show a larger concentration of subsidence values in Fig. 7, where the evolution of LOS displacements along the cross-
the range from −2.5 to −10 mm/yr (Fig. S2K in Supplementary mate- sections indicated in Fig. 6A (EA′, EB′ and ED′) and B (AA′, AB′ and AD′)
rial) than in the case of ERS-1/2 (Fig. S2E in Supplementary material) are depicted. Each profile represents an isochronous cumulative LOS dis-
due to the increased size of the subsidence bowl in this period, as evi- placement line with respect to the first data. Plotted in Fig. 8 are displace-
dent in Fig. 6. However, a large population percentage of PS still lies ment time series for some of the areas with the highest subsidence rates
within the interval of ±2 mm/yr (70%) (Fig. S2K in Supplementary ma- in Fig. 6A and C. Most noteworthy are areas EA-3, EA-4, EB-2, ED-2, ED-3,
terial). In the case of the mean LOS velocity standard deviations for ED-5, and ED-8 for the ERS-1/2 period (Fig. 8A), and AA-5, AA-6, AA-7,
Envisat dataset, 84% of PS population is inside a ±2 mm/yr interval AB-2, AB-3, AB-4, AD-4, AD-5, AD-7, and AD-8 areas for Envisat
(Fig. S2L in Supplementary material). In our case, a stability threshold (Fig. 8B). They represent the average deformation of an area with a radius
around ±2 mm/yr is assumed, as it is the value commonly considered of 50 m.
in the literature for C-band radar data (Meisina et al., 2008; Cigna In either period, the subsidence pattern is quite similar for most of
et al., 2014; Del Ventisette et al., 2014; Notti et al., 2014; Ciampalini the areas. Yet in the case of ERS-1/2, the subsidence trend is represented
et al., 2015; Frangioni et al., 2015; Novellino, 2015). Therefore, a PS by two clusters (Fig. 6A). The first, close to Benalmádena, comprises
with a mean LOS velocity outside this threshold is classified as unstable. areas EA-3, EA-4, ED-2, ED-3, and ED-5. They present a first period,
Spatial analysis of the subsidence distribution shows the subsiding from 1992 to mid-1995, with a subsidence of about −40 mm and
sectors to be located over highly urbanized areas parallel to the coast. with rates in the order of −15 mm/yr; and a second period, from mid-
An overview of the sectors subsiding the most is more clearly seen in 1995 to the end of 2000, with a deformation profile showing a gentle

Fig. 10. Two dimensional gravity models and mean SAR line-of sight (LOS) velocity curves for ERS-1/2 (blue line) and Envisat (red line) data. The position of the profiles is marked in Figs. 2,
6, 9, and 12.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 681

and almost linear slope of roughly −10 mm of subsidence at a rate of the main reliefs of the Betic Cordillera (−160 mGal), which increase to-
−2 to −3 mm/yr. The second cluster, located at the intersection of wards the Málaga coast, where they become positive (60–80 mGal) as a
cross lines ED-ED′ and EB-EB′, contains areas EB-2 and ED-8, showing result of the thinning continental crust and the presence of peridotite
a more linear subsidence of −20 to −25 mm over the entire period bodies (Torné et al., 1992; Ayala, 2013; Pedrera et al., 2016). On the
(1992–2000), and with a rate in the order of −3 to −4 mm/yr. In the other hand, relative minima correspond to the presence of low-
case of Envisat, all the areas displayed in Fig. 8B show a similar subsi- density rocks, as those constituting sedimentary basins (Insua-Arévalo
dence trend, almost linear for the whole period (2002–2009), of ap- et al., 2004; Ruiz-Constán et al., 2009).
proximately −30 to −50 mm, with a rate of −5 to −9 mm/yr. Estimation of the regional anomaly in this zone is complex, as it
varies from E-W at Sierra de Mijas to a NE-SW trend northwards, and
5. Gravity data again E-W in the surroundings of Málaga (Torné et al., 1992). In this
study, an E-W trending regional anomaly was determined and
Gravity prospecting is a convenient technique for defining the ge- subtracted from the Bouguer anomaly (Fig. 9B) by kriging gravity data
ometry of a basin, if sufficient density contrast exists between sedi- located at the basement rocks. Residual anomaly values range between
ments and the basement rocks. It entails measurement of the −14 and 4 mGal. Five gravity profiles (Fig. 10) were performed crossing
gravitational field of the Earth at different points of its surface in order the subsidence maxima obtained through the InSAR technique (EA, EB,
to identify anomalies due to local high- or low-density bodies. In this ED, AA, AB and AD). Density contrast considered between the sedimen-
study, acquisition was performed using a Scintrex Autograv CG-5 grav- tary infill and the basement rocks is in the range of 0.52–0.72 g/cm3,
ity meter. Raw data were corrected and processed to calculate complete based on the average value for each lithology (Telford et al., 1991) and
Bouguer anomalies (Fig. 9A), by applying standard corrections to the previous gravity studies in other sedimentary basins of the Spanish
measured values at each station. We applied a standard density of southern coast (Duque et al., 2008 and Martínez-Martos et al., 2016), re-
2.67 g/cm3 and considered the GRS67 Geodetic Reference System spectively. The sedimentary infill shows an irregular morphology with
(Eckerman et al., 2017; Mandal et al., 2015). The absolute gravity high thickness changes, which were not reported before due to the ab-
value was calculated taking as reference the Granada base station of sence of previous geophysical data in the study area. The main feature is
the Instituto Geográfico Nacional's gravimetric network (IGN, 2017b). a furrow parallel to the coast that, depending on the applied density
The 5 m IGN digital terrain elevation model (IGN, 2017a) and a 200 contrast, shows an estimated maximum thickness of 450–800 m at
m-resolution bathymetric database (General Bathymetric Chart of the the northern part of the study area (profiles EB-EB′ and AB-AB′),
Oceans, GEBCO) were used. whereas the average thickness of the basin is around 200–400 m
We acquired a total of 209 gravity data (Figs. 2 and 9) along several (Fig. 10). There are few boreholes in the sedimentary infill, most of
profiles covering the whole study area. Data spacing was established them very shallow. Most lie near the contact with the carbonate aquifer
taking into account the subsidence pattern, with data spaced 300 m and show thicknesses of 30–40 m (C7 and C12, Fig. 3). The highest sed-
on average, but around 200 m at the subsidence maxima. The 2D gravity imentary thicknesses are seen in C13 (300 m) and C1 (190 m), but do
models (Fig. 10) were obtained by means of GRAVMAG V.1.7 software, not cut the top of the carbonate aquifer.
from the British Geological Survey (Pedley et al., 1993). It allows for re-
moval of the regional trend (yellow lines, Fig. 9A) caused by the regional 6. Discussion
geological structure of Sierra de Mijas, and isolation of the residual
anomaly related to the sedimentary rocks (Fig. 9B). To better discern the relationship between subsidence and water
The Bouguer anomaly map of the Iberian Peninsula (IGN, 1976; withdrawal, we superposed the water table evolution in the carbonate
Ayala, 2013) reveals that the area has negative values associated with aquifer (C10) to the deformation trends respectively observed for the

Fig. 11. (A) Coefficient of determination R2 (expressed as a percentage) showing correlation between subsidence at PS points located in Fig. 6. Correlation among the normalized
piezometric level and the normalized ERS-1/2 (B) and Envisat (C) deformation trends.
682
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687
Fig. 12. Residual anomaly isolines and depocenters location superimposed over the mean SAR LOS velocity for ERS-1/2 (A) and Envisat (B), assuming a linear deformation rate.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 683

ERS-1/2 (1992–2000) and the Envisat (2003–2009) satellites (Fig. 8A associated with confined carbonate aquifers are less prominent
and B). In the first case, the profile (Fig. 8A) shows a sharp decline in (Meinzer, 1923; Li et al., 2006), although some cases have been linked
water level during the period 1992–1995 that correlates with the defor- to gas extraction in mouldic limestone reservoirs (Dudley et al., 2009;
mation trend at all the control points analyzed (located in Fig. 6). After Van Ditzhuijzen et al., 1984) or intensive use of groundwater in con-
1996 the piezometric level ascends, but the deformation pattern re- fined carbonate aquifers (Davis et al., 1963; Stringfield and Rapp,
mains fairly stable or just moderately descending, with fluctuations 1976; Ruiz-Constán et al., 2016). In this context, subsidence usually
linked to summer periods. For the Envisat time span (Fig. 8B), both happens by the decrease of pore pressure and compaction of the highly
the piezometric level and the deformation pattern show a negative compressible fine-grained sediments that constitute the aquitard
slope with a total subsidence of 55 mm (AD-5) for a water table descent sealing of the aquifer, as consequence of water extraction in the con-
of 73 m over the whole period. fined carbonate aquifer (Poland and Davis, 1969). In our study area,
We furthermore compared the temporal subsiding trend of several the lithological column of the Bajo Guadalhorce detritic aquifer com-
points located at areas of maximum subsidence (Fig. 11) using the coef- prises up to 400 m of upper Pliocene marls, depending on the particular
ficient of determination (R2; square of the Pearson correlation coeffi- site, which could be compressed due to exploitation of the confined Si-
cient), following Ruiz-Constán et al. (2017). The high correlation erra de Mijas carbonate aquifer. Hence, the subsidence maxima de-
observed among all the points (R2 values N76%) indicates a common tected would be the result of interplay between depocenter location
triggering factor for deformation throughout the study area. In addition, and fine grained-size facies distribution within the infilling of the
we compared the normalized value of the subsiding trend (AA-1, AA-3, depocenters.
AB-4, ED-3, ED-4 and ED-8) with the piezometric level evolution at the The gravity study was intended to locate sedimentary depocenters
carbonate aquifer (C10), in order to eliminate the dimensional differ- that could justify the relative subsidence observed. Results point to a
ences between variables of different datasets (Luque-Espinar et al., heterogeneous basin geometry conditioned by the presence of a NNE-
2008). The normalized value (Z) was calculated by means of the expres- SSW furrow limited by a basement high next to the coastline (Fig. 12).
sion Z = (X − μ) / σ (where X is the analyzed variable, μ is the mean Its orientation may be originally conditioned by the NNE-SSW fault set
value and σ is the standard deviation). Results are consistent for both described by Sanz de Galdeano and López Garrido (1991) at the eastern
variables (Fig. 11B and C), pointing to the fact that the subsidence de- border of Sierra de Mijas. The superposition of the deformation and
tected is largely associated with water withdrawal in the confined Sierra gravity profiles (Fig. 10) indicates that subsidence maxima coincide
de Mijas aquifer that produced the decreasing of the interstitial pressure with the location of the greatest sedimentary depocenters only in pro-
at the overlying detritic sediments. file AA-AA′. In the rest of the basin, the high subsidence values appear
Deformation related to karstic aquifers is broadly attributed to col- to be displaced seawards (EA-EA′, EB-EB′ and AB-AB′) and towards
lapse and sinkhole formation in unconfined aquifers (Waltham et al., the north (ED-ED′ and AD-AD′) in the Envisat with regard to the ERS-
2007; Salvati and Sasowsky, 2002). Examples of land subsidence 1/2 time span. In other words, subsidence first takes place at the borders

Fig. 13. Scheme representing the piezometric and subsidence evolution at the study area before 1992 (A), during 1992–2000 (B) and during 2002–2009 (D). Pumping wells are
represented by red tower symbols.
684 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

(Fig. 13B) where the exploitation boreholes are located and the sedi- consolidation process (usually N20 years) can be also seen, but mostly
mentary thickness is lower, then migrates towards the center of the related to the presence of organic soils. Differential subsidence distribu-
basin (Fig. 13C), involving a greater amount of compressible sediments. tion may be conditioned by the spatio-temporal evolution of the urban-
Furthermore, a common ascending pattern at the end of all the profiles ized areas. For example, urbanization N-NW of Benalmádena Port
(usually 0.5–1 km far from the coast) reveals the interference of the sea, (Fig. 14A) started in the 70s, although the vast majority took place in
apparently tied to the fairly constant piezometric level of the detritic the first half of the 1980s. The ERS-1/2 maximum subsidence located
aquifer owing to a good hydraulic connection, probably through the Pli- around this area should be emphasized by the building load, but its in-
ocene sandy levels (Carrasco et al., 2012). cidence is unexpected for the Envisat period as the consolidation pro-
These results suggest that several factors are responsible for the gen- cess should have finished. Other areas, e.g. the NE end of Sierra de
eral compaction of the sedimentary infill. The initial triggering factor of Mijas (in Torremolinos), experienced late urban development that
subsidence is the pore pressure reduction linked to intensive exploita- roughly started in the last years of the 90s (Fig. 14B), so the main
tion of the carbonate aquifer. To this regard, a homogeneous basin stage of subsidence was reported during the Envisat time span. How-
would entail a similar subsidence rate for the whole area. However, ever, further detailed studies should be performed to confirm urbaniza-
other conditioning factors blur this trend: the aforementioned sedimen- tion as a contributing factor of the observed subsidence.
tary thickness, the exploitation borehole locations and the fine-grained Future extended drought periods (Mishra and Singh, 2010) pre-
facies distribution, together with the influence of the Mediterranean Sea. dicted by climate change models presume significant challenges for
Man-made constructions such as skyscrapers or highways are the water management of aquifers subjected to intensive use in arid and
origin of subsidence in many urban settlements worldwide (InSARap, semiarid areas (Wanders et al., 2010; Van Lanen et al., 2013) like
2016; Hu et al., 2004; Xue et al., 2005; Wu et al., 2008; Perissin et al., those of southern Spain. Recurring water table lows will produce irre-
2012; Dong et al., 2014). Between the late 20th century and the onset versible damage in the form of permanent decreases in porosity and
of the economic crisis in 2008, the Spanish shoreline underwent very in- structural building damage (Wilson and Gorelick, 1996). Our study
tense urbanization. This process was not governed by any strategic points to the dire need to dedicate additional effort in monitoring subsi-
planning in the context of the territorial model adopted, or analysis of dence around confined karstic aquifers subjected to intensive exploita-
its environmental and socioeconomic consequences, evoking the unbri- tion. In the Torremolinos area, more research regarding the lithology
dled development of Florida (USA) (Galacho Jiménez, 2005; Fernández column of affected areas or/and the location and flow rate of wells is
Muñoz and Barrado Timón, 2011). The rapid superposition of external necessary to assess the risks involved and to better understand the sub-
loads is therefore an additional contributing factor that should be sidence mechanism. Sound knowledge regarding these parameters is
taken into account. The induced settlement can be essentially broken invaluable for the design and proposal of appropriate mitigation mea-
down into a contemporaneous compression (2–5 years after urbaniza- surements to prevent or alleviate further damage in future drought pe-
tion) and a primary consolidation process, possibly taking up to riods. Two possible means of intervention would be the cessation of
20 years, depending on the hydraulic conductivity and thickness of withdrawal of fluids, or the increase/restoration of fluid pressure by
the fine-grained sediments (Stramondo et al., 2008). A secondary means of artificial recharge.

Fig. 14. Orthophotos showing the urbanization evolution N-NW of the Benalmádena Port (A) and N of Torremolinos (B). The area locations are marked on Fig. 12A and B, respectively.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 685

7. Conclusions Andreo, B., 2007. Sierra de Mijas. In: Durán-Valsero, J.J. (Ed.), (Coord), Atlas
hidrogeológico de la provincia de Málaga. Instituto Geológico y Minero de España
and Diputación de Málaga, pp. 173–178.
To understand the effects of the vast and rapid urbanization along Andreo, B., Sanz de Galdeano, C., 1994. Structure of the Sierra de Mijas (Alpujarride Com-
the Costa del Sol over the past decades, we applied InSAR to two inde- plex, Betic Cordillera). Annales Tectonicae 3, 23–35.
Andreo, B., Carrasco, F., Vadillo, I., Liñán, C., 1996. Características hidrogeológicas de las Si-
pendent sets of SAR scenes (ERS-1/2 and Envisat) and analyzed the erras Blanca y Mijas (Provincia de Málaga, Cordillera Bética). Geogaceta 20 (6),
spatio-temporal evolution of subsidence in a highly urbanized area in 1267–1270.
Costa del Sol (S Spain) during a 17-year period (1992–2009). This Andreo, B., Carrasco, F., Bakalowicz, M., Mudry, J., Vadillo, I., 2002a. Use of hydrodynamic
and hydrochemistry to characterise carbonate aquifers. Case study of the Blanca-
coastal area underwent a huge growth in population together with a Mijas unit (Málaga, southern Spain). Environ. Geol. 43, 108–119.
massive arrival of tourists who relied heavily on groundwater pumping Andreo, B., Carrasco, F., Catalán, F., Durán, J.J., Fernández, G., Linares, L., López, G., López,
to satisfy water demand. An accumulated deformation of −102 mm, J.A., Mayorga, R., Trenado, L., Vadillo, I., 2002b. Características hidrogeológicas de las
Sierras Blanca y Mijas y del Bajo Guadalhorce. In: Rubi, J.C., López, J.A. (Eds.),
with maximum displacement rates of −6 mm/yr (1992–2000) and
Aportaciones al conocimiento de los acuíferos andaluces. Libro homenaje a Manuel
− 11 mm/yr (2003–2009), were recorded. Satellite information com- del Valle, pp. 395–411.
bined with hydrogeological data led us to improved knowledge about Ayala, C., 2013. A new compilation of gravity data over the Iberian Peninsula and sur-
the effects of intensive exploitation of the confined Sierra de Mijas car- rounding areas. Internal Report Topolberia Project (Consolider-Ingenio). IGME,
Madrid http://www.ictja.csic.es/images/Documents/FinalReportGravityTopolB.pdf,
bonate aquifer in terms of ground deformation detected at the detritic Accessed date: 15 September 2017.
sedimentary infill that overlies it and experiences the main urbanization Bamler, R., Hartl, P., 1998. Synthetic aperture radar interferometry. Inverse Probl. 14,
pressure. Furthermore, during the whole period, an accumulated de- R1–R54.
Sinkholes: their geology, engineering, and environmental impact. In: Beck, B.F. (Ed.), Pro-
scent of the confined water level of 140 m and several fluctuations of ceedings of the First Multidisciplinary Conference on Sinkholes, 15–17 October 1984,
N80 m were recorded. Correlation of these parameters points to a com- Orlando, Florida. Balkema, A.A., Rotterdam/Boston.
mon triggering factor related to the decrease in interstitial pressure due Hydrogeology and engineering geology of sinkholes and karst. In: Beck, B.F., Pettit, A.J.,
Herring, J.G. (Eds.), Proceedings of the Seventh Multidisciplinary Conference on Sink-
to water withdrawal. Gravity data reveal the presence of great hetero- holes and the Environmental and Engineering Impacts of Karst, 10–14 April 1999,
geneities in the basement geometry, such as the presence of a NNE- Harrisburg-Hershey, Pennsylvania. Balkema, A.A., Rotterdam, Netherlands.
SSW furrow parallel to the coast, up to 800 m thick. Location of the sed- Berardino, P., Fornaro, G., Lanari, R., Sansosti, E., 2002. A new algorithm for surface defor-
mation monitoring based on small baseline differential SAR interferograms. IEEE T.
imentary depocenters coincides only partially with subsidence maxima, Geosci. Remote 40, 2375–2383.
as deformation may be also conditioned by the interplay of factors such BOE, 2016. Real Decreto 11/2016, de 8 de enero, por el que se aprueban los Planes
as lateral-spatial changes of lithofacies, the proximity of the sea, and Hidrológicos de las demarcaciones hidrográficas de Galicia-Costa, de las Cuencas
Mediterráneas Andaluzas, del Guadalete y Barbate y del Tinto, Odiel y Piedras. Boletín
rapid urbanization. The findings put forth here underline the dire
Oficial del Estado 2016 19:6082–6084. https://www.boe.es/boe/dias/2016/01/22/
need for additional efforts in monitoring subsidence linked to intensive pdfs/BOE-A-2016-602.pdf.
exploitation of confined karstic aquifers as one step towards adopting Cardenal, J., Benavente, J., Cruz-Sanjulian, J.J., 1994. Chemical evolution of groundwater in
water management policies that help to prevent or alleviate further Triassic gypsum-bearing carbonate aquifers (Las Alpujarras, southern Spain).
J. Hydrol. 161 (1–4), 3–30.
damage during drought periods to come. Caro Cuenca, M., 2012. Improving Radar Interferometry for Monitoring Fault-Related Sur-
face Deformacion: Application for the Roer Valley Graben and Coal Mine Induced Dis-
placements in the Southern Netherlands. Ph. D. Thesis. Delft University of
Acknowledgements Technology.
Caro Cuenca, M., Hooper, A.J., Hanssen, R.F., 2011. A new method for temporal phase
unwrapping of persistent scatterers InSAR time series. IEEE T. Geosci. Remote 49
SAR data are provided by the European Space Agency (ESA) in the (11), 4606–4615.
scope of 7629 CAT-1 project. Piezometric data are accessible from Carrasco, F., Andreo, B., Vadillo, I., 2003. Consideraciones hidrogeológicas sobre el sector
IGME (http://info.igme.es/BDAguas/) and the DHCMA (http://sig. costero del acuífero del Bajo Guadalhorce. Tecnología de la intrusión de agua de
mar en acuíferos costeros: países mediterráneos 67–76.
magrama.es/recursossub/visor.html?herramienta=Piezometros) Carrasco, F., Andreo, B., Benavente, J., Vadillo, I., Jiménez, P., Urresti, B., Argamasilla, M.,
websites. Orthophotos were obtained from IDEMAP (http://www. 2012. Acuíferos costeros de Málaga. Demarcación Mediterránea Andaluza. Nuevas
idemap.es). This research was supported by PRX 12/00297, ESP2006- aportaciones al conocimiento de los acuíferos costeros. Serie Hidrogeología y Aguas
Subterráneas 29(III), pp. 117–132.
28463, CGL2016-80687-R AEI/FEDER and CGL2015-71510-R projects Ciampalini, A., Raspini, F., Moretti, S., 2015. Landslide back monitoring and forecasting by
from the Ministerio de Economía, Industria y Competitividad (Spain), using PSInSAR technique: the case of Naso (Sicily, southern Italy). Atti della Società
PAIUJA 2017-2018 and CEACTierra projects from the University of Jaén Toscana di Scienze Naturali, Memorie Serie A 122:19–31. https://doi.org/10.2424/
ASTSN.M.2015.16.
(Spain). Additional support came from the RNM-148 and the RNM- Cigna, F., Novellino, A., Jordan, C.J., Sowter, A., Ramondini, M., Calcaterra, D., 2014. Inter-
282 research groups of the Junta de Andalucía (Spain). Interferometric mittent SBAS (ISBAS) InSAR with COSMO-SkyMed Xband high resolution SAR data
data were processed using the public domain SAR processor DORIS for landslide inventory mapping in Piana degli Albanesi (Italy). Proceedings of the
SPIE 9243, SAR Image Analysis, Modeling, and Techniques XIV, 92431B. (22 October
and StaMPS/MTI. The DEM is freely provided by Instituto Geográfico
2014) https://doi.org/10.1117/12.2067424.
Nacional de España. The satellite orbits used are from Delft University Colesanti, C., Ferretti, A., Novali, F., Prati, C., Rocca, F., 2003. SAR monitoring of progressive
of Technology and ESA. This work was also supported by the Ministry and seasonal ground deformation using the permanent scatterers technique. IEEE
Trans. Geosci. Remote Sens. 4, 1685–1701.
of Education, Youth and Sports from the National Programme of Sus-
Crosetto, M., Monserrat, O., Cuevas-González, M., Devanthéry, N., Crippa, B., 2016.
tainability (NPU II) project “IT4Innovations excellence in science - Persostemt scatterer interferometry: a review. ISPRS J. Photogramm. 115:78–89.
LQ1602” (Czech Republic). https://doi.org/10.1016/j.isprsjprs.2015.10.011.
Custodio, E., 2002. Aquifer overexploitation: what does it mean? Hydrogeol. J. 10 (2),
254–277.
Appendix A. Supplementary data Davis, G.H., Small, J.B., Counts, H.B., 1963. Land subsidence related to decline of artesian
pressure in the Ocala Limestone, at Savannah, Georgia. In: Trask, P.D., Kiersek, G.A.
Supplementary data to this article can be found online at https://doi. (Eds.), Engineering Geology. Case Histories. 4. Geol. Soc. America.
Del Ventisette, C., Righini, G., Moretti, S., Casagli, N., 2014. Multitemporal landslides inven-
org/10.1016/j.scitotenv.2018.04.280. tory map updating using spaceborne SAR analysis. Int. J. Appl. Earth Obs. 30:238–246.
https://doi.org/10.1016/j.jag.2014.02.008.
DHCMA, 2017. Demarcación Hidrográfica de las Cuencas Mediterráneas Andaluzas.
References http://sig.mapama.es/redes-seguimiento/visor.html?herramienta=Piezometros,
Accessed date: 15 September 2017.
AEMET, 2017. Agencia Estatal de Meteorología. Seasonal Analysis. http://www.aemet.es/ Dong, S., Samsonov, S., Yin, H., Ye, S., Cao, Y., 2014. Time-series analysis of subsidence as-
en/serviciosclimaticos/vigilancia_clima/analisis_estacional, Accessed date: 17 Sep- sociated with rapid urbanization in Shanghai, China measured with SBAS InSAR
tember 2017. method. Environ. Earth Sci. 72 (3), 677–691.
Amelung, F., Galloway, D.L., Bell, J.W., Zebker, H.A., Laczniak, R.J., 1999. Sensing the ups Dudley, J.W., van der Linden, A., Mah, K.G., 2009. Predicting accelerating subsidence above
and downs of Las Vegas: InSAR reveals structural control of land subsidence and the highly compacting Luconia carbonate reservoirs, offshore Sarawak Malaysia. SPE
aquifer-system deformation. Geology 27, 483–486. Reserv. Eval. Eng. 12 (01), 104–115.
686 A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687

Duque, C., Calvache, M.L., Pedrera, A., Martín-Rosales, W., López-Chicano, M., 2008. Com- IGN, 2017b. Instituto Geográfico Nacional de España. Red gravimétrica. http://www.ign.
bined time domain electromagnetic soundings and gravimetry to determine marine es/ign/main/index.do, Accessed date: 15 September 2017.
intrusion in a detrital coastal aquifer (Southern Spain). J. Hydrol. 349 (3–4), 536–547. INE, 2015. Instituto Nacional de Estadística. http://www.ine.es, Accessed date: 15 Sep-
Eckerman, L., Agüero, A., Spagnotto, S., Martinez, P., Nacif, S., 2017. Seismic-gravimetric tember 2017.
analysis of the subducted Nazca plate between 32° S and 36° S. Geodesy and InSARap, 2016. Sentinel-1 InSAR Performance Study with TOPS Data. Satellites Confirm
Geodynamics 9 (1), 57–66. Sinking of San Francisco Tower. http://www.insarap.org, Accessed date: 22 Septem-
Estévez-González, C., Chamón Cobos, C., 1978. Mapa Geológico de España 1:50.000 ber 2017.
MAGNA. Hoja n° 1067, Torremolinos. Instituto Geológico y Minero de España, Insua-Arévalo, J.M., Martín-González, F., Capote, R., Martínez-Díaz, J.J., 2004. Análisis
Madrid. tectónico de los mapas de anomalía gravimétrica de la Cuenca de Málaga (Cordillera
Fernández Muñoz, S.F., Barrado Timón, D.A., 2011. El desarrollo turístico-inmobiliario de Bética Occidental). Bol. Geol. Min. 115 (3), 521–536.
la España mediterránea e insular frente a sus referentes internacionales (Florida y Li, C., Tang, X., Ma, T., 2006. Land subsidence caused by groundwater exploitation in the
la Costa Azul): un análisis comparado. Cuadernos de Turismo. 27 pp. 373–402. Hangzhou-Jiaxing-Huzhou Plain, China. Hydrogeol. J. 14 (8), 1652–1665.
Ferretti, A., Prati, C., Rocca, F., 2001. Permanent scatterers in SAR interferometry. IEEE T. Llamas, M.R., Custodio, E., 2003. Intensive use of groundwater: a new situation which de-
Geosci. Remote 39 (1), 8–20. mands proactive actions. In: Llamas, M.R., Custodio, E. (Eds.), Intensive Use of
Ferretti, A., Savio, G., Barzaghi, R., Borghi, A., Musazzi, S., Novali, F., Prati, C., Rocca, F., Groundwater, Challenges and Opportunities. A.A. Balkema Publishers, Lisse,
2007a. Submillimeter accuracy of InSAR time series: experimental validation. IEEE pp. 13–34.
T. Geosci. Remote 45 (5):1142–1153. https://doi.org/10.1109/TGRS.2007.894440. Luque-Espinar, J.A., Chica-Olmo, M., Pardo-Igúzquiza, E., García-Soldado, M.J., 2008. Influ-
Ferretti, A., Monti-Guarnieri, A., Prati, C., Rocca, F., Massonnet, D., 2007b. InSAR Principles: ence of climatological cycles on hydraulic heads across a Spanish aquifer. J. Hydrol.
Guidelines for SAR Interferometry Processing and Interpretation. ESA Publications, 354 (1–4), 33–52.
TM-19. ESTEC. Mandal, A., Gupta, S., Mohanty, W.K., Misra, S., 2015. Sub-surface structure of a craton–
Ferretti, A., Fumagalli, A., Novali, F., Prati, C., Rocca, F., Rucci, A., 2011. A new algorithm for mobile belt interface: evidence from geological and gravity studies across the Rengali
processing interferometric data-stacks: SqueeSAR. IEEE T. Geosci. Remote 49 (9): Province–Eastern Ghats Belt boundary, eastern India. Tectonophysics 662, 140–152.
3460–3470. https://doi.org/10.1109/TGRS.2011.2124465. Martínez-Martos, M., Galindo-Zaldivar, J., Lobo, F.J., Pedrera, A., Ruano, P., Lopez-Chicano,
Foster, S.S.D., 1992. Unsustainable development and irrational exploitation of groundwa- M., Ortega-Sánchez, M., 2016. Buried marine-cut terraces and submerged marine-
ter resources in developing nations—an overview. IAH Hydrogeology Selected Papers built terraces: the Carchuna-Calahonda coastal area (southeast Iberian Peninsula).
3, 321–336. Geomorphology 264, 29–40.
Frangioni, S., Bianchini, S., Moretti, S., 2015. Landslide inventory updating by means of Martos-Rosillo, S., Rodríguez-Rodríguez, M., Moral, F., Cruz-Sanjulián, J.J., Rubio, J.C., 2009.
Persistent Scatterer Interferometry (PSI): the Setta basin (Italy) case study. Geomat. Analysis of groundwater mining in two carbonate aquifers in Sierra de Estepa (SE
Nat. Haz. Risk 6 (5–7):419–438. https://doi.org/10.1080/19475705.2013.866985. Spain) based on hydrodynamic and hydrochemical data. Hydrogeol. J. 17 (7), 1617.
Galacho Jiménez, F.B., 2005. El planteamiento urbanístico municipal de la Costa del Sol: Meinzer, O.E., 1923. Outline of ground-water hydrology, with definitions. U.S. Geol. Sur-
procesos de formulación y criterios de ordenación. Baética. Estudios de Arte, vey Water-Supply Paper 494, 71.
Geografía e Historia. 27 pp. 97–114. Meisina, C., Zucca, F., Notti, D., Colombo, A., Cucchi, A., Savio, G., Giannico, C., Bianchi, M.,
Galloway, D.L., Hudnut, K.W., Ingebritsen, S.E., Phillips, S.P., Peltzer, G., Rogez, F., Rosen, 2008. Geological interpretation of PSInSAR data at regional scale. Sensors 8 (11):
P.A., 1998. Detection of aquifer system compaction and land subsidence using inter- 7469–7492. https://doi.org/10.3390/s8117469.
ferometric synthetic aperture radar, Antelope Valley, Mojave Desert, California. Mishra, A.K., Singh, V.P., 2010. A review of drought concepts. J. Hydrol. 391 (1):202–216.
Water Resour. Res. 34 (10), 2573–2585. https://doi.org/10.1016/j.jhydrol.2010.07.012.
Galve, J.P., Pérez-Peña, J.V., Azañón, J.M., Closson, D., Caló, F., Reyes-Carmona, C., Jabaloy, Nieto, J.M., Andreo, B., Mudarra, M., 2016. Estimación de parámetros hidrogeológicos en
A., Ruano, P., Mateos, R.M., Notti, D., Herrera, G., Béjar-Pizarro, M., Montserrat, O., los acuíferos costeros del Bajo Guadalhorce (Málaga) mediante el análisis de la
Bally, P., 2017. Evaluation of the SBAS InSAR Service of the European Space Agency's influencia de las mareas. Geogaceta 59, 39–42.
Geohazard Exploittion Platform (GEP). Remote Sens. 9:1291. https://doi.org/10.3390/ Notti, D., Herrera, G., Bianchini, S., Meisina, C., García-Davalillo, J.C., Zucca, F., 2014. A
rs9121291. methodology for improving landslide PSI data analysis. Int. J. Remote Sens. 35 (6):
García-Moreno, I., Mateos, R.M., 2011. Sinkholes related to discontinuous pumping: sus- 2186–2214. https://doi.org/10.1080/01431161.2014.889864.
ceptibility mapping based on geophysical studies. The case of Crestatx (Majorca, Novellino, A., 2015. Assessment of Landslide Susceptibility in Structurally Complex For-
Spain). Environ. Earth Sci. 64 (2), 523–537. mations by Integration of Different A-DInSAR Techniques. PhD Thesis. Federico II Uni-
Hanssen, R.F., 2001. Radar Interferometry: Data Interpretation and Error Analysis. Kluwer versity of Naples, Department of Earth, Environment and Resources Sciences, Naples,
Academic Publishers, Dordrecht. Italy.
Hetland, E.A., Musé, P., Simons, M., Lin, Y.N., Agram, P.S., DiCaprio, C.J., 2012. Multiscale Pedley, R.C., Busby, J.P., Dabeck, Z.K., 1993. GRAVMAG User Manual-Interactive 2.5D Grav-
InSAR Time Series (MInTS) analysis of surface deformation. J. Geophys. Res. 117, ity and Magnetic Modelling. British Geological Survey, Technical Report WK/93/26/R.
B02404. https://doi.org/10.1029/2011JB008731. Pedrera, A., Galindo-Zaldívar, J., Acosta-Vigil, A., Azor, A., González-Menéndez, L.,
Hoffmann, J., Zebker, H., Galloway, D., Amelung, F., 2001. Seasonal subsidence and re- Rodríguez-Fernández, L.R., Ruiz-Constán, A., 2016. Serpentinization-driven extension
bound in Las Vegas Valley, Nevada, observed by synthetic aperture radar interferom- in the Ronda mantle slab (Betic Cordillera, S Spain). Tectonics 37, 205–215.
etry. Water Resour. Res. 37, 1551–1566. Pepe, A., Calò, F., 2017. A review of Interferometric Synthetic Aperture RADAR (InSAR)
Hooper, A.J., 2006. Persistent Scatterer Radar Interferometry for Crustal Deformation Multi-Track Approaches for the retrieval of Earth's surface displacements. Appl. Sci.
Studies and Modelling of Volcanic Deformation. Ph.D. Thesis. Stanford University. 7 (12):1264. https://doi.org/10.3390/app7121264.
Hooper, A.J., 2008. A multi-temporal InSAR method incorporating both persistent scat- Perissin, D., Wang, Z., Lin, H., 2012. Shanghai subway tunnels and highways monitoring
terer and small baseline approaches. Geophys. Res. Lett. 35 (16), L16302. https:// through Cosmo-SkyMed Persistent Scatterers. ISPRS J. Photogramm. 73, 58–67.
doi.org/10.1029/2008GL034654. Piles Mateo, E., Estévez-González, C., Barba Martin, A., 1978. Mapa Geológico de España 1:
Hooper, A., 2010. A Statistical-cost Approach to Unwrapping the Phase of InSAR Time Se- 50.000 MAGNA. Hoja n° 1066, Coín. Instituto Geológico y Minero de España, Madrid.
ries. European Space Agency (Special Publication)ESA SP-677. Poland, J.F., Davis, G.H., 1969. Land subsidence due to withdrawal of fluids. In: Varnes, D.J.,
Hooper, A., Zebker, H., Segall, P., Kampes, B., 2004. A new method for measuring deforma- Kiersch, G. (Eds.), Reviews in Engineering Geology 2. Geol. Soc. America, pp. 187–269.
tion on volcanoes and other natural terrains using InSAR persistent scatterers. Pulido-Bosch, A., Morell, I., Andreu, J.M., 1995. Hydrogeochemical effects of groundwater
Geophys. Res. Lett. 31 (23), L23611. https://doi.org/10.1029/2004GL021737. mining of the Sierra de Crevillente Aquifer (Alicante, Spain). Environ. Geol. 26 (4),
Hooper, A., Segall, P., Zebker, H., 2007. Persistent scatterer InSAR for crustal deformation 232–239.
analysis, with application to Volcán Alcedo. Galápagos. J. Geophys. Res. 112 (B7), Rodríguez-Estrella, T., 2012. The problems of overexploitation of aquifers in semi-arid
B07407. https://doi.org/10.1029/2006JB004763. areas: the Murcia Region and the Segura Basin (South-east Spain) case. Hydrol.
Hooper, A., Bekaert, D.P.S., Spaans, K., Arikan, M., 2012. Recent advances in SAR interfer- Earth Syst. Sci. Discuss. 9 (5):5729–5756. https://doi.org/10.5194/hessd-9-5729-
ometry time series analysis for measuring crustal deformation. Tectonophysics 2012.
514–517:1–13. https://doi.org/10.1016/j.tecto.2011.10.013. Rosen, P.A., Hensley, S., Joughin, I.R., Li, F.K., Madsen, S.N., Rodriguez, E., Goldstein, R.M.,
Hooper, A., Bekaert, D., Spaans, K., 2013. StaMPS/MTI manual. Version 3.3b1. School of 2000. Synthetic aperture radar interferometry. Proc. IEEE 88 (3), 333–382.
Earth and Environment, University of Leeds, UK Available at:. https://homepages. Rosenthal, E., 1988. Hydrochemical changes induced by overexploitation of groundwater
see.leeds.ac.uk/~earahoo/stamps/StaMPS_Manual_v3.3b1.pdf, Accessed date: 22 Sep- at common outlets of the Bet Shean-Harod multiple-aquifer system, Israel. J. Hydrol.
tember 2017. 97 (1–2), 107–128.
Hu, R.L., Yue, Z.Q., Wang, L.U., Wang, S.J., 2004. Review on current status and challenging Rucker, M.L., Panda, B.B., Meyers, R.A., Lommler, J.C., 2013. Using InSAR to detect subsi-
issues of land subsidence in China. Eng. Geol. 76 (1), 65–77. dence at brine wells, sinkhole sites, and mines. Carbonates Evaporites 28, 141–147.
IGME, 1983a. Investigación hidrogeológica de las cuencas del Sur de España. Sistema Ruiz, A.M., Caro Cuenca, M., Sousa, J.J., Gil, A.J., Hanssen, R.F., Perski, Z., Galindo-Zaldívar, J.,
acuífero n° 38 “Mármoles de Sierra Blanca y Sierra de Mijas”. Technical report n° 6. Sanz de Galdeano, C., 2011. Land subsidence monitoring in the southern Spanish
http://www.igme.es/sistemas_infor/Sid.htm, Accessed date: 22 September 2017. coast using satellite radar interferometry. Proceedings of the FRINGE 2011 Workshop,
IGME, 1983b. Investigación hidrogeológica de las cuencas del Sur de España. Sistema Frascati, Italy, 19–23 September 2011, pp. 19–23.
acuífero n° 37 “Detrítico de Málaga”. Technical report n° 5. http://www.igme.es/ Ruiz-Constán, A., Galindo-Zaldívar, J., Pedrera, A., Sanz de Galdeano, C., 2009. Gravity
sistemas_infor/Sid.htm, Accessed date: 22 September 2017. anomalies and orthogonal box fold development on heterogeneous basement in
IGME, 2017. Instituto Geológico y Minero de España. http://info.igme.es/bdaguas/, the Neogene Ronda Depression Western Betic Cordillera. J. Geodyn. 47 (4), 210–217.
Accessed date: 15 September 2017. Ruiz-Constán, A., Ruiz-Armenteros, A.M., Lamas-Fernández, F., Martos-Rosillo, S., Delgado,
IGN, 1976. Spanish Gravimetric Map of Bouguer Anomalies. Scale: 1/1.000.000. J.M., Bekaert, D.P.S., Sousa, J.J., Gil, A.J., Caro Cuenca, M., Hanssen, R.F., Galindo-
IGN, 2017a. Instituto Geográfico Nacional de España. Centro de descargas. http:// Zaldivar, J., Sanz de Galdeano, S., 2016. Multi-temporal InSAR evidence of ground sub-
centrodedescargas.cnig.es/CentroDescargas/index.jsp, Accessed date: 15 September sidence induced by groundwater withdrawal: the Montellano aquifer (SW Spain).
2017. Environ. Earth Sci. 75 (3), 1–16.
A. Ruiz-Constán et al. / Science of the Total Environment 636 (2018) 670–687 687

Ruiz-Constán, A., Ruiz-Armenteros, A.M., Galindo-Zaldívar, J., Lamas-Fernández, F., Sousa, Urresti-Estala, B., Jiménez-Gavilán, P., Vadillo-Pérez, I., Carrasco-Cantos, F., 2016. Assess-
J.J., Sanz de Galdeano, C., Pedrera, A., Martos-Rosillo, S., Caro Cuenca, M., Delgado, J.M., ment of hydrochemical trends in the highly anthropised Guadalhorce River Basin
Hanssen, R.F., Gil, A.J., 2017. Factors determining subsidence in urbanized floodplains: (Southern Spain) in terms of compliance with the European Groundwater Directive
evidences from MT-InSAR in Seville (Southern Spain). Earth Surf. Process. Landf. 42: for 2015. Environ. Sci. Pollut. Res. 23 (16), 15990–16005.
2484–2497. https://doi.org/10.1002/esp.4180. Vadillo, I., Carrasco, F., Sánchez, D., 2007. Bajo Guadalhorce. In: Durán-Valsero, J.J. (Ed.),
Salvati, R., Sasowsky, I.D., 2002. Development of collapse sinkholes in areas of groundwa- Atlas hidrogeológico de la provincia de Málaga. Instituto Geológico y Minero de
ter discharge. J. Hydrol. 264, 1–11. España and Diputación de Málaga, Madrid.
Sanz de Galdeano, C., López Garrido, A.C., 1991. Tectonic evolution of the Málaga Basin Van der Gun, J., Lipponen, A., 2010. Reconciling groundwater storage depletion due to
(Betic Cordillera). Regional implications. Geodin. Acta 5 (3), 173–186. pumping with sustainability. Sustainability 2 (11), 3418–3435.
Schmidt, D.A., Burgmann, R., 2003. Time-dependent land uplift and subsidence in the Van Ditzhuijzen, P.J.D., Berhad, S.S., de Waal, J.A., 1984. Reservoir compaction and surface
Santa Clara valley, California, from a large interferometric synthetic aperture radar subsidence in the Central Luconia gas bearing carbonates, offshore Sarawak, East
data set. J. Geophys. Res. 108:2416–2428. https://doi.org/10.1029/2002JB002267. Malaysia. Offshore South East Asia Conference, 4-27-4-40.
Sousa, J., Ruiz, A., Hanssen, R., Bastos, L., Gil, A., Galindo-Zaldívar, J., Sanz de Galdeano, C., Van Lanen, H.A.J., Wanders, N., Tallaksen, L.M., Van Loon, A.F., 2013. Hydrological drought
2010. PS-InSAR processing methodologies in the detection of field surface deforma- across the world: impact of climate and physical catchment structure. Hydrol. Earth
tion – study of the Granada Basin (Central Betic Cordilleras, Southern Spain). Syst. Sc. 17 (5):1715–1732. https://doi.org/10.5194/hess-17-1715-2013.
J. Geodyn. 49 (3–4):181–189. https://doi.org/10.1016/j.jog.2009.12.002. Waltham, A.C., Bell, F., Culshaw, M.G., 2007. Sinkholes and Subsidence: Karst and Cavern-
Sousa, J., Hooper, A., Hanssen, R., Bastos, L., Ruiz, A., 2011. Persistent Scatterer InSAR: a ous Rocks in Engineering and Construction. Springer, Berlin.
comparison of methodologies based on a model of temporal deformation vs spatial Wanders, N., Van Lanen, H.A.J., Van Loon, A.F., 2010. Indicators for drought characteriza-
correlation selection criteria. Remote Sens. Environ. 115 (10), 2652–2663. tion on a global scale. WATCH Technical Report No. 24 Available at:. http://www.
Stramondo, S., Bozzano, F., Marra, F., Wegmuller, U., Cinti, F.R., Moro, M., Saroli, M., 2008. eu-watch.org/publications/technical-reports, Accessed date: 1 September 2017.
Subsidence induced by urbanisation in the city of Rome detected by advanced InSAR Weary, D.J., 2015. The cost of karst subsidence and sinkhole collapse in the United States
technique and geotechnical investigations. Remote Sens. Environ. 112 (6), compared with other natural hazards. 14th Sinkhole Conference NCKRI Symposium,
3160–3172. pp. 433–445.
Stringfield, V.T., Rapp, J.R., 1976. Land subsidence resulting from withdrawal of ground Wilson, A.M., Gorelick, S., 1996. The effects of pulsed pumping on land subsidence in the
water in carbonate rocks. Publication n°121 of the International Association of Hydro- Santa Clara Valley, California. J. Hydrol. 174 (3), 375–396.
logical Sciences. Proceedings of the Anaheim Symposium, December 1976. U.S. Geo- Wu, J., Shi, X., Xue, Y., Zhang, Y., Wei, Z., Yu, J., 2008. The development and control of the
logical Survey, Reston, Virginia. land subsidence in the Yangtze Delta, China. Environ. Geol. 55 (8), 1725–1735.
Telford, W.M., Geldart, L.P., Sheriff, R.E., 1991. Applied Geophysics (Second Edition). Cam- Xue, Y.Q., Zhang, Y., Ye, S.J., Wu, J.C., Li, Q.F., 2005. Land subsidence in China. Environ. Geol.
bridge University Press. 48, 713–720.
Torné, M., Banda, E., García-Dueñas, V., Balanyá, J.C., 1992. Mantle-lithosphere bodies in Zhou, X., Chang, N., Li, S., 2009. Applications of SAR interferometry in earth and environ-
the Alboran crustal domain (Ronda peridotitas, Betic-Rif orogenic belt). Earth Planet. mental science research. Sensors 9 (3):1876–1912. https://doi.org/10.3390/
Sc. Lett. 110, 163–171. s90301876.

You might also like