You are on page 1of 17

Engineering Geology 283 (2021) 105998

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

Modeling historical subsidence due to groundwater withdrawal in the Alto


Guadalentín aquifer-system (Spain)
J.A. Fernández-Merodo a, b, P. Ezquerro a, c, *, D. Manzanal c, M. Béjar-Pizarro a, R.M. Mateos a,
C. Guardiola-Albert a, d, J.C. García-Davalillo a, J. López-Vinielles a, c, e, R. Sarro a, G. Bru f,
J. Mulas a, R. Aragón a, C. Reyes-Carmona a, g, P. Mira h, M. Pastor c, G. Herrera a, i
a
Geohazards InSAR Laboratory and Modeling Group (InSARlab), Geoscience Research Department, Geological Survey of Spain (IGME), Alenza 1, 28003 Madrid, Spain
b
Universidad Politécnica de Madrid, ETSI Minas y Energía C/ Ríos Rosas, 21, 28003 Madrid, Spain
c
Universidad Politécnica de Madrid, ETSI Caminos, Canales y Puertos C/ Profesor Aranguren s/n, 28040 Madrid, Spain
d
Environmental Geology and Geomathematics, Geoscience Research Department, Geological Survey of Spain (IGME), Alenza 1, 28003 Madrid, Spain
e
HEMAV SL, Carrer d’Esteve Terrades 1, 08860, Castelldefels, Barcelona, Spain
f
Instituto de Geociencias (UCM-CSIC), Calle del Dr. Severo Ochoa, 7, 28040 Madrid, Spain
g
Department of Geodynamics, University of Granada, Campus Universitario Fuentenueva, 18071 Granada, Spain
h
Centro de Estudios y Experimentación de Obras Públicas (CEDEX), Calle de Alfonso XII, 3, 28014 Madrid, Spain
i
EuroGeoSurveys: Earth Observation and Geohazards Expert Group (EOEG), Rue Joseph II 36-38, 1000 Brussels, Belgium

A R T I C L E I N F O A B S T R A C T

Keywords: The Alto Guadalentín Basin (Spain) is widely recognized as an area of major anthropogenic land subsidence due
Subsidence to groundwater extraction. This paper presents a numerical methodology to quantify the severe subsidence of
Groundwater withdrawal this basin over history. First, a 3D groundwater model is proposed to reproduce groundwater evolution in the
Generalized plasticity model
regional Alto Guadalentín aquifer system since 1960, leading to an average piezometric level drop of 150 m.
MODFLOW
GEHOMADRID
Secondly, a generalized plasticity state parameter-based model is calibrated to reproduce the mechanical
Alto Guadalentín behavior, observed in oedometer laboratory tests, of compressible materials extracted from a 300-m drilled
borehole located in the area of maximum subsidence. The strength of this constitutive model is that a single set of
material parameters can be used to reproduce the mechanical behavior of material located at different depths,
hence having different confining pressures and void ratio states. Afterwards, subsidence is assessed through a
partially saturated 1D vertical finite element model, solving Biot equations that reproduce the slow vertical
drainage and vertical consolidation processes, taking into account the calibrated constitutive model and pre­
scribing the previously computed groundwater evolution of the aquifer. Finally, the subsidence model is adjusted
with different displacement data available from 1992: datasets acquired by ERS, ENVISAT, Cosmo-SkyMed
satellites and the global positioning system GNSS. The proposed calibrated subsidence model reproduces the
3.1 m subsidence monitored in the period 1992–2018, and quantifies historical subsidence (since 1960) in the
Alto Guadalentín Basin area at around 5.8 m. Moreover, the model predicts subsidence of up to 7.3 m by 2100 for
an assumed constant hydraulic head from 2012 onward.

1. Introduction to buildings and infrastructure, changes in flood risk areas, and loss of
water storage capacity. It occurs when the withdrawal of groundwater
Ground subsidence due to long-term groundwater withdrawal is an from an aquifer system or subsurface reservoir is faster than its natural
important geological-anthropogenic hazard that may produce damage or artificial recharge. In a context of global climate change, the

* Corresponding author at: Geohazards InSAR Laboratory and Modeling Group (InSARlab), Geoscience Research Department, Geological Survey of Spain (IGME),
Alenza 1, 28003 Madrid, Spain.
E-mail addresses: jose.fernandez@igme.es, jose.merodo@upm.es (J.A. Fernández-Merodo), p.ezquerro@igme.es (P. Ezquerro), m.bejar@igme.es, d.manzanal@
upm.es, manuel.pastor@upm.es (D. Manzanal), rm.mateos@igme.es (R.M. Mateos), c.guardiola@igme.es (C. Guardiola-Albert), jc.garcia@igme.es (J.C. García-
Davalillo), jlvinielles@hemav.com (J. López-Vinielles), r.sarro@igme.es (R. Sarro), guadalupe.bru@igeo.ucm-csic.com (G. Bru), j.mulas@igme.es (J. Mulas), r.
aragon@igme.es (R. Aragón), cristinarc@ugr.es (C. Reyes-Carmona), pablo.mira@cedex.es (P. Mira), g.herrera@igme.es (G. Herrera).

https://doi.org/10.1016/j.enggeo.2021.105998
Received 20 April 2020; Received in revised form 30 December 2020; Accepted 7 January 2021
Available online 9 January 2021
0013-7952/© 2021 Elsevier B.V. All rights reserved.
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

increasing agricultural and urban use of aquifer resources is a growing (Tessitore et al., 2016; Mateos et al., 2017; Ezquerro et al., 2019).
problem, especially in arid and semiarid areas. In addition to its sig­ A commonly accepted geomechanic framework describes subsidence
nificant economic impact (Holzer and Galloway, 2005), subsidence in view of Terzaghi’s Principle and the 1D consolidation theory or the
poses a risk for society that must be controlled. Recent plans for sub­ improved Biot’s 3D consolidation theory. Still, a reliable predictive tool
surface resource management should therefore include a study of related is elusive. Some key reasons are: 1) the complexity of mixed formula­
environmental impacts and incorporate numerical predictions of the tions that represent the solid-pore fluid interaction, especially in the case
expected land settlement above (and near) the exploited system of non-saturated materials with liquid and gaseous phases, 2) the

Fig. 1. (a) Geologic map, (b) Soft soil thickness distribution from (Bonì et al., 2015) and (Béjar-Pizarro et al., 2016), (c) Plio-quaternary bottom derived from
(Cerón, 1995).

2
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

complexity of realistic non-linear constitutive models for soil 2. The Alto Guadalentín Basin
compressibility, 3) the complexity of numerical models to resolve mixed
formulations, especially with the Finite Element Method. Moreover, 2.1. Geomorphological and geological description
regional studies are hampered by: 1) the heterogeneity of the 3D geol­
ogy, the horizontal size of the porous medium involved can be orders of The Alto Guadalentín Basin is an intramontane tectonic depression
magnitude greater than the vertical size, 2) difficulty in defining located in Murcia province, in the southern part of the Spanish Medi­
boundary conditions and estimating historical pumping rates as well as terranean Arc (Fig. 1a). Since no permanent watercourses cross the basin
aquifer recharges, 3) difficulty in defining realistic initial sates, 4) dif­ and precipitation is low (average annual rainfall of 250 mm and alter­
ficulty in choosing suitable values for the soil parameters, and 5) limi­ nate dry years with less than 150 mm), the aquifer-system, covering an
tation of the computing power in the case of major 3D problems. area of approximately 277 km2, is a strategic resource for nearly
Therefore, different modeling approaches can be followed to address the 100.000 people (annual census, Spanish Statistical Office) whose agri­
numerical simulation of subsidence due to groundwater withdrawal; a cultural sector is very important. Bounded to the north by the active
shortened classification with associated references is available in Gam­ Alhama de Murcia fault (Masana et al., 2004), this basin is made up of
bolati and Teatini (2015). Predictive numerical models have been Miocene deposits (300–900 m depth) with conglomerates, sandstones
developed for many sites strongly affected by land subsidence over the and marls, and Quaternary sediments from alluvial fan systems that
past decades: Mexico City in Mexico (Ortega-Guerrero et al., 1999; overlap previous materials (Cerón and Pulido-Bosch, 1996; Cerón,
Ortiz-Zamora and Ortega-Guerrero, 2010), Shanghai in China (Shi et al., 1995). Plio-Quaternary sediments (Fig. 1c) constitute the main aquifer
2007; Ye et al., 2012), Central Valley, Santa Clara Valley, and Antelope layer, with a complex structure of sand and gravel lenses embedded in a
Valley in California (Leake, 1990; Wilson and Gorelick, 1996; Leake and silt/clay matrix. Recent studies (Bonì et al., 2015; Béjar-Pizarro et al.,
Galloway, 2010), Houston in Texas (Kasmarek and Strom, 2002), Hanoi 2016) reinterpreted the geological information from boreholes to define
in Vietnam (Thu and Fredlund, 2000), and Tokyo in Japan (Aichi, 2008). a new layer in the upper part of the Plio-Quaternary deposits charac­
In this work, a new numerical model is presented to quantify his­ terized by the presence of continuous fine sediments (silts and clays)
torical subsidence in the Alto Guadalentín Basin, located in southeastern (Fig. 1b). The location and thickness of this layer, closely related to the
Spain (Fig. 1). In this area, subsidence rates of up to 13 cm/year as a magnitude of subsidence observed in the area, were introduced in the
direct consequence of long-term aquifer exploitation have been recog­ new hydrogeological models developed in the basin (Ezquerro et al.,
nized by González and Fernández (2011), Rigo et al. (2013), Bonì et al. 2017). The pre-orogenic Paleozoic basement has a horst and graben
(2015) and Béjar-Pizarro et al. (2016) using different Interferometric pattern and is composed by metamorphic complexes reaching maximum
synthetic aperture radar (InSAR) techniques and processing multi-sensor depths of 1000 m. The aquifer system’s most productive layer would be
SAR images with different resolutions over a monitoring time span of 20 the Plio-Quaternary sediments, disposed in the typical alluvial fan multi-
years (1992–2012). These observations were confirmed by Fernandez layered pattern. Below them, Miocene materials act as a low perme­
et al. (2018) using a GNSS network consisting of 33 stations which ability or impervious limit, depending on the area.
densely covers the Alto Guadalentín Basin, in three geodetic campaigns Limited natural recharge and intensive use of the aquifer system’s
carried out in 2015, 2016 and 2017. Continuous subsidence rates of resources since 1960s led to an average groundwater level drop of some
11–13 cm/year and accumulated land settlement of 3.1 m over the last 150 m in 50 years. The aquifer was officially declared overexploited in
27 years makes the Alto Guadalentín Basin one of the most relevant 1988.
examples worldwide of anthropogenic land subsidence due to ground­
water extraction (Gambolati and Teatini, 2015). As this aquifer-system 2.2. Monitoring description
has been intensively exploited since 1960, historical subsidence quan­
tification is needed to understand not only the past but also the future After detection of the Alto Guadalentín subsidence phenomenon
behavior of the Alto Guadalentín Basin. Ezquerro et al. (2017) proposed using satellite data (González and Fernández, 2011), available InSAR
a simple empirical subsidence model through regression functions, data were processed to define its magnitude and extension (Rigo et al.,
relating groundwater changes, soft soil thickness and surface deforma­ 2013; Bonì et al., 2015; Béjar-Pizarro et al., 2016). First, ERS
tion measurements from ENVISAT images of the study area calibrated (1992–2000) and ENVISAT (2003− 2010) C-band images presented by
for the period 2003–2010. Despite the limited temporal range, the Bonì et al. (2015) and Béjar-Pizarro et al. (2016), respectively, were
model roughly estimates that the maximum subsidence in the aquifer used to generate Line Of Sight (LOS) displacement velocities and time
system is 5.5 m since the beginning of groundwater extraction (period series (Fig. 2a, b). ERS results showed a maximum displacement of 12
1960–2012). cm/year over the northern side of the basin and a secondary local
This paper presents a new numerical methodology to calculate his­ maximum in the south (Fig. 2a). Displacement decreased toward the
torical subsidence in the Alto Guadalentín Basin, based on realistic basin border, reaching its stability threshold (±1 cm/year). This stability
hydro-mechanical coupling behavior and advanced mechanical stress- threshold was assumed to be 1.5–2 times the standard deviation of the
strain relationships. First, a 3D groundwater model reproduces the velocity of stable areas. The ENVISAT displacement rate was slightly
groundwater evolution in the regional Alto Guadalentín aquifer system lower than ERS, with a maximum of 11 cm/year (Fig. 2b). At the end of
since 1960. Secondly, a generalized plasticity state parameter-based the ENVISAT mission (2010), the C-band’s lack of images until Sentinel-
model is calibrated to reproduce the mechanical behavior, observed in 1 deployment (2015) was compensated with Cosmo-SkyMed X-band
oedometer laboratory tests, of compressible materials that were data (CSK). Whereas ERS and ENVISAT datasets were processed using a
extracted from a 300 m borehole drilled in the area of maximum sub­ PSI approach, CSK data were processed by means of SBAS in order to
sidence. After calibrating the constitutive model and computing the improve the X-band performance in non-urban areas. CSK results, with a
groundwater evolution, subsidence is modeled through a 1D vertical higher resolution (3–9 m) than previous data over the basin (ERS and
finite element model, solving the unsaturated Biot equations that ENVISAT, 30–90 m), have revealed a lower subsidence rate, with a
reproduce slow vertical desaturation (percolation due to bottom maximum of 8.2 cm/year (Fig. 2c). The displacement spatial pattern
drainage) and associated vertical consolidation in the aquitard. The remained similar however, the deformation footprint resembling an el­
model is calibrated with displacement data from ERS, ENVISAT, Cosmo- lipse centered to the SE of Lorca city, having a width of 25 km and height
SkyMed satellites and the global positioning system GNSS. In light of the 7 km. Despite the difference between the displacement rates, it should
results, the proposed model and the historical subsidence prediction are be noted that all results are expressed in LOS geometry. ERS and
discussed. ENVISAT acquisition geometries have an incidence angle of 23◦ ,
whereas the CSK angle is 39◦ . Assuming that subsidence generates a

3
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Fig. 2. InSAR-derived deformation in the Alto Guadalentín Basin. LOS deformation velocities from (a) ERS data (C-band, 1992–2000) (Bonì et al., 2015), (b)
ENVISAT data (C-band, 2003–2010) (Béjar-Pizarro et al., 2016), and (c) Cosmo-SkyMed data (X-band, 2011–2017).

mainly vertical displacement, all the measured InSAR data have been metamorphic substratum that represents the lower impermeable limit.
projected to the vertical component, obtaining subsidence rates of 13, Detailed geologic logging analysis of the continuous core led to the
11.9 and 10.55 cm/year for each period, that sign a declining trend over stratigraphic column depicted in Fig. 4. Based on eye inspection and the
time. percentage of core drilling recovery and basic laboratory tests, the col­
Two continuous GNSS stations (LOR1 and LORC) located in the NW umn was divided in two layers. The first, to a depth of 150 m, comprises
sector of the Alto Guadalentín Basin (see location in Fig. 1) have fine clays and silts that make up the aquitard. The second, at a depth of
recorded 3D displacements since 2008. Though not located in the area of 150 to 300 m, is composed by larger grain size materials (sands and
maximum subsidence, the recorded vertical displacement they register gravels) that make up the confined aquifer.
coincides with the deformation monitored by means of INSAR tech­ Four representative samples were taken at depths of 62.8-64 m
niques (Bonì et al., 2015). A new permanent GNSS station, continuously (TP2), 85.7–86.3 m (TP3), 221.9–222.2 m (TP10) and 291.3–291.8 m.
monitoring, was installed by the Geological Survey of Spain (IGME) in The initial void ratios of the four samples, respectively, were 0.60, 0.86,
the area of maximum subsidence in February 2016 (see location in 0.49 and 0.40. Upper samples (TP2 and TP3) are classified as clay and
Fig. 1). The information from this permanent station (named ORCA) is silt with low plasticity (CL and ML) according to the Unified Soil Clas­
processed daily by the National Geographic Institute (IGN) and confirms sification System (USCS) with a liquid limit (LL) between 21.2% and
the deformation trend observed through INSAR techniques. Fig. 3 shows 45.4%, and a plastic limit (PL) between 20.7% and 32.1%. The fine
the subsidence measured by the different techniques since 1992 in the content of the upper samples (<74 μm) is between 90% and 100%, and
area of maximum subsidence. The accumulated subsidence recorded is its clay-sized fraction (<2 μm) is between 30% and 60%. Deeper samples
around 3.1 m over the last 27 years. This curve is used later to calibrate (TP10 and TP13) are classified as sand with fines and silt with low
the subsidence model. plasticity (SL and ML). The fine content is 20% for TP10 and 75% for
TP13, whereas the clay-sized fraction ranges between 5% and 35%,
respectively. Details regarding the index properties derived from labo­
2.3. Geotechnical description ratory tests can be checked in Table 1.
The four samples underwent a series of laboratory tests for their
A research borehole was drilled in the area of maximum subsidence mechanical characterization. Shear strengths were determined through
in January 2015 (see location in Fig. 1). The borehole —300 m in depth, consolidated drained and undrained triaxial tests. Volumetric behaviors
providing a continuous core 24 cm in diameter— failed to reach the were determined through oedometric tests on saturated samples. Results
are described in section 3.2.2.
Determining the phreatic level during the borehole drill operation
was complicated by the non-stop rotary drilling process and the non-
casing and collapse of the walls. Still, it was estimated at around 270-
m depth. This estimation matches the degree of saturation in samples
TP10 and TP13. Average values of saturated conductivity kw, based on
many measurements in triaxial cells do not allow one to distinguish
precise hydraulic behaviors among the samples.
For the partial saturation characterization, soil water retention
curves (SWRC) and unsaturated hydraulic conductivities were esti­
mated, respectively via the (van Genuchten, 1980) function and the
Mualem model (Mualem, 1976):

(1)
1/n− 1
Sr (s) = (1 + (α∙s)n )
( ( )1− )2
1/n
(2)
n
kw (Sr ) = ks ∙Sr0.5 ∙ 1 − 1 − Srn− 1

The parameters α, n, ks, were determined in view of a basic


Fig. 3. Subsidence recorded by different techniques since 1992 in the area of geotechnical description (the particle size distribution and bulk density
maximum subsidence.

4
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Fig. 4. (a) Simplified stratigraphic column. (b) Adopted soil water retention curves. (c) Unsaturated hydraulic conductivities.

Table 1
Basic geotechnical description obtained from laboratory tests. LL: liquid limit, PL: plastic limit, PI: plastic index, Gs: specific gravity, e0: initial void ratio, n0: initial
porosity, W: humidity, Sr: degree of saturation, kw: conductivity, CL: clay with low plasticity, ML: silt with low plasticity, SM: sand with non-plastic silty fines, SiC: silty
clay, SiCL: silty clay loam, LS: loamy sand, CL: clay loam The average value of conductivity kw is based on a large number of measurements.
Sample Depth Particle size LL PL PI USCS Soil type USDA Soil G0 e0 n0 γd W Sr kw
Name (m) distribution (%) (%) (%) description texture class (Kg/ (%) (%) (m/s)
m3)

TP2 62.85–64 60% clays 45.4 32.1 13.3 CL Si C 2.72 0.60 0.37 1703 20 91 2.25e-
40% silts 10
TP3 85.75–86.35 30%clays 21.2 20.7 10.5 ML Si C L 2.79 0.86 0.46 1501 29 95 5.07e-
60% silts 10
10% sands
TP10 221.90–222.5 5% clays / / / SM LS 2.72 0.49 0.33 1829 10 56 1.70e-
15% silts 10
80% sands
TP13 291.30–291- 35% clays 34.9 19.2 15.7 ML CL 2.77 0.40 0.28 1984 14 99 1.39e-
80 40% silts 10
23% sands
2% gravel

of TP2 and TP10 for the aquitard and aquifer layer, respectively) and the 3. Modeling
pedotransfer functions (PTFs) implemented in Roseta software (Schaap
et al., 2001), as shown in Table 2. Adopted soil water retention curves 3.1. Conceptual model
and unsaturated hydraulic conductivities are plotted in Fig. 4.
The modeling approach is based on a single observation (Fig. 4):
surface subsidence monitoring indicates that the deformation trend has
Table 2 been nearly constant since 1995, and no deceleration pattern has been
Van Genuchten (1980) water retention and Mualem unsaturated hydraulic observed since 2012, after the stabilization of the hydraulic head
conductivity model parameters. confirmed by recent installed piezometers. Three mechanical processes
Material (USDA soil texture class) α (1/cm) n ks (m/s) and modeling alternatives serve to explain this observation: i) delayed
Aquitard (Si C) 1.51 e-4 1.23 2.10 e-10 deformation at constant load due to a viscoplastic mechanical behavior
Aquifer (L S) 4.94 e-4 1.64 5.49 e-9 of the involved materials, ii) softening behavior and progressive decay of

5
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

material stiffness, or iii) slow consolidation due to slow downward of nodal displacements and pore pressures. The importance of coupling
percolation of water in the 150 m thick overlying clayey unit. Deep has been discussed in Lewis and Schrefler (1998) and Gambolati et al.
structural deformation controlled by active faulting has been discarded. (2000). According to the latter reference, uncoupled solutions would
Of the three above alternatives, given the absence of viscous and/or appear to be fully warranted in a typical normally consolidated and
softening evidence in oedometer and triaxial tests, the authors deemed pressurized basin on any timescale of practical interest, and would
the third one to be most appropriate. The conceptual model is as follows: preferably be used by groundwater hydrologists and petroleum engi­
water is extracted from the deep confined aquifer in the period of time neers. The simplified uncoupled approach is applied in the codes most
1960–2012. The confined aquifer depressurizes and empties, and the often used to simulate regional-scale land subsidence due to ground­
water head at the aquifer base goes down. We assume a constant water water withdrawal: MODFLOW-based packages IBS1 (Interbed Storage
head since 2012, as the aquifer has not recharged since. From 1960 to Package, version 1) (Leake and Prudic, 1991), IBS2 (Leake, 1990), and
2100, the water contained in the aquitard would slowly percolate SUB-WT (Leake and Galloway, 2007).
downwards due to bottom drainage, producing slow consolidation In this paper, in fitting with the proposed conceptual model, an
process and delayed subsidence. Fig. 5 summarizes the subsidence alternative formulation is put forward:
conceptual model.
1. A simplified 3D groundwater model is developed to describe historic
3.2. Modeling strategy regional changes in the hydraulic head of the Alto Guadalentín
confined aquifer Basin since 1960.
Soils and rocks are geomaterials with voids that may be filled by 2. Changes in the water head located in the area of maximum subsi­
water, air, and other fluids. Hence they are multiphase materials, their dence, coinciding with the location of the performed borehole, are
mechanical behavior being governed by the coupling between phases. selected from this previous 3D groundwater model.
The classic model describing coupling between pore fluid and soil 3. A transient fully-coupled approach is applied to a 1D vertical column
skeleton was proposed by Biot (1941). It has been modified by the group by prescribing the previously computed head at the bottom aquifer
of Prof. Zienkiewicz at the University of Swansea (UK), who proposed boundary. Detailed slow vertical desaturation processes in the
formulations that proved more efficient from a computational point of aquitard, percolation due to bottom drainage, and associated vertical
view (Zienkiewicz et al., 2000). Also worth mentioning is works of Lewis consolidation are then computed using the non-linear Biot equations.
and Schrefler (1998), Coussy (1995) and de Boer (2000).
The numerical solution of two-phase consolidation problems in This numerical methodology aims to capitalize advantages of each
quasi-static conditions (i.e., neglecting inertial effects) to compute model. The simplified 3D groundwater model reproduces the global
ground subsidence in pumped aquifers relies on one of two main ap­ regional trend of water flows and changes in the hydraulic head of the
proaches to describe the interaction between the flow field and the stress aquifer. Water flows are basically horizontal, driven by recorded yearly
field. The first, referred to as the “uncoupled” approach, neglects the water discharge and recharge. Discharges can be attributed to local,
effect of the volumetric strain on pore pressure. This allows reducing the known pumping rates, while recharges rely on natural input from lateral
free variable to the pore pressure only. The computational effort is discontinuous water streams. In this simplified model, for practical
similar to the one required for time-dependent seepage analyses. Once calibration purposes, partial saturation is not taken into account, and the
pore pressure is obtained, it may be used as an external source of aquifer parameters, including hydraulic conductivity and specific stor­
strength in a strain-stress analysis of the porous system to provide the age, are keep constant.
medium deformation and land subsidence. Conversely, the 1D hydro-mechanical model aims to reproduce in
The so-called “coupled” approach takes into account the effect of detailed the slow vertical saturated-unsaturated flow in deformable
volumetric strain on the pore pressure and considers as unknowns, at the porous media due to bottom drainage. This so-called subsidence model
same time, both the effective stress and the pore pressure distributions. accounts for full coupling of the effective stress and pore pressures, and
In its most popular finite element formulation, the free variables consist includes capillary pressure, partial saturation and non-linear

Fig. 5. Conceptual model (a) 1960: the aquifer is confined in equilibrium; (b) 1960–2012: pumping period with no surface recharge, the water head in the aquifer
goes down; (c) 2012–2100: no pumping, no recharge, the water contained in the aquitard slowly percolates downwards due to bottom drainage, and there is slow
consolidation and delayed subsidence.

6
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

conductivity of the materials. An advanced non-linear constitutive


model is moreover proposed to reproduce accurate inelastic soil
compressibility.
The following paragraphs present a basic description and the results
of the groundwater model, constitutive model and subsidence model.

3.3. Groundwater model

Our study relied on a previously published hydrogeological model of


the Alto Guadalentín Basin (Ezquerro et al., 2017), updated to reflect the
aquifer’s groundwater flow evolution since 1960. The proposed model is
solved using MODFLOW software (Harbaugh, 2005).

3.3.1. Groundwater model description


For a confined aquifer, MODFLOW simulates the groundwater flow
in a saturated medium using a form of the Navier–Stokes equation
written as:
[ ] [ ] [ ]
∂ ∂h ∂ ∂h ∂ ∂h ∂h
kxx + kyy + kzz + W = Ss (1 − n) (3)
∂x ∂x ∂y ∂y ∂z ∂z ∂t

where:

- kxx, kyy, kzz are the values of hydraulic conductivity along the x, y,
and z coordinate axes (L/T);
- h is the potentiometric head (L), which can be related to the pore-
pressure: pw = ρw ∙ g ∙ h (ρw is the density of the fluid and g the
gravitational acceleration);
- W is a volumetric flux per unit volume, representing sources and/or
sinks of water, where negative values are extractions and positive
values are injections (T− 1);
- Ss is the specific storage of the porous material (L− 1);
- n is the porosity; and
- t is time (T).

MODFLOW solves the partial differential eq. (3) using a finite dif­
ference approximation (Harbaugh, 2005). The 243.5 km2 basin was
discretized using square 100-m grid cells. In view of previous work (Bonì
et al., 2015; Béjar-Pizarro et al., 2016), and according to the geotech­
nical description and the simplified stratigraphic column adopted, the
hydrogeological model comprises two hydrogeological units/layers. The
first layer is composed by the fine-grained sediments with low conduc­
tivity. The second layer is the most productive part of the aquifer-
system, made up of the Plio-Quaternary deposits. A third layer acting
as the bottom of the model and representing the very low conductivity
Miocene materials is also introduced for the MODFLOW computation.
Values and spatial distribution of the material parameters (hydraulic
conductivity and specific storage) were calibrated based on the
UCODE_2014 code (Poeter et al., 2014) and the 47 available piezometric
level measurement points, spatially distributed. The boundary and
initial conditions of the model are depicted in Fig. 6: groundwater ex­
tractions from wells are known/prescribed and constitute the main
output of the aquifer-system, whereas the Alto-Bajo Guadalentín Basin
exchange is almost negligible (red line in Fig. 6a). Natural recharge of
the system comes from the lateral discontinuous water streams (Fig. 6a)
and surface infiltration from rainfall and irrigation return (Fig. 6b). It
should be noted that no significant infiltration is considered in the yel­
low area defined in Fig. 6b because of the very low surficial aquitard
conductivity and high evapotranspiration in this area. See (Ezquerro
et al., 2017) for a complete description and quantitative estimation of
yearly water recharge and discharge.
The hydrological model was implemented through two steps; first, a
steady-state simulation was performed using the 1960 data, after which
a transient simulation for the period 1960–2012 was run using the Fig. 6. Groundwater model (a) Lateral boundary conditions; (b) Superficial
boundary conditions; (c) Initial conditions. From (Ezquerro et al., 2017).
steady-state solution as the initial condition and fixing the time step as
equal to one year.

7
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

3.3.2. Groundwater model calibration and results


Considering the long duration of the simulated time span (52 years)
and the large global piezometric level change (between 170 and 210 m),
calibration of the material parameters (hydraulic conductivity and
specific storage) was performed successfully. The average absolute dif­
ference between simulated and observed piezometric levels is 14.9 m,
and the Root Mean Square Error (RMSE) around the X=Y line in a
simulated-observed plot is 17.4 m. Assuming an average piezometric
level drop of 190 m, both errors are under 10% of the total water head.
See (Ezquerro et al., 2017) for a complete description of the calibration
procedure and time and spatial error analysis.
Fig. 7 shows the groundwater model results, and more precisely, the
hydraulic head spatial distribution on four specific and relevant dates.
The maximum water level decline, above 200 m, was calculated in the
central-west area, where a major concentration of deep wells is found.
The area where the borehole is located is less depressed due to a lower
density of wells. Fig. 8 shows the time evolution of the hydraulic head in Fig. 8. Groundwater model results. Computed time evolution of the hydraulic
the borehole cell —a continuously declining trend that accelerates head at the bottom of the borehole grid cell.
during the massive extractions of 1970s and 1980s. After declaration of
the aquifer system as partially overexploited in 1988, the level was at the bottom boundary of the assumed simplified stratigraphic column.
stabilized for 3 years. But then the declining trend picked up again as a Recently installed piezometers indicate that the level is again stabilizing,
result of three drought periods occurring between 1990 and 2012 (CHS, hence a constant hydraulic head is assumed after the end of the
2014). The model calculates a 160 m head drop over the period 2012–2100 groundwater model computation, for subsidence prediction
1960–2012. The time evolution of the hydraulic head plotted in Fig. 7 is purposes.
used in the forthcoming subsidence analysis as a pore-water prescription

Fig. 7. Groundwater model results. Computed hydraulic head spatial distribution in: (a) 1972; (b) 1988; (c) 1993; (d) 2012. From (Ezquerro et al., 2017).

8
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

3.4. Constitutive model: generalized Plasticity state parameter-based ( )ζc


model ec = eΓ − λ
p′
(9)
p′atm
In this study, the generalized plasticity/critical state-based model
proposed by Manzanal et al. (2010, 2011a, 2011b) is used to describe eΓ, λ and ζc are constant parameters and patm′ is the atmospheric pressure
the mechanical behavior of the materials involved. The new model equal to 101 kPa.
stands as an enhancement of the basic Generalized Plasticity model by
Pastor-Zienkiewicz (Pastor et al., 1990), which is particularly suited to (iii) The unit tensor n discriminating loading and unloading condi­
deal with cyclic loading. Here, a single set of material parameters can be tions is defined as:
used to reproduce the mechanical behavior of soils under different ⎛ ⎞T
confining pressures and void ratio states through the definition of a state df
⎜√̅̅̅̅̅̅̅̅̅̅̅̅̅ 1 ⎟ d0 ( )
n=⎝ , √̅̅̅̅̅̅̅̅̅̅̅̅̅ ⎠ df = Mf ∙emψ s − η (10)
parameter. Therefore, it can be used at different depth locations, and the 1 + d2 1 + d2 Mf
model automatically stiffens the material across the assumed simplified
f f

stratigraphic column.
where:
3.4.1. Constitutive model description Mf
( )β
e
The elastoplastic behavior of a material can be defined by specifying = h1 − h2 (11)
Mg ec
the relationship between the increments of strain and stress as:
and h1, h2 and β new model constants.
dε = Cep ep
L/U : d σ or dσ = DL/U : dε (4)

(iv) iv) The loading and unloading plastic module HL/U is postulated
where dσ and dε are, respectively, the increments of the stress and strain
as:
tensors, and Cep
L/U is the fourth order constitutive tensor which depends
on stress level, past history of material, direction of applied loading ( )β
increment and state variables: √̅̅̅̅̅̅̅̅̅̅̅̅̅̅

− β0 eec
(12)

HL = H0 ∙e ∙ p′ ∙p′atm ∙HDM ∙Hf (Hv + Hs )
1 ( )
Cep e
L/U = C + ngL/U ⊗ n (5)
HL/U where each plastic function is defined as:
Therefore, in order to fully characterize the material behavior, we
must provide the following items (Pastor et al., 1990): (i) The elastic Hv = Hv0[Mg ∙ e− β vψ s
− η] (ξ
MAX
)γ [ ( α ) η ]− 1/α
HDM =
constitutive tensor Ce, (ii) The unit tensor describing the direction of ξ ξ = p′ 1 − ∙
1 + α Mf
plastic flow ngL/U in loading and unloading, (iii) The unit tensor n Hs = β1 ∙ e− β0∙ξdev ( )
η 4
(
1
)
Hf = 1− ηf = 1 + ∙Mf
discriminating loading and unloading conditions, and (iv) The loading ηf α
and unloading plastic moduli HL/U.
The generalized plasticity critical state based model proposed by
(Manzanal et al., 2011; Manzanal et al., 2011b) was developed assuming
that the material is isotropic, for which reason it was formulated in terms
and H0′ , β0′ , β, Hv0, βv, β0, β1, γ are parameters of the model.
of the three invariants of the effective stress tensor p′ , q and θ together
Unloading is supposed to be elastic, adopting an infinite unloading
with the work conjugate strain invariants εv and εs. The generalized
plastic modulus:
plasticity state parameter-based model provides the following defini­
tions for each item that are void ratio dependent: HU = ∞ (13)
For a complete description of the constitutive model, see Manzanal
(i) The elastic constitutive tensor Ce is defined using as shear and
et al. (2010, 2011a, 2011b). The model is governed by two elastic and
bulk modulus the expressions:
four critical state parameters, four parameters related with dilatancy
(2.97 − e)2
√̅̅̅̅̅̅̅̅̅̅̅
2(1 + ν) and plastic flow, and eight parameters related with the plastic modulus.
(6)

G = Geso ∙ ∙ p′ ∙p0 K = G
1+e 3(1 − 2ν)
3.4.2. Constitutive model calibration
where Geso and ν are the unidimensional elastic shear modulus and The constitutive parameters were calibrated using the step-by-step
Poisson’s ratio, respectively, while e is the void ratio. procedure described in [(Manzanal, 2008)] and [Cuomo et al., 2018].
Four oedometer tests and four drained triaxial tests were performed on
(ii) The unit tensor describing the direction of plastic flow ngL/U in the representative samples described in section 2.3.
loading and unloading is defined as a function of dilatancy dg: Fig. 9 represents the comparison between model predictions and
⎛ ⎞T experimental data, plotting the recorded void ratio vs applied axial
stress. The proposed model fits fairly well with the experimental data,
⎜ dg 1 ⎟ d0 ( )
ng = ⎝√̅̅̅̅̅̅̅̅̅̅̅̅̅ , √̅̅̅̅̅̅̅̅̅̅̅̅̅ ⎠ dg = ∙ Mg ∙emψ s − η (7) differing just at the end of the unloading branches, where unloading
2 2 M
1+d 1+d stresses are no longer realistic. It must be underlined that high-pressure
g
g g

oedometer tests (up to 16 MPa) were performed on the two deeper


where d0 and m are model constants. samples, TP10 and TP13.
ψ s = e − ec (8)
3.5. Subsidence model
represents the state parameter.
η is the stress ratio and Mg is the Critical State Line in the p′ -q plot; ec In this study, the finite element code GEHOMADRID (Fernández-
is the void ratio at the critical state for a given mean effective stress p′ : Merodo, 2001; Mira, 2002), where the proposed constitutive model was
implemented, is used to describe the subsidence phenomena. The basic
elements of the adopted mathematical and numerical models are
recalled below.

9
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Fig. 9. Comparison between generalized plasticity critical state-based model predictions and experimental data in oedometer tests.

{ ( )}
3.5.1. Subsidence model description
- 1
= n KSww +1−KSnSw Sw +pw CnS with CS = n dSw
Hydro-mechanical coupling between pore fluids (air and water) and Q* dpw

the soil skeleton is represented by the Biot equations (Biot, 1941). It is - KS: compressibility of the solid
assumed that: i) for a consolidation analysis describing slow subsidence - Kw: compressibility of the water
phenomena, inertia effects and second derivatives with respect to time - S: matrix defined with the kinematic strain-displacement relation:
( )
are neglected, ii) for partially saturated soils, the pore air pressure pa is
εij = 12 ∂∂uxij +∂∂uxji ≡ S∙du
at constant atmospheric pressure, and iii) velocities and accelerations of
pore fluid relative to soil skeleton are small. The governing equations - m: vector (1, 1, 1, 0, 0, 0)T
that describe the balance of linear momentum of the soil mixture, the
balance of linear momentum of the pore water, and the balance of mass The effective stress σ’ in (14) must be replaced by the stress-strain
of the pore water can be simplified as: constitutive relationship defined in eq. (4) (see section 3.3.1). Coupled
′ partial differential eqs. (14) must be complemented by suitable bound­
ST (σ − Sr pw m) + ρb = 0
ary and initial conditions. The main advantage of this formulation,
(14)
Sr mT Su̇ − ∇T kw ∇pw +
ṗw
+ ∇ T k w ρw b = 0 called coupled u-pw formulation, is that field variables are reduced to
Q* two: displacements u and pore pressures pw. It is worth mentioning that
partial saturation is taken into account through saturation-capillary
where the following notation was used: pressure and conductivity-saturation relationships defined in eq. (1)
and (2).
- pw: pore water pressure and ṗw : first time derivative of pore water GEHOMADRID solves these equations using standard Galerkin
pressure techniques for the space discretization, Generalized Newmark time
- u: solid displacements and u̇: first time derivative of displacements stepping scheme for the time discretization, and Newton Raphson al­
- σ’: effective stress that is related to the total stress σ with σ ’ = σ − Sr gorithms for the non-linearities. For a complete description of numerical
∙ pwm resolution strategies see Fernández-Merodo (2001), Mira (2002) or
- Sr (s): degree of saturation as a function of suction s (defined in eq. Zienkiewicz et al. (2000).
(1)). In that case pa = 0 and suction s = pa − pw = − pw (suction is The simplified stratigraphic column Fig. 4 was discretized using 20
positive for negative values of pw) quadratic quadrilateral elements and 103 nodes. A coupled u-pw
- kw: non-linear conductivity is a function of the degree of saturation, formulation is implemented in the quadratic quadrilateral elements
kw(Sr) (defined in eq. (2)) (eight nodes for the displacements field variable and four nodes for the
- ρ: density of the mixture defined as ρ = (1 − n)ρS + nSrρw, where n is pore pressure field variable). Horizontal displacements are fixed in the
the porosity, ρS is the density of the solid grains and ρw is the water whole column (1D analysis), while vertical displacements are fixed in
density. the bottom boundary, and pore water pressure is prescribed in the
- b: external forces (gravity)

10
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

bottom boundary according to the hydraulic head computed by the subsidence increase of 1 m (from 5.8 m to 6.8 m) in the period
groundwater model (Fig. 8). No significant superficial water recharge 1960–2020.
due to surface infiltration from rainfall of irrigation return was consid­ Hydraulic coupling between the 3D groundwater model and the 1D
ered for the top boundary, as it was in the groundwater model (the subsidence model is analyzed prescribing the computed groundwater
borehole is located in the yellow area defined in Fig. 6b). Only gravity hydraulic head in three different ways. In the first one, the hydraulic
load is considered. head is fixed at the bottom boundary, allowing for vertical drainage and
The calibrated generalized plasticity critical state-based model pa­ vertical consolidation in the aquitard and in the aquifer. In the second
rameters (Table 3) and the assumed Van Genuchten water retention one, the hydraulic head is fixed in the aquifer layer, allowing for vertical
curve and Mualem unsaturated hydraulic conductivity model parame­ drainage and vertical consolidation only in the aquitard; in this case,
ters (Table 2) describe the hydro-mechanical behavior of the aquitard- aquifer conductivity does not intervene in the computation, and the pore
aquifer system in eq. (14). water pressure in the aquifer corresponds to the hydrostatic solution. In
The subsidence computation entailed two steps. After performing a the third, the hydraulic head is fixed in the whole column, i.e. in the
static analysis to derive an initial equilibrium state corresponding to the aquifer and in the aquitard, in which case the aquifer and aquitard
year 1960, a time-dependent analysis for the period 1960–2100 was permeabilities do not intervene in the computation, and pore water
performed based on the 1960 steady state solution as the initial condi­ pressure along the column corresponds to the hydrostatic solution.
tion, then fixing the time step as equal to one year. Fig. 13 shows the time evolution of subsidence for each of the three
cases. The computed subsidence when fixing the hydraulic head at the
3.5.2. Subsidence model calibration and results bottom boundary fits well with the observed subsidence trend, by means
The first static analysis computation leads to an initial equilibrium of a slow consolidation process, since 1995.
state where the computed effective stress along the column increases In turn, Fig. 14 shows the evolution along the column of pore pres­
with depth whereas the computed void ratio decreases with depth, sure, vertical displacement and vertical deformation. The more realistic
meaning a more compacted state for deeper materials (see Fig. 10). pore pressure profiles correspond to the first case, where non-linear
Different compaction states can be assumed in the early 1960’s, aiming conductivities (see eq. (2) and Fig. 4) are updated in the whole col­
to represent the historical long-term deposition of the quaternary sedi­ umn. Hydrostatic pore pressure profiles in the third case (instataneous
ments. In this work, three initial compaction states were considered, consolidation) are not realistic, and this case is shown only for illus­
adopting as initial void ratio profiles from the top to the bottom of the trative purposes. The time evolution of vertical displacement and ver­
column: (0.56–0.51), (0.61–0.54), and (0.66–0.58). These three initial tical deformation along the column show where consolidation takes
states correspond to equilibrium situations computed through the static place over time.
analysis, looking to fit the void ratio 0.6 of the surficial sample TP2. Finally, the computed subsidence corresponding to an initial
As the adopted constitutive model is stress- and void-ratio- compaction making the void ratio index decrease from 0.61 to 0.54,
dependent, the assumed initial compaction has a strong influence on given a 300-m-depth column and prescribing the groundwater hydraulic
the final subsidence results. Fig. 11 shows the evolution of the computed head at the bottom boundary, is the best result overall, closely fitting the
subsidence for the three different initial states. The computed subsi­ subsidence monitored for the period 1995–2012. This calibrated solu­
dence is greater for higher values of the void ratio, i.e. for lower initial tion leads to an historical subsidence of 5.8 m from 1960 to 2020, and a
compacted materials. Computed subsidence for an initial void ratio prediction of up to 7.3 m in 2100 (for an assumed constant hydraulic
between 0.61 and 0.54 fits well with the observed subsidence trend since head since 2012).
1995.
Inspection of vertical displacement and vertical deformation profiles 4. Discussion
in Fig. 14 shows that, despite stiffening of the material with depth due to
the generalized plasticity state parameter constitutive model, the ma­ The proposed historical subsidence quantification depends heavily
terial deforms at great depth. Thus, the assumed thickness of the on the assumed hypotheses. In the following, the most important ones
compressible material should bear an influence on the final subsidence are discussed, and the associated errors analyzed.
results. As the exact depth of the bedrock was not determined in the 300- First of all, the proposed methodology for modeling subsidence leans
m-depth research borehole executed, the influence of aquifer thickness on the conceptual model presented in Fig. 5. The subsidence phenom­
was analyzed (see Fig. 12). An increase in thickness of 300 m to 350 m (i. enon is controlled by hydro-mechanical coupling in porous material,
e. an increase of the aquifer layer from 150 m to 200 m) implies a and therefore depends on groundwater flow and pore water pressure
changes. The conceptual model decomposes the groundwater flow into
two components according to the velocity and the direction of the flow.
Table 3 The first is the regional groundwater flow, governed by discharges due
Calibrated model parameters. to pumping and recharges due to natural lateral discontinuous water
Parameter Value streams. This flow is mainly horizontal and located in the deep aquifer. It
should be noted that no effective surface infiltration from rainfall and
Elasticity Geso (kPa) 49.
ν 0.25 irrigation return is considered in the yellow area defined in Fig. 5b, thus
Critical State Mg 0.9 reducing the vertical flow in this area as well. The regional groundwater
eΓ 0.63 flow is computed through a coarse 3D MODFLOW model where the Alto
λ 0.0085 Guadalentín Basin is discretized using a simplified two-layered stratig­
0.6
raphy (Fig. 4). The overlying aquitard, composed by fine-grain materials
ζC
Plastic flow d0 1.
m 0.9 of low permeability, acts as a physical hydraulic barrier. The second
h1 / h 2 1. / 0. component is the vertical flow. This flow in the aquitard is not only
Plastic modulus H0′ 1. controlled by gravity but above all by bottom drainage because of pore
0.
β0 ′
water pressure lowering in the aquifer. Recalling that the vertical
β 1.8
Hv0 21. dimension of the aquitard (150 m) is much smaller than the horizontal
βv 7. one (tens of kilometers), bottom boundary control appears to prevail.
β0 20. The vertical flow is modeled in the second coupled hydro-mechanical
β1 5. 1D GEHOMADRID model, where the bottom drainage and column
8.
desaturation are represented via a bottom boundary condition that
γ

11
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Fig. 10. (a) Finite element mesh. Computed initial state: (b) vertical effective stress in Pa., (c) void ratio, (d) pore pressure in Pa, and (e) degree of saturation.

Fig. 11. Subsidence computation. Influence of initial conditions. Fig. 13. Subsidence computation. Influence of hydraulic boundary conditions.

3D, the conceptual model based on decoupling fast horizontal flow and
slow vertical flow is supported by the difference in permeability of the
simplified two-layered stratigraphy and the size of the aquitard-aquifer
system.
The reliability of the regional 3D MODFLOW model is partially
addressed in section 3.1.2. The average absolute differences between the
simulated and recorded piezometric levels is 14.9 m; and the Root Mean
Square Error (RMSE) around the X=Y line in a simulated-observed plot
is 17.4 m. A total of 47 piezometric measurement points, spatially
distributed, were used for the calibration. Unfortunately, no pore pres­
sure measurements were available for the hotspot, i.e. the research
borehole cell; installation of electric piezometers in the borehole failed
due to collapse of the borehole walls. According to eye-inspection during
the borehole drill operation in 2015, the water table height was roughly
estimated around 30 m a.s.l. (270 m b.g.s.), indicating than the
computed hydraulic head in the aquifer, 120 m a.s.l. (180 m b.g.s), could
have been lower estimated. Moreover, the degree of saturation of sample
Fig. 12. Subsidence computation. Influence of aquifer system thickness.
TP10, 80 m a.s.l. (220 b.g.s) was determined by laboratory tests to be
around 56% (Table 1), far from saturated conditions, indicating that the
prescribes the pore pressure drop due to pumping. Importantly, this real hydraulic head should be below this mark. The computed 160 m
model —unlike the regional MODFLOW— takes into account ground­ hydraulic head decline (from 280 to 120 m a.s.l.) transforms the initially
water flow in non-saturated material. Therefore, although the flow is confined aquifer into a non-confined aquifer. It must be underlined that

12
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Fig. 14. Computed pore pressure, vertical displacement and vertical deformation time evolution along the assumed simplified stratigraphic column. Influence of
hydraulic boundary prescription: column (a) hydraulic head fixed at the bottom, column (b) hydraulic head fixed in the aquifer, column (c) hydraulic head fixed fixed
in the entire column (aquifer+aquitard).

Fig. 15. (a) Computed degree of saturation time evolution along the assumed simplified stratigraphic column. (b) Computed void ratio time evolution along the
assumed simplified stratigraphic column.

13
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

no pore pressure information above the water table, i.e. suction, is given representative samples collected in 2015 (see Fig. 15b). Only compac­
by the saturated MODFLOW model. tion in TP3 differs substantially from the proposed solution and indicates
The reliability of the 1D non-saturated flow model, associated with that the initial void ratio assumed for 1960 (Fig. 10c) should have been
the coupled hydro-mechanical GEHOMADRID model, can be partially larger. Yet Figs. 15b and 14a indicate that, according to the proposed
addressed by comparing the computed degree of saturation along the model, deformation was located at the base of the column in the period
column with the saturation determined by laboratory tests on the four 1960–2012 (pumping period) and deformation is located nowadays
representative samples collected in 2015 (see Fig. 15a). The degree of (since 2012) in the central area, between the lower part of the aquitard
saturation computed by GEHOMADRID depends directly on the 120 m a. and the upper part of the aquifer. These results are in agreement with
s.l. hydraulic head computation performed by MODFLOW (the latter pore pressures changes along the column predicted by the model. Un­
considered a bottom boundary condition in GEHOMADRID); it could fortunately, no deformation measurements along the column are avail­
explain not only the saturation difference at TP10 but also at TP2 and able to confirm those results, since the installation of a deep magnetic
TP3. It should be stressed that the pore pressure profiles of Fig. 14a are extensometer (300-m-depth) in 2015 failed due to technical difficulties
realistic because of the non-linearity of permeability. Pore pressure (collapse of the borehole walls).
tends to an equilibrium state. At the top of the column, in the aquitard, In sum, the proposed subsidence model was calibrated to fit the su­
negative pore pressure (suction) remains nearly constant, around 200 perficial subsidence trend observed since 1995 and gives a historical
KPa, without reaching large suction values. This is because material subsidence of 5.8 m from 1960 to 2020. The model also predicts sub­
becomes much more impermeable as the column desaturates, i.e. as the sidence up to 7.3 m in 2100 for an assumed constant hydraulic head
suction increases (see definition of non-linear permeability in eq. (2) and since 2012, though this may be a conservative estimate, as the natural
Fig. 4). Deserving mention as well is the fact that a better partial satu­ recharge and hydraulic head should recover now with the end of over­
ration characterization could be achieved using ad hoc non-saturated exploitation. If such water recovery is confirmed in the aquifer, the
laboratory tests instead of the approximated water retention functions. proposed mechanical constitutive model could take into account the
No suction measurements were carried out to confirm this hydraulic associated elastic rebound (see unloading branches of oedometer tests in
behavior in real conditions. To the authors’ best knowledge, very few Fig. 9).
studies have introduced capillary pressures and non-linear permeability Despite visible damages could be expected due to the large accu­
into subsidence quantification. The Liakopulos drainage column exper­ mulated subsidence, no significant damages have been detected. The
iment in Del Monte sand (Liakopulos, 1965) has served as a reference for studied basin is mainly an agricultural area, with no reports of wide­
different authors (Narasimhan and Witherspoon, 1978; Zienkiewicz spread or severe building damage due to angular distortion thus far. The
et al., 1990; Lewis and Schrefler, 1998; Fernández-Merodo, 2001) to sediment thickness and homogeneity may prevent differential subsi­
check their numerical models. The role of air pressure in the transition dence. The main problems reported are changes in the flood charac­
from a fully saturated to a partially saturated condition during the teristics of the basin, becoming more prone to inundation: two recent
desaturation process in the Liakopoulos experiment was highlighted by floods followed strong storm episodes in September 2012 and September
Lewis and Schrefler (1998) using a three-phase flow approach (u-pw-pa 2019. These events submerged thousands of hectares of flat agricultural
formulation), but it remains an open question in confined aquifer sys­ land that coincides with the reported maximum subsidence area (red
tems. Here, the proposed two-phase analysis (u-pw formulation with area in Fig. 2c). It is also worth mentioning that a large earth fissure,
pa=0) appears sufficient to reproduce the recorded subsidence and 400-m long and 2–3-m deep (Pastor et al., 2020), appeared in the area
explain the desaturation process. after the 2012 flood. However, further studies are needed to determine
From the mechanical point of view, the particle size distribution of whether this earth fissure and the observed/computed subsidence phe­
the four representative samples collected at different depths is consistent nomena are related.
with the two-layer stratigraphic simplification adopted, TP2 being a clay In view of the literature (Gambolati and Teatini, 2015), the
and TP10 sand (see Table 1). Nevertheless, the oedometer tests did not computed historical subsidence of 5.8 m since 1960 makes Alto Gua­
allow us to distinguish differences in mechanical behavior between the dalentín Basin one of the world’s most important land subsidence spots
aquitard and aquifer materials. On the one hand, only four samples due to groundwater withdrawal. In closing this discussion section, an
represent the overall 300-m deep column, while on the other hand, it attempt is made to contextualize the phenomena by comparing repre­
was difficult to recover sand and gravel material during the borehole sentative data (average maximum subsidence, subsidence rate, depth of
drill operations, and mechanical characterization errors are possible. pumping, hydraulic head drop and geology) in selected areas of major
Still, the proposed 1D coupled hydro-mechanical GEHOMADRID model land subsidence worldwide. Table 4 summarizes the comparison. Due to
incorporates a new advanced constitutive model that was able to the uniqueness of each case, no clear deductions can be made. The
reproduce the mechanical behavior of the four representative samples. highest accumulated subsidence is seen for historically overexploited
Very few constitutive models are able to reproduce material mechanical and monitored basins like Mexico City and San Joaquín Valley (13 m
behavior under different confining pressures and void ratio states using and 8.5 m, respectively). These areas present a great difference between
a single set of material parameters. In our case, 18 parameters were their maximum compressible thickness (more than 250 m) and
needed to reproduce the results of the four oedometer tests (Fig. 9). For groundwater change (100 m), supporting the great disparity of ground
the sake of comparison, twenty parameters would be needed to repro­ subsidence behavior. Moreover, recently detected ground subsidence
duce the same results from the four oedometers with a basic Modified areas like Beijin or Jakarta present very high displacement rates yet
CamClay model (five parameters for each oedometer). minor groundwater changes.
Regarding the adopted 1D approximation column and the assumed
purely vertical deformation, Fernandez et al. (2018), using a spatially 5. Conclusion
distributed GNSS network, made it evident that horizontal displace­
ments are not negligible in the Alto Guadalentín Basin. In the future, Numerically simulating subsidence due to groundwater withdrawal
regional 3D subsidence analysis will take into account the “arc effect” at a regional scale is a complex problem that can be approached through
due to surrounding stiffer material and the 3D distributed groundwater different models. In this paper, an original numerical methodology is
model to reproduce 3D strain and horizontal displacements. proposed to quantify historical subsidence in the Alto Guadalentín Basin
The reliability of the 1D subsidence model, associated with the since 1960.
coupled hydro-mechanical GEHOMADRID model, can be partially The confined aquifer is firstly simplified as a two-layered aquitard-
addressed comparing the computed void ratio evolution along the col­ aquifer system. The new approach relies on decomposition of the
umn with void ratios determined by laboratory tests in the four groundwater flow into two components, according to the velocity and

14
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Table 4
Selected areas of major land subsidence due to groundwater withdrawal worldwide.
Location Accumulated Maximum subsidence Depth of Groundwater Maximum References
Subsidence (m) velocity (cm/year) pumping (m) (Piezometric) level compressible
change (m) thickness (m)

Mexico Chalco Basin 13 (1960–2006) N/A 450 25 (1985–2005) 300 (Ortiz-Zamora and
8 (1985–2006) Ortega-Guerrero, 2010)
East Mexico 6.5 (1891–1952) 37 (ERS 1999–2000) N/A N/A 100 (Osmanoğlu et al., 2011;
City 2.5 (1952–1973) 28 (ENVISAT Cabral-Cano et al., 2008)
9.7 (1940–2007) 2004–2006)
Villa de Arista N/A 18.4 N/A N/A N/A (Chaussard et al., 2014)
Zamora N/A 12.8 N/A N/A N/A (Chaussard et al., 2014)
USA San Joaquin 8.5 (1926–1970) 19 (1926–1970) N/A 120 (1860–1961) 48 (Borchers and Carpenter,
2.86 (1960–2004) 30 (2007–2011) 2014)
1.2 (2007–2011)
Santa Clara 4.2 (1910–1967) N/A 245 72 (1910–1967) 18 (Borchers and Carpenter,
2014)
Iran Mashhad 1 (1995–2005) 22 (GNSS 2005–2006) N/A 30 (1980–2004) N/A (Motagh et al., 2007)
valley 25 (ENVISAT
2003–2005)
Japan Tokio 4.6 (1910–1970) N/A 160 60 (1910–1970) N/A (Poland, 1984)
China Beijing 0.8 (1955–2000) 11 (2003–2010) 200 30 (1955–2000) N/A (Zhang et al., 2014)
Indonesia Jakarta 2 (1960–2013) 26 N/A N/A N/A (Hay-Man et al., 2012;
Chaussard et al., 2013;
Erkens et al., 2015;
Bandung 0.8 (2004–2008) 20 N/A 17 (2004–2008) 132 Chatterjee et al., 2013;
Chaussard et al., 2013)
Spain Alto 3.1 (1992–2018) 13 280 200 (1961–2012) 200 (Bonì et al., 2015;
Guadalentín Béjar-Pizarro et al., 2016;
Ezquerro et al., 2019)

the direction of the flow. A rough 3D MODFLOW model reproduces Authors statement
depressurization and fast horizontal flow due to pumping in the
confined aquifer at the regional scale. A coupled hydro-mechanical 1D J.A.F.M. conceived the conceptualization of the study; J.A.F.M. and
GEHOMADRID model reproduces in detail the slow vertical flow and P.E worked on the analysis of the data and wrote the original draft. D.M.,
desaturation processes in the aquitard due to bottom drainage. The 1D P.M. and M.P. were involved in the data modeling; M.B.P., C.G.A, J.C.G.
GEHOMADRID model, unlike the regional MODFLOW, takes into ac­ D., J.L.V. R.S, G.B, R.A. and C.R.C worked in collecting data; R.M.M., J.
count a full coupling of effective stress and pore pressures, as well as the M., and G.H. supervised and reviewed the manuscript. All authors have
capillary pressure, partial saturation, and non-linear conductivity of the read and agreed to the published version of the manuscript.
materials. These key features are often overlooked in subsidence
modeling; nevertheless, they are needed to reproduce column desatu­
Declaration of Competing Interest
ration processes and obtain realistic pore water pressure profiles.
Moreover, an advanced general plasticity constitutive model is proposed
The authors declare that they have no known competing financial
to accurately reproduce inelastic soil compressibility across the column
interests or personal relationships that could have appeared to influence
in a uniform way, through the definition of a state parameter. This new
the work reported in this paper.
constitutive model is applied for the first time in modeling a subsidence
case study.
Acknowledgments
After a successful regional calibration, the 3D MODFLOW model
calculates a 160 m head drop in the area of interest over the period
This work was developed in the framework of AQUARISK Project:
1960–2012. Despite the assumed simplified hypothesis (simplified 1D
Analysis of Geological-Geotechnical Risks due to Groundwater Exploi­
two-layer stratigraphic column, initial void ratio state, compressible
tation Using Space and Terrestrial Techniques. Application on urban
thickness of materials and partial saturation characterization), the
structures and infrastructures was funded by Spanish Research Program
calibrated GEHOMADRID model reproduces the monitored 3.1 m sub­
from Ministry of Economy and Competitiveness (ESP2013-47780-C2-2-
sidence in the time period 1992–2018, and predicts subsidence of 5.8 m
R). The second author expresses gratitude for Ph.D. Student Contract
for the period 1960–2020. Accordingly, its historical subsidence makes
BES-2014-069076. Further support comes from the e-Shape project,
the Alto Guadalentín Basin one of the most important land subsidence
with funding from the European Union’s Horizon 2020 research and
spots due to groundwater withdrawal worldwide. The model further­
innovation program under grant agreement 820852, and RESERVOIR
more predicts subsidence of up to 7.3 m by 2100 for an assumed con­
Project, part of the PRIMA Programme Horizon 2020 European Union’s
stant hydraulic head since 2012 owing to the slow desaturation and
programme for research and innovation under grant agreement number
consolidation processes.
1924.
In the wake of these results, an upcoming challenge is to analyze
updated piezometric and subsidence data, and a 3D fully coupled model
will be executed to fine-tune the proposed methodology. These models References
should prove very useful for water resource management. In addition to
Aichi, M., 2008. Coupled groundwater flow/deformation modelling for predicting land
providing the basis of past exploited system behavior, they serve to subsidence. In: Takizawa, S. (Ed.), Groundwater Management in Asian Cities:
predict environmental impacts (land settlements) according to defined Technology and Policy for Sustainability. Springer, Tokyo, pp. 105–124.
Béjar-Pizarro, M., Guardiola-Albert, C., García-Cárdenas, R.P., Herrera, G., Barra, A.,
water extraction/injection scenarios, with the ultimate aim of achieving
López Molina, A., Tessitore, S., Staller, A., Ortega-Becerril, J.A., García-García, R.P.,
the sustainability and reliability of groundwater supply in a given area. 2016. Interpolation of GPS and Geological Data using InSAR Deformation Maps:
Method and Application to Land Subsidence in the Alto Guadalentín Aquifer (SE
Spain). Remote Sens. 8, 965. https://doi.org/10.3390/rs8110965.

15
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Biot, M.A., 1941. A general theory of three-dimensional consolidation. J. Appl. Phys. 12 Leake, S.A., Prudic, D.E., 1991. Documentation of a computer program to simulate
(2), 155–164. aquifer-system compaction using the modular finite-difference ground-water flow
de Boer, R., 2000. Theory of Porous Media. Springer-Verlag, Berlin. model. U.S. Geol. Surv. Tech. Water Resour. Invest, Book 6, chap. A2, 68.
Bonì, R., Herrera, G., Meisina, C., Notti, D., Béjar-Pizarro, M., Zucca, F., González, P.J., Lewis, R.L., Schrefler, B.A., 1998. The Finite Element Method in the Static and Dynamic
Palano, M., Tomás, R., Fernández, J., Fernández-Merodo, J.A., Mulas, J., Aragón, R., Deformation and Consolidation of Porous Media. John Wiley and Sons.
Guardiola-Albert, C., Mora, O., 2015. Twenty-year advanced DInSAR analysis of Liakopulos, A.C., 1965. Transient flow through unsaturated porous media. In: Ph.D.
severe land subsidence: The Alto Guadalentín Basin (Spain) case study. Eng. University of California at Berkley, Thesis.
Geology. https://doi.org/10.1016/j.enggeo.2015.08.014. Manzanal, D., 2008. Constitutive model based on Generalized Plasticity incorporating
Borchers, J.W., Carpenter, M., 2014. Land Subsidence from Groundwater Use in state parameter for saturated and unsaturated sand (Spanish). PhD Thesis School of
California. 151 (California Water Foundation. Civil Engineering, Polytechnic University of Madrid.
Cabral-Cano, E., Dixon, T.H., Miralles-Wilhelm, F., Díaz-Molina, O., Sánchez-Zamora, O., Manzanal, D., Pastor, M., Fernández Merodo, J.A., Mira, P., 2010. A state parameter
Carande, R.E., 2008. Space geodetic imaging of rapid ground subsidence in Mexico based Generalized Plasticity model for unsaturated soils. Cmes-Comp. Model. Eng.
City. GSA Bulletin. https://doi.org/10.1130/B26001.1. 55 (3), 293–317. https://doi.org/10.3970/cmes.2010.055.293.
Cerón, J.C., 1995. Estudio hidrogeoquímico del acuífero del Alto Guadalentín (Murcia) Manzanal D., Fernández-Merodo J.A., Pastor M. 2011a. Generalized plasticity state
[Ph.D. thesis]. University of Granada, Granada, p. 265. parameter-based modelling of saturated and unsaturated soils. Part I saturated state.
Cerón, J.C., Pulido-Bosch, A., 1996. Groundwater problems resulting from CO2 pollution Int. J. Numer. Anal. Methods Geomech. (2011) Vol 35:1347–1362. https://doi.org/
and overexploitation in Alto Guadalentín aquifer (Murcia, Spain). Environ. Geol. 28 0.1002/nag.961.
(4), 223–228. https://doi.org/10.1007/s002540050096. Manzanal, D., Pastor, M., Fernández-Merodo, J.A., 2011b. Generalized plasticity state
Chatterjee, R.S., Syafiudin, M.F., Abidin, H.Z., 2013. Land Subsidence Characteristics in parameter-based modelling of saturated and unsaturated soils. Part II unsaturated
Bandung City, Indonesia as Revealed by Spaceborne Geodetic Techniques and soil modelling. Int. J. Numer. Anal. Met. 35, 1899–1917. https://doi.org/10.1002/
Hydrogeological Observations. Photogramm. Eng. Remote Sens. 7, 639–652. nag.983.
Chaussard, E., Amelung, F., Abidin, H., Hong, S.-H., 2013. Sinking cities in Indonesia: Masana, E., Martínez-Díaz, J.J., Hernández-Enrile, J.L., Santanach, P., 2004. The Alhama
ALOS PALSAR detects rapid subsidence due to groundwater and gas extraction. de Murcia fault (SE Spain), a seismogenic fault in a diffuse plate boundary:
Remote Sensing of Environment 128, 150–161. https://doi.org/10.1016/j. Seismotectonic implications for the Ibero-Magrebian region. J. Geophys. Res. 109,
rse.2012.10.015. B01301. https://doi.org/10.1029/2002JB002359.
Chaussard, E., Wdowinski, S., Cabral-Cano, E., Amelung, F., 2014. Land subsidence in Mateos, R.M., Ezquerro, P., Luque, J.A., Béjar-Pizarro, M., Notti, D., Azañón, J.M.,
Central Mexico detected by ALOS InSAR time-series. Remote Sens. Environ. 140, Montserrat, O., Herrera, G., Fernández-Chacón, F., Peinado, T., Galve, J.P., Pérez-
94–106. Peña, V., Fernández-Merodo, J.A., Jiménez, J., 2017. Multiband PSInSAR and long-
CHS, 2014. Plan Hidrológico de la Cuenca del Segura 2015/2021. In: Análisis period monitoring of land subsidence in a strategic detrital aquifer (Vega de
piezométrico histórico y de los últimos 25 años (1990-2014) de las masas de Agua Granada, SE Spain): an approach to support management decisions. Journal of
subterráneas de la demarcación Hidrográfica del Segura. 070.057 Alto Guadalentín. Hydrology 533, 71–87. https://doi.org/10.1016/j.jhydrol.2017.07.056.
Tech. rep. http://www.chsegura.es/chs/cuenca/sequias/pes/eeapes.html#doc Mira, P., 2002. Análisis por Elementos Finitos de Problemas de Rotura de Geomateriales.
_completa. PhD. In: ETS de Ingenieros de Caminos. Universidad Politécnica de Madrid, Canales
Coussy, O., 1995. Mechanics of Porous Media. John Wiley and Sons, Chichester. y Puertos.
Cuomo, S., Moscariello, M., Manzanal, D., Pastor, M., Foresta, V., 2018. Modelling the Motagh, M., Djamour, Y., Walter, T.R., Wetzel, H.U., Zschau, J., Arabi, S., 2007. Land
mechanical behaviour of a natural unsaturated pyroclastic soil within Generalized subsidence in Mashhad Valley, Northeast Iran: results from InSAR, levelling and
Plasticity framework. Comput. Geotech. 99, 191–202. https://doi.org/10.1016/j. GPS. Geophysical Journal International 168 (2), 518–526.
compgeo.2018.03.006. Narasimhan, T., Witherspoon, P.A., 1978. Numerical model for saturated-unsaturated
Erkens, G., Bucx, T., Dam, R., de Lange, G., Lambert, J., 2015. Sinking coastal cities. In: flow in deformable porous media: 3. Applications. Water Resources Research 14,
Proceedings of the International Association of Hydrological Sciences, 372, 1017–1034. https://doi.org/10.1029/WR014i006p01017.
pp. 189–198. https://doi.org/10.5194/piahs-372-189-2015. Ortega-Guerrero, A., Rudolph, D.L., Cherry, J.A., 1999. Analysis of long-term land
Ezquerro, P., Guardiola-Albert, C., Herrera, G., Fernández-Merodo, J.A., Béjar- subsidence near Mexico City: Field investigations and predictive modelling, Water
Pizarro, M., Bonì, M., 2017. Groundwater and subsidence modelling combining Resour. Res. 35 (11), 3327–3341. https://doi.org/10.1029/1999WR900148.
geological and multi-satellite SAR data over the Alto Guadalentín aquifer (SE Spain). Ortiz-Zamora, D., Ortega-Guerrero, A., 2010. Evolution of long-Eterm land subsidence
Geofluids 2017. https://doi.org/10.1155/2017/1359325. near Mexico City: Review, field investigations, and predictive simulations. Water
Ezquerro, P., Tomás, R., Béjar-Pizarro, M., Fernández-Merodo, J.A., Guardiola-Albert, C., Resour. Res. 46, W01513. https://doi.org/10.1029/2008WR007398.
Staller, A., Sánchez-Sobrino, J.A., Herrera, G., 2019. Improving multi-technique Osmanoğlu, B., Dixon, T.H., Wdowinski, S., Cabral-Cano, E., Jiang, Y., 2011. Mexico City
monitoring using Sentinel-1 and Cosmo-SkyMed data and upgrading groundwater subsidence observed with persistent scatterer InSAR. Int. J. Appl. Earth Obs. Geoinf.
model capabilities. Science of the Total Environment 703 (2020), 134757. https:// 13, 1–12. https://doi.org/10.1016/j.jag.2010.05.009.
doi.org/10.1016/j.scitotenv.2019.134757. Pastor, J.L., Mulas, J., Tomás, R., Herrera, G., Fernández-Merodo, J.A., Béjar, M.,
Fernandez, J., Prieto, J.F., Escayo, J., et al., 2018. Modeling the two- and three- Jordá, L., López-Davalillos, J.C., Aragón, R., Mateos, R.M., 2020. Geometrical and
dimensional displacement field in Lorca, Spain, subsidence and the global geotechnical characterization of the earth fissures appeared in the Guadalentín
implications. Sci. Rep. 8, 14782. https://doi.org/10.1038/s41598-018-33128-0. Valley after the September 2012 flooding. In: TISOLS 2020 conference.
Fernández-Merodo, J.A., 2001. Une approche à la modélisation des glissements et des Pastor, M., Zienkiewicz, O.C., Chan, A.H.C., et al., 1990. Generalized plasticity and the
effondrements de terrains: Initiation et propagation. PhD. In: École Centrale Paris. modelling of soil behaviour. Int. J. Numer. Anal. Meth. Geomech. 14, 151–190.
Gambolati, G., and Teatini, P. 2015. Geomechanics of subsurface water withdrawal and https://doi.org/10.1002/nag.1610140302.
injection, Water Resour. Res., 51, 3922–3955, http://doi:https://doi.org/10.1002/ Poeter, E.P., Hill, M.C., Lu, D., Tiedeman, C.R., Mehl, S., 2014. UCODE 2014, with new
2014WR016841. capabilities to define parameters unique to predictions, calculate weights using
Gambolati, G., Teatini, P., Bau, D., Ferronato, M., 2000. The importance of poro-elastic simulated values, estimate parameters with SVD, evaluate uncertainty with MCMC.
coupling in dynamically active aquifers of the Po river Basin, Italy, Water Resour. In: And More: Integrated Groundwater Modeling Center ReportNumber:
Res. 36 (9), 2443–2459. https://doi.org/10.1029/2000WR900127. GWMI2014–02,GWMI.
González, P.J., Fernández, J., 2011. Drought-driven transient aquifer compaction imaged Poland, J.F., 1984. Guidebook to studies of land subsidence due to ground-water
using multitemporal satellite radar interferometry. Geology 39, 551–554. https://do withdrawal. United Nations Educational, Scientific and Cultural Organization,
i.org/10.1130/G31900.1. Chelsea.
Harbaugh, A.W., 2005. MODFLOW-2005, the U.S. Geological Survey modular ground- Rigo, A., Béjar-Pizarro, M., Martínez-Díaz, J., 2013. Monitoring of Guadalentín valley
water model – the Ground-Water Flow Process: U.S. Geological Survey Techniques (southern Spain) through a fast SAR Interferometry method. J. Appl. Geophys. 91,
and Methods 6- A16. 39–48. https://doi.org/10.1016/j.jappgeo.2013.02.001.
Hay-Man, G.A., et al., 2012. Mapping land subsidence in Jakarta, Indonesia using Schaap, Marcel G., Leij, Feike J., Martinus, T.H., Genuchten, Van, 2001. Rosetta: a
persistent scatterer interferometry (PSI) technique with ALOS PALSAR, 18. computer program for estimating soil hydraulic parameters with hierarchical
International Journal of Applied Earth Observation and Geoinformation, pedotransfer functions. Journal of Hydrology Vol. 251 (3–4), 163–176. https://doi.
pp. 232–242. https://doi.org/10.1016/j.jag.2012.01.018. org/10.1016/S0022-1694(01)00466-8.
Holzer, T.L., Galloway, D.L., 2005. Impacts of land subsidence caused by withdrawal of Shi, X., Xue, Y., Ye, S., Wu, J., Zhang, Y., Yu, J., 2007. Characterization of land
underground fluids in the United States. Rev. Eng. Geol. XVI 87–99. subsidence induced by groundwater withdrawals in Su-XiChang area, China,
Kasmarek, M.C., Strom, E.W., 2002. Hydrogeology and simulation of groundwater flow Environ. Geol. 52 (1), 27–40.
and land-surface subsidence in the Chicot and Evangeline aquifers, Houston area, Tessitore, S., Fernández-Merodo, J.A., Herrera, G., Mulas, J., Calcaterra, D., 2016.
Texas. U.S. Geol. Surv. Water Resour. Invest. Rep. 024022, 61. Comparison of water-level, extensometric, DInSAR and simulation data for
Leake, S.A., 1990. Interbed storage changes and compaction in models of regional quantification of subsidence in Murcia City (SE Spain). Hydrogeol. J. 24 (3),
groundwater flow. Water Resour. Res. 26 (9), 1939–1950. https://doi.org/10.1029/ 727–747. https://doi.org/10.1007/s10040-015-1349-8.
WR026i009p01939. Thu, T.M., Fredlund, D.G., 2000. Modelling subsidence in the Hanoi City area. Vietnam,
Leake, S.A., Galloway, D.L., 2007. MODFLOW ground-water mode: User guide to the Can. Geotech. J. 37, 621–637. https://doi.org/10.1139/t99-126.
subsidence and Aquifer-System Compaction Package (SUB-WT) for water-table van Genuchten, M. Th., 1980. A closed-form equation for pre-dicting the hydraulic
aquifers. U.S. Geol. Surv. Tech. Methods, Book 6, chap. A23, 42. conductivity of unsaturated soils. SoilSci. Soc. Am. J. 44, 892–898.
Leake, S.A., Galloway, D.L., 2010. Use of the SUB-WT package for MODFLOW to simulate Wilson, A.M., Gorelick, S., 1996. The effects of pulsed pumping on land subsidence in the
aquifer-system compaction in Antelope Valley, California, USA. In: Carreon- Santa Clara Valley, California. J. Hydrol. 174, 375–396. https://doi.org/10.1016/
Freyre, D., et al. (Eds.), Land Subsidence, Associated Hazards and the Role of Natural 0022-1694(95)02722-X.
Resources Development, IAHR Publ, 339. Int. Assoc. for Hydro-Environ, pp. 61–67.

16
J.A. Fernández-Merodo et al. Engineering Geology 283 (2021) 105998

Ye, S., Xue, Y., Wu, J., Li, Q., 2012. Modeling visco-elastic-plastic deformation of soil Zienkiewicz, O.C., Xie, Y.M., Schrefler, B., Ledesma, A., Bićanić, N., 1990. Static and
with modified merchant model. Environ. Earth Sci. 66 (5), 1497–1504. https://doi. dynamic behavior of soils: a rational approach to quantitative solutions. II. Semi-
org/10.1007/s12665-011-1389-x. saturated problems. Proceedings of the Royal Society of London. A. Mathematical
Zhang, Y., et al., 2014. Characterization of land subsidence induced by groundwater and Physical. Sciences 429, 311–321.
withdrawals in the plain of Beijing city. China. Hydrogeology Journal 22, 397–409. Zienkiewicz, O.C., Chan, A.H.C., Pastor, M., Shrefler, B.A., Shiomi, T., 2000.
https://doi.org/10.1007/s10040-013-1069-x. Computational Geomechanics. John Wiley and Sons.

17

You might also like