Vous êtes sur la page 1sur 15

Groundwater flow modelling of the regional aquifer of the Pampa

del Tamarugal, northern Chile


Rodrigo Rojas & Alain Dassargues

Abstract The Pampa del Tamarugal Aquifer (PTA) is an Nord du Chili. Dans cette étude, un modèle d’écoulement
important source of groundwater in northern Chile. In this souterrain de cet aquifère est réalisé et calibré pour la
study, a groundwater flow model of this aquifer is période 1983–2004. Celui-ci reproduit le champs
developed and calibrated for the period 1983–2004. The d’écoulement observé et les composantes du bilan d’eau
model reproduces the observed flow-field and the water raisonnablement bien. Cinq scénarios sont définis pour
balance components reasonably well. Five scenarios are évaluer la réponse du système à des conditions de
defined to evaluate the response to different pumping pompage différentes. D’après ces scénarios, les niveaux
situations. These scenarios show that groundwater heads piézomètriques vont continuer à baisser sous l’effet des
will continue to decrease with the present pumping taux de pompage actuels. Pour tenir compte de la
discharge rates. To account for variations in the model variabilité des résultats du modèle due aux incertitudes
results due to uncertainties in average recharge rates, sur les taux de recharge moyens, des épisodes de recharge,
randomly generated recharge realizations with different générés aléatoirement et présentant différents niveaux
levels of uncertainty are simulated. Evaporation flow rates d’incertitude sont simulés. Les taux d’évaporation et l’eau
and groundwater flowing out of the modelled area seem souterraine s’écoulant hors de la zone modélisée semblent
unaffected by the recharge uncertainty, whereas the non affectés par l’incertitude sur la recharge, les termes
storage terms can vary considerably. For the most d’emmagasinement en revanche peuvent varier considéra-
intensive pumping scenario under the generated random blement. Dans le cas du scénario avec le pompage le plus
recharge rates, it is unlikely that the cumulative discharged important et des taux de recharge aléatoires, il est peu
volume from the aquifer, at the end of the simulation probable que le volume cumulé pompé hors de l’aquifère
period, will be larger than 12% of the estimated à la fin de la simulation atteigne une valeur supérieure à
groundwater reserve. Fluctuations in simulated groundwa- 12% de la réserve d’eau souterraine estimée. Les fluctua-
ter heads due to uncertainties in the average recharge tions des niveaux piézomètriques simulés liées aux
values are more noticeable in certain areas. These incertitudes sur les recharges moyennes sont plus percep-
fluctuations could explain unusual behaviour in the tibles dans certaines zones. Ces fluctuations pourraient
observed groundwater heads in these areas. expliquer un comportement inhabituel des niveaux piézo-
mètriques observés dans ces zones.
Résumé L’aquifère de la Pampa del Tamarugal (PTA)
représente une importante source d’eau souterraine dans le Resumen El Acuífero de la Pampa del Tamarugal (PTA)
es una fuente importante de agua subterránea en el norte de
Chile. En este estudio se desarrolla y calibra un modelo de
flujo de agua subterránea para el periodo 1983–2004. El
Received: 25 November 2005 / Accepted: 26 June 2006
Published online: 10 August 2006
modelo reproduce razonablemente bien el campo de flujo
observado y los componentes del balance hídrico. Se
© Springer-Verlag 2006 definen cinco escenarios para evaluar la respuesta a
diferentes situaciones de bombeo. Estos escenarios mues-
R. Rojas ()) : A. Dassargues
tran que con las tasas de descarga de bombeo actuales las
Hydrogeology and Engineering Geology Group, presiones de agua subterránea continuarán en descenso.
Department of Geography-Geology, Para explicar las variaciones en los resultados del modelo
Katholieke Universiteit Leuven, debidas a incertidumbres en tasas de recarga promedio se
Celestijnenlaan 200 E, 3001, Leuven (Heverlee), Belgium han simulado realizaciones de recarga generadas aleator-
e-mail: Rodrigo.RojasMujica@geo.kuleuven.be
Tel.: +32-16-326449 iamente con diferentes niveles de incertidumbre. Las tasa
Fax: +32-16-326401 de flujo por evaporación y el agua subterránea que fluye
fuera del área modelizada parecen no haber sido afectadas
A. Dassargues
Hydrogeology, Department of Georesources, por la incertidumbre en recarga mientras que los términos
Geotechnologies and Building Materials, de almacenamiento pueden variar considerablemente. Para
Université de Liège, Liège, Belgium el escenario de bombeo más intenso, bajo las tasas de

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


538
recarga generadas aleatoriamente, no es probable que el groundwater sources. In the capital of the most northern
volumen descargado acumulado del acuífero al final del region of Chile, drinking water is supplied by means of
periodo de simulación sea mayor del 12% de la reserva de two well fields known as Canchones and El Carmelo
agua subterránea estimada. Las fluctuaciones en las which are located in the Pampa del Tamarugal Aquifer
presiones simuladas de agua subterránea debido a las (PTA; Fig. 1). Due to the pumping discharge from these
incertidumbres en los valores de recarga promedio son más well fields, a steady decrease in the groundwater heads
notorios en ciertas áreas. Las fluctuaciones podrían explicar recorded in the monitoring network controlled by the
el comportamiento inusual en las presiones de agua Dirección General de Aguas (DGA) of Chile, is observed.
subterránea observadas en estas áreas. The study of the PTA has been a concern since the
early 1960s and many local institutions and international
Keywords Arid regions . Numerical modelling . agencies, e.g. DGA, have attempted to describe the PTA.
Groundwater recharge . Stochastic . Atacama Desert Based on a water balance at regional scale, Grilli et al.
(1986), on behalf of the DGA, estimated an average
recharge flow rate for the PTA of 1,002 l/s distributed over
Introduction seven eastern sub-basins (Fig. 1b). Subsequently, the
DGA (1987) corrected this value to 990 l/s in the
Due to extreme arid conditions, groundwater in northern framework of a general study of the water balance of
Chile is a vital resource. In the Pampa del Tamarugal Chile.
basin (Fig. 1), annual precipitation is nil in the lower One of the first efforts to numerically model the
areas, but reaches values of about 200 mm/yr at altitudes groundwater flow in the PTA was reported in 1988
above 3,500 m above sea level (asl). This spatial (DGA-UChile 1988). A groundwater flow model for
precipitation distribution controls the hydrology of the steady-state conditions for the year 1960 and a transient-
region. Groundwater and surface water originating in the state model were developed. The overall model was able
Andes Mountains are the main water source for human to reproduce the general flow-field in the area. Neverthe-
activities (Aravena 1995). less, recharge coming from the most northern sub-basin
In this region, many coastal cities and interior towns as (Aroma) was not included in this study since it was not
well as most of the mining industry entirely depend on possible to reproduce the observed groundwater head in

Fig. 1 Location of the study area. a General location in Chile, b Pampa del Tamarugal Basin and sub-basins delimitation (shaded area is
the Pampa del Tamarugal Aquifer), c Observation wells and vegetation pattern in the study area

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


539
the northern part of the modelled area during the estimation of the recharge is used for modelling purposes.
calibration process. To account for the response of the aquifer to different
The Japanese International Cooperation Agency (JICA) pumping discharge situations, a series of five scenarios
jointly with the DGA and Pacific Consultants International were considered. In terms of predictions, simulated
(PCI) developed a large-scale study in the year 1995 scenarios start from the current discharge situation
(JICA-DGA-PCI 1995). This study focused on the (2004–2005) which allows for the assessment of possible
development of the water resources at a regional scale, future strategies for management purposes. In addition,
including the two most northern regions of Chile. Within since the recharge magnitude has largely been the subject
the studied areas, the PTA was intensively surveyed. For of conjecture rather than proof and calculation (Houston
that study, a groundwater flow model for steady-state 2002), another important objective, which has not been
conditions was developed for the year 1993; the model included in previous studies, is to account for variations in
was able to roughly reproduce the general observed the average recharge values. To accomplish this, recharge
groundwater flow-field. However, two major limitations is treated as a random variable and the model is run for a
of this model can be mentioned. First, the calibration large number of stochastically generated random recharge
process was based on unverified recharge processes where values. In this way, the influence of the uncertainty in the
a significant amount of groundwater recharge was as- groundwater recharge on the water balance components
sumed to come from deep fissures in basement rocks. This and groundwater heads in the PTA is determined.
recharge mechanism was based on results that demon-
strated the presence of fresh and recent groundwater at
shallow levels in the centre part of the Pampa del Materials and methods
Tamarugal (Margaritz et al. 1990). Nevertheless, recent
studies suggest that recharge at relatively shallow levels is Study area
taking place, as a result of infiltrating runoff in the apex of Located in the Pampa del Tamarugal basin, the PTA
alluvial fans originating from the eastern sub-basins (Grilli covers an area of ca. 5,000 km2 and it is limited in the
et al. 1999; Houston 2002). This certainly could also west by the Coastal Range and in the east by the Chilean
explain the presence of fresh and recent groundwater in Pre-Cordillera. It is almost 160 km long and its width
the centre part of the PTA. Second, the assumed steady- varies between 20 and 60 km with an average elevation of
state conditions for the year 1993 seem not to be valid, as 1,000 m asl (Fig. 1).
suggested by present-day data (Rojas 2005). Direct precipitation on the lower areas of the PTA is nil
Recently, Houston (2002) described the recharge and thus no recharge through this mechanism can be
mechanism in an alluvial fan located in the PTA expected. On the other hand, the eastern sub-basins
(Chacarilla sub-basin). A methodology for the estimation (Fig. 1b) receive recharge from precipitation coming from
of recharge magnitudes due to flash-flood events was high altitudes in the east. These sub-basins lie in a well-
presented and discussed. The obtained recharge value was developed rainshadow, and, as a result, there is a rapid
approximately 20% larger than the values reported in decrease in rainfall as air masses move west and descend
JICA-DGA-PCI (1995) for the same sub-basin. (Houston 2002). It is also noted that in a hydrologic
In this context, the objective of this study is to develop average year, surface watercourses disappear before
an up-to-date regional-scale groundwater flow model for reaching most of the alluvial fans located in the PTA,
PTA that integrates the current knowledge and the hydro- suggesting that part of this water is recharging the aquifer
geological information available for the region. To achieve system through infiltration and lateral groundwater flows
this, a much longer period of observation data is used in (Aravena 1995).
the calibration process to account for transient conditions The study area is part of the Atacama Desert and
compared to previous studies. Also, the most reliable therefore arid conditions are extreme. According to the

Fig. 2 Stratigraphic classifica-


tion and lithology description of
Pampa del Tamarugal Aquifer
(after JICA-DGA-PCI 1995)

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


540
DGA (1987) pan evaporation rates vary between PCI 1995; Digert et al. 2003). As described in Houston
2,000 mm/yr and 2,500 mm/yr. Aridity limits the presence (2002), the basin started to form in the early Oligocene and
of vegetation to very localized places in the area. These it is a complex asymmetric graben bounded in the west and
places correspond to natural or reforested areas located in in the east by N–S regional fault zones. The lowermost
the north (Dolores), centre (Salar de Pintados) and south sediments are coarse conglomerates and gravels of the
(Salar de Bellavista) of the study area (Figs. 1 and 5). Sichal and Altos de Pica Formations, eroded from the
During the 1960s, these areas were formally declared part adjacent uplifted Coastal Range and Pre-Cordillera.
of the Pampa del Tamarugal Natural Forest Reserve and Throughout the Miocene, coarse clastic sediments contin-
from 1964 an intensive plan of reforestation started (FAO ued to be deposited over wide areas. Some events of
1989). volcanic activity took place producing andesitic tuffs and
ignimbrites from the eruptive centres located on the east. A
series of large alluvial fans began to develop towards the
Geology end of the Miocene. Since the Pliocene only minor alluvial
The geology of the area has been extensively studied and evaporitic sediments have been deposited in the basin.
(Dingman and Galli 1965; DGA-UChile 1988; JICA-DGA- The lithology of the geological units is described in Fig. 2.

Fig. 3 a Geological longitudinal profile (X-X′) of Pampa del Tamarugal Aquifer. White dashed lines indicate main faults (adapted from
JICA-DGA-PCI 1995) b W-E geological cross-section (Y-Y′) in the area of Pica (adapted from DGA-CCHEN 1998)

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


541
Figure 3 shows a longitudinal geological profile (XX’) the study area are depicted. Well 30 is located in the north
of the study area and a geological cross-section (YY’) in whereas well 72 is located in the western part of the centre
the area of Pica. In Fig. 3a, the basement rocks of zone of the study area. This last area is directly influenced
Longacho Formation are differentiated in the whole area. by the pumping discharge of the Canchones and El
Uplifting of this formation is observed in the north and Carmelo well fields. This figure shows the difference in
south limits. Also in the west, in the Coastal Range, and in the rate of decrease of groundwater heads for both wells.
the east, outcroppings of this formation are observed. In Fig. 5, the groundwater elevation map for the year
Overlying the Longacho Formation, is the Altos de Pica 1960 is depicted from nearly 60 measurement points. The
Formation, which is differentiated into lower and upper main groundwater flow direction is from north to south
layers. In the uppermost strata, recent sediments mainly with an east-west component in the Pica area. In the north
composed of saline alluvial deposits, gravel, sand and clay area, a groundwater divide is observed with part of the
have been deposited (JICA-DGA-PCI 1995). In Fig. 3b, groundwater flowing towards the forested area in Dolores.
the local situation of the geological configuration in the Towards the south, in the Huara sector, the hydraulic
area of Pica is depicted. From this figure it can be seen gradient is considerably steeper, which is explained by the
that the N–S fault system present in this area can be lower hydraulic conductivity of the deposits (DGA-
considered as a boundary between PTA and the Pica UChile 1988). In the centre, groundwater flows from east
Aquifer. to west and flow is directed towards the western forested
Tamarugo areas and the Salar de Pintados. Both corre-
spond to the main natural discharge areas of the aquifer
Hydrogeology for the year 1960. In the southern area (Oficina Victoria-
In JICA-DGA-PCI (1995), more than 400 boring logs Cerro Gordo) groundwater flow is mainly to the south-
were interpreted and a transient electromagnetic (TEM) west with a well-defined discharge zone in the Salar de
survey was performed. The information from the geo- Bellavista.
electrical profiles was complemented with the drilling of
eleven wells ranging in depth between 150 m and 300 m.
This analysis showed that the main aquifer system was Water balance
composed of units Q4 and Q3 (Fig. 3a). According to Based on data from DGA-UChile (1988), JICA-DGA-PCI
JICA-DGA-PCI (1995), unit Q3 is composed of sand and (1995) and the records of legally constituted wells
gravel and is underlain by a thick clayey layer (Q2). Unit provided by the DGA, an estimation of the pumping
Q4 consists of sand and gravel with mud, and/or discharge is presented in Fig. 6. This pumping discharge is
intercalated with mud layers and is overlain by unit Q3 mainly localized in the centre part of the modelled
from Huara to the Salar de Bellavista area. Also, in the domain. Although this artificial discharge has been
context of this study, a series of pumping tests and a reported in the previously cited studies, still there is some
complete re-analysis of the existing information were degree of uncertainty in the quantification of the real
carried out. A summary of the aquifer parameters obtained groundwater extraction. Nevertheless, the most important
for the PTA is presented in Table 1. groundwater extraction (drinking water) has been reason-
At this time, DGA controls 38 observation wells in the ably estimated. In Fig. 6, other uses are defined such as
study area (Fig. 1c). In Fig. 4, two representative wells for mining, industrial and irrigation discharges. The ratio

Fig. 4 Observed groundwater


head in wells 30 and 72. Depth
in meters below ground level
(m bgl) and head (h) in meters
above sea level (m asl)

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


542
Table 1 Summary of aquifer parameters of the Pampa del Tamarugal Aquifer (PTA) (adapted from JICA-DGA-PCI 1995)
Zone Transmissivity (m2/d) Hydraulic conductivity (m/d) Storativity (–)
Dolores 1–1031 0.02–101.1 3.0e–04–0.05
Huara 8–1506 0.11–20.3 5.7e–07–0.08
Pozo Almonte, Pintados 9–1094 0.3–91.6 9.4e–06–0.04
Oficina Victoria, Cerro Gordo 81–420 0.8–12.0 3.0e–07–0.33

(other uses/drinking water use) for the period 1961–1993


is ca. 0.2. From year 1994 onwards, this ratio has
increased from 0.6 to 1.2 reflecting an intensive pumping
discharge for other uses. This could be partially explained
by the very favourable results of previous studies, which
resulted in an augmentation of water uses, particularly for
mining activities, and by the observed economic growth
by the industrial and mining sector, especially in northern
regions of Chile.
As previously mentioned, there is a general agreement
that recharge is mainly by groundwater flowing from the
eastern sub-basins in the study area (DGA-UChile 1988;
JICA-DGA-PCI 1995; Aravena 1995; Grilli et al. 1999;
Houston 2002). According to DGA-UChile (1988), the
groundwater recharge for the period 1960–1981 is
estimated to be 1,002 l/s. On the other hand, according
to JICA-DGA-PCI (1995), the groundwater recharge for
the period 1960–1993 is estimated as 976 l/s. Most recent
estimations confirm groundwater recharges between 925
and 976 l/s for the period 1960–1993 (DGA 1996;
DICTUC 2005). For this study, it is assumed that the
yearly average recharge value will not deviate significant-
ly from the previous estimations, if the period 1994–2004
is included in the calculations. Hence, the recharges
estimated in JICA-DGA-PCI (1995) are used. The lateral
recharges from each sub-basin are presented in Table 2.
As already mentioned, vegetative cover in the study
area is limited to three zones: Dolores, Salar de Pintados
and Salar de Bellavista (Figs. 1 and 5). These zones are
composed of trees highly adapted to extreme arid and
saline conditions and, therefore, they can sustain them-
selves by absorbing air moisture and by extracting
groundwater with their massive and well-developed root
systems (FAO 1989). As a consequence, it can be
considered that transpiration (but not evaporation) occurs
in these forested areas. The estimated transpiration rates
are presented in Table 3 and reflect the policy of

Fig. 5 Groundwater elevation (m asl) map for the PTA in the year
1960 (adapted from JICA-DGA-PCI 1995). The shaded areas
represent forested areas Fig. 6 Estimated artificial groundwater discharge in the study area

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


543
Table 2 Estimation of recharge (l/s) for PTA (After JICA-DGA- Groundwater flowing out of the study area was
PCI 1995) maintained as constant, given that the main pumping
Location Recharge (l/s) discharge zone is located in the Canchones-Pintados area
Aroma sub-basin 310 and that Salar Viejo does not evaporate. Therefore,
Tarapaca sub-basin 318 hydraulic gradient at the south boundary is not affected
Sagasca sub-basin 89 either by the regional cone of depression produced in the
Quipisca sub-basin 72 Canchones-Pintados area or by the evaporation.
Quisma sub-basin 21
Chacarilla sub-basin 159
Ramada sub-basin 7
Total Recharge 976 Groundwater flow model
Conceptual model
The model domain is limited in the west by the contact zone
reforestation that started during the 1960s. This can be
defined between the basement rocks of Longacho Formation
verified when analyzing the surface area of the forested
(J) and the sedimentary materials (Fig. 3). The east
regions: for year 1960 ca. 6,840 ha, for the period 1980–
boundary is defined as a supposed line roughly following
1985 ca. 21,400 ha, and for the years 1995–2000 ca.
the main faults lineaments and the outcroppings of the
28,000 ha. (Rojas 2005).
Longacho Formation. The northern limit corresponds to
Another important component of the water balance is
Dolores area, whereas the southern limit is located
the evaporation from salares. Salares can be defined as
southwards, at the Cerro Gordo outcropping. This defines
saline surface deposits where evaporation of groundwater
an area of 4,303 km2, which is 160 km long and between
can take place (Risacher et al. 1998). In the study area,
18 and 40 km wide. A one-layer model is assumed by
there are three salares: Salar de Pintados, Salar de
combining units Q3 and Q4 in one hydrostratigraphic unit.
Bellavista and Salar Viejo (Fig. 5), the last one being a
This is based on the fact that most of the wells are open to
non-active evaporative salar, i.e. groundwater evaporation
both units and the available information on hydrogeological
is not expected from it. Based on data from Grilli et al.
parameters is obtained jointly for both units. The depth of
(1986), DGA-UChile (1988) calculates an evaporation
the modelled area varies between 50 and ca. 300 m (Fig. 3).
flow rate, jointly for Salar de Pintados and Salar de
At the northern limit, a negligible groundwater outflow is
Bellavista, of 542 l/s and 286 l/s for the years 1960 and
observed as mentioned in DGA-UChile (1988) and,
1987, respectively. On the other hand, JICA-DGA-PCI
therefore, a no-flow condition is defined. The groundwater
(1995) calculates an evaporation flow rate of 145 l/s for
elevation map for years 1960 and 1995 (JICA-DGA-PCI
the year 1993 also based on data from Grilli et al. (1986).
1995; Rojas 2005) shows that the north groundwater divide
The groundwater flowing out at the south boundary is
depicted in Fig. 5 has undergone little change, although the
estimated between 164 and 356 l/s (Rojas 2005).
water usage has significantly increased in that period.
Results for the 1960, 1987 and 1993 water balance are
In the south, a specified head boundary, constant in
shown in Table 4. For calculating the 1960 water balance, the
time, is defined based on observed groundwater head data
evaporation flow rate from the salares was used to close the
(Rojas 2005). Since evaporation is not expected from the
balance assuming steady-state conditions. Recharge from
Salar Viejo and the groundwater extraction is localized in
the eastern sub-basins was assumed to be constant and equal
the centre area of the modelled domain, it is reasonable to
to the values estimated by JICA-DGA-PCI (1995). This
assume that the south boundary condition will remain
value (976 l/s) was adopted since it was considered the most
constant over years.
reliable estimation of recharge given the longer and more
A no-flow condition is defined in the west, whereas in
recent period of analysis. Transpiration from forested areas
the east, specified flux conditions are defined
increases due to reforestation strategies. Evaporation from
corresponding to the contribution of each of the seven
the Salar de Pintados and Salar de Bellavista decreases over
sub-basins. This is done by incorporating point recharge
time, as expected, due to the lowering of the groundwater
wells in the apex of the alluvial fan for each of the
heads. The higher values of evaporation for the year 1960
corresponding sub-basins. No significant pumping dis-
are justified in DGA-UChile (1988) where the occurrence of
charge is included for the year 1960. Evaporation and
shallow ponds and, therefore, higher groundwater heads in
transpiration processes are also incorporated; transpiration
these salares, is reported.
takes place in the Tamarugo areas, whereas evaporation
occurs in the salares areas. To implement and numerically
solve the groundwater flow equation subject to the
Table 3 Annual transpiration (l/s) for the Tamarugo zones in the respective boundary conditions, MODFLOW (McDonald
study area (adapted from DGA-UChile 1988 and JICA-DGA-PCI
1995) and Harbaugh 1988) was used.
Year Transpiration (l/s)
1960 210 Model calibration
1980–1985 690 A two-step calibration process was performed for the steady-
1985–1993 904
state model. First, a trial-and-error calibration was done and,

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


544
Table 4 Water balance (l/s) for years 1960, 1987, 1993
Flow components 1960 1987 1993
In Out In Out In Out
Recharge from sub-basins 976 976 976
Transpiration Tamarugo areas 210 690 904
Evaporation from salares 410–602 286 145
Groundwater outflow 164–356 164–356 164–356
Pumping discharge 0 716 730
Total 976 976 976 1,856–2,048 976 1,943–2,135

subsequently, the best approach reached by trial-and-error Simulation scenarios


was improved using the “estimation mode” of the PEST The model was run for five future scenarios of ground-
(parameter estimation) algorithm (Doherty et al. 1994). water consumption in the PTA. In the fifth scenario,
The steady-state model was calibrated for the year recharge was treated as a random variable and the model
1960. An initial spatial distribution of the hydraulic was run for a large number of stochastically generated
conductivity was obtained from the available data (35 random recharge values. Initial conditions were obtained
measurements). Data were spatially interpolated using the from the last year (2004) in the transient calibration. The
kriging algorithm and the corresponding results were simulated period corresponds to 2005–2050. The
divided into arbitrary ranges obtaining 12 zones. During corresponding scenarios are as follows:
the trial-and-error calibration, different spatial configura-
tions and values for the hydraulic conductivity were 1. Scenario 2005: pumping discharge situation of the year
adopted. The preliminary “zonation” was manually mod- 2004–2005 (1,585 l/s=714 l/s for drinking water
ified (subdividing the original zones) to improve the requirements +871 l/s for other uses)
calibration, obtaining 24 final zones of hydraulic conduc- 2. Scenario 2005+20%: pumping discharge situation of
tivity. This final “zonation” was in conformity with the the year 2004–2005 increased by 20% (1,900 l/s)
observed ranges of the hydraulic conductivities and it 3. Scenario 2005–20%: pumping discharge situation of
followed the spatial trend of the initially interpolated the year 2004–2005 decreased by 20% (1,268 l/s)
values. Following the trial-and-error calibration, the 4. Scenario drinking water projection (DWP): This
estimation mode of PEST was used to improve the scenario corresponds to the projection of the drinking
preliminary results of the steady-state calibration. water demands for the city of Iquique. In this scenario,
The transient-state model was calibrated on the period the pumping discharges for drinking water usage were
1983–2004 following only the trial-and-error approach for linearly projected from 714 up to 1,235 l/s, whereas the
adjusting the storage coefficients. An initial spatial distribu- other uses demands were maintained constant to the
tion with four zones was adopted for the storage coefficients 2005 situation (871 l/s). Thus, the final pumping
based on the spatial interpolation of available data using the discharge for the 2050 year totaled 2,106 l/s. This
kriging algorithm. Following a similar approach for the water demand for drinking water purposes was
calibration of the hydraulic conductivity in the steady-state estimated based on the projection of population and
model (i.e. manually modifying the preliminary “zonation”), the corresponding water requirements (JICA-DGA-PCI
storage coefficients were calibrated comparing calculated 1995; Arrau 1998; Brown 2003; Rojas 2005). It was
groundwater heads with historical data derived from the also assumed that drinking water for Iquique would be
monitoring network of the DGA. For the transpiration from largely extracted from the PTA with the pumping
the Tamarugo areas, a linear variation was assumed infrastructure of the year 2005.
according to the values defined in Table 3. Parameters for 5. Scenario DWP + randomly generated recharge: Based
the calculation of the evaporation flow rates from the on the scenario DWP, the effects of the different
salares were maintained constant (extinction depth = 1.5 m average recharge flow rates on the groundwater heads
and evaporation rate at surface level=1,500–2,000 mm/yr). and the flow components of the water balance were
Recharge flow rates were defined according to the values analyzed. A series of 100 purely random realizations of
defined in Table 2. Finally, to improve the calibration of the average recharge values were generated for three levels
transient-state model, minor adjustments to the spatial of uncertainty expressed as a percentage of the average
distribution of the hydraulic conductivity zones obtained recharge (R) values presented in Table 2. The level of
from the steady-state model were made. These modifica- uncertainty was set arbitrarily at a minimum
tions included just a few grid cells of the numerical model (σ1=0.05R), an average (σ2=0.15R) and a maximum
(ca. 30 grid cells out of 2,770 active grid cells) and they did (σ3=0.35R) value. Random recharge flow rates were
not alter the results of the steady-state model. Given the generated for each sub-basin using a random number
scale of the study area, calibration targets were arbitrarily generator honouring a log-normal distribution to avoid
defined as: normalized root mean squared (NRMS) lower negative values. Spatio-temporal structure of correla-
than 5%, absolute residual mean (ARM) lower than 1.5 m, tion for the recharge values were not considered in the
root mean squared (RMS) lower than 2 m. synthetic generation.

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


545
Results of decreasing evaporation flow rates from the salares is
confirmed, simulated values for the evaporation (Table 5)
Steady-state and transient-state calibration overestimate the values of Table 4. This difference can be
The calibrated values for hydraulic conductivities (24 explained by the generalized overestimation of ground-
zones) vary between 3×10−2 m/d and 6.7×101 m/d with a water heads in the sector of Salar de Pintados (Fig. 8) and
geometric mean KGM=3.5 m/d. Whereas the calibrated the uncertainty in the definition of the active evaporative
values for the storage coefficient vary from 5×10−3 up to surface of the salares.
3×10−1. In Fig. 9, some representative wells of the study area
The observed vs. simulated groundwater heads for the are presented. The overall trend of groundwater head
steady-state and transient-state models are shown in decrease is reasonably well simulated in the north area
Fig. 7. For the case of the steady-state model (Fig. 7a), (well A8; Fig. 9a) and in the centre-north area (well 60–
NRMS, ARM and RMS values show an acceptable 59; Fig. 9b). However, short-term variations are not
calibration when compared with the previously defined reproduced by the model.
calibration targets. In the case of the transient model, In well 134 (Fig. 9c), located in the centre area of the
Fig. 7b shows the results for the complete period of modelled domain (Canchones), the overall tendency of
calibration (1983–2004). From Fig. 7b it is possible to see decrease is fairly modelled until the beginning of the year
that the calibration targets are also met, but with slightly 1999. After that, a significant increase in the groundwater
poorer quality than the steady case. heads is observed, which is also properly simulated. This
The spatial distribution of the errors which are defined reflects the effect of Canchones well field being replaced
as the difference between the observed and simulated during the period 1998–1999 by the El Carmelo well field
groundwater heads is depicted in Fig. 8. From this figure, (Table 6). This is also noticed in observation well 60–59
it is possible to see that the +4 m and −4 m error terms (Fig. 9b), located near El Carmelo well field, which shows
represent an almost negligible area of the modelled a change in the slope of the overall trend for that year. In
domain. Most of the residuals are contained in the range the centre-south area (well 249), observed values are
between −2 and +2 m for a large part of the modelled area. clearly overestimated and the trend in groundwater heads
For the northern area, which corresponds to the thinnest is not properly simulated (Fig. 9d). In general, this area
part of the aquifer, the range of the error terms correspond presents an overestimation of the observed heads for a
to −1 to +1 m (2 m) and this represents 4% of the large number of observation wells. This indicates that the
thickness of the aquifer. For the rest of the modelled extension of the cone of depression produced by the
domain, the range of the error terms is less than 2% of the Canchones well field is not properly reproduced towards
average aquifer thickness (thickness ca. 250–300 m). the south-west areas in the middle of the model domain.
Simulated groundwater heads clearly overestimate the In the south area (Well 276) the modelled results are in
observed values in the middle of the domain (Canchones- concordance with the overall trend (Fig. 9e). Differences
Pintados) where head residuals are on the order of 2–4 m. between simulated and observed groundwater heads are
In the west of the domain, observed values are under- smaller than 1.5 m.
estimated on the order of 1–2 m. In the north of the model
area, the residuals are on the order of 1 m.
The simulated water balance for years 1960, 1987 and Sensitivity analysis
1993 is presented in Table 5. All flow terms are in A sensitivity analysis was performed for the following
agreement with estimations made in Table 4. Some minor calibrated parameters: constant head at the south boundary
discrepancies are observed in the transpiration term from condition (SBC), evaporation rate, extinction depth for the
Tamarugo areas which are derived from the interpolation evaporation process, recharge flow rates and hydraulic
of the values of Table 3. Groundwater outflows from the conductivities. The applied head at the SBC was varied
modelled area are nearly constant around 160 l/s for the between 20 m below and above the calibrated value. The
whole period (1960 and 1983–2004). Although the trend performance of the model in terms of the calibration

Fig. 7 Observed vs. simulated


groundwater heads for
the groundwater model of
the PTA: a results for the
steady-state model (year 1960),
b results for the transient-
state model (period 1983–2004).
NRMS normalized root
mean squared, ARM absolute
residual mean and RMS
root mean squared

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


546
components, results were insensitive to changes on both
of these parameters. For the recharge flow rates, an
arbitrary range of variation was selected (R±0.5R).
NRMS, RMS and ARM were rather sensitive to changes
in recharge rates. Evaporation in the Salar de Pintados was
the most sensitive flow component with a 50% increase in
recharge producing an 80% increase in evaporation.
Each of the 24 calibrated hydraulic conductivity (K)
zones was multiplied by the following factors: 0.01, 0.1,
0.5, 5, 10 and 100. Changes in the spatial configuration or
combined sensitivity of two or more K-parameters were not
evaluated. Model results were particularly sensitive to
changes in the hydraulic conductivity in Pintados-Can-
chones area since most of the wells are located in that area.
In terms of flow components, a variation from 0.5 up to
10 K resulted in groundwater outflows through the SBC
within the defined range (164–356 l/s). For the evaporation
flow rate, a variation between 0.1 up to K resulted in
evaporation flow rates in the estimated range (410–602 l/s).
If the values of hydraulic conductivity are larger than the
values defined in the calibrated case, i.e. (5, 10 or 100 K),
the values obtained for the evaporation flow rate from the
salares shoot up to unrealistic values.

Results simulation scenarios


Water balance results
Figure 10 shows the simulated flow components of the
PTA for different scenarios on an annual basis. In Fig. 10a,
evaporation flow rates for different scenarios are depicted.
The variations between scenarios 2005, 2005+20% and
DWP are negligible. However, for scenario 2005–20%,
evaporation flow rates are slightly larger suggesting that
actual pumping discharge is consuming groundwater that
otherwise would be lost through evaporation. This is quite
reasonable since most of the production wells are not
located near salares areas and, therefore, any increase in
pumping discharge is partially not reflected in evaporation
flow rate decreases.
Figure 10b shows the annual groundwater flow to
storage. The peak produced in the groundwater flow to
storage during the year 1998–1999 is explained by the
cessation of the pumping discharge from the well field.
This cessation induced a considerable recovery of the
groundwater heads (Fig. 9c). After year 2005, groundwa-
ter going to storage for scenarios 2005, 2005+20% and
2005–20% decreases to zero value between years 2020
Fig. 8 Spatial distribution of error terms for the steady-state model and 2030. Scenario DWP accelerates the depletion of this
(year 1960) component reaching zero value by year 2013. This is
explained by the fact that when drinking water demands
exceed the El Carmelo well field pumping capacity,
targets was stable up to SBC±10 m. For values greater Canchones well field capacity is used. Figure 10c shows
than SBC+10 m, the groundwater head increased in such a the loss of groundwater storage, i.e. groundwater storage
way that evaporation took place in the Salar Viejo which, entering the “mobile” flow system to supply the demand.
based on field observations, is unrealistic. For scenarios 2005, 2005+20% and 2005–20% this compo-
Evaporation rate and extinction depth were varied nent smoothly decreases and remains roughly constant after
between 1,250 and 2,500 mm/yr, and 0.2 and 1.8 m, year 2040. On the contrary, for scenario DWP, the increase
respectively. In terms of calibration targets and flow in the loss of groundwater storage is due to the increase in

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


547
Table 5 Simulated water balance results (l/s) for years 1960, 1987, 1993
Flow components 1960 1987 1993
In Out In Out In Out
Recharge from sub-basins 976 976 976
Transpiration Tamarugo areas 210 676 890
Evaporation from salares 603 362 275
Groundwater outflow 163 156 170
Pumping discharge 0 716 730
Total 976 976 976 1910 976 2065

pumping discharge as a response to increasing groundwater groundwater flowing out of the modelled area are
demands for drinking water purposes. imperceptible for σ1, σ2 and σ3. Results for the loss of
In the case of the DWP scenario under random groundwater storage under random recharge flow rates for
recharge flow rates, the main results are presented in σ3 and 100 recharge realizations are shown in Fig. 11a. In
Fig. 11. Changes in the evaporation flow rates and

Fig. 9 Observed and simulated groundwater heads in time: a well A8 (north sector), b well 60– 59 (centre-north sector), c well 134
(centre-centre sector), d well 249 (centre-south Sector), e well 276 (south sector)

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


548

2004
this figure, the black line represents the scenario DWP

650
695
45
under average recharge flow rates (976 l/s).
Fluctuations of the average and standard deviation of
2003

650
695
45 the loss of groundwater storage for the last year of
simulation (2050) vs. number of realizations, are shown in
Fig. 11b and c, respectively. From these figures, tenden-
2002

650
695
cies are clearly seen for σ1, σ2 and σ3. The loss of
45

groundwater storage is on the order of 2,060 l/s±23 l/s,


2001

2,085 l/s±65 l/s and 2,100 l/s±110 l/s, respectively.


650
695
45

On the other hand, for σ1, σ2 and σ3, the cumulative


loss of groundwater storage for the last year of the
2000

simulation period (2005–2050) was obtained for each


650
695
45

recharge realization. From these results, three probability


distributions were obtained for the cumulative loss of
1999

650
695

groundwater storage at year 2050. In Fig. 11d, the


45

probability distribution of the cumulative loss of ground-


water storage for σ3 is depicted. Based on the estimated
1998
670

670

groundwater storage reserve made in JICA-DGA-PCI


0

(1995) for PTA (26.9×109 m3) and from Fig. 11d, it is


unlikely that the cumulative loss of storage for supplying
1997
670

670

the demand of the system in the year 2050 will be larger


0

than 12% of the estimated groundwater storage reserves.


1996
670

670
0

Groundwater heads
In the north sector, none of the wells seem to present large
1995
Table 6 Yearly pumping rates for Canchones and El Carmelo well fields (l/s; Adapted from Rojas 2005)

670

670

impact on the groundwater heads for the different


0

scenarios. For the centre-north sector, the largest decrease


in groundwater heads is produced by scenario 2005+20%.
1994
670

670

By the end of the simulation period, the drawdown is on


0

the order of 2 m in comparison to scenario 2005. On the


other hand, in the centre-sector near Canchones well field,
1993
650

650

the largest drawdown is produced by scenario DWP


0

(ca. 18 m), whereas for El Carmelo well field, the largest


1992

drawdown is 3 m. For the south sector, no major effect on


650

650

groundwater heads for scenarios 2005+20% and DWP in


0

comparison to scenario 2005 are observed.


1991

To assess the effect of random recharge rates on the


650

650
0

simulated groundwater heads, seven artificial observation


wells are located at different representative places. For the
1990
650

650

artificial observation well P1 located in the lower area of


0

the alluvial fan produced by the Aroma sub-basin (north),


fluctuations (for σ1, σ2 and σ3) in groundwater heads are
1989
650

650

between 0 and 1.5 m (Fig. 12). For observation wells


0

located in the lower part of the alluvial fan of Tarapaca


1988

and Sagasca creeks (north and centre), significant ground-


630

630

water heads fluctuations are observed only for σ3 and they


0

are on the order of 0.4–1 m at the end of the simulated


1987

period. These areas present the largest variation from the


630

630
0

average recharge situation. This could be explained


because they are located in the confluence of Aroma and
1986
610

610

Tarapaca sub-basins, which jointly generate ca. 50% of


0

the total incoming groundwater recharge.


Observation wells located in the centre areas present
1985
Year

580

580

minor fluctuations (0.1 m) for the largest uncertainty level


0

(σ3). For an observation well located in the alluvial fan of


Chacarilla creek (south), fluctuations from the average
El Carmelo
Canchones
Well field

recharge situation are on the order of 0.5 m.


From this, it is clear that the northern Aroma and
Total

Tarapaca sub-basins influence the groundwater heads in

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


549

Fig. 10 Annual simulated flow components for scenarios 2005, 2005+20%, 2005–20% and the drinking water projection (DWP):
a evaporation from salares (l/s), b groundwater flow to storage (l/s), and c loss of groundwater storage (l/s)

the north sector of the study area. If the groundwater calibrated parameters contained in realistic ranges that do
recharge coming from these sub-basins changes in not substantially differ from values obtained in past
magnitude, it will be reflected in areas as far as 19 km studies. Also, groundwater balance results are in agree-
from Aroma alluvial fan apex or 15 km from Tarapaca ment with previous estimations made in past studies.
alluvial fan apex. Also, the southern Chacarilla sub-basin The impossibility of estimating pumping discharges at
has minor influence on the groundwater heads located in a time scale less than 1 year without exhaustive fieldwork,
the lower areas of the corresponding alluvial fan. makes it difficult to reproduce short-term fluctuations in
the observed groundwater heads, particularly in those
wells affected by local pumping conditions. However,
the observed overall trend of the groundwater heads and
Discussion and conclusions the observed flow-field are properly reproduced with the
exception of a limited area towards Salar de Pintados.
PTA is a strategic source of groundwater which has been Simulated evaporation flow rates overestimate calcu-
intensively used since the 1960s. The groundwater extrac- lations made in past studies. This can be attributed to two
tion for different uses has a clear effect on the observed main factors: first, the simulated groundwater heads
groundwater heads, especially, in those zones affected by overestimate the observed groundwater heads in the Salar
the Canchones and El Carmelo well fields. In these areas, a de Pintados area and, second, there is uncertainty
more pronounced drawdown is observed compared to other regarding the actual extent to which evaporation from
areas. Although Houston (2002) suggested that these long- the surface of salares takes place. These two points
term drawdowns in groundwater heads may partly result represent the major limitations of the groundwater flow
from long-term climatic fluctuations, it is likely that these model. To alleviate these limitations, it is proposed to
drawdowns are simply due to overexploitation. fully describe this area in a three-dimensional conceptual
This study has developed an up-to-date groundwater model and to define, in a more accurate way, the active
flow model, calibrated for steady-state conditions (1960) surface for the evaporation process.
and transient-state conditions (1983–2004). Different Scenarios show that groundwater heads will continue
groundwater extraction situations are considered and, to decrease with the actual pumping discharge (2005).
more importantly, a sensitivity study on the average This decrease is slightly more severe for an increase of
recharge values is performed. The calibration criteria 20% in the artificial discharge. For the case of the DWP
define an acceptable performance of the model with final scenario, areas located near the Canchones well field are

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


550

Fig. 11 Results for scenario DWP+randomly generated recharge: realization for three uncertainty levels (year 2050) and d probability
a annual loss of groundwater storage for σ3 and for 100 recharge distribution for the simulated cumulative loss of groundwater
realizations; b average loss of groundwater storage vs. number of storage for σ3 (year 2050). σ1, σ2 and σ3 are the minimum,
realization for three uncertainty levels (year 2050); c standard average and maximum levels of uncertainty, respectively
deviation of the loss of groundwater storage vs. number of

more drastically affected than areas located near the El ing that part of the present system’s demand is consumed by
Carmelo well field. Although this is strictly related to the the evaporation process. This presents two conflicting
assumptions that defined the respective scenarios, even a courses of action: to reduce the present pumping discharge
reduction in the actual pumping discharges shows this to reduce the demand on the system, knowing that part of
trend of increasing drawdown. It is clear that in the mid this reduced discharge will be lost through evaporation from
and long-term, groundwater users of the PTA will be salares, or to increase the actual pumping discharge to the
facing a worse situation than the present one. point where no groundwater is lost through evaporation.
On the other hand, if pumping discharges are reduced, This analysis is only valid if there is no other source or sink
the evaporation flow rates tend to slightly increase, suggest- present in the modelling system.
For scenario DWP under random recharge values,
evaporation flow rates from salares and SBC groundwater
outflows are invariable to all uncertainty levels in recharge
(σ1, σ2 and σ3). This could be related to: (1) the time span
(45 years) used to evaluate the flow components under
random recharge values, which could suggest that the
simulation period might be too short to observe more
pronounced effects, or (2) since evaporation is strong and
recharge is weak in this area, the insensitivity might also be
due to the evaporation overpowering the recharge effect.
Storage flow components vary according to the level of
uncertainty. The loss of groundwater storage in supplying
the system demand increases when the uncertainty in the
Fig. 12 Simulated groundwater heads for a synthetic observation recharge increases. Despite this, the cumulative loss of
well (P1) in the Aroma creek for σ3=0.35R and for 100 recharge groundwater storage represents less than 12% of the
realizations (the black line represents the average recharge value) assessed groundwater total reserves.

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6


551
Uncertainties in the simulated groundwater heads due to Dirección General de Aguas–Comisión Chilena de Energía Nuclear–
uncertainties in average recharge rates are more noticeable DGA-CCHEN (1998) Aplicación de técnicas hidrológicas no
convencionales en zonas áridas y semiáridas. Evaluación de
in areas near Aroma, Tarapaca and Chacarilla creeks. These recursos hídricos en el sector de Pica, Hoya de la Pampa del
sub-basins represent ca. 81% of the total incoming Tamarugal, I Región [Application of non-conventional hydro-
recharge. Although these fluctuations in groundwater head logical techniques in arid and semi-arid zones. Water resources
are minor, they certainly could explain observed unusual evaluation in Pica area, Pampa del Tamarugal basin, I Region].
Technical report S.I.T. No. 48, Centro de Información de
behaviour in wells located in these areas, where some local Recursos Hídricos, Dirección General de Aguas, Ministerio de
recharge events could be expected. Obras Públicas, Santiago, Chile, pp 101
Finally, it is clear that temporal and spatial variation of the Dirección General de Aguas–Universidad de Chile-DGA-UChile
recharge events may play an important role. The use of a (1988) Modelo de simulación hidrogeológico de la Pampa del
stochastic framework in the groundwater modelling has Tamarugal [Hydrogeological simulation model of the Pampa del
Tamarugal]. Technical Report, Centro de Información de
demonstrated this fact, especially, in those zones where the Recursos Hídricos, Dirección General de Aguas, Ministerio de
random nature, spatial and temporal, of precipitation (assum- Obras Públicas, Santiago, Chile, pp 98
ing it is the main source of groundwater recharge) is much Dirección de Investigaciones Científicas y Tecnológicas Universidad
more pronounced, e.g. arid zones. Undoubtedly, the northern Católica de Chile–DICTUC (2005) Análisis técnico de solicitud de
derechos de aprovechamiento de ACF minera [Technical analysis of
part of Chile is one of these zones and similar approaches can the groundwater rights request of mining company ACF] Technical
be extended to other aquifers in similar arid zones. report, Dirección de Investigaciones Científicas y Tecnológicas
Universidad Católica de Chile, Santiago, Chile, pp 107
Acknowledgements The authors wish to thank to the Dirección Food and Agriculture Organization–FAO (1989) Role of Forestry
General de Aguas, Chile for providing all the necessary information in combating desertification. FAO Conservation Guide 21,
for the development of this study. Also, the comments of two Food and Agriculture Organization of the United Nations,
anonymous reviewers are greatly appreciated for improving the Rome
original manuscript. Grilli A, Vidal F, Garín C (1986) Balance hidrológico nacional, I
Región [National hydrologic balance, I Region]. Dirección
General de Aguas. Technical Report EH/86/2, Centro de
References Información de Recursos Hídricos, Dirección General de
Aguas, Ministerio de Obras Públicas, Santiago, Chile, pp 20
Aravena R (1995) Isotope hydrology and geochemistry of northern Grilli A, Aguirre E, Duran M, Townsend F, Gonzales A (1999)
Chile groundwaters. Bull Inst Français Études Andines 24 Origen de las aguas subterráneas del sector Pica-Salar del
(3):495–503 Huasco, Provincia de Iquique, I Región de Tarapaca [Origin of
Arrau F (1998) Distribución y comercialización de las aguas en the groundwater in the sector of Pica-Salar del Huasco,
Chile [Water distribution and commercialization in Chile]. Province of Iquique, I Region of Tarapacá]. In: AIDIS-CHILE
Study Series No. 178, The Library of the National Congress, Congress, Capítulo chileno Asociación Interamericana de
Santiago, Chile Ingeniería Sanitaria y Ambiental, Santiago, Chile, pp 18
Brown E (2003) Uso eficiente del recurso hídrico [Efficient use of Houston J (2002) Groundwater recharge through an alluvial fan in
the water resource]. In: National Workshop of Chile, Towards a the Atacama Desert, northern Chile: mechanisms, magnitudes
national plan of integrated water resources management and causes. Hydrol Process 16:3019–3035
CEPAL, Chile. December 2003, CEPAL/GWP-SAMTAC, Japanese International Cooperation Agency, Dirección General de
Santiago, Chile, pp 22 Aguas, Pacific Consultants International (JICA-DGA-PCI)
Doherty J, Brebber L, Whyte P (1994) PEST: Model-independent (1995) The study on the development of Water Resources in
parameter estimation. User’s manual. Watermark Computing, northern Chile. Technical reports. Supporting reports B and C,
Corinda, Australia Centro de Información de Recursos Hídricos, Dirección
Digert F, Hoke G, Jordan T, Isacks B (2003) Subsurface General de Aguas, Ministerio de Obras Públicas, Santiago,
stratigraphy of the neogene Pampa del Tamarugal basin, Chile, pp 259
northern Chile. X Congreso Geológico Chileno, Concepción, Margaritz M, Aravena R, Peña H, Suzuki O, Grilli A (1990) Source
Chile, October 2003, pp 8 of groundwaters in the deserts of northern Chile: evidence for
Dingman R, Galli C (1965) Geology and groundwater resources of deep circulation of groundwaters from the Andes. Ground Water
the Pica area, Tarapaca Province, Chile. Bull United States 28:523–517
Geological Survey No. 1189, USGS, Reston, VA McDonald M, Harbaugh A (1988) A modular three-dimensional
Dirección General de Aguas–DGA (1987) Balance hídrico de Chile finite-difference groundwater flow model. USGS Technical
[Water balance of Chile]. Technical Report–Digital data, Centro Report on Modelling Techniques Book 6, USGS, Reston, VA,
de Información de Recursos Hídricos, Dirección General de pp 596
Aguas, Ministerio de Obras Públicas, Santiago, Chile Risacher F, Alonso H, Salazar C (1998) Geoquímica de aguas en
Dirección General de Aguas–DGA (1996) Determinación de la cuencas cerradas I, II y III regiones–Chile [Water geochemistry in
disponibilidad de recursos hídricos para constituir nuevos closed basins I, II and III regions–Chile]. Technical Report DGA-
derechos de aprovechamiento de aguas subterráneas en el sector UCN-ORSTOM, Centro de Información de Recursos Hídricos,
del acuífero de la Pampa del Tamarugal [Determination of water Dirección General de Aguas, Ministerio de Obras Públicas,
resources availability for granting new groundwater rights in Santiago, Chile, pp 84
the area of the Pampa del Tamarugal aquifer]. Technical report Rojas R (2005) Groundwater flow model of Pampa del Tamarugal
S.D.T. No. 68, Centro de Información de Recursos Hídricos, Aquifer–Northern Chile. MSc Thesis Water Resources Engi-
Dirección General de Aguas, Ministerio de Obras Públicas, neering. Katholieke Universiteit Leuven and Vrije Universiteit
Santiago, Chile, pp 27 Brussel, Belgium, pp 103

Hydrogeology Journal (2007) 15: 537–551 DOI 10.1007/s10040-006-0084-6

Vous aimerez peut-être aussi