Saltwater Wedge Variation in A Non Anthropogenic Coastal Kars - 2014 - Journal o

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Hydrology 519 (2014) 2350–2365

Contents lists available at ScienceDirect

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

Review Paper

Saltwater wedge variation in a non-anthropogenic coastal karst aquifer


influenced by a strong tidal range (Burren, Ireland)
Marie Perriquet a,b,⇑, Véronique Leonardi a,1, Tiernan Henry b,2, Hervé Jourde a,3
a
Université Montpellier 2, HydroSciences Montpellier, Case MSE, Place Eugène Bataillon, 34095 Montpellier Cedex 5, France
b
Earth and Ocean Sciences, School of Natural Sciences, National University of Ireland, Galway, University Road, Galway, Ireland

a r t i c l e i n f o s u m m a r y

Article history: Spatial and temporal changes in saltwater wedges in coastal karst aquifers are still poorly understood,
Received 28 August 2014 largely due to complex mixing processes in these heterogeneous environments, but also due to anthro-
Accepted 2 October 2014 pogenic forcing such as pumping, which commonly affect natural variations in wedges. The purpose of
Available online 16 October 2014
this study was first to characterize the hydrodynamic functioning of a karst aquifer in an oceanic temper-
This manuscript was handled by Peter K.
Kitanidis, Editor-in-Chief, with the
ate climate with little anthropogenic pressure but strongly influenced by a high tidal range and second, to
assistance of Barbara Mahler, Associate evaluate the extent and movements of a saltwater wedge influenced by both the tide and the natural
Editor recharge of the aquifer. Variations in specific conductivity combined with water chemistry results from
six boreholes and two lakes located in the Bell Harbour catchment (western Ireland) enabled us to assess
Keywords: the extent of the intrusion of the saltwater wedge into the aquifer as a function of both karst recharge and
Karst tidal movements at high/low and neap/spring tidal cycles. The marked spatial disparity of the saltwater
Coastal aquifer wedge was analysed as a function of both the hydrodynamic and the structural properties of the karst
Saltwater wedge aquifer. Results showed that the extent of the saltwater wedge depended not only on the intrinsic prop-
Tidal influence erties of the aquifer but also on the relative influence of the recharge and the tide on groundwater levels,
Burren (Ireland) which have opposite effects. Recharge in the Burren area throughout the year is large enough to prevent
Hydraulic gradient
saltwater intruding more than about one kilometre from the shore. A strong tidal amplitude seems to be
the motor of sudden saltwater intrusion observed in the aquifer near the shore while the position of the
groundwater level seems to influence the intensity of the salinity increase. Competition between
recharge and the tide thus controls the seawater inputs, hence explaining temporal and spatial changes
in the saltwater wedge in this coastal karst aquifer.
Ó 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2351
2. Study area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2352
2.1. Location and climatic conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2352
2.2. Geology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2353
2.3. Hydrogeology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2353
2.4. Recent hydrogeological results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2354
2.5. Field sampling and data available . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2354
3. Assessment of karst hydrodynamic properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2355
3.1. Hydrodynamic response to recharge events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2355
3.2. Hydrodynamic response to tidal variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2356

⇑ Corresponding author at: Université Montpellier 2, HydroSciences Montpellier, Case MSE, Place Eugène Bataillon, 34095 Montpellier Cedex 5, France. Tel.: +33 7 61 22 26
30.
E-mail addresses: mperriquet@gmail.com (M. Perriquet), veronique.leonardi@univ-montp2.fr (V. Leonardi), tiernan.henry@nuigalway.ie (T. Henry), herve.jourde@
univ-montp2.fr (H. Jourde).
1
Tel.: +33 4 67 14 32 45.
2
Tel.: +353 91 495 096.
3
Tel.: +33 4 67 14 90 80.

http://dx.doi.org/10.1016/j.jhydrol.2014.10.006
0022-1694/Ó 2014 Elsevier B.V. All rights reserved.
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2351

Nomenclature

d delay factor MSL mean sea level


f tidal efficiency factor NT neap tide
HT high tide P period of influence of a recharge event
i hydraulic gradient S slope of a recession curve
imin lowest hydraulic gradient observed during a low/high SpC specific conductivity
tidal cycle SST strong spring tide
iampl hydraulic gradient difference between high and low tide ST spring tide
IPCC Intergovernmental Panel on Climate Change WL water level
LT low tide

3.3. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2356


3.4. Discussion: Consequences in terms of intrinsic functioning of the karst aquifer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2357
4. Spatial and temporal variability of the saltwater wedge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2358
4.1. Spatial extent of the saltwater wedge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2358
4.2. Temporal variability of the extent of the saltwater wedge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2360
4.2.1. Temporal variation in saltwater wedge at high/low and neap/spring tidal scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2360
4.2.2. Conditions for larger intrusion of the saltwater wedge into the karst aquifer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2361
4.3. Discussion on the temporal saltwater wedge variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2363
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2364
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2365
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2365

1. Introduction 2002; Blavoux et al., 2004; Bonacci and Roje-Bonacci, 1997;


Cooper, 1959; Henry, 1959; Herzberg, 1901; Hubbert, 1940). The
Understanding the interaction between freshwater and seawa- hydraulic gradient depends on (i) tidal fluctuations, (ii) natural
ter in coastal karst aquifers is critical for current and future water recharge of the aquifer and (iii) spatial heterogeneities of its hydro-
management, especially in the context of various EU Directives dynamic properties. Tidal fluctuations may be a major factor in
(Water Framework Directive, Floods Directive, Marine Strategy determining variations in boundary conditions. For example,
Framework Directive) and given the potential impacts of changing Beddows et al. (2007) showed that variation in pressure heads dri-
weather patterns and sea levels (Ferguson and Gleeson, 2012). The ven by the tide can increase the rate of mixing. Martin et al.
natural equilibrium between seawater and freshwater is still (2012) quantified the exchanges between the matrix and conduits
poorly understood, largely due to the complex mixing processes as a function of tidal amplitude. However to date, no study has
in these heterogeneous environments (Dörfliger and Willwer, explained the spatial and temporal changes in the saltwater wedge
2005). The presence of karst conduits that allow discharge of fresh- in a karst aquifer taking into account both water chemistry (major
water and ingress of sea water make it difficult to predict spatial cation/anion analysis) and tidal influence (at high/low tide and
and temporal changes in the saltwater wedge (Bakalowicz, 2005). neap/spring tide scales) on variations in SpC in an aquifer with very
In these aquifers, the extent of the saltwater wedge thus depends low anthropogenic forcing. Natural recharge of the aquifer may also
on the flow path network, resulting in irregular patterns of saltwa- be a major factor driving variations in inland hydraulic heads: the
ter intrusion into the aquifer (Arfib et al., 2002). Previous work on extension of the saltwater wedge over time could depend on the
saltwater wedges in karst aquifers show the effect of areas of high equilibrium between groundwater recharge and the influence of
hydraulic conductivity on the distribution of the mixing zone tides that modify the boundary conditions. While a lot of published
(Moore et al., 1992; Stringfield and LeGrand, 1971). The morphol- studies (Arfib et al., 2002; Bayari et al., 2010; Fleury et al., 2007;
ogy and geometry of conduits have been shown to be first order Maramathas et al., 2003; Tulipano, 2005) focus on coastal karst
controls on groundwater temperature and specific conductivity aquifers in arid to semi-arid areas such as the Mediterranean region,
(SpC) gradients in the aquifer (Beddows et al., 2007). Knowledge this study focuses on the Bell Harbour karst catchment, which is
of the spatial variations in the hydrodynamic properties of the characterised by a strong tidal range, a consistently high recharge
aquifer is therefore required to understand the extent of the salt- throughout the year, and a low population density in a rural agricul-
water wedge. Hydraulic diffusivities can be estimated through tural setting. Because of the low water demand in the Burren area,
groundwater responses to tides. Jacob (1950) and Ferris and this zone is very well suited to study the saltwater intrusion mech-
Branch (1952) provided analytical solutions to assess hydraulic dif- anisms in a close to natural context (little anthropogenic forcing).
fusivity from changes in groundwater levels, while considering a The Intergovernmental Panel on Climate Change (IPCC) forecast glo-
one-dimensional (1D), homogeneous, isotropic, confined, and bal sea level rise of up to 1 m by 2100, and this will directly affect the
semi-infinite aquifer with a sharp boundary condition subject to hydrogeological behaviour of the karst aquifer. These changes are
oscillating forcing. In this study, analyses of water table fluctua- examined at the light of the presented results.
tions in boreholes highlighted three types of hydrodynamic The aim of this study was to test the hypothesis that recharge
response that were confirmed with the Jacob–Ferris method. and the tide control spatial and temporal changes in the saltwater
Previous studies showed that saltwater intrusion is driven by the intrusion. To this end, the intrinsic properties of the karst aquifer
hydraulic gradient, which in turn, is controlled by the difference in (hydrodynamic properties, structure) were characterised, and then
hydraulic heads between the sea and the aquifer (Arfib et al., an attempt was made to link these properties and spatial and tem-
2352 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Fig. 1. (a) Location of the study area in Ireland; (b) location of Bell Harbour catchment in the Burren region; (c) aerial photograph of Bell Harbour catchment with location of
the monitoring points, the main fault, and intertidal springs. Spatial extent of the saltwater wedge and its temporary and tidally influences over the catchment are
represented; the maximum, minimum and average SpC values in six boreholes and two lakes are also shown.

poral changes in the saltwater intrusion into the karst system. The 2. Study area
relative influence of recharge and of the tide on the water level
(WL) and variations in SpC measured in six boreholes and two 2.1. Location and climatic conditions
lakes is discussed and a general conceptual model explaining the
extent and movements of the saltwater wedge in a karst aquifer lit- The Bell Harbour catchment, covering an area of around 50 km2
tle affected by anthropogenic pressure is proposed. in County Clare, is located in a large karst area called the Burren on
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2353

Fig. 2. Geological cross-section of the Bell Harbour catchment.

the west coast of Ireland and is defined by a valley surrounded by 2.3. Hydrogeology
upland areas to the west, south and east that reach an altitude of
about 300 m. Discharge from the catchment is to the north into Bell The landscape in the Burren region (and in the study area) is
Harbour, which opens onto Galway Bay (Fig. 1). The Burren has an among the best examples of karst landscape in Europe and is the
oceanic climate, with annual precipitation averaging 1500 mm finest example of karst terrain in Ireland (Drew, 2001). The pure
and annual effective rainfall estimated at 980 mm (Walsh, 2012). bedded limestones are exposed to weathering and are susceptible
The rainfall occurs throughout the year, although the spring and to dissolution, leading to the formation of a well-developed shal-
early summer months tend to be drier. low epikarst of 5–10 m thickness, and distinctive karst topographic
features such as swallow holes, sinking streams, limestone pave-
2.2. Geology ments, caves and large springs. There are three main groundwater
flow systems operating in this region (from the shallowest to the
The Burren region is one of the most extensive limestone karst deepest): (1) a strongly altered, rapidly draining shallow epikarstic
areas in north-western Europe (600 km2) (Gallagher et al., 2006) system present in the upper 5–10 m; (2) an unsaturated (or infil-
dominated by massive or bedded Carboniferous limestone of sev- tration) zone where vertical flows pass through the rock via system
eral hundred metres thickness and bounded to the east by lime- of fissures or joints; (3) a phreatic zone where three distinct com-
stones of the Gort Lowlands and to the south by Namurian partments are linked together: conduits which allow rapid ground-
sandstones and shales (Fig. 1b). In the Bell Harbour catchment, water flows, fissures which allow relatively rapid flows, and the
gently dipping (2–3° to the south) pure-bedded limestones of the limestone matrix which allows slow movement of groundwater.
Burren Formation underlie the entire area. These rocks are charac- Flow paths through the unsaturated zone may also be along similar
terised by pale grey and thickly to massively bedded limestone with flow paths.
occasional cherty intervals of shale and dolomite horizons (Pracht Diffuse recharge dominates in the Bell Harbour catchment
et al., 2004). These limestones are underlain by impure limestones across the large area of exposed limestone pavement in the south-
(which are not expressed at the surface) (400 m in thickness), ern (higher elevation) portion of the catchment. The main flow
which overlie Devonian Old Red Sandstones and the Galway Granite path of the water in the limestone is controlled by (i) the horizon-
pluton (Fig. 2). In the Burren, faults can be identified both on the sur- tal bedding partings (Fig. 2) as the impermeable shale layers or
face (Simms, 2001) or underground (Judd and Mullan, 1994): only chert lenses limit the vertical movement of the groundwater
two major faults are mapped in the area, one of which is in the wes- (Drew, 2003), and by (ii) the hydraulic head boundary conditions
tern part of the Bell Harbour catchment study area (Fig. 1c). This (groundwater basin divide to the south, and the boundary imposed
fault, known as MacDermott’s Fault, runs approximately north– by the tide in Bell Harbour). Water drains almost wholly by under-
south and shows a slight sinistral displacement of members of the ground conduits directly into Bell Harbour via submarine or littoral
Burren Formation. Pracht et al. (2004) assume the apparent dis- diffuse springs (Fig. 2). Groundwater flow rates in the range 50–
placement on the fault is minimal, less than 200 m. Geophysical 150 m/h have been recorded from water tracer tests under base
investigation of this fault is currently being completed, using elec- flow conditions in the Burren, and flow velocities are assumed to
trical resistivity tomography, as part of this ongoing research. Initial increase up to four fold in flood conditions and to halve under very
indications suggest that the fault extends beyond its current low flow conditions (Drew and Daly, 1993). Aquifer storage is low
mapped extent running under Bell Harbour bay (O’Connell, 2012, and during extended wet periods the conduit systems can back up,
personal communication). Joints and veins are extensive through- leading to the development of two temporary lakes or turloughs
out the area (Gillespie et al., 2001). (Luirk and Gortboyheen lakes), which either drain or feed the karst
2354 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Fig. 3. Example of continuous data (temperature, SpC, WL) monitored at borehole B05, plotted with the tide, precipitation and hydraulic gradient from August 27, 2010 to
January 4, 2011.

aquifer, and may persist on the landscape for weeks or even min intervals from September 2010 to September 2011 and at
months (Fig. 1c). five-minute intervals from December 2011 to January 2013
(Fig. 1c). Data were not recorded at every measurement point for
2.4. Recent hydrogeological results the full periods. In Situ Aqua TROLL 200 loggers were installed in
the six boreholes and CTD-Divers were sited at two lakes. The In
At the start of this study, field investigations were undertaken Situ loggers were connected to a vented cable to account for atmo-
to locate karst features and to assess and select measurement loca- spheric pressure. A Baro-Diver installed adjacent to borehole B05
tions (boreholes, springs etc.). Six private boreholes were selected allowed for compensation of the pressure measurements from
for the study. Five intertidal springs were identified along the east- the CTD-Divers. An RTK (Real Time Kinematic) GPS survey of all
ern shore of Bell Harbour and there are two turloughs and one per- boreholes and collection points was undertaken to relate all WL
manent lake (Muckinish Lake) in the catchment (Fig. 1). Discrete data collected to mean sea level (MSL). In addition, vertical profiles
SpC measurements at the five intertidal springs were done at low of salinity and temperature were recorded on June 6, 2012 in four
tide (LT), and ranged from 40 mS/cm to 400 lS/cm depending on boreholes: B03, B05, B08 and B57 during a rainy period. Data on
the tidal stage and the hydrological period. A LIDAR map of the tide height were provided by the Marine Institute and were
area showed circular features on the bed of the Bay which are recorded at Galway Port station at a six minute time interval. This
aligned with MacDermott’s fault. These were assumed to be either data set was then converted into a five minute step by linear inter-
submarine springs discharging fresh or brackish water, or swallow polation to enable direct comparison with data collected in the
holes allowing seawater inflow into the aquifer (Fig. 1c). These fea- boreholes and lakes. Continuous measurements of the WL taken
tures have not previously been described; a salinity survey of the over a five month period at an intertidal spring in Bell Harbour
bay performed during this project, as well as continuous SpC data allowed estimation of a tidal time lag of 1 h in the bay (which
collected at different depths above one of these circular features, was used in this study), compared to the tide height recorded at
suggest that brackish water outflow from the aquifer only occurs Galway Port station. Rainfall data were measured at National Uni-
during large recharge events. Brackish water outflows and saltwa- versity of Ireland, Galway (NUIG) using a daily read rain gauge:
ter inflows may therefore alternate, depending on the hydraulic readings are totals for the 24-h period from 9 am to 9 am the next
gradient within the aquifer, similar to the karst system of the Betic day. The hydraulic gradient was calculated to estimate the poten-
Cordilleras in Southern Spain (Fleury et al., 2008). tial outflow at each measurement point, distinguishing the load
variation in the aquifer from the load variation at the limit (tidal
2.5. Field sampling and data available variations in the bay). Tidal height was subtracted from the WL
and the result divided by the linear distance between the measure-
This project is the first detailed groundwater study to be under- ment point and the shore. An example of the data from B05 is
taken in this catchment. Temperature, SpC and water level (WL) shown in Fig. 3. Additionally, discrete water chemical measure-
were recorded in six boreholes (B03, B05, B08, B15, B57 and B59) ments were performed: water samples were collected for major
and two lakes (Muckinish lake, L01 and Luirk lake, L02) at 15- cation/anion analysis from four boreholes (B05, B08, B15, B57)
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2355

Fig. 4. Examples of hydrodynamic responses for each of the six boreholes from the typology we established. A different period has been selected for B15 as no WL data have
been collected at this borehole from the January 12 to 27, 2011. Slopes (S) and period of influence (P) of each recession curve have been drawn.

and two of the turloughs (L01, L02) on March 21, 2011 during a rel- hydraulic diffusivities assessed using the Jacob–Ferris equation
atively dry period. Before raw groundwater samples were collected for each borehole.
from existing water taps connected directly to the pumps in bore-
holes, field parameters were monitored to indicate geochemical
3.1. Hydrodynamic response to recharge events
stabilization (5–10 min).
Observations of changes in the WL in each of the six boreholes
during and after a recharge event enabled determination of the
3. Assessment of karst hydrodynamic properties typical hydrodynamic behaviour of each borehole using the inten-
sity of the variation in WL (amplitude and period of influence (P))
WL variations were analysed first in response to precipitation and the slope (S) on the WL recession curve. The steepness of the
and second in response to tidal fluctuations to better understand slope was assumed to reflect the drainage of the karst system
the hydrodynamic behaviour of boreholes and the overall system. (Powers and Shevenell, 2000; Shevenell, 1996): a steep slope rep-
However, even though the analysis of the functioning of the aquifer resents the dominant effects of the larger karst features (i.e. con-
using recharge events implies consideration of direct infiltration duit/large fissures) on drainage, but also includes the effects of
through sinkhole and the pathway between the surface and the the other regimes. A break and decrease in the slope illustrates
borehole, most of the analysis relies on the interpretation of WL the storage of the aquifer that results in the emptying of well-con-
recession curves, that immediately follow the recharge event. nected karstified fissures and/or the matrix. Three segments in a
These WL recession curves are driven by both the hydrodynamic borehole recession curve in a multiple porosity karst system can
properties and the connectivity of the saturated zone, allowing thus be interpreted as the consequence of three types of flow
comparison with the analysis of WL variations in boreholes due occurring in: (i) conduits/large fissures; (ii) fissures; and (iii) the
to the effect of tide. matrix. One example of a typical hydrograph recession curve was
Moreover, these two methods were used during different peri- selected for each of the six boreholes and classification was based
ods: (i) during ‘‘high flow periods’’, when the main process driving on the range of responses starting from the borehole with the high-
WL variations is recharge as the tidal variations have little effect on est amplitude and steepest slope (when several segments were
WL; (ii) during ‘‘low flow periods’’ when the main process driving present, only the steepest one was used for the classification)
WL variations is the tide as there is no recharge. These two differ- (D1) to the borehole with the lowest amplitude and flattest slope
ent methods were used to ascertain the coherency between the (D6) (Fig. 4). The period of influence P is indicative only and was
various types of hydrodynamic environment estimated from not used for the classification as it is not the most representative
hydrodynamic responses to recharge events and the different in estimating hydraulic diffusivities. This analysis was applied to
2356 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

the boreholes but not to the lakes (L01, L02). (Changes in the WL of of sinusoidal oscillations of pressure in one-dimensional flow
the lakes are strongly dampened by the capacity of the lake and (implying neither vertical nor flow parallel to the shoreline). How-
consequently cannot be compared to changes in WL in the bore- ever, we can assume that some karst features (especially conduits),
holes). For example, the hydrodynamic reaction of B57 exhibits a where pressure variation propagates between the bay and the
high amplitude peak (>10 m) and a steep recession curve aquifer, are confined; for unconfined karst features the range of
(S57 = 6.3 m/d) (Fig. 4). Moreover, its period of influence was short variation of the water table level (in response to the tidal ampli-
(P57 = 1.6 days). This indicates a large value of hydraulic diffusivity tude) is small compared to the saturated thickness of the aquifer,
(D1) and hence high connectivity of this borehole with the karst which makes it possible to use the Jacob–Ferris equation to assess
drainage network, due to well-connected fissures. its hydrodynamic properties (Erskine, 1991), which corresponds to
Eq. (1):
3.2. Hydrodynamic response to tidal variations pffiffiffiffiffiffiffiffiffiffiffiffi  t pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

h ¼ h0 eðx pS=t0 TÞ sin 2p  x pS=t0 T ð1Þ
t0
The Jacob–Ferris method (Ferris and Branch, 1952; Jacob,
1950) for tidal propagation through a coastal aquifer was used where h is WL (m) above MSL; x is distance from the sea (m); t is
to estimate hydraulic diffusivity (D) values, although a variety time (d); t0 is the period of tidal oscillation (d) which for the west
of more sophisticated analytical solutions exist for one and of Ireland is 0.52 days; h0 is the amplitude of tidal oscillation (m);
two-dimensional systems (Li et al., 2000; Townley, 1995; Trefry, T is the transmissivity of the aquifer (m2/d); and S is aquifer storage.
1999) and for vertical section systems (Guo et al., 2010; Li and Tidal oscillations remain sinusoidal with a time lag (delay = d)
Jiao, 2002). The relatively simple Jacob–Ferris equation remains and decrease exponentially in amplitude with distance from the
useful for estimating aquifer hydraulic properties from measured sea (tidal efficiency factor = f). An analytical solution to Eq. (1) pro-
groundwater level variations (Rotzoll et al., 2013), even though vides expressions for d and f:
problems can arise from the application of this simple homoge- rffiffiffiffiffiffiffiffiffi
neous tidal propagation models. Differences in hydraulic diffusiv- St0
d¼x ð2aÞ
ities calculated from either tidal efficiency values (Df: ratio of 4pT
amplitude between the tidal level in the bay and the WL data col-
pffiffiffiffi
lected at the monitoring point) or delay values (Dd: time lag x pS
f ¼e Tt 0
ð2bÞ
between the tidal level in the bay and the WL data collected at
the monitoring point) have been noted in numerous studies These solutions can be rearranged into expressions for aquifer
(Drogue et al., 1984; Erskine, 1991; Ferris and Branch, 1952; hydraulic diffusivity (D) in terms of the delay (Dd) and tidal effi-
Smith, 1999). Normally, for a confined homogenous aquifer ciency (Df):
where head losses are isotropic, Df/Dd value should be 1 (Ferris
and Branch, 1952). These inconsistencies of observed hydraulic T 2
Dd ¼ ¼ ðxt0 Þ=ð4pd Þ ð3aÞ
diffusivities (tidal efficiency-delay inconsistencies) have only S
been partially explained because of the various influences at play:
vertical flows, phreatic influences (Jha et al., 2003), geometric T xp
Df ¼ ¼ ð3bÞ
effects (e.g. variable aquifer thickness), non linearities associated S ðln f Þ2 t 0
with capillarity and density-driven flow and spatial heterogeneity
(e.g. horizontal layering) (Trefry and Bekele, 2004). These tidal The delay is the time lag between the tidal signal and WL signal
efficiency-delay inconsistencies were also identified in this study and the tidal efficiency factor corresponds to the ratio of the tidal
but the assessed hydraulic diffusivities are in agreement with the signal with the amplitude of the WL. For each monitoring point,
heterogeneity of the aquifer; these biases have been used to sup- the delay and the tidal efficiency factor were calculated for sepa-
port the different types of proposed hydrodynamic environments rate 15-day periods without the influence of recharge and for dif-
and are discussed in detail below. In any case, this method gives ferent tidal amplitudes during spring tide (ST) and neap tide
a general idea of the hydraulic diffusivity between the shoreline (NT). The average hydraulic diffusivity was then deduced: Df from
and the monitoring point and thus allows identification of the the tidal efficiency factor and Dd from the delay factor. This method
spatial distribution of the hydraulic diffusivity field at the catch- was not used for B15 as no tidal influence was detected in this
ment scale. borehole, probably due to its low hydraulic diffusivity and its dis-
The estimated values correspond to the hydraulic diffusivity of tance from the shore.
an equivalent porous medium between the sea and the borehole,
which may be representative of either one of these hydrodynamic 3.3. Results
environment (conduit, fracture and matrix), or more likely a com-
bination of these environments. Use of the Jacob–Ferris equation The boreholes can be ranked on their hydraulic diffusivities
generally assumes a confined aquifer with homogeneous hydrody- (from highest to lowest) based on their hydrodynamic response
namic properties and considered the propagation along the aquifer to recharge (Fig. 4 and Table 1):

Table 1
Type of hydrodynamic environment deduced from the hydrodynamic response of the borehole to recharge events. Borehole’s typology is illustrated in Fig. 4. Amplitude (in
metres), period of influence ‘‘P’’ (in days) and slope of recession curve ‘‘S’’ (in metres/day) were estimated from recession curves. x (m) correspond to the distance from the shore
in metres.

Borehole x (m) Typology Amplitude (m) Period of influence P (d) Slope of recession curve S (m/d) Type of hydrodynamic environment
B57 2400 D1 10.4 1.6 6.3 Conduits
B08 1065 D2 6.25 4.3 2.6 Conduits/large fissures
B59 2300 D3 5 >10.4 sa = 1.5/sb = 0.5/sc = 0.4 Fissures and matrix
B05 265 D4 3.3 4.5 0.9 Small fissures
B15 4490 D5 4.4 6.4 sa = 0.7/sb = 0.3 Matrix
B03 560 D6 3.75 9.1 sa = 0.7/sb = 0.1 Matrix
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2357

 B57 is located in a zone of high hydraulic diffusivity (D1) (Sec- 3.4. Discussion: Consequences in terms of intrinsic functioning of the
tion 3.1), representative of a conduit-dominated flow karst aquifer
environment.
 The amplitude (6.25 m) and the slope (S08 = 2.6 m/d) of the The hydraulic diffusivities calculated using the efficiency factor
recession curve at B08 were lower than B57 but the slope and the delay factor are largely similar (Tables 1 and 2). However,
remained relatively steep in comparison with those of other as noted previously, important differences have been observed
boreholes. B08 is located in a zone of relatively high hydraulic between the hydraulic diffusivity estimated on the basis of the
diffusivity (D2), representative of conduits or large fissures- tidal efficiency factor (Df) and hydraulic diffusivity calculated by
dominated flow environment. the delay equation (Dd) for B59, B15, B03, L01 and L02, showing
 B59 showed three different slopes on its recession curve rep- a ratio Df/Dd decreasing with hydraulic diffusivity values (Table 2).
resentative of an environment with different hydraulic diffu- From Eq. (1) it is assumed that head losses can affect the tidal effi-
sivities. The steepest one (S59a = 1.5 m/d) was however lower ciency factor (f) or the delay (d). For a confined homogenous aqui-
than at B08 (S08 = 2.6 m/d) and the two others were particu- fer where head losses are isotropic, Df/Dd value should be 1 (Ferris
larly low and almost equal (S59b = 0.5 and S59c = 0.4 m/d) sug- and Branch, 1952): a difference between Dd and Df at the same dis-
gesting that borehole hydrodynamics are characteristic of tance from the shore would indicate heterogeneity of the environ-
both fissures and matrix-dominated flow environments (D3). ment. Head losses in the matrix are supposed to be higher than in
These three successive slopes on its recession curve for which the conduits: at each tidal cycle, the tidal wave enters the aquifer
two are representative of a poorly hydraulic diffusive zone preferentially via the conduits that feed the matrix. Head losses
explain the particularly long observed period of influence P occur in the conduits but additional head losses arise when the
(P59 > 10.4 d). tidal signal propagates from the conduit through the matrix; in
 D4, D5 and D6 have been assigned to B05, B15 and B03 respec- turn, this leads to a larger dampening of the sinusoidal wave and
tively: despite lower amplitude observed at B05 than at B15 and a higher delay of the tidal signal recorded in the matrix than in
B03, it has been assumed its hydraulic diffusivity is higher as the conduits. However, in the matrix environment, the dampening
the slope of the recession curve (S05 = 0.9 m/d) was steeper than of the tidal signal amplitude seems comparatively much more
the ones at B15 (S15a = 0.7 and S15b = 0.3 m/d) and B03 (S03a = 0.7 important than the increase of the delay. As a consequence, in a
and S03b = 0.1 m/d), and its period of influence much shorter matrix environment, the relative change of the tidal efficiency fac-
(similar to the one of B08). B05 is therefore considered to be tor (f) differs from the relative change of the delay (d). In this case,
moderately connected with the karst drainage network (D4): two boreholes located at the same distance from the shore, one
B05 hydrodynamics would be considered as small fissures- drilled in a matrix-dominated flow environment (e.g. B59 located
dominated flow environment. 2300 m from the shore), and the other in conduit-dominated flow
 B15 and B03 both have smaller slopes on their recession curve environment (e.g. B57 located 2400 m from the shore), would have
which are representative of an environment with different the ratio fconduit/fmatrix > 1 and the ratio dconduit/dmatrix < 1: e.g., fB57/
hydraulic diffusivities. Their recession curves were however fB59 = 29 and dB57/dB59 = 0.4. There is a difference of two orders of
very flat and expanded (P15 = 6.4 and P03 = 9.1 d). B15 and B03 magnitude between these two ratios. Nonetheless, the Jacob–Ferris
are assumed to be poorly connected with the main drain karst equation accounts for these differences in estimated hydraulic dif-
system (D5 and D6) and representative of a matrix-dominated fusivities by attenuating them: e.g. ratio DfB57/DfB59 = 7.7 and
flow environment. DdB57/DdB59 = 6.5 (Dfconduit/Dfmatrix > 1 and Ddconduit/Ddmatrix > 1).
f and d are influenced by head losses but in a different way in
The outputs from the Jacob–Ferris equations are given in Table 2. matrix-dominated flow environment than in conduit-dominated
Boreholes and lakes are ranked from highest hydraulic diffusivities flow environment. Also, given that Dfconduit/Dfmatrix > Ddconduit/
(B57) (only Dd has been taken account; the reasons are explained Ddmatrix this confirms that the tidal efficiency factor (f) is more
in the discussion below) to lowest hydraulic diffusivities (B03). affected by head losses in matrix-dominated flow environment
Df and Dd decrease from B57 to B03 (except Df of L01 and L02) than the delay (d). Df thus appears to be more sensitive to the het-
but the values are different: Df/Dd ratio of 0.3, 0.47 and 0.65 was erogeneity of the aquifer than Dd. The latter may give a mean
found for boreholes B03, B05 and B59, respectively, and even lower hydraulic diffusivity value of the area around the borehole and
at L01 and L02 (0.15 and 0.14 respectively). When this ratio is the aquifer between the monitoring point and the shore by taking
lower than 1, Df appears to be under-estimated in comparison with into account all flow environments encountered (matrix, fissure
Dd (Table 2). and conduit); while Df may give a hydraulic diffusivity value more
representative of low flow dominated environment such matrix or/
and fissure flow environments (if they are encountered). Thus,
where Df is not affected by either matrix or fissure flow environ-
ments, Df and Dd determined from the hydraulic signal in bore-
Table 2 holes with high hydraulic diffusivities (e.g., B57 and B08
Hydrodynamic response of the borehole to tidal variations. Hydraulic diffusivities (Df supposed to be located in conduit-dominated flow environment)
and Dd in metres square/second) were calculated using tidal efficiency factor (f) and
delay factor (d, in hours).
are similar. The Dd value is considered as the hydraulic diffusivity
reference, because it is more representative of the overall hydraulic
Borehole x (m) f Df (m2/s) d (h) Dd (m2/s) Ratio Df/Dd diffusivity of the aquifer between the monitoring point and the
B57 2400 0.131 97.3 3h35 124 0.8 shore than Df.
B08 1065 0.4036 98.2 1h55 91.5 1.07 Strong propagation biases have been also noticed in unconfined
B59 2300 0.0045 12.7 8h45 19.1 0.65
L01 485 0.0637 2.8 2h20 18.2 0.15
aquifers (Jha et al., 2003) or layered aquifers when the lower layer is
L02 900 0.0052 2 3h55 14.6 0.14 more conductive than the upper layer (Trefry and Bekele, 2004) and
B05 265 0.4 5.8 1h25 12.4 0.47 both concluded that Df is more reliable than Dd. The Bell Harbour
B03 560 0.0407 2.1 3h30 7.1 0.3 aquifer may be compared to these aquifers as it is also a layered
Hydraulic diffusivity values in bold (Df) are assumed to be more reliable than aquifer; however, the presence of drainage system conduit/matrix
Dd values. Values at L01 and L02 are in italics as they are not comparable with makes it confined through the saturated conduits and unconfined
borehole values. in the surrounding matrix. Before it can reach a borehole located
2358 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Fig. 5. Hydrogeological profile of the catchment with depth of the boreholes and the loggers. Surface water levels (WL) deduced from WL measured on February 20, 2011 at
the 6 boreholes are in red for the high tide and in blue for the low tide. Dotted lines are the virtual surface WL between the WL measured at B57 which appears to be
disconnected from the main karst system and the bay. Vertical SpC profiles for B03, B05, B08 and B57 are in green and are plotted at the same scale as the depth of boreholes.
The maximum, minimum and average SpC values are given for each measurement point together with the vertical SpC profiles for four boreholes. Different colors are used for
each karst feature associated to the six boreholes. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

in unconfined matrix flow environment, the tidal flow firstly prop- of the lake cushions the tidal amplitude and may be added to the
agates through saturated conduits, before it can be expressed in the dampening effect of conduit/matrix on tidal amplitude explained
unconfined matrix. This mix of confined/unconfined systems and above. The Df inferred from the lakes’ hydrodynamics and reces-
the fact that head losses in conduit flow environment differ from sion curves are therefore not comparable with the Df inferred from
head losses in matrix flow environment, makes it difficult to com- borehole hydrodynamics. The delay does not appear to be affected
pare karst aquifers with more heterogeneous systems. In this case by the capacity of the lake and Dd estimated for the lakes can there-
Df is not necessarily more reliable than Dd. fore be compared with the Dd observed in the boreholes (Table 2).
From these observations, the Dd value is considered as repre- However, the calculated Dd value was considered to be insufficient
sentative of the aquifer diffusivity, as it varies more homoge- to assess the type of hydrodynamic environment for the two lakes.
neously than Df; furthermore, the ratio between Dd and Df allows Fig. 5 shows a hydrogeological profile of the catchment where
characterising the type of hydrodynamic environment at each depths of the boreholes and the loggers and their karst features
measurement point. The lower the Df/Dd ratio the greater the com- associated are shown. WLs at HTs and LTs on 20th of February
ponent of matrix flow environment is present in the aquifer system 2011 are shown in Fig. 5. This figure shows that WL in boreholes
between the point measurement and the shore. located close to the shore in an environment of a high to interme-
Consequently, when the Dd value is larger than the Df, this sug- diate hydraulic diffusivity (B05 and B08) can differ considerably
gests matrix flow dominates over conduit flow; in this case, bore- between HT and LT. These differences are less obvious in a matrix
hole hydrodynamics are assumed to be characteristic of a matrix- environment (B03), in areas well inland from the coast (B59) or
dominated flow environment, such as B03 (Table 2); B15 is consid- where lake effects occur (L01). The WL at B57 was low compared
ered to be characteristic of the same environment as B03 due to to the other boreholes and was only characteristic of a conduit flow
the similarity of their recession curves. Conversely, when Dd and environment after a recharge event: at this location, water from
Df values area similar, this means that water flows preferentially the recharge was flushed very rapidly through the conduits
through conduits and consequently that the connection (between towards the bay.
either the borehole or the lake) and the karst drainage network
corresponds to conduits (e.g., B57 and B08; Table 2). The fissure- 4. Spatial and temporal variability of the saltwater wedge
dominated flow environment is intermediate between conduit and
matrix-dominated flow environments and can be illustrated by 4.1. Spatial extent of the saltwater wedge
interim Df/Dd ratios such as those observed for B05 and B59. These
hydraulic diffusivity values are therefore consistent with the previ- The water chemical data collected in boreholes and lakes in the
ous interpretation based on the typology of hydrodynamic study site showed that the groundwater was characterised by a
responses to recharge events (Tables 1 and 2). From these results, Ca–HCO3 facies representative of water flowing within limestone
generalised karst features (matrix, fissure and conduit) have been except at L01, where the water contained a significantly high pro-
associated with each borehole and are shown in Fig. 5. portion of Na, Cl, SO4, Mg and K, i.e. representative of a seawater
For the two lakes L01 and L02, the Df/Dd ratios were 0.15 and signature (Fig. 6). Water sampled from B05, B08 and L02 are all
0.14 respectively, which can be attributed to head losses due to freshwater. Slightly higher concentrations of Na, Cl, SO4 and K were
exchanges between the karst aquifer and the lake; the capacity recorded in B05, B08 and L02 (Fig. 6 and Table 3) compared to the
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2359

Fig. 6. Schöeller–Barkaloff diagram of the water chemistry results for four boreholes and two lakes. Water samples were collected for major cation/anion analysis on March
21, 2011 during a relatively dry period. Before raw groundwater samples were collected from existing water taps connected directly to the pumps in boreholes, field
parameters were monitored to indicate geochemical stabilization (5–10 min).

Table 3
Concentration of Na, Cl, SO4, Mg and K measured at the three monitoring points
described in above, there is a strong hydraulic connection between
slightly influenced by seawater, at L01, which is strongly influenced by seawater, and
at B15, considered as Ca–HC0 the lake and the bay that may be due to the development of pref-
3 background water.
erential conduits in the weathered zone parallel to the fault core of
Monitoring point Na+ Cl SO2 Mg2+ K+
4
MacDermott’s fault (Bense et al., 2013).
(mg/l) (mg/l) (mg/l) (mg/l) (mg/l)
In the other boreholes and in L02, the average SpC varied
L01 1380 4874 848 136 56
between 600 and 720 lS/cm except at B15 (430 lS/cm) which is
B05 10 29 9 6 5
B08 8 21.4 9.4 6 3
situated furthest inland. Water chemistry data suggested that
L02 7 16.4 <5 3 2 B05, B08 and potentially L02 can possibly be slightly influenced
B15 7 14 <5 2 1 by seawater, and this was confirmed by maximum SpC values that
reached 1240 lS/cm at B05, 1290 lS/cm at B08 and 830 lS/cm at
L02 (the lower maximum value at L02 may be due to the dilution
Ca–HCO3 background water. The proportions of these ions are rep- effect of the lake). The values at B05 and B08 are two to three time
resentative of a seawater signature; indeed, Na/Cl ratios of all the higher than the SpC at B57, B59 and B15 (maximum values of 720,
samples follow the freshwater-seawater mixing line confirming 780 and 550 lS/cm respectively), considered as the background
that the main contribution of the salinity is due to a saltwater SpC in the area (Fig. 1c). B05 and B08 are assumed to be located
intrusion (Pulido-Leboeuf, 2004). in a variable mixing zone (Fig. 1c). The lack of water chemistry data
The minimum, maximum, and average SpC values recorded by from B03 and the largest recorded SpC value of 800 lS/cm from
the loggers at the monitoring locations are shown in Figs. 1c and this borehole make it difficult to estimate if B03 is influenced by
5 and confirm the inferences from the water chemical analysis: the saltwater wedge.
the high average SpC value at L01 (13,000 lS/cm) representative The vertical variation in the SpC recorded at B05 and B03 (Fig. 5)
of brackish water with variations up to 28,600 lS/cm, confirms made it possible to determine the vertical variation in the halo-
the presence of the saltwater wedge at L01 most of the time. As clines: 10 m below MSL, the increase in the SpC from 500 lS/cm
2360 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Table 4
Comparison of the influence of the tide in Bell Harbour bay for different amplitudes on SpC at B05, B08 and L01. The time lag is measured between the highest tidal level in the bay
and the highest SpC value recorded for each low/high tidal cycle. x (m) corresponds to the distance from the shore in metres. Tampl corresponds to the amplitude of the tide in Bell
Harbour. NT, ST and SST are respectively, neap tide, spring tide and strong spring tide periods.

Borehole/lake x (m) SpC variations (lS/cm)


SST: Tampl > 3.8 m ST: 2.5 < Tampl < 3.8 m NT: Tampl < 2.5 m
(time lag) (time lag) (time lag)
L01 485 1500–5000 0–1500 0
(5h to 5h40) (5h to 5h40 or no tidal signal) (no tidal signal)
B05 265 80–150 50–70 30
(2h) (11h) (11h)
B08 1065 40–60 30 20–30
(8h) (8h) (8h)

Fig. 7. (a) WL data at B05 as a function of the tidal water level; (b) SpC data at B05 as a function of the tidal water level. Hysteresis cycles are observed during the strong
spring tide (SST), spring tide (ST) and neap tide (NT) periods. Delay corresponds to the time lag between the high tide in the bay and the highest WL (or SpC) data observed for
one tidal cycle.

to 1000 lS/cm at B05 and from 500 lS/cm to 750 lS/cm at B03 (tidal amplitude between 2.5 m and 3.8 m) and neap tide (NT) peri-
indicates the presence of the mixing zone at B05 below 10 m below ods (tidal amplitude <2.5 m) as well as the time lag between the
MSL; despite an increase of 250 lS/cm at 10 m below MSL in B03, tidal signal in the bay and SpC signal are shown in Table 4.
the maximum SpC values are well below a saltwater influence sig- Tidal SpC variations at the three monitoring points increased
nal. Lower SpC values observed close to the base of B05 (850 lS/ with the amplitude of the tide (Table 4) suggesting the tidal ampli-
cm) are assumed to be due to a karst conduit where freshwater tude may have a direct effect on the positioning of the saltwater
flows occur (Fig. 5). The vertical salinity profile at B08 showed only wedge in the aquifer. The range of tidal SpC variations obtained
a slight increase in SpC close to the surface (1 m above MSL) from at L01 was by far the highest, confirming the high connectivity
450 to 550 lS/cm, probably due to the epikarst, which may allow between this lake and the bay. The time lag between the tidal wave
rapid drainage of recharge water into the borehole (Fig. 5). This and the SpC oscillation varied between 5 h and 5 h 40 min during
borehole is relatively shallow (depth is to 9 m below MSL) and SST and most of ST periods; note an extra time lag should be
it may not be deep enough to reflect the halocline. The profile at accounted for due to the inertia effect of the lake generated by
B57 exhibited a stable SpC value of 720 lS/cm throughout the ver- its large volume. Tidal SpC variations during NT and some ST peri-
tical profile, and this SpC is representative of typical groundwater ods may not be observed due to this inertia effect. Tidal SpC vari-
in the Burren karst aquifer not affected by saltwater wedge (Fig. 5). ations were higher in B05 than in B08, suggesting that the
impact of the tidal variation and of the saltwater wedge might be
4.2. Temporal variability of the extent of the saltwater wedge stronger in B05 than in B08, which is located farther from the
bay (Table 4). The time lag at B05 varied with the tidal amplitude
4.2.1. Temporal variation in saltwater wedge at high/low and neap/ unlike at B08 for which the time lag stayed constant for all tidal
spring tidal scales amplitude periods.
The impact of high tide/low tide fluctuations on the SpC was To identify the cause of the differences in time lags, the WL and
assessed for different tidal amplitude periods to evaluate how far SpC values at B05 versus the tide levels in Bell Harbour were ana-
the saltwater wedge moved at each tidal cycle. This analysis was lysed during one tidal cycle for each NT, ST and SST periods. Nota-
performed on B05, B08 and L01 data sets which are the only ble hysteresis cycles were observed caused by the remnant effect
monitoring points where tidal variations of SpC were observed. of the tidal level on the WL and SpC values, confirming the tidal
The range of the SpC variations observed during strong spring tide impact on WL and SpC values which increased and decreased with
(SST) periods (tidal amplitude >3.8 m), spring tide (ST) periods the rising and falling tide respectively (Fig. 7a and b). Differences in
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2361

Fig. 8. WL and SpC data at B05, tidal level in the bay and hydraulic gradient (i) are plotted from December 6 to 10, 2011. Periods of 2 h or 11 h of time lag between the HT in
the bay and the highest SpC value of a tidal cycle at B05 have been represented. Tampl corresponds to the tidal amplitude in the bay.

the time lag have been observed in hysteresis cycles of SpC Hysteresis cycles (SpC values versus the tide levels in the bay)
between different tidal amplitude periods while the time lag of were also observed at L01 and B08:
WL versus tidal level in Bell Harbour remains constant over time
(Fig. 7a and b). Hysteresis cycles of SpC are different according to – At L01, they were only observed during some ST and SST periods
the tide periods: when the tidal amplitude was sufficiently high; no time lags
were observed in NT and some ST periods between the tide in
– During ST or NT periods, SpC values started to increase slowly at the bay and SpC signal in the lake. When hysteresis cycles were
rising tide in B05: the SpC was highest at the beginning of the observed, the SpC increased rapidly at HT and decreased slowly
following rising tide, which explained the long time lag from the beginning of the rising tide.
observed between the high tide in the bay and this highest – At B08, hysteresis cycles revealed no significant differences for
value of SpC observed in the borehole (11 h). SpC values are the different periods of the tide confirming that the SpC at
higher during rising rather than falling tides which explains B08 was not influenced by tidal amplitude. SpC increased
the clockwise direction of the loops. The SpC decreased rapidly slowly during the falling tide and decreased slowly during the
at HT and continued to decrease slowly during the falling tide rising tide; the tidal influence was much more subtle at this
(Fig. 7b). monitoring point, due to its distance from the shore. Thus, hys-
– During the SST period, SpC values increased rapidly at HT resulting teresis cycles allow direct observation of variations in SpC
in the short time lag observed under these conditions (2 h). SpC depending on the stage and the amplitude of the tide.
values are lower during rising rather than falling tides and the loop
changes to an anticlockwise direction. SpC also decreased rapidly 4.2.2. Conditions for larger intrusion of the saltwater wedge into the
at low tide (LT) and values remained relatively stable during rising karst aquifer
and falling tides (Fig. 7b). Occasional, temporary and significant increases in SpC were
recorded at B05 (Fig. 3) and L01. During these temporary larger
These cycles highlighted the rate of variation in the SpC saltwater wedge intrusions, the range of SpC values and variations
depending on the stage of the tide (for a high/low tidal cycle); were different at B05 and L01:
for example, the rate of increase in SpC was slower than the rate
of decrease during ST or NT periods, and slower than the rate of  SpC values were higher than 1000 lS/cm at B05 while they
increase during SST periods. However, some hysteresis cycles were higher than 15,000 lS/cm at L01; the mixing zone
observed in SpC data from B05 showed a time lag of 11 h despite extended across the entire water column at B05 while L01
SST periods. was still located in the saltwater wedge but with higher SpC
By comparing SpC, WL, tide levels in the bay and hydraulic gra- values than normal.
dient (i) variations over time for several periods, all of the condi-  Variations of SpC were higher than 500 lS/cm at B05 whereas
tions required to get the shorter time lag (2 h) have been they were higher than 2000 lS/cm, and up to 15,000 lS/cm at
identified (Fig. 8): L01.

– tidal amplitude (Tampl) > 3.8 m (SST periods), As above, the conditions required to get more extensive intru-
– WL at HT > 3.8 m above MSL, without being affected by the sions of the saltwater wedge (and the mixing zone) have been
recharge. identified by comparing SpC, WL, tidal levels in the bay and
hydraulic gradient (i) variations over time for several periods at
If these conditions are met, the short time lag of 2 h remains B05 and L01:
between the tide in the bay and the tidal SpC variations even if
the Tampl in the bay decreases. However, even if the Tampl is high,  At B05, WL at LT should be <2 m above MSL and the tidal
if the WL decreases below 3.8 m at HT, the process stops and amplitude in the bay should be >4 m (WL in the bay at HT
the time lag reverts to 11 h (Fig. 8). is then >2 m above MSL) to start the process (Fig. 9). The
2362 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Fig. 9. Example of a temporary increase of SpC (red box) at B05. WL and SpC data at B05, tidal level in the bay and hydraulic gradient (i) are plotted from April 9 to 13, 2012.
Tampl corresponds to the tidal amplitude in the bay, iampl is the amplitude of the hydraulic gradient between the high and low tide. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 10. Example of a temporary increase of SpC (red box) at L01. WL and SpC data at L01, tidal level in the bay and hydraulic gradient (i) are plotted from April 27, 2012 to
May 18, 2012. Tampl corresponds to the tidal amplitude in the bay. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

lowest hydraulic gradient observed during a low/high tidal variations in the hydraulic gradient between the aquifer and the
cycle (imin) was between 0% and 0.5% and occurred close to bay could facilitate intrusion of the seawater and thus could drive
the HT (Fig. 9). the saltwater wedge further inland. The temporary increase of SpC
 At L01, WL at LT should be <1.8 m above MSL and the tidal ceased at B05 and L01 when the tidal amplitude in the bay
amplitude in the bay >3.5 m (WL in the bay at HT is then decreases. The latter seems to be the motor of the process by
>1.75 m above MSL) to start the process (Fig. 10). imin was then inverting the hydraulic gradient at L01 and by inducing strong
close to 0 or slightly negative and occurred at HT (Fig. 10): an amplitudes in the hydraulic gradient (iampl > 1%) at B05.
inversion of the hydraulic gradient occurred between the WL From the study of several temporary increases of SpC at L01 and
in the bay and L01 for a short period during the process. B05, it appears that the intensity of SpC increases varies with the
altitude of the WL at the monitoring point: the lower the WL, the
Additionally, a large difference in the hydraulic gradient higher the observed SpC value. The elevation of the WL in the bore-
between HT and LT (iampl) has been observed at B05 over a number hole or the lake appears to have a large influence on the range of
of HT/LT cycles before the increase of SpC (iampl > 1%; Fig. 9). These SpC increases.
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2363

Table 5
Order of influence of the saltwater wedge at the three monitoring points affected by the tide and summary of the results explaining the temporal extent of the saltwater wedge.
Tampl corresponds to the amplitude of the tide in the bay. imin is the lowest hydraulic gradient and iampl the highest amplitude of the hydraulic gradient of a high/low tidal cycle
observed during a saltwater intrusion.

Order of influence of saltwater wedge L01>> B05> B08


Geochemistry results SpC values in lS/ Brackish water Freshwater Freshwater
cm
(min, max) (930, 28,620) (320, 1480) (200, 1290)
Presence/influence of saltwater wedge Presence Temporary influence Temporary influence?
Tidal influence on SpC values Very high (up to 5000 lS/cm) Intermediate (up to 150 lS/cm) Low (up to 60 lS/cm)
Hysteresis cycles – SpC versus tidal level Only during ST and SST periods: quick and Stronger and quicker at SST Stable: slow at all tidal periods
in the bay very strong periods
(time lag) (5 h to 5 h 40 min) (11 h at NT, ST; 2 h at SST) (8 h)
SpC variation of saltwater intrusions in 2000 up to 15,000 500 –
lS/cm
(SpC values) (>15,000) (>1000) (>1000) during a period of recharge
Conditions for saltwater intrusion  WL at LT < 1.8 m  WL at LT < 2 m iampl too low (<0.2) or/and borehole not
 Tampl > 3.5 m  Tampl > 4 m enough deep
 imin 6 0  iampl > 1
 Good hydraulic diffusivity with the  Good hydraulic diffusivity
bay with the bay

4.3. Discussion on the temporal saltwater wedge variations  Larger saltwater wedge intrusions occurred:
– During SST periods when WL at LT was <1.8 m for L01 and
The maximal extent of the saltwater wedge and mixing zone <2 m for B05. B08 is too shallow to record the direct influ-
intrusions into the aquifer is shown in Fig. 1. All the monitoring ence of the saltwater wedge.
points affected by the saltwater wedge – tidally (L01, B05 and – When the hydraulic gradient was very low or slightly
B08), continuously (L01) or temporarily (B05) – are located less reversed; this phenomenon has been already observed in
than two kilometres from the shore (Fig. 1). A hierarchy of influ- other coastal karst aquifers (Bonacci and Roje-Bonacci,
ence of the saltwater wedge at the monitoring points has been 1997; Drogue, 1989; Fleury et al., 2008).
established from these observations and from the hysteresis cycles – Strong tidal amplitude seems to be the motor of the process
(Table 5). The saltwater wedge (> 10,000 lS/cm) was observed by driving seawater into the aquifer, while the position of the
only at L01 (SpC values were mainly brackish, i.e. > 10,000 lS/ WL in the aquifer may influence the intensity of the SpC increase.
cm). B05, and possibly B08, appear to be reflective of a temporary
mixing zone (1000 lS/cm < SpC < 10,000 lS/cm) rather than The temporal extent of the saltwater wedge is therefore
directly by the saltwater wedge, because of the low observed max- dependant on a balance between the recharge which controls the
imum SpC values (<1500 lS/cm) (Fig. 1). WL in the aquifer, and the tidal amplitude which controls the
The main results of this study are summarised in Table 5 sug- water level in the bay and the range of seawater tidal movements.
gesting that for L01, B05 and B08: Similar behaviours have been observed in a Turkish karst aquifer
where SGDs occur essentially through the fracture system
 Tidal influences on the SpC values increased with the tidal (Ozyurt, 2008).
amplitude. The monitoring points affected by the saltwater wedge are
 Hysteresis cycles of SpC versus tidal level in the bay have been those with a relatively high hydraulic diffusivity located less than
observed: 1000 m from the shore. L01 is assumed to be well connected with
– At L01: for ST and SST periods: the increase of SpC due to the the bay through MacDermott’s Fault and B05 and B08 are consid-
tide was rapid and large which confirms a high hydraulic dif- ered to be characteristic of fissured and conduit hydrodynamic
fusivity between the lake and the bay due almost certainly to environments, respectively (Table 1). The extent of the intrusion
the development of conduits along the large MacDermott’s of the saltwater wedge into the karst aquifer is therefore tightly
fault. linked with the geological structure (fault) of the aquifer and its
– At B05: the rate of variation in SpC with the tide was more intrinsic properties (hydraulic diffusivity).
rapid during some SST periods (time lag of 2 h) than during In order to study saltwater wedge variations in coastal karst
NT and ST periods (time lag of 11 h). Short time lags (2 h) aquifers subject to an oceanic climate and a high tidal range like
and larger SpC changes (>100 lS/cm) at B05 are related to at Bell Harbour, it is therefore recommended firstly to focus on
a large tidal amplitude in the bay when the WL in the bore- strong spring tide periods during periods of low or absent recharge
hole is above a given threshold during a period of no near the shore and preferentially in high hydraulic diffusivity zones.
recharge. When water level is >3.8 m MSL at high tide this Temporal variations of the saltwater wedge and the mixing
temporarily activates a more direct hydraulic path (probably zone have been proposed through a conceptual cross-section
a conduit) that allows seawater intrusion. The large ampli- model including B57, B08 and B05 (Fig. 11a and b). Only the most
tude of the tide may facilitate the seawater intrusion along pertinent monitoring points have been shown for clarity:
this path when the hydraulic gradient (between the bay
and B05) i < 0.5% at high tide (Fig. 9).  For a large flood event during NT period (Fig. 11a), conduits
– At B08: despite its high estimated hydraulic diffusivity along the MacDermott’s Fault connected to B57 are only
(Table 2), the hysteresis cycles of SpC versus tidal level in affected by the mixing zone near the shore (<500 m) from
the bay showed the increase in SpC due to the tide was slow which freshwater to brackish water flows through submarine
and relatively weak for any tidal periods. The tidal influence springs in the middle of the bay. The aquifer may only be
of the saltwater wedge at B08 is thus less important than at slightly affected by the saltwater wedge located at least 20 m
L01 and B05 (Table 5). below MSL at the shore.
2364 M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365

Fig. 11. Conceptual cross-section model of the freshwater-seawater interaction in the coastal aquifer of the Bell Harbour area; (a) location of the saltwater wedge (SWW) and
the mixing zone during a strong flood event and a neap tide (NT) period; (b) location of the SWW and the mixing zone during a dry hydrological and a strong spring tide (SST)
period: intrusion of SWW occurs preferentially in the conduit via submarine swallow hole and via a fissure-dominated flow environment of the intertidal spring. SpC isoclines
of 1000 lS/cm and 10,000 lS/cm have been represented in order to distinguish the freshwater zone of the aquifer, the mixing zone and the SWW.

 During low recharge and SST periods (Fig. 11b), the saltwater Mediterranean coast (e.g. extensive saline intrusions of more than
wedge may enter further into the aquifer, preferentially several kilometres from the coast in Crete in Greece (Arfib et al.,
through the fissured environment via intertidal springs which 2007); permanent brackish inland and/or submarine springs
become sinkholes, and through the conduit along the MacDer- (Fleury et al., 2007); or continuous inflow of seawater observed
mott’s Fault. Submarine springs in the middle of the bay behave simultaneously with outflow of brackish water (Drogue, 1989)).
like intertidal springs where seawater from the bay can intrude
in the aquifer. In this case, the saltwater wedge may enter in the 5. Conclusion
deep part of B05 located in a fissure-dominated flow environ-
ment at 265 m from the shore. The mixing zone would expand In this non-anthropogenic coastal karst aquifer with a strong
further and shallower in the aquifer and the halocline observed tidal influence the spatial variability of the saltwater wedge has
at 10 m below MSL at B05 may rise. It is assumed the mixing been found to be tightly linked to the geological structure and
zone could expand at 20 m below MSL at 1000 m from the intrinsic properties of the karst aquifer. Values of SpC (which
shore. Under these conditions it would not reach B08 (9 m reflected the extent of the saltwater wedge) were relatively low
below MSL) and its associated conduit(s) (Fig. 11b). and stable throughout the catchment, except at L01. A large fault
connecting the land and the bay underlies L01, explaining why
Submarine groundwater discharge in Bell Harbour occurs the lake is within the saltwater wedge. Three types of hydrody-
mostly in a fissure-dominated flow environment via intertidal namic environment (conduits, fissures and matrix) were defined
springs and through conduits developed along MacDermott’s Fault for each of the six boreholes by comparing two different methods:
only during large flood events (Fig. 11a). While the karst is well analysis of the recession curves of the WL data, and use of the Fer-
developed and the conduits open directly to the sea, the ground- ris equation to assess the impact of the tide on WL data. These
water discharge is mainly fresh (<1000 lS/cm) and the saltwater approaches, however, did not work for the data collected from
wedge remains close to the shore (<500 m). Even during low the lakes. Analysis of the impact of the tidal amplitude on SpC val-
recharge and SST periods, the mixing zone at 1000 m from the ues measured at the monitoring points confirmed the influence of
shore is only present at 20 m below MSL (Fig. 11b) which differs the tide on SpC values at B05, B08 and L01. Sudden temporary
greatly from other karst aquifers such as those found along the increases in SpC (>1000 lS/cm at B05 and >15,000 lS/cm at L01)
M. Perriquet et al. / Journal of Hydrology 519 (2014) 2350–2365 2365

in the aquifer during low recharge periods were observed at B05 Drew, D., Daly, D., 1993. Groundwater and karstification in mid-Galway, south
Mayo and north Clare. A joint Report: Department of Geography, Trinity College
and L01. They occurred when (i) the tidal amplitude was large
Dublin and Groundwater Section, Geological Survey of Ireland, Dublin, 86.
(>3.8–4 m), (ii) and the WL was low (<1.8–2 m). The temporal var- Geological Survey of Ireland.
iability of the extent of the saltwater wedge was shown to depend Drogue, C., 1989. Continuous inflow of seawater and outflow of brackish water in
on the balance between the influence of recharge and the influence the substratum of the karstic island of Cephalonia, Greece. J. Hydrol. 106 (1),
147–153.
of the tide on the hydrodynamics of the aquifer. The WL in the Drogue, C., Razack, M., Krivic, P., 1984. Survey of a coastal karstic aquifer by analysis
aquifer seems to control the intensity of the SpC increase while of the effect of the sea-tide: Example of the Kras of Slovenia, Yugoslavia.
the amplitude of the tide seems to be the motor of the process. Environ. Geol. 6 (2), 103–109.
Erskine, A., 1991. Effect of tidal fluctuation on a coastal aquifer in the UK. Ground
The WL in the aquifer has been almost always higher (or briefly Water 29 (4), 556–562.
equal at B05 or slighlty lower at L01) than the tidal level in the Ferguson, G., Gleeson, T., 2012. Vulnerability of coastal aquifers to groundwater use
bay and the saltwater wedge has never been observed farther than and climate change. Nature Climate Change 2 (5), 342–345.
Ferris, J., Branch, G.S.G.W., 1952. Cyclic fluctuations of water level as a basis for
one kilometre inland. The constant and relatively large recharge determining aquifer transmissibility. US Dept. of the Interior, Geological Survey,
typical of this coastal zone maintains aquifer water levels high Water Resources Division, Ground Water Branch.
enough to protect it against extensive saltwater intrusions. The Fleury, P., Bakalowicz, M., de Marsily, G., 2007. Submarine springs and coastal karst
aquifers: a review. J. Hydrol. 339 (1–2), 79–92.
presence of the saltwater wedge at L01 is assumed to be due to Fleury, P., Bakalowicz, M., Marsily, G., Cortes, J.M., 2008. Functioning of a coastal
the direct connection to the bay via conduits developed along a karstic system with a submarine outlet, in southern Spain. Hydrogeol. J. 16 (1),
large fault. This work highlighted the nature of the balance 75–85.
Gallagher, S.J., MacDermot, C.V., Somerville, I.D., Pracht, M., Sleeman, A.G., 2006.
between tidal levels and water levels in the aquifer and provides
Biostratigraphy, microfacies and depositional environments of upper Viséan
a starting point for assessing likely impacts in these coastal zones limestones from the Burren region, County Clare, Ireland. Geol. J.
in coming decades. Modest increases in recharge coupled with the 41 (1), 61–91.
predicted rises in sea levels in coming decades may lead to increase Gillespie, P., Walsh, J., Watterson, J., Bonson, C., Manzocchi, T., 2001. Scaling
relationships of joint and vein arrays from The Burren, Co., Clare, Ireland. J.
occurrences of inland flooding under non-extreme conditions and Struct. Geol. 23 (2–3), 183–201.
may lead to shifts in the type and extent of the saltwater wedge. Guo, H., Jiao, J.J., Li, H., 2010. Groundwater response to tidal fluctuation in a two-
Characterising coastal karst catchments and collection of sufficient zone aquifer. J. Hydrol. 381 (3–4), 364–371.
Henry, H.R., 1959. Salt intrusion into fresh-water aquifers. J. Geophys. Res. 64 (11),
and appropriate data from the terrestrial and the marine environ- 1911–1919.
ments will be critical to predicting and managing these dynamic Herzberg, D., 1901. Die wasserversorgung einiger nordseebäder. J. Gasbeleucht.
environments. Verw. Beleuchtungsarten Wasserversorg. 44 (815–819), 842–844.
Hubbert, M.K., 1940. The theory of ground-water motion. J. Geol., 785–944.
Jacob, C., 1950. Flow of groundwater. Chap, pp. 321.
Acknowledgments Jha, M.K., Kamii, Y., Chikamori, K., 2003. On the estimation of phreatic aquifer
parameters by the tidal response technique. Water Resour. Manage 17 (1), 69–
88.
Funding for this project was provided by the National Univer- Judd, B., Mullan, G., 1994. Poll Na Gcéim, County Clare Ireland. In: Proc. Univ. Bristol
sity of Ireland (NUI), Galway, Griffith Geoscience Award, which Spelacol. Sot.
was based on research grant-aided by the Department of Commu- Li, H., Jiao, J., 2002. Analytical solutions of tidal groundwater flow in coastal two-
aquifer system. Adv. Water Resour. 25 (4), 417–426.
nications, Energy and Natural Resources under the National Geo-
Li, L., Barry, D., Cunningham, C., Stagnitti, F., Parlange, J., 2000. A two-dimensional
science Programme 2007–2013. We gratefully acknowledge the analytical solution of groundwater responses to tidal loading in an estuary and
field and logistical support provided by John Coyne, Colin Brown, ocean. Adv. Water Resour. 23 (8), 825–833.
Yvonne O’Connell and the research students and EOS staff of NUIG. Maramathas, A., Maroulis, Z., Marinos-Kouris, D., 2003. Brackish karstic springs
model: application to almiros spring in crete. Ground Water 41 (5), 608–619.
Our thanks also go to the owners of the boreholes. Martin, J.B., Gulley, J., Spellman, P., 2012. Tidal pumping of water between
Bahamian blue holes, aquifers, and the ocean. J. Hydrol. 416–417, 28–38.
References Moore, Y., Stoessell, R., Easley, D., 1992. Fresh-water/sea-water relationship within
a ground-water flow system, northeastern coast of the Yucatan Peninsula.
Ground Water, 30.
Arfib, B., De Marsily, G., Ganoulis, J., 2002. Les sources karstiques côtières en
Ozyurt, N.N., 2008. Analysis of drivers governing temporal salinity and temperature
Méditerranée: étude des mécanismes de pollution saline de l’Almyros variations in groundwater discharge from Altug Submarine Karst Cave (Kas-
d’Héraklion (Crète), observations et modélisation. Bull. Soc. géol. France, 245– Turkey). Environ. Geol. 54 (4), 731–736.
253. Powers, J.G., Shevenell, L., 2000. Transmissivity estimates from well hydrographs in
Arfib, B., De Marsily, G., Ganoulis, J., 2007. Locating the zone of saline intrusion in a karst and fractured aquifers. Ground Water 38 (3), 361–369.
coastal karst aquifer using springflow data. Ground Water 45 (1), 28–35.
Pracht, M. et al., 2004. Geology of Galway Bay. A geological description to
Bakalowicz, M., 2005. Karst groundwater: a challenge for new resources. Hydrogeol. accompany the Bedrock Geology, 1(100,000).
J. 13 (1), 148–160.
Pulido-Leboeuf, P., 2004. Seawater intrusion and associated processes in a small
Bayari, C.S. et al., 2010. Submarine and coastal karstic groundwater discharges coastal complex aquifer (Castell de Ferro, Spain). Appl. Geochem. 19 (10), 1517–
along the southwestern Mediterranean coast of Turkey. Hydrogeol. J., 1–16. 1527.
Beddows, P.A., Smart, P.L., Whitaker, F.F., Smith, S.L., 2007. Decoupled fresh-saline
Rotzoll, K., Gingerich, S.B., Jenson, J.W., El-Kadi, A.I., 2013. Estimating hydraulic
groundwater circulation of a coastal carbonate aquifer: spatial patterns of properties from tidal attenuation in the Northern Guam Lens Aquifer, territory
temperature and specific electrical conductivity. J. Hydrol. 346 (1–2), 18–32.
of Guam, USA. Hydrogeol. J., 1–12.
Bense, V., Gleeson, T., Loveless, S., Bour, O., Scibek, J., 2013. Fault zone hydrogeology. Shevenell, L., 1996. Analysis of well hydrographs in a karst aquifer: estimates of
Earth Sci. Rev. 127, 171–192. specific yields and continuum transmissivities. J. Hydrol. 174 (3–4), 331–355.
Blavoux, B., Gilli, E., Rousset, C., 2004. Alimentation et origine de la salinité de la Simms, M., 2001. Exploring the limestone landscapes of the Burren and the Gort
source sous-marine de Port-Miou (Marseille-Cassis). Principale émergence d’un lowlands. Burrenkarst. com. Belfast.
réseau karstique hérité du Messinien. Comptes Rendus Geosci. 336 (6), 523–
Smith, A.J., 1999. Application of a Tidal Method for Estimating Aquifer Diffusivity:
533. Swan River. CSIRO Land and Water, Western Australia.
Bonacci, O., Roje-Bonacci, T., 1997. Sea water intrusion in coastal karst springs: Stringfield, V.T., LeGrand, H.E., 1971. Effects of karst features on circulation of water
example of the BlaÅ3=4 Spring (Croatia). Hydrol. Sci. J. 42 (1), 89–100. in carbonate rocks in coastal areas. J. Hydrol. 14 (2), 139–157.
Cooper, H., 1959. A hypothesis concerning the dynamic balance of fresh water and Townley, L.R., 1995. The response of aquifers to periodic forcing. Adv. Water Resour.
salt water in a coastal aquifer. J. Geophys. Res. 64 (4), 461–467.
18 (3), 125–146.
Dörfliger, N., Willwer, C., 2005. Introduction, Hydraulics and hydrodynamics in Trefry, M.G., 1999. Periodic forcing in composite aquifers. Adv. Water Resour. 22 (6),
Groundwater management of coastal karstic aquifers. European Cooperation in
645–656.
the Field of Technical Research: Environment, EUR21366EN (Cost Action 621), Trefry, M., Bekele, E., 2004. Structural characterization of an island aquifer via tidal
EC, Brussels. methods. Water Resour. Res. 40 (1), W01505.
Drew, D., 2001. Classic landforms of the Burren karst. Geographical Association in
Tulipano, L., 2005. Groundwater management of coastal karstic aquifers. European
Conjunction With the British Geomorpological Research Group. Cooperation in the Field of Technical Research: Environment, EUR21366EN
Drew, D., 2003. The hydrology of the Burren and of Clare and Galway lowlands.
(Cost Action 621), EC, Brussels.
Caves of County Clare and South Galway. University of Bristol Spelaeological Walsh, S., 2012. A Summary of Climate Averages for Ireland. Met Eireann, Dublin.
Society, Bristol, United Kingdom, pp. 31–42.

You might also like