You are on page 1of 11

Environmental Modelling and Software 114 (2019) 129–139

Contents lists available at ScienceDirect

Environmental Modelling & Software


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

A coupled surface-subsurface hydrologic model to assess groundwater flood T


risk spatially and temporally
Xuan Yua, Daniel Moraetisb, Nikolaos P. Nikolaidisc, Bailing Lid, Christopher Duffye,
Bingjun Liua,∗
a
School of Civil Engineering, Sun Yat-sen University, Guangzhou, 510275, China
b
College of Science, Earth Science Department, Sultan Qaboos University, P.O. Box 36, Muscat, 123, Oman
c
School of Environmental Engineering, Technical University of Crete, University Campus, 73100, Chania, Greece
d
Hydrological Sciences Laboratory, NASA Goddard Space Flight Center, Greenbelt, 20771, MD, USA
e
Department of Civil and Environmental Engineering, Pennsylvania State University, University Park, PA, 16802, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Floods introduced by rainfall events are responsible for tremendous economic and property losses. It is important
Groundwater flooding to delineate flood prone areas to minimize flood damages and make mitigation plans. Recently, there has been
Coupled surface-subsurface model recognition of the need to understand the risk from groundwater flooding caused by the emergence of
PIHM groundwater at the ground surface. Mapping groundwater flood risk is more challenging compared to river
Risk assessment
flooding, especially in karst systems, where subsurface preferential flowpaths can affect groundwater flow. We
Critical zone observatory
Koiliaris river basin
developed a coupled surface-subsurface modelling framework resolving spatial information and hydrologic data
to assess the groundwater flood risk at a karstic watershed: Koiliaris River Basin, Greece. The simulated
groundwater table was used to delineate groundwater flooding. Modelled results revealed the role of faults in
groundwater flooding generation. We anticipate the coupled surface-subsurface approach to be a starting point
for more sophisticated flooding risk assessment, including magnitude and temporal duration of groundwater
flooding.

1. Introduction structures could significantly improve overbank flows. These studies


yield important insights into adequate design and development of flood
Flooding is one of the most frequently occurring natural disasters management strategies for river management. However, the previous
that cause significant economic damages and affect people's life around approach neglect a significant factor, groundwater flood, which can be
the world (Jonkman et al., 2008; UNISDR, 2015; Kundzewicz et al., critical in flood management, especially in karstic watersheds (Finch
2014; Huntingford et al., 2014). Further more, it has now been widely et al., 2004; Macdonald et al., 2008).
noticed that the risk of life and economic losses caused by flood may Groundwater flooding (Fig. 1) happens when groundwater tables
increase due to social-development, land subsidence, and climate rise above the land surface away from perennial river channels or in the
change (Arnell and Gosling, 2016; Hirabayashi et al., 2013; Jongman floodplain before the river goes out of bank (Finch et al., 2004;
et al., 2015; Ward et al., 2017; Archfield et al., 2016). It is urgent to Macdonald et al., 2008). Groundwater flooding receives less attention
delineate flood prone areas under changing environment and to reduce compared to river flooding because it happens less often with less se-
or prevent the detrimental effects of flood waters. To battle the problem verity (e.g. Ascott et al., 2017; Morris et al., 2018). However the
of delineating flood prone areas most of the research focus on the study management practices to prevent groundwater flooding are limited
of parameters of surface river flooding (Fig. 1, Yamazaki et al., 2011; Li compared to river flooding, which can be reduced by building en-
and Duffy, 2011). In our area of interest, Vozinaki et al. (2015) and gineering structures (embankments, walls, dams, etc.). Once it happens,
Kourgialas and Karatzas (2013) estimated crop loss due to overland groundwater flooding will take time to dissipate because groundwater
flooding by considering several parameters like flood depth, flow ve- moves much more slowly than surface water (Pinault et al., 2005). Such
locity, and vegetation types. Costabile and Macchione (2015) showed long-lasting groundwater flood can cause not only significant damage
that implementing high-resolution topographic data and man-made to infrastructure (Abboud et al., 2017) but also more severe river


Corresponding author.
E-mail address: liubj@mail.sysu.edu.cn (B. Liu).

https://doi.org/10.1016/j.envsoft.2019.01.008
Received 18 September 2018; Received in revised form 27 December 2018; Accepted 12 January 2019
Available online 23 January 2019
1364-8152/ © 2019 Elsevier Ltd. All rights reserved.
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

and land use. The novelty of this work is to understand groundwater


flooding through spatial and temporal dynamics of groundwater table
simulated by coupled surface-subsurface models. We anticipate this
work to be a starting point for more sophisticated flooding risk as-
sessment. The incorporation of geological, climate and soil data into an
affordable computation load could explain spatial vulnerability as-
sessment of magnitude and temporal duration of groundwater flooding.

2. Data and methods

2.1. Koiliaris River Basin

The Koiliaris River Basin (KRB) is a Critical Zone Observatory


(Menon et al., 2014) that represents severely degraded soils due to
heavy agricultural impact over many centuries. The 130 km2 watershed
is located 15 km east of Chania, Greece (Fig. 2a). Chania has a popu-
lation of 53,910 and is the second largest city of Crete. The elevations of
KRB range from 0 to 2041 m above mean sea level and the slopes range
from 0° to 62°. Climate in the watershed is Mediterranean and char-
acterized by a separation of seasons. The average rainfall is 652 mm of
which 20% in the dry season with an average temperature of 24˚ Celsius
Fig. 1. Illustration of river flooding and groundwater flooding at a watershed.
while in the wet season 80% of the average rainfall occurs with an
average temperature of 13˚ Celsius (Stamati et al., 2013). The land
flooding (Huntingford et al., 2014). Mapping groundwater flood prone crops include citrus trees and olive trees which are cultivated
area is more challenging compared to river flooding, due to many throughout the year, while vegetables and wheat are cultivated con-
controlling factors and spatial uncertainties throughout the whole wa- tinuously in alterations appropriate for each period (Moraetis et al.,
tershed, including topography, land use, geology, etc. Kourgialas and 2015). We used hourly climate data (including precipitation, tem-
Karatzas (2014) integrated 6 factors: flow accumulation, slope, land perature, relative humidity, wind velocity, solar radiation, and vapour
use, rainfall intensity, geology, and elevation to map flood hazard pressure) from three meteorological stations (i.e. Psychro Pigadi, Sa-
across the Koiliaris River Basin (KRB), Greece. The static GIS informa- monas, and Kalyves. Fig. 2b). Four soil types (Fig. 2c) covers the KRB,
tion of vegetation, topography, and soil can provide valuable in- with the majority is Eutric leptosols (56%) and Calcaric regosols (24%).
formation for locations with potential flooding risk. For example the Leptosol and regosol soils are expecting to show some differentiation in
vegetation cover is considerably affecting the response of the water hydrologic properties such as those mentioned by Kirchen et al. (2017)
infiltrating depth in the soil (Rossi et al., 2018). Furthermore, soil type with leptosols water holding capacity to be the lower. The watershed
like e.g. silt loam can be water repellent and delay water infiltration (Li consists of 13 land cover types (Fig. 2d) based on the Corine Land Cover
et al., 2018). Thus, temporal dynamics of groundwater flooding is still types (Bossard et al., 2000), and 28% of the watershed is covered by
unclear, which requires continuous hydrologic modelling of ground- natural grassland (pink colour in Fig. 2d), while 23% of the total area is
water table elevation. covered by sclerophyllous vegetation (yellow colour in Fig. 2d).
Groundwater flooding is rarely modelled spatially throughout a
watershed. Especially, in areas with complex geological karst setting 2.2. Model description
like this in KRB, groundwater flooding is prone to be developed due to
the combination of low storage, channelized preferential flow, and A coupled surface-subsurface hydrologic model, Penn State
rapid lateral inflow from steep uphill area (Naughton et al., 2012; Integrated Hydrologic Model (PIHM), was selected to perform hydro-
Upton & Jackson, 2011). Butscher and Huggenberger (2007) used a logic simulations of streamflow and groundwater dynamics. PIHM
conceptual karst model to approximate subsurface flow for hydrologic (Kumar, 2009; Qu and Duffy, 2007) simulates the hydrological cycle
simulation, and found that 32% of the springs are discharging from an including interception, throughfall, infiltration, recharge, evapo-
uphill area with slopes greater than 20°. Nikolaidis et al. (2013) coupled transpiration, overland flow, unsaturated soil water, groundwater flow,
a distributed hydrologic model and a conceptual karst model for in- and channel routing, in a fully coupled scheme. Evapotranspiration
tegrated water management at KRB. Explicit representation of karsts in (ET) is calculated using the Penman-Monteith approach adapted from
a hydrologic model will be helpful to examine hydrologic links between Noah_LSM (Chen and Dudhia, 2001). Overland flow is described in 2-D
karst and the rest part of the watershed (e.g. Shoemaker et al., 2008; estimated of St. Venant equations. PIHM uses diffusive wave approx-
Reimann et al., 2011; Rugel et al., 2016), improve understanding of imation for overland flow and streamflow. The model conceptualizes
groundwater flooding, and achieve sophisticated water management. each subsurface column into two parts, unsaturated and saturated
In this study, we examine the spatial and temporal response of layers. Water balance equation based on hydraulic head, which is
groundwater conditions of a karst watershed to assess the risks of capable of representing the state in both unsaturated and saturated
groundwater flooding. Our first objective is to test the ability of a zones. While only considering vertical moisture movement in the un-
coupled surface-subsurface hydrologic model to capture groundwater saturated zone, the model simulates 2-D (lateral) groundwater flow in
flooding. In the model, the spatial heterogeneity of the watersheds is unconfined aquifer based on the Dupuit approximation without ad-
explicitly resolved in the watershed responses, which includes vegeta- dressing the karstic conduits network and the diffuse flow system in
tion, geology, and faults. The second objective is to understand how deep (e.g. Butscher and Huggenberger, 2007). Interception and snow-
faults development would affect the groundwater flooding processes. melt are obtained using a Bucket Model and a Temperature Index
The methodological steps of the study are (1) application of a coupled Model, respectively. Horizontally, the modelling domain is decomposed
surface-subsurface, spatially distributed hydrologic model to delineate into Delaunay triangles. The governing equation for each process is
spatial distribution and temporal dynamics of groundwater flooding; discretized on the triangles from canopy to bedrock, forming partial
(2) examination of the sensitivity of fault conductivity on flood risk; differential equations (PDEs). PIHM uses a semi-discrete finite volume
and (3) calculate agricultural loses based on groundwater flooding area formulation for solving the set of the coupled PDEs, resulting in a

130
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Fig. 2. a) KRB (shown as red area) in Greece (shown as dark gray), b) historic groundwater flooding sites modified from (Kourgialas and Karatzas, 2011), c) soil map
and d) land cover. Code of land cover is explained in Table S1. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web
version of this article.)

system of ordinary differential equations (ODEs) representing all pro- dynamics of surface water and groundwater interactions (e.g. Yu et al.,
cesses of each triangle. The ODE system is solved using the CVODE 2015a; Zhang et al., 2018; Zhang et al., 2019), hence allowing spatial
implicit solver (Cohen and Hindmarsh, 1996). Detailed descriptions of and temporal delineation of floods and 3) geologic faults with sub-
the modelling theory and mathematical formulation can be found at the stantial impacts on groundwater flow can be explicitly represented in
PIHM website (http://www.pihm.psu.edu/) and associated publication the model as large scale preferential flow through high permeable re-
(Kumar, 2009; Qu, 2005; Qu and Duffy, 2007). gions.
In this study, we applied the coupled surface-subsurface hydrologic
modelling method to explicitly simulate groundwater table elevation
2.3. Model parameterization, calibration
throughout KRB for groundwater flood assessment. There are several
advantages for PIHM modelling for this study: 1) the easy setup for the
PIHMgis (Bhatt et al., 2014) was used to parameterize the model
physical properties of the surface and subsurface layers including faults
domain using the aforementioned datasets (i.e. DEM, meteorological,
by PIHMgis (Bhatt et al., 2014) 2) the model captures fully coupled
soil, and land cover data). The KRB and the extended karstic area were

131
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

discretized into 914 triangles, and the river was discretized to 92 linear streamflow variation between observed and modelled results were in
segments (Fig. S1). The extended karstic area has been assigned by reasonable agreement and are considered acceptable. Similar modelled
others as contributing area in the water budget of our study area and observed runoff ratios of 0.60 and 0.62, respectively, also indicated
(Kourgialas et al., 2010). We estimated subsurface hydrologic para- reliable partitioning of the water budget. Furthermore, 8-day average
meters using texture data and pedotransfer functions (Wösten et al., ET (ORNL DAAC, 2017) of this watershed was in good agreement with
1999). The pedotransfer functions used are generally applied for the MODIS data (Fig. 3b).
different soil textures within Europe (Wösten et al., 1999) and they are
not site specific. Therefore, it is suitable for a regional study such as the 2.4. Delineation of groundwater flooding map
present work. The distribution of major faults was implemented ac-
cording to the description in Tataris and Christodoulou (1969). An Since the focus of this paper is on the dynamics of groundwater
analysis of the tectonic phases in Crete has revealed different phases of response, further confidence in the model result in this context was
deformation which have definitely influence the faults hydraulic built by evaluating the ability of the model to simulate the historic
properties. For example the mega- and meso-folds with fold axis W-E flooding locations. We extracted daily groundwater table depths of each
and NeS which were created during the beginning of the orogenesis in finite volume (i.e. triangle of the modelling mesh) from the calibrated
Crete. Perpendicular to these folds, dip-slip W-E and NeS striking faults PIHM simulation results. Groundwater flooding locations and areas
were created during the extensional phase in Crete (Fassoulas et al., were delineated by analysing the temporal dynamics of PIHM simulated
2004). The previous faults considered the oldest generation of faults groundwater tables according to Eq. (2).
which led to intense karstification with vast and deep sinkholes and
shafts like Gourgoutakas and Liontari Caves (Moraetis et al., 2010; ⎧ High MaximumGW > 0
Groundwater flood vulnerability = Moderate SkewnessRGW < 0
Adamopoulos, 2013). These W-E and NeS are of high hydraulic con- ⎨
nectivity compare to the other fault zone of NW-SE striking direction ⎩ Low SkewnessRGW ≥ 0 (1)
which are much younger with less karstification and extensive pre-
GW − MinimumGW
servation of faults’ slickenlines (Fassoulas, 2001; Caputo et al., 2010). RGW =
MaximumGW − MinimumGW (2)
Both these generations of faults exist in Koiliaris. Hydraulic con-
ductivity is simulated as being 10 times in the oldest faults (W-E and Groundwater flooding may happen under two conditions: 1)
NeS) and 2 times in the youngest faults (NW-SE) (Fig. 2c). The hy- groundwater table rises above land surface during the simulation
draulic conductivity is selected higher in the older faults since the in- period, and 2) groundwater table is shallow during the wet season and
tensely karstified areas are along the W-E faults. Kartsic features, the area can be potentially flooded during rainfall events. We used
especially sinkholes have been observed also in other areas along ex- maximum groundwater table elevation (MaximumGW) to quantify the
tensional faults like in our case (Prasad et al. 2017). These karstified first situation, where groundwater table height was above surface.
areas are largely affecting the water flow and the hydraulic con- These places are of high groundwater flooding risk. The second situa-
ductivities along the faults. The hydraulic conductivity can change from tion was examined by the frequency of groundwater getting close to the
10−7-10−6 (m/s) in massive limestones (Kovács, 2003) to 10−5 in the surface. We rescaled the groundwater table elevation to 0 to 1 as re-
epikarst (Hartmann et al., 2012) and to 10−3 (m/s) in karstified faults lative groundwater table elevation (RGW, Eq. (2)) and then analysed
(PolomčiĆ et al., 2013). Finally, the past speleological expeditions in the temporal dynamics using the skewness of the rescaled groundwater
the deep sinkholes in our study area shows chaotic vertical shafts from table elevations (SkewnessRGW) (Fig. 4). If the skewness is less than 0, it
masl (meters above sea level) of 1500 m down to 300 masl. The karstic means the groundwater table is often close to the land surface, which
shafts correspond to what Moraetis et al. (2010) and Nikolaidis et al. can potentially result in groundwater flooding (i.e. moderate ground-
(2013) presented as fast responding karstic system or upper reservoir. water flooding risk). If the skewness is positive, the groundwater is
The land cover parameters were obtained from monthly GLDAS often far below the land surface, hence less likely to cause groundwater
vegetation parameters (Rodell et al., 2004) and mapped according to flooding (i.e. low groundwater flooding risk).
Table S1. The depth of groundwater flow was estimated based on the Time lag between river flooding peak and groundwater table height
depth of the groundwater in the low land areas approximately 20 m peak was used to identify the temporal dynamics of groundwater
below the land surface. The depth is consistent with the depth of the flooding. We selected 8 events during the 3 years. In each event, the
epikarst around the world (Jones, 2013). The value was used uniformly time of peak streamflow and peak groundwater table height was ana-
throughout the watershed. lysed.
Our simulations were performed for 3 years (July 2007–July 2010, We varied the fault hydraulic conductivity for sensitivity analysis to
when flooding was severe, Fig. S2) using PIHM v2.2 (2015). Initial assess the impacts of faults on groundwater flooding, since the fault
condition was achieved using spinup of two previous years (July hydraulic conductivity can vary significantly based on field experi-
2005–June 2007). Initialized groundwater tables were below the land ences. The hydraulic conductivity of faults was changed at different
surface and above bedrock (20 m below the land surface). Observed magnitudes to simulate karst development scenarios. We analysed both
daily streamflow data at Stylos Spring (Fig. 2b) and MODIS remotely spatial locations of groundwater flooding and the time delays of
sensed ET dataset (MOD16) (Running et al., 2017) were used for model groundwater flooding.
calibration. According to the hydrologic characteristics of parameter,
we calibrated two groups separately: event and seasonal scale groups 2.5. Cost estimation of agricultural losses due to groundwater flooding
(Table 1, Yu et al., 2013). The event scale parameters were optimized
through an evolutionary algorithm (Yu et al., 2013). We selected The estimation of groundwater flooding cost was based on the cal-
streamflow observation data from Feb 25, 2009 to March 4, 2009, in culations demonstrated by Vozinaki et al., (2015). The designated area
which highest streamflow happened, for event scale parameter cali- of high risk was ascribed with the land uses data from Corine database
bration. The seasonal scale parameters were tuned based on evapo- (European Environment Agency, 2004). The intersection of land cover
transpiration (ORNL DAAC, 2017) using observations from July 1, 2008 and high risk flooding areas produced the percentage of different land
to June 30, 2009. The rest of the simulation was used for validation. We uses in the high risk flooding areas. The agricultural flood damage was
used the relative error (E), Pearson product-moment correlation coef- calculated by the total area of each crop type under high flooding risk,
ficient (R), and Nash-Sutcliffe coefficient of efficiency (NSE) (Nash and the weight yield per area and the cost per unit weight of yield. We
Sutcliffe, 1970) to evaluate the model performances (Table 2). retrieved the weight yield per area and the cost per unit weight of yield
Streamflow calibration results are shown in Fig. 3a. The dynamics of from Vozinaki et al. (2015). In the previous study the depth and flow

132
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Table 1
PIHM Model Parameter Attributes and calibration factors.
Parameter description Unit Initial value Calibrated factor Group

Matrix conductivity (horizontal) (m/day) 4.2–6.8 0.3 Event scale group


Matrix conductivity (vertical) (m/day) 0.4–0.7 0.7
Macropore conductivity (horizontal) (m/day) 42–68 3.2
Macropore conductivity (vertical) (m/day) 4–7 6.9
Infiltration rate (m/day) 0.7–1.5 5.2
Macropore depth (m) 0.1 0.5
Porosity (m3/m3) 0.49–0.53 0.5
Van Genuchten parameter α (m−1) 2.9–3.5 0.3
Van Genuchten parameter β 1.2–1.3 2.6
River Mannings roughness (day/m1/3) 4.63 × 10−7 3.2
River bed conductivity (horizontal) (m/day) 0.6 1.8
River bed conductivity (vertical) (m/day) 0.1 1
Root zone depth (m) 0.6 1.5 Seasonal scale group
Vegetation fraction (m2/m2) 0.1–0.7 1.1
Field capacity (m3/m3) 0.8 × porosity 1.1
Wilting point (m3/m3) 0.01 1
Maximum interception storage capacity factor (m) 2.0 × 10−4 1.6

velocity of the flood were mostly 0–1.2 m and 0–0.3 m/s respectively. one day, the peaks of groundwater flooding delayed longer, and showed
Thus, our estimations are based on these calculations and we adopted diverse spatial patterns (events 2, 3, 4, and 6, Fig. 7). Apart from the
the flood damage cost standard deviation given therein. time lags in the emergence of the event, there were also spatio-temporal
changes in the decay of the phenomena (Figs. 6 and 7). Such in-
3. Results formation is critical to flood mitigation.

3.1. Spatial distribution of groundwater flooding


3.3. Sensitivity to hydraulic conductivity of faults
Groundwater flooding map was created as three types of risk: high,
Increase fault hydraulic conductivity created wider areas of
moderate, and low with areas of 100, 21, and 8.7 km2, respectively
groundwater flooding (Fig. 8). The delineated groundwater flooding
(Fig. 5). The spatial distribution of groundwater flooding was located
areas were 20.3, 29.4, and 32.0 km2 for low, medium, and high K of
around low elevation regions and faults. This was probably because
faults. The high flooding risk areas were 8.0, 8.7, and 9.2 km2 for low
groundwater recharge from sinkholes in mountainous area could
medium, and high K of faults. The moderate flooding risk areas were
quickly discharge through highly fractured faults. Eleven of the 30
12.0, 20.7, and 22.8 km2 for low medium, and high K of faults.
historic flooded sites were captured by high flood risk, and 14 historic
Increase fault hydraulic conductivity reduced responding time of
flooded sites were captured by moderate flood risk. The rest 5 historic
groundwater flooding. The average time lags between groundwater
sites not captured by high or moderate risks were located next to the
flooding and river flooding were 10.4, 8.9, and 6.7 h for low medium,
riverbanks. Previous studies (Vozinaki et al., 2015; Kourgialas and
and high K of faults (Fig. 9).
Karatzas 2014) used high resolution modelling of riverine areas to
capture flooding locations. Since, our model grid is an unstructured
mesh, it will be easy to increase spatial resolution around the river. 3.4. Damage cost calculation

3.2. Temporal dynamics of groundwater flooding In our study the total area of high flooding risk is 7% of the total
watershed and 23% of the cultivated area. The land uses distribution in
Groundwater flooding was generally less abrupt than river flooding. the high-risk areas is given in Table 4. The olive grooves have the
In all of the 9 selected rainfall-runoff events (Fig. 6, Table 3), there was highest coverage in the high-risk areas (39%), while agricultural land
a significant delay of the peak between groundwater flooding and river has the second highest coverage (24%) (Table 4). The highest cost of
flooding indicated by model simulation. The average time lag (Fig. 6b) damage derives from olive groves and the second comes from fruit trees
varied from 4.4 to 14.2 h with a mean of 8.9 h. (Table 4). The total damage cost is 2.1 × 106 €. The calculation of the
Time lags between groundwater flooding and river flooding (i.e. cost followed the methodology of Vozinaki et al., (2015), where the
peak flow in the river) showed spatial heterogeneity. The time lags synthetic damage curves were used. Even the number of data was not
were generally small around the faults, and large in the flood plain (i.e. large in the previous study, it was considered representative for our
the northeast part of KRB). This was due to the high hydraulic con- study area since a questionnaire survey had be conducted specifically
ductivity of the faults. When the precipitation event lasted more than for Koiliaris watershed.

Table 2
Model performance of daily streamflow.
E (%) R NSE

Calibration (2008/7/1–2009/6/30) 2.3 0.8 0.61


Validation (2007/7/1–2008/6/30, 2009/7/1–2010/6/30) 3.7 0.77 0.56
Equationa S −O ∑i (Oi − O¯ )(Si − S¯ ) ∑i (Oi − Si)2
× 100% 1−
O
∑i (Oi − O¯ )2 ∑i (Si − S¯ )2 ∑i (Oi − O¯ )2

Ideal value 0 1 1
Range -∞,+∞ −1, 1 -∞,1

a
S and O represent simulated value and observed value respectively.

133
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Fig. 3. Comparison of modelled and observed (a) discharge at Stylos spring, (b) daily evapotranspiration. In subplot (a) gray area denotes the calibration period,
while white area denotes the validation area.

Fig. 4. Histogram of relative groundwater table elevation a) positive skewness and b) negative skewness.

134
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Fig. 5. Groundwater flooding risk map for the Koiliaris River basin. Red dot
indicates historic groundwater flooding sites. (For interpretation of the refer-
ences to colour in this figure legend, the reader is referred to the Web version of
this article.)

4. Discussion

4.1. Data interpretation of coupled surface-subsurface modelling

The advantage of coupled surface-subsurface modelling is to ex-


plicitly simulate the head above ground. Our simulation is distributed
throughout the watershed and processes of infiltration and discharge
are dynamically considered through coupled surface-surface equations.
Therefore, identification of groundwater flooding according to model-
ling results is relatively easy. Groundwater models (e.g. MODFLOW) are
often used to identify location of groundwater spring (e.g. Reed et al.,
2016; Gallegos et al., 2013). Similarly, groundwater flooding can be
simulated through groundwater flow modelling to avoid realistic in-
teractions between surface water and groundwater. However, such
strategy may have limitation around riverbank. We note that Drainage
Pack (Harbaugh, 2005) of MODFLOW can be a potential solution for
groundwater flooding. It will helpful to compare our results with
groundwater modelling results.
The purpose of coupled surface-subsurface modelling is to improve
understanding on hydrologic processes rather than runoff prediction
efficiency. Previous watershed modelling studies obtained higher NSE
of runoff (e.g. 0.91 and 0.87 in Kourgialas et al., 2010; 0.62 in
Nikolaidis et al., 2013). Ideally, coupled surface-subsurface modelling
has more parameters and gives better fit. Practically, the parameters are
bounded by physical meanings and the calibration of coupled surface-
subsurface models is usually difficult (Yu et al., 2013). In addition,
calibration to different types of data may result poorer fit as the para-
meters try to serve different dynamics that dominate different data sets.
Both uncertainties in spatial parameters and temporal forcing can
consistently create model discrepancy at the stream gauge. It is ac-
ceptable that our simulated runoff was degraded comparing to previous
modelling studies. Notably, the application of coupled surface-subsur-
face hydrologic models is not only on streamflow dynamics (Chen et al.,
2015; Seo et al., 2016; Zhang et al., 2017) but also on spatial and
temporal dynamics of soil moisture (Shi et al., 2015), groundwater Fig. 6. Temporal dynamics of groundwater flooding and river flooding in-
table elevation (Yu et al., 2016), snow accumulation (Kumar et al., dicated by model simulation. a) Rainfall runoff processes, solid blue line re-
2013; Yu et al., 2015b), which can create continuous spatial products of presents simulated river runoff, dashed black line represents groundwater table
elevation relative to the land surface, red bar represents rainfall rate. b) Peak
hydrologic and ecologic variables (Yu et al., 2015a; Condon et al.,
time of river flooding and groundwater flooding. Time starts from previous no-
2015; Shi et al., 2018).
precipitation date. (For interpretation of the references to colour in this figure
Interpretation of coupled surface-subsurface modelling results re- legend, the reader is referred to the Web version of this article.)
quires combination of localized information and targeted problem.

135
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Table 3
Nine largest precipitation events in KRB during 2007/7/1–2010/6/30.
No. Event date Return Peak of river Peak of groundwater
period (year) flooding (hour) flooding (hour)

Mean Std. Mean Std.

1 2008/1/24 <1 22.3 4.8 26.7 7.0


2 2008/2/11 2 9.2 5.5 23.4 10.7
3 2008/2/16 1 20.0 0.9 27.4 7.9
4 2009/1/10 <1 26.7 2.2 35.6 6.9
5 2009/2/25 5 24.4 0.5 31.7 7.0
6 2009/4/5 <1 14.2 1.7 25.3 8.0
7 2009/5/4 <1 20.3 0.4 26.8 8.0
8 2010/1/17 3 13.0 0.9 23.1 12.4
9 2010/3/15 <1 5.7 0.6 16.2 10.8

Coupled surface-subsurface models usually simulate many hydrologic


fluxes and state variables across spatial and temporal scales. These si- Fig. 8. Sensitivity of groundwater flooding area to fault hydraulic conductivity.
mulated results are usually validated only by streamflow or ground- Young and old refer to fault categories shown in Fig. 2c.
water table elevation at limited locations, which leaves large un-
certainties in spatial distribution of groundwater table (Yu et al., 2016)
and soil moisture (Shi et al., 2015). Without calibration on local
groundwater table, we showed that relative dynamics of groundwater
table could be reliable in terms of spatial information, when absolute
groundwater table height showed deviated pattern. This is probably
due our model confidence in ET performance. Further calibration on
groundwater table and assessment of uncertainties are needed to refine
the spatial information on flood risk.

4.2. Implications for flood management

In this basin, identification of faults is critical for prediction of


groundwater flooding. Macropore has been found to contribute
groundwater flooding due to accumulated wetness over several years
(Pinault et al., 2005). Our results showed that distribution of faults
Fig. 9. Sensitivity of groundwater flooding time lag to fault hydraulic con-
could potentially increase groundwater flood risk. Fault development
ductivity.
was not incorporated in vast majority of hydrology models. At KRB,
Kourgialas et al. (2008) and Nikolaidis et al. (2013) have introduced

Fig. 7. Spatial distribution of time lag between river flooding and groundwater flooding indicated by model simulation.

136
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

Table 4
Land uses coverage (m2) and % percentage of land uses in high flooding risk areas. The cost of flooding damage was calculated based on the cost per unit of yield
weight and yield per land use.
Land uses Area (km2) Percentage of different land use Cost per unit weight Yield (kg/km2) 106€
(%) (€/kg)

Fruit trees and berry plantations 1.4 16 0.18 3 × 106 0.8


Olive groves 3.6 39 0.48 5 × 105 0.8
Complex cultivation patterns 0.2 2 0.51 4 × 106 0.4
Land principally occupied by agriculture, with significant areas of natural 2.2 24 – – –
vegetation
Natural grasslands 1.2 13 – – –
Sclerophyllous vegetation 0.5 6 0.18 1 × 106 0.1
Total 9.1 100 – – 2.1

fault data. However, these faults were incorporated to support the de- damages both in economy and mortality (Hu et al., 2018).
lineation of extended watershed area for the karstic aquifer and forcing The groundwater modelling of the present study is significant in the
the routing of water from the extended karst to the springs inside the understanding of the soil degradation in karstic mountainous areas. It
watershed. Our study suggested that including faults in a physically- has been presented from other studies that the karstic groundwater is
based 3D hydrologic model automatically the extended karst is deli- conveying dissolved pollutants during flood events like e.g. nitrates
neated and thus the model provides additional information for flood from mountainous areas toward lowland areas at Koiliaris watershed
management. (Nikolaidis et al., 2013; Moraetis et al., 2010). The previous impact is
Faults of high hydraulic conductivity are vulnerable to abrupt related to livestock grazing in the higher land where grasses are
groundwater flooding when the rain intensity exceeds the infiltration growing during summer. The identification of groundwater flood areas
rate of the karst (Gutiérrez et al., 2014). The groundwater table can rise will designate the areas accepting probably low quality water. On the
quickly due to high connectivity to surface water. The previous explains other hand the quality analysis of the groundwater flood water will
probably also the observed time lag of groundwater flooding and river offer the opportunity to understand the soil degradation reflection in
flooding. The higher lag in the rain events where precipitation is dis- the karst water.
tributed in more days could be possibly related to more effective in-
filtration from the karst. On the other hand more intense events lower
the infiltration capacity and increase the abrupt fault response. Not to 4.3. Sources of uncertainty
neglect also, the effect of the spatial distribution of the rain event, close
or not in the major faults area, which can affect the lag between the Model uncertainty was created by the simplified vertical re-
groundwater flooding and the river flooding. presentation of geology. The geology was represented by homogeneous
Faults of low hydraulic conductivity are vulnerable to slowly de- texture from ground surface to bedrock in each triangle. Such simpli-
veloped groundwater flooding. Groundwater storage can be accumu- fication may likely results in high spikes in simulated streamflow.
lated during the whole wet seasons, years before groundwater dis- Consequently, the groundwater flooding damage may be over-pre-
charge happens (Pinault et al., 2005). This type of groundwater dicted. Also, we used 20 m as the thickness from ground surface to
flooding may last long time and difficult to mitigate. bedrock depth, though we have noticed that the geologic layer can
Flooding risks associated with faults behaviours in the groundwater reach as high as several kilometres in the karst area and few meters in
flooding events may impose significant costs in the agriculture sector. the alluvial. Since spatial thickness of geology is currently not available,
For example, the flood risk designation for KRB from Vozinaki et al. our simplification could be valid in terms of relative groundwater ele-
(2015) revealed areas of high flooding risk around the river with an vation and frequency analysis. In addition, uniform depth of geologic
estimation of 0.2 × 106 € damages for an inundated area of 0.4 km2 layer can significantly reduce the computational cost.
(one order of magnitude smaller compared to 2.1 × 106 € in this study Subsurface heterogeneity in 3D is another overlooked aspect. In
for the whole river basin). The percentage of the study area in the karstic regions groundwater flowpaths can be long and one can imagine
previous work is restricted to river over-bank flooding and correspond that groundwater flooding in one place could results from hydro-
only in the 0.003% of the total watershed and 0.02% of the agricultural dynamics some distance away (Butscher and Huggenberger, 2007;
land which is very low compare to our estimations of the affected areas Gutiérrez et al., 2014; Hartmann et al., 2014; Rugel et al., 2016). Our
(7 and 24% respectively). The reduction of the affected area is due to model PIHM can only represent very local and shallow near surface
the fact that the previous study considered the flooding caused by dynamics of groundwater tables. New generations of models should
Koiliaris River (a 5 Km stretch from the springs to the sea located in the address multilayer subsurface dynamics at timescales ranging from
alluvial plain) without considering groundwater overland flow. The floods to seasonal long distance cavernous drainage. Though dynamic
cost of the flood damage in our case is expanding in million of euros, modelling of karstic system is computationally expensive, the greatest
compare to the hundred thousand obtained by Vozinaki et al. (2015). limitation to the development at present is a scarcity of field observa-
Despite, we have not considered the depth and the velocity of the flood tions.
in our cost estimation we believe that the cost will be maintained in the Precipitation can greatly influence groundwater flooding. We only
same magnitude. Other critical factors which are lately gaining large simulated 3 years (annual runoff was 4.93 m3/s) hydrologic processes
attention is the flood frequency and the flood duration (Ward et al., due to intensive computational cost of coupled surface-subsurface
2017). It has been stressed that El Niño Southern Oscillation (ENSO) is modelling (The execution time was 4120 min on Intel Core i5-8500 @
affecting the flood duration more than flood frequency (Ward et al., 3.00 GHz). We selected relatively wet years since 1972 (annual runoff
2017). However, the recent studies in Crete showed that climate be- was 2.96 m3/s from 1972 to 2010) (Fig. S2). Therefore, we probably
comes drier with higher frequency of extreme events (Tsanis et al., overestimated the cost of flood damage. Long-term hydrologic model-
2011). The same work shows an increase of 35% of events with high ling is necessary to reveal the impacts of regional climate mechanisms
precipitation which a clear indication that a minimum 35% of damages (Koutroulis et al., 2010) on groundwater flooding when computation
is likely to occur. The flood frequency definitely will increase the power is improved.
Improvement of temporal and spatial ET prediction is important to

137
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

modelling flooding processes (Abiodun et al., 2018). Our simulated ET flooding using hydrograph analysis and impulse response functions. Hydrol. Process.
was less confident during winter and spring season. This was due to 31, 4586–4599.
Bhatt, G., Kumar, M., Duffy, C.J., 2014. A tightly coupled GIS and distributed hydrologic
uncertainties in spatial parameters of land cover and soil properties. modeling framework. Environ. Model. Softw 62, 70–84.
Spatially, the ET was primarily determined by vegetation type, special Bossard, M., Feranec, J., Otahel, J., 2000. CORINE Land Cover Technical Guide:
attention should be paid to the dominant vegetation types to improve Addendum 2000.
Butscher, C., Huggenberger, P., 2007. Implications for karst hydrology from 3D geological
spatial accuracy of model simulation. In addition, the simulated ET modeling using the aquifer base gradient approach. J. Hydrol. 342 (1–2), 184–198.
deviation in winter and spring may cause flood prediction error trig- Caputo, R., Catalano, S., Monaco, C., Romagnoli, G., Tortorici, G., Tortorici, L., 2010.
gered by snow melting. Thus, observation of snow accumulation is re- Active faulting on the island of Crete (Greece). Geophys. J. Int. 183, 111–126.
Chen, F., Dudhia, J., 2001. Coupling an advanced land surface-hydrology model with the
quired to improve flood prediction. Penn State-NCAR MM5 modeling system. Part I: model implementation and sensi-
Finally, the flood damage cost which has been referred in the pre- tivity. Mon. Weather Rev. 129, 569–585.
sent study is only an indicative estimation to show the magnitude of the Chen, X., Kumar, M., McGlynn, B.L., 2015. Variations in streamflow response to large
hurricane-season storms in a Southeastern US watershed. J. Hydrometeorol. 16,
problem of groundwater flood. Critical factors like flood depth and
55–69.
velocity were followed from other published work as already men- Cohen, S.D., Hindmarsh, A.C., 1996. CVODE, a stiff/nonstiff ODE solver in C. Comput.
tioned. Therefore these factors need to be elaborated for more accurate Phys. 10, 138–143.
estimation of flood damage cost. Accordingly, the coupled surface- Condon, L.E., Hering, A.S., Maxwell, R.M., 2015. Quantitative assessment of groundwater
controls across major US river basins using a multi-model regression algorithm. Adv.
subsurface model can be adopted in regard to grid resolution, process Water Resour. 82, 106–123.
integration, result post-processing. Costabile, P., Macchione, F., 2015. Enhancing river model set-up for 2-D dynamic flood
modelling. Environ. Model. Softw 67, 89–107.
European Environment Agency, 2004. Corine Land Cover 2000 - Mapping a Decade of
5. Summary Change.
Fassoulas, C., 2001. The tectonic development of a Neogene basin at the leading edge of
We developed a coupled surface-subsurface model for groundwater the active European margin: the Heraklion basin, Crete, Greece. J. Geodyn. 31,
49–70.
flood risk assessment. The model can be used to delineate flood-hazard Fassoulas, C., Rahl, J.M., Ague, J., Henderson, K., 2004. Patterns and conditions of de-
areas and risk-levels and predict flood time, which are fundamental formation in the Plattenkalk nappe, Crete, Greece: a preliminary study. Bull. Geol.
components of flood management. The methodology presented in this Soc. Greece XXXVI, 1626–1635.
Finch, J.W., Bradford, R.B., Hudson, J.A., 2004. The spatial distribution of groundwater
paper could become a useful tool for the prediction of groundwater flooding in a chalk catchment in southern England. Hydrol. Process. 18, 959–971.
flooding areas and for the better understanding on the mechanisms of Gallegos, Josue Jacob, Hu, Bill X., Davis, Hal, 2013. Simulating flow in karst aquifers at
groundwater flooding. laboratory and sub-regional scales using MODFLOW-CFP. Hydrogeol. J. 21 (8),
1749–1760.
This paper demonstrated a solution for improving environmental
Gutiérrez, F., Parise, M., De Waele, J., Jourde, H., 2014. A review on natural and human-
assessment dealing with groundwater dynamics. In particular, coupled induced geohazards and impacts in karst. Earth Sci. Rev. 138, 61–88.
surface and subsurface modelling can offer sophisticated solutions to Harbaugh, A.W., 2005. MODFLOW-2005, the U.S. Geological survey modular ground-
the users regarding the coupling of hydrological information to not only water model – the ground-water flow process. In: U.S. Geological Survey Techniques
and Methods 6-A16, . https://pubs.usgs.gov/tm/2005/tm6A16/.
flooding but also other groundwater related problems (e.g. ecological Hartmann, A., Lange, J., Weiler, M., Arbel, Y., Greenbaum, N., 2012. A new approach to
conservation, agricultural management). In this sense, the techniques model the spatial and temporal variability of recharge to karst aquifers. Hydrol. Earth
proposed here about the distributed groundwater dynamics may be Syst. Sci. 16, 2219–2231.
Hartmann, A., Goldscheider, N., Wagener, T., Lange, J., Weiler, M., 2014. Karst water
easily implemented in diverse environmental management, since PIHM resources in a changing world: review of hydrological modeling approaches. Rev.
is open source software. Geophys. 52 (3), 218–242.
Hirabayashi, Y., Mahendran, R., Koirala, S., Konoshima, L., Yamazaki, D., Watanabe, S.,
Kim, H., Kanae, S., 2013. Global flood risk under climate change. Nat. Clim. Change
Acknowledgement 3, 816.
Hu, P., Zhang, Q., Shi, P., Chen, B., Fang, J., 2018. Flood-induced mortality across the
The authors owe thanks to the four anonymous reviewers for their globe: spatiotemporal pattern and influencing factors. Sci. Total Environ. 643,
171–182.
constructive comments which greatly improved the manuscript.
Huntingford, C., Marsh, T., Scaife, A.A., Kendon, E.J., Hannaford, J., Kay, A.L., Lockwood,
Financial support for this study was provided by the European M., Prudhomme, C., Reynard, N.S., Parry, S., Lowe, J.A., 2014. Potential influences
Commission 7th Framework Programme as a Large Integrating Project on the United Kingdom's floods of winter 2013/14. Nat. Clim. Change 4 (9), 769.
Jones, W.K., 2013. Physical structure of the epikarst. Acta Carsol. 42 (2–3), 311–314.
(SoilTrEC, Grant Agreement no. 244118), the National Natural Science
Jongman, B., Winsemius, H.C., Aerts, J.C., de Perez, E.C., van Aalst, M.K., Kron, W.,
Foundation of China (Grant No. 51879289 and No. 91547108), and the Ward, P.J., 2015. Declining vulnerability to river floods and the global benefits of
National Key Research and Development Program of China adaptation. Proc. Natl. Acad. Sci. 112, E2271–E2280.
(2017YFC0405900). Jonkman, S.N., Vrijling, J.K., Vrouwenvelder, A., 2008. Methods for the estimation of loss
of life due to floods: a literature review and a proposal for a new method. Nat.
Hazards 46, 353–389.
Appendix A. Supplementary data Kirchen, G., Calvaruso, C., Granier, A., Redon, O.P., Van der Heijden, G., Bréda, N.,
Turpault, P.M., 2017. Local soil type variability controls the water budget and stand
productivity in a beech forest. For. Ecol. Manag. 390, 89–103.
Supplementary data to this article can be found online at https:// Kourgialas, N.N., Karatzas, P.G., Nikolaidis, P.N., 2008. Simulation of the flow in the
doi.org/10.1016/j.envsoft.2019.01.008. Koiliaris river basin (Greece) using a combination of GIS, the HSPF model and a
karstic – snow melt model. In: 4th Biennial International Congress of iEMSs,
Barcelona Spain, vol. 1. pp. 512–520.
References Kourgialas, N.N., Karatzas, G.P., 2011. Flood management and a GIS modelling method to
assess flood-hazard areas—a case study. Hydrol. Sci. J.–J. Sci. Hydrol. 56, 212–225.
Abboud, J.M., Ryan, M.C., Osborn, G.D., 2017. Groundwater flooding in a riv- Kourgialas, N.N., Karatzas, G.P., 2013. A hydro‐economic modelling framework for flood
er‐connected alluvial aquifer. J. Flood Risk Manag. e12334. damage estimation and the role of riparian vegetation. Hydrol. Process. 27, 515–531.
Abiodun, O.O., Guan, H., Post, V.E.A., Batelaan, O., 2018. Comparison of MODIS and Kourgialas, N.N., Karatzas, G.P., 2014. A hydro-sedimentary modeling system for flash
SWAT evapotranspiration over a complex terrain at different spatial scales. Hydrol. flood propagation and hazard estimation under different agricultural practices. Nat.
Earth Syst. Sci. 22, 2775–2794. Hazards Earth Syst. Sci. 14, 625–634.
Adamopoulos, K., 2013. 1000 and 1 caves in “Lefka Ori” massif, on Crete, Geece. In: 16th Kourgialas, N., Karatzas, G., Nikolaidis, N., 2010. An integrated framework for the hy-
International Congress of Speleology, Czech Republic, Brno. drologic simulation of a complex geomorphological river basin. J. Hydrol. 381,
Archfield, S.A., Hirsch, R.M., Viglione, A., Blöschl, G., 2016. Fragmented patterns of flood 308–321.
change across the United States. Geophys. Res. Lett. 43 (10), 239. https://doi.org/10. Koutroulis, A.G., Tsanis, I.K., Daliakopoulos, I.N., 2010. Seasonality of floods and their
1002/2016GL070590. 232–10. hydrometeorologic characteristics in the island of Crete. J. Hydrol. 394, 90–100.
Arnell, N.W., Gosling, S.N., 2016. The impacts of climate change on river flood risk at the Kovács, A., 2003. Geometry and Hydraulic Parameters of Karst Aquifers: a Hydrodynamic
global scale. Climatic Change 134, 387–401. Modeling Approach. Faculté des Sciences, Université de Neuchâtel Thesis: 131 pp.
Ascott, M.J., Marchant, B.P., Macdonald, D., McKenzie, A.A., Bloomfield, J.P., 2017. Available in pdf at. www.unine.ch/biblio/.
Improved understanding of spatio‐temporal controls on regional scale groundwater Kumar, M., 2009. Toward a Hydrologic Modeling System. PhD Diss. The Pennsylvania

138
X. Yu et al. Environmental Modelling and Software 114 (2019) 129–139

State University, pp. 251pp. groundwater/surface water interaction in a karst watershed: lower Flint River Basin,
Kumar, M., Marks, D., Dozier, J., Reba, M., Winstral, A., 2013. Evaluation of distributed southwestern Georgia, USA. J. Hydrol.: Reg. Stud. 5, 1–19.
hydrologic impacts of temperature-index and energy-based snow models. Adv. Water Running, S., Mu, Q., Zhao, M., 2017. MOD16A2 MODIS/Terra Net Evapotranspiration 8-
Resour. 56, 77–89. Day L4 Global 500m SIN Grid V006. NASA EOSDIS Land Process. DAAC. https://doi.
Kundzewicz, Z.W., Kanae, S., Seneviratne, S.I., Handmer, J., Nicholls, N., Peduzzi, P., org/10.5067/MODIS/MOD16A2.006.
Mechler, R., Bouwer, L.M., Arnell, N., Mach, K., 2014. Flood risk and climate change: Seo, S.B., Sinha, T., Mahinthakumar, G., Sankarasubramanian, A., Kumar, M., 2016.
global and regional perspectives. Hydrol. Sci. J. 59, 1–28. Identification of dominant source of errors in developing streamflow and ground-
Li, S., Duffy, C.J., 2011. Fully coupled approach to modeling shallow water flow, sedi- water projections under near‐term climate change. J. Geophys. Res. Atmos. 121,
ment transport, and bed evolution in rivers. Water Resour. Res. 47, W03508. 7652–7672.
Li, Y., Ren, X., Hill, R., Malone, R., Zhao, Y., 2018. Characteristics of water infiltration in Shi, Y., Baldwin, D.C., Davis, K.J., Yu, X., Duffy, C.J., Lin, H., 2015. Simulating high‐-
layered water-repellent soils. Pedosphere 28 (5), 775–792. resolution soil moisture patterns in the Shale Hills watershed using a land surface
Macdonald, D.M.J., Bloomfield, J.P., Hughes, A.G., MacDonald, A.M., Adams, B., hydrologic model. Hydrol. Process. 29, 4624–4637.
McKenzie, A.A., 2008. Improving the Understanding of the Risk from Groundwater Shi, Y., Eissenstat, David M., He, Yuting, Davis, Kenneth J., 2018. Using a spatially-dis-
Flooding in the UK. tributed hydrologic biogeochemistry model with a nitrogen transport module to
Menon, M., Rousseva, S., Nikolaidis, N.P., et al., 2014. SoilTrEC: a global initiative on study the spatial variation of carbon processes in a Critical Zone Observatory. Ecol.
critical zone research and integration. Environ. Sci. Pollut. Res. 21, 3191. Model. 380, 8–21.
Moraetis, D., Efstathiou, D., Stamati, F., Tzoraki, O., Nikolaidis, N.P., Schnoor, J.L., Shoemaker, W.B., Kuniansky, E.L., Birk, Steffen, Bauer, Sebastian, Swain, E.D., 2008.
Vozinakis, K., 2010. High-frequency monitoring for the identification of hydrological Documentation of a Conduit Flow Process for MODFLOW-2005: U.S. Geological
and bio-geochemical processes in a Mediterranean river basin. J. Hydrol. 389, Survey Techniques and Methods Report, Book 6. Chapter A24. http://pubs.usgs.
127–136. gov/tm/tm6a24/.
Moraetis, D., Paranychianakis, N.V., Nikolaidis, N.P., Banwart, S.A., Rousseva, S., Stamati, E.F., Nikolaidis, P.N., Schnoor, L.J., 2013. Modeling topsoil carbon sequestration
Kercheva, M., Nenov, M., Shishkov, T., de Ruiter, P., Bloem, J., 2015. Sediment in two contrasting crop production to set-aside conversions with RothC – calibration
provenance, soil development, and carbon content in fluvial and manmade terraces at issues and uncertainty analysis. Agric. Ecosyst. Environ. 165, 190–200.
Koiliaris River Critical Zone Observatory. J. Soils Sediments 15, 347–364. Tataris, A.A., Christodoulou, G., 1969. Geological Map 1:50000, Sheet Alikianos. Greek
Morris, S.E., Cobby, D., Zaidman, M., Fisher, K., 2018. Modelling and mapping ground- Institute of Geology and Mineral Exploration, Athens.
water flooding at the ground surface in Chalk catchments. J. Flood Risk Manag. 11, Tsanis, I.K., Koutroulis, A.G., Daliakopoulos, I.N., Jacob, D., 2011. Severe climate-in-
S251–S268. duced water shortage and extremes in Crete. Climatic Change 106 (4), 667–677.
Nash, J.E., Sutcliffe, J.V., 1970. River flow forecasting through conceptual models part I– UNISDR, 2015. Global Assessment Report on Disaster Risk Reduction.
A discussion of principles. J. Hydrol. 10, 282–290. Upton, K.A., Jackson, C.R., 2011. Simulation of the spatio‐temporal extent of ground-
Naughton, O., Johnston, P.M., Gill, L., 2012. Groundwater flooding in Irish Karst: the water flooding using statistical methods of hydrograph classification and lumped
hydrological characterisation of Ephemeral lakes (Turloughs). J Hydrol. 470–471, parameter models. Hydrol. Process. 25, 1949–1963. https://doi.org/10.1002/hyp.
82–97. 7951.
Nikolaidis, N.P., Bouraoui, F., Bidoglio, G., 2013. Hydrologic and geochemical modeling Vozinaki, A.-E.K., Karatzas, G.P., Sibetheros, I.A., Varouchakis, E.A., 2015. An agri-
of a karstic Mediterranean watershed. J. Hydrol. 477, 129–138. cultural flash flood loss estimation methodology: the case study of the Koiliaris basin
ORNL DAAC, 2017. MODIS Collection 6 Land Products Global Subsetting and (Greece), February 2003 flood. Nat. Hazards 79, 899–920.
Visualization Tool. https://doi.org/10.3334/ornldaac/1379. Ward, P.J., Jongman, B., Aerts, J.C., Bates, P.D., Botzen, W.J., Loaiza, A.D., Hallegatte, S.,
PIHM, 2015. Pennsylvania State Integrated Hydrologic Model Version 2.2, Zenodo. Kind, J.M., Kwadijk, J., Scussolini, P., 2017. A global framework for future costs and
https://doi.org/10.5281/zenodo.31438. benefits of river-flood protection in urban areas. Nat. Clim. Change 7, 642.
Pinault, J.-L., Amraoui, N., Golaz, C., 2005. Groundwater‐induced flooding in macro- Wösten, J.H.M., Lilly, A., Nemes, A., Le Bas, C., 1999. Development and use of a database
pore‐dominated hydrological system in the context of climate changes. Water Resour. of hydraulic properties of European soils. Geoderma 90, 169–185.
Res. 41. Yamazaki, D., Kanae, S., Kim, H., Oki, T., 2011. A physically based description of
PolomčiĆ, D., Dragišić, V., Živanović, V., 2013. Hydrodynamic modeling of a complex floodplain inundation dynamics in a global river routing model. Water Resour.
karst-alluvial aquifer: Case study of Prijedor groundwater source, Republic of Srpska, Res. 47.
Bosnia and Herzegovina. Acta Carsol. 42 (1), 93–107. Yu, X., Bhatt, G., Duffy, C., Shi, Y., 2013. Parameterization for distributed watershed
Prasad, M., Ramakrishna Reddy, M., Sunitha, V., 2017. Bedrock structural controls on the modeling using national data and evolutionary algorithm. Comput. Geosci. 58,
occurrence of sinkholes : a case study from chintakommadinne area, part of cuddapah 80–90. https://doi.org/10.1016/j.cageo.2013.04.025.
basin, south India. J Ind. Geophys. Union 21 (2), 124–139. Yu, X., Bhatt, G., Duffy, C.J., Wardrop, D.H., Najjar, R.G., Ross, A.C., Rydzik, M., 2015a. A
Qu, Y., 2005. An Integrated Hydrologic Model for Multi-Process Simulation Using Semi- coupled surface–subsurface modeling framework to assess the impact of climate
discrete Finite Volume Approach. Pennsylvania State University. change on freshwater wetlands. Clim. Res. 66, 211–228.
Qu, Y., Duffy, C.J., 2007. A semidiscrete finite volume formulation for multiprocess Yu, X., Lamačová, A., Duffy, C.J., Krám, P., Hruška, J., White, T., Bhatt, G., 2015b.
watershed simulation. Water Resour. Res. 43, W08419. Modeling the long term water yield effects of forest management in a Norway spruce
Reed, Erin M., Wang, Dingbao, Duranceau, Steven J., 2016. Modeling anthropogenic forest. Hydrol. Sci. J. 60, 174–191.
boron in groundwater flow and discharge at Volusia Blue Spring (Florida, USA). Yu, X., Lamačová, A., Duffy, C., Krám, P., Hruška, J., 2016. Hydrological model un-
Hydrogeol. J. 25 (1), 91–101. certainty due to spatial evapotranspiration estimation methods. Comput. Geosci. 90,
Reimann, T., Geyer, T., Shoemaker, W.B., Liedl, R., Sauter, M., 2011. Effects of dyna- 90–101.
mically variable saturation and matrix‐conduit coupling of flow in karst aquifers. Zhang, M., Chen, X., Kumar, M., Marani, M., Goralczyk, M., 2017. Hurricanes and tropical
Water Resour. Res. 47, W11503. https://doi.org/10.1029/2011WR010446. storms: a necessary evil to ensure water supply? Hydrol. Process. 31, 4414–4428.
Rodell, M., et al., 2004. The global land data assimilation system. Bull. Am. Meteorol. Soc. Zhang, Y., Li, W., Sun, G., Miao, G., Noormets, A., Emanuel, R., King, J.S., 2018.
85 (3), 381–394. Understanding coastal wetland hydrology with a new regional scale process-based
Rossi, M.J., Ares, J.O., Jobbágy, E.G., Vivoni, E.R., Vervoort, R.W., Schreiner-McGraw, hydrologic model. Hydrol. Process.
A.P., Saco, P.M., 2018. Vegetation and terrain drivers of infiltration depth along a Zhang, Y., Li, W., Sun, G., King, J.S., 2019. Coastal wetland resilience to climate varia-
semiarid hillslope. Sci. Total Environ. 644, 1399–1408. bility: a hydrologic perspective. J. Hydrol.
Rugel, K., Golladay, S.W., Jackson, C.R., Rasmussen, T.C., 2016. Delineating

139

You might also like