You are on page 1of 12

Journal of Hydrology 397 (2011) 93–104

Contents lists available at ScienceDirect

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

A simple thermal mapping method for seasonal spatial patterns


of groundwater–surface water interaction
Christian Anibas a,⇑, Kerst Buis b, Ronny Verhoeven c, Patrick Meire b, Okke Batelaan a,d
a
Department of Hydrology and Hydraulic Engineering, Vrije Universiteit Brussel (VUB), 1050 Brussels, Pleinlaan 2, Belgium
b
Department of Biology, Ecosystem Management Research Group (ECOBE), University of Antwerp, 2610 Antwerp, Universiteitsplein 1, Belgium
c
Civil Engineering Department, Hydraulics Laboratory, University of Ghent, 9000 Ghent, Sint-Pieternieuwstraat 41, Belgium
d
Department of Earth and Environmental Sciences, K.U. Leuven, 3001 Leuven, Celestijnenlaan 200e, Belgium

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

Article history: A simple thermal mapping method for simulating seasonal and spatial patterns of groundwater–surface
Received 2 October 2009 water interaction is developed and tested for a segment of the Aa River, Belgium. Spatially distributed
Received in revised form 14 November 2010 temperature profiles in the hyporheic zone of the river are measured in winter and summer seasons of
Accepted 25 November 2010
three consecutive years. Inverse modeling of the one-dimensional heat transport equation is applied to
This manuscript was handled by P. Baveye,
estimate vertical advective fluxes using the numerical STRIVE model and an analytical model. Results
Editor-in-Chief, with the assistance of D.P. of the study show that seasonal flux estimates for summer and winter can be derived with a minimum
Ahlfeld, Associate Editor data input and simulation effort. The estimated fluxes are analyzed via non-parametric statistical tests,
while spatial interpolation techniques are used to generate maps of distributed flux exchange. The esti-
Keywords: mated seepage is compared with volumetric flux obtained from piezometer measurements and output of
Heat transport modeling a groundwater model. The thermal method shows higher discharge rates in winter and that the relative
Temperature contribution of exfiltration to the river discharge is higher in summer. A higher flux and a more hetero-
Hyporheic zone geneous flow pattern are observed in the upper reach of the river compared to the lower reach. This spa-
Groundwater–surface water interaction tial difference shows the importance of the local geomorphology and to a lesser extent the hydrogeologic
FEMME setting on hyporheic flux exchange in the river. A significantly higher flux is noted on the banks than in
Belgium
the center of the river, which is driven by the relatively high hydraulic conductivity of the river banks. It is
concluded that bank flow in groundwater–surface water interaction deserves more attention. The main
channel of the Aa River alone accounts for about 15% of the total river discharge at its outlet. As the devel-
oped thermal method is cost-effective, simple and fast, it is recommended for use in identifying zones of
interest in initial stages of field investigations of groundwater–surface water interaction.
Ó 2010 Elsevier B.V. All rights reserved.

1. Introduction Directive (European Commission, 2000, 2006) mandate the protec-


tion of the links between groundwater and surface water systems.
An important process in ecohydrology is groundwater–surface To do this, reliable and transferable conceptual models (NRC, 2004;
water interaction at the interface of lake or riverbeds (Stanford, Smith, 2005; Schmidt et al., 2008) are required to assess mass
1998; Woessner, 2000). The accurate quantification of exchange fluxes and related biochemical processes across groundwater–sur-
of water and energy between groundwater and surface water sys- face water interfaces.
tems remains a challenging task due to the problems of heteroge- Different methods have been discussed and used to quantify the
neity and scale (Becker et al., 2004; Kalbus et al., 2008). exchange flux between groundwater and surface water systems
The saturated zone beneath and beside streams, rivers, lakes (Kalbus et al., 2006). Seepage meters are the only means by which
and wetlands where groundwater and surface water actively mix groundwater–surface water exchange is directly measured (Lee,
and exchange is called hyporheic zone (Brunke and Gonser, 1977; Rosenberry, 2008). Although seepage meters are technically
1997; Boulton et al., 1998; Hayashi and Rosenberry, 2002; Smith, simple, there are uncertainties in the measured flux due to opera-
2005). The processes in the hyporheic zone are variable in space tional problems in the field (Shaw and Prepas, 1990; Murdoch and
and time on relatively small scales (McClain et al., 2003). National Kelly, 2003). Indirect methods estimate the groundwater–surface
and international regulations like the European Water Framework water interaction by deriving the fluxes across the interface of
the hyporheic zone from other parameters. Groundwater head
⇑ Corresponding author. Tel.: +32 26495877; fax: +32 26293022. measurements from piezometers and boreholes in riverbeds or
E-mail address: canibas@vub.ac.be (C. Anibas). near streams (Cey et al., 1998; Baxter et al., 2003) can be used to

0022-1694/$ - see front matter Ó 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhydrol.2010.11.036
94 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

derive vertical flow estimates. But the exact estimation of hypor- glected for our analysis hence the one dimensional, vertical, aniso-
heic-zone flux is hampered by difficulties in measuring in situ thermal transport of heat through homogeneous, porous media is
hydraulic conductivity (Chen, 2000). Piezometers installed in river- described as (Suzuki, 1960; Stallman, 1965; Lapham, 1989).
beds are also prone to scouring and damage by floods or debris.
Other indirect methods of estimating groundwater–surface water @2T @T @T
k  v z cw qw ¼ cq ð1Þ
exchange include differential discharge gauging (Becker et al., 2004), @z2 @z @t
numerical modeling (Cardenas and Zlotnik, 2003; Fleckenstein
et al., 2004; Saenger et al., 2005) and the use of conservative trac- where k is the thermal conductivity of the soil–water matrix in
ers like dye, salt, chloride or stable isotopes (Carey and Quinton, J s1 m1 K1, T the temperature at point z at time t in the sediment
2005). Remote sensing, a relatively new technique in this field in K (°C), cw the specific heat capacity of the fluid in J kg1 K1, qw
(Becker, 2006), offers a spatially continuous qualitative estimation the density of the fluid in kg m3, vz the vertical component of
of hyporheic-zone flux exchange (Loheide and Gorelick, 2006). the groundwater velocity in the sediment in m s1, c the specific
Heat can be used as a tracer to estimate groundwater–surface heat capacity of the rock-fluid matrix in J kg1 K1, and q the
water interaction (Stonestrom and Constantz, 2003; Anderson, wet-bulk density (density of the rock-fluid matrix) in kg m3. The
2005; Constantz, 2008; Constantz et al., 2008). The exchange pat- terms cwqw and cq represent the volumetric heat capacity of the
terns can be inferred qualitatively from temperature variations at fluid and the rock-fluid matrix in J m3 K1 respectively. The first
the groundwater–surface water interface (Lowry et al., 2007) or term of the left hand side of Eq. (1) represents the conductive and
quantified by empirically relating mapped streambed tempera- the second term the advective part of the heat transport. A positive
tures to volumetric fluxes (Conant, 2004). Hatch et al. (2006) and sign of the vertical groundwater velocity stands for water moving
Keery et al. (2007) presented methods to determine streambed from the surface into the hyporheic zone (e.g. groundwater re-
seepage rates using phase shifts and amplitude damping of ther- charge or losing stream reach) and a negative sign represents water
mal time series data between pairs of sensors. Exchange rates have moving from the hyporheic zone into the river (e.g. groundwater
also been successfully quantified by inverse modeling of tempera- discharge or gaining stream reach). The groundwater velocity is ex-
ture profiles (Arriaga and Leap, 2006; Schmidt et al., 2006; Anibas pressed in mm d1, equivalent with a unit flux in L m2 d1.
et al., 2009). In the case of a thermal steady state, the temperature distribu-
Suzuki (1960) and Stallman (1965) presented analytical solu- tion in the riverbed is supposed to be constant over time, which re-
tions for using temperature as a natural tracer. Bredehoeft and duces the right hand side of Eq. (1) to 0:
Papadopoulos (1965) introduced a graphical type curve solution
@2T @T
method based on temperature–depth profiles for the estimation k  v z cw qw ¼0 ð2Þ
of vertical groundwater fluxes under steady-state thermal condi- @z2 @z
tions. As temperature distributions in the subsurface are transient Notice that in Eq. (2), the thermal properties of the fluid–sedi-
by nature, Lapham (1989) developed a numerical model that al- ment matrix are only described by the thermal conductivity k.
lows quantification of groundwater–surface water interaction For sediments as different as peat or gravel the thermal conductiv-
using an annual temperature variation simplified by a sine func- ity k varies less than an order of magnitude (Freeze and Cherry,
tion. Arriaga and Leap (2006) and Schmidt et al. (2006, 2007) pre- 1979; Domenico and Schwartz, 1998); e.g. saturated sands usually
sented applications of the estimation of groundwater–surface have k values between 1.4 and 2.2 J s1 m1 K1 (Lapham, 1989;
water interaction based on steady state heat transport, while Ani- Schön, 1998; Stonestrom and Blasch, 2003). This small range of k
bas et al. (2009) showed when the steady-state thermal assump- is a major advantage of the thermal method compared to the
tion can be used. hydraulic head method in which the hydraulic conductivity can
This paper presents a simple and fast method for quantifying vary over orders of magnitude (Chen, 2000). Hence, in many cases,
seasonally distributed groundwater–surface water exchange. The the thermal conductivity can be parameterized by taking its value
proposed approach is a robust, adjustable and first-hand field from literature.
investigation tool for detailed monitoring of zones of interest. The thermal method is an indirect method, field data must be
The underlying hypothesis of the methodology is that steady-state processed with a model in order to derive quantitative estimates
solutions for simultaneous vertical transport of heat and water in of the flow velocity. We apply inverse modeling of the thermal
riverbeds are able to detect groundwater–surface water interaction steady state by solving Eq. (1) numerically and Eq. (2) analytically.
with sufficient spatial and temporal resolution. Measured riverbed Fig. 1 shows the concept of the steady state 1D vertical fluid and
temperature profiles from the Aa River in Belgium are used to test heat transport for a saturated sediment column of the riverbed,
the proposed method. Furthermore, the analytical solution method in which measured temperatures define the upper and lower
of Arriaga and Leap (2006) is used in conjunction with the numer- boundary conditions. The concept of the model does not allow
ical heat transport model STRIVE to statistically verify the pro- to determine lateral or longitudinal flow vectors (Fairley and
posed method. The STRIVE model is based on the approach of Nicholson, 2005; Hoehn and Cirpka, 2006). One or more measured
Lapham (1989) and the FEMME ecosystem modeling platform temperatures within the domain are needed in order to fit the sim-
(Soetaert et al., 2002). ulated temperature distribution to the measured profile (Fig. 1a).
When groundwater flow occurs, the thermal profile shows a stron-
ger curvature with increasing flow velocity (Fig. 1b). For practical
2. Methodology reasons the presented methodology is limited to discharging flow
conditions if applied to shallow groundwater flow systems
The temperature in the hyporheic zone varies seasonally and (Schmidt et al., 2007). Mathematically Eqs. (1) and (2) allow esti-
diurnally as a consequence of heating and cooling of the land sur- mation of infiltration, a reversal of the flux under steady state
face. During summer months the groundwater temperature is gen- conditions however is producing a uniform temperature distribu-
erally cooler than the river temperature, whereas in winter it is tion with no significant thermal gradient. Hence it is not possible
generally the opposite. The groundwater flow in the riverbed is as- to obtain a realistic fit of the model with the applied boundary
sumed to obey Darcy’s law consequently the natural temperature conditions (Fig. 1c).
distribution in the riverbed is influenced by the movement of To solve the thermal steady state problem a hydraulic
water. The geothermal gradient in the shallow subsurface is ne- steady state has to be assumed, diurnal, seasonal and irregular
C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104 95

goodness of fit between the simulated versus measured tempera-


ture profile (Anibas et al., 2009). Hence, it is important that the
processed profiles selected for analysis respect the steady state
assumption.
Two models based on the differential Eqs. (1) and (2) have been
used in this paper, called ‘STRIVE’ and ‘Arriaga and Leap (2006)’
respectively.

2.1. Arriaga and Leap (2006) model

We applied the analytical type curve method of Bredehoeft and


Papadopulos (1965) as described by Arriaga and Leap (2006) using
Microsoft Excel with the Solver add-in. This constitutes a simple
and fast tool to derive quantitative estimates of the vertical
groundwater–surface water exchange.

2.2. STRIVE model

A vertical heat transport model was implemented in STRIVE


Fig. 1. Concept of vertical temperature distribution and simulation of vertical 1D (STReam RIVer Ecosystem), which is a modular package of FOR-
water and heat transport in a river or lake bed under thermal steady state TRAN subroutines for the dynamic ecosystem modeling platform
assumption of constant upper and lower boundary temperatures; (a) medium FEMME (Soetaert et al., 2002). STRIVE simulates different physical
upward groundwater flux; (b) high upward groundwater flux; (c) downward flux.
and biogeochemical transport and exchange processes including
vertical 1D heat and water transport (Eq. (1)). The heat transport
model obtains a fit by minimizing the differences between mea-
(meteorological) transient thermal influences in the measured data sured and simulated temperature distributions according to inter-
should be insignificant, which is a limitation for very dynamic nal user defined integration and fitting routines (Soetaert et al.,
river–aquifer systems. Seasonal thermal influences can be approx- 2002) by adapting the flow velocity (vz).
imated by a sine wave with a wavelength of an astronomical year; For this study the STRIVE model is descritizised as a vertical,
close to the maximum (summer) and minimum (winter) of the one-dimensional saturated and homogeneous sediment column
sine a steady state assumption is valid (Anibas et al., 2009). The of 5 m length and is composed of 100 layers. To reduce descritiza-
application of the thermal steady state assumption therefore is tion errors, the spacing of the model layers follows a sinusoidal dis-
limited to these periods. Diurnal temperature changes have a lim- tribution, resulting in layer thicknesses of 0.001 m at the upper and
ited penetration depth in the sediment due to the high frequency lower boundary, while the thickness of the layers is maximum
of its temperature signals. Schmidt et al. (2006) advises therefore about 0.08 m in the center of the model domain.
to exclude measurements close to the surface water–groundwater The simulation results from STRIVE and Arriaga and Leap (2006)
interface from further analyses. Irregular transient influences, allowed the use of statistical tests for the comparison of spatial,
caused by meteorological events, cannot be corrected and will temporal and conceptual differences of the models, as well as for
result in noise in the vertical heat distribution. Violations of the comparison of solution procedures of STRIVE and Arriaga and Leap
thermal steady state assumption cannot be observed from the (2006).

Fig. 2. Location of the Aa River study site.


96 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

Fig. 3. Digital elevation model showing the examined section of the Aa River with the location of 26 measurement points and three piezometer nests. The river flow is from
weir 3–4. The dotted line indicates the boundary between the occurrence of the Formation of Kasterlee in the upstream part of the river section and the Formation of Diest in
the downstream part (DOV, 2010).

3. Field site and measurements and riffle structures. The local geology is determined by tertiary
Formations of Diest and Kasterlee; the former occurs in the down-
We collected temperature versus depth profiles along a 1425 m stream part while it is overlain by the latter in the upstream part
long stretch of the lower part of the Aa River, Flanders, Belgium (Figs. 3 and 4) (DOV, 2010). The Formation of Kasterlee consists
(Fig. 2). The river has a total length of 36.7 km and a drainage area of fine sands and fractions of clay. The Formation of Diest is de-
of 235 km2; its catchment comprises 15% of the Nete basin. The scribed as a heterogeneous sand layer. In regional groundwater
measurement site is located around 5.5 km upstream of the conflu- models the hydraulic conductivity of the Kasterlee and Diest For-
ence of the Aa with the Kleine Nete River. At the site the Aa flows in mation was parameterized as 12.5 and 9.8 m d1 respectively
WSW direction and has an average water depth of around 1.1 m, a (Woldeamlak, 2007). Both formations together with the underlying
width of 14 m, an average discharge of about 1.9 m3 s1 and a mea- Berchem Formation constitute one aquifer system. It is underneath
sured slope of 0.48‰. The Aa is a typical Flemish lowland river, concealed by the Boom Aquitard (Fig. 4; Marechal and Laga, 1988).
which was stretched and canalized in the 1960s. This section of Temperature can be measured rapidly and easily as sensors are
the Aa contains three meanders and is upstream and downstream technically simple, cheap, widely available and can be handled by
limited by weirs, respectively numbered 3 and 4 (Fig. 3). Ground- unskilled personnel. We performed ‘roaming surveys’ (Keery and
water tables in the area are shallow, about 1 m below the land sur- Binley, 2007) of temperature profile measurements, using a 2.0 m
face. The cross-section of the river is rectangular with a flat river long self-made instrument, the ‘Temperature stick’ (‘T-stick’)
bottom. An exception is the area around points 10 and 11 (Anibas et al., 2009). Fig. 5 shows the measurement setup and
(Fig. 3), where erosion and sedimentation processes established the instrument, consisting of a metal tube, a T-shaped handle at
pronounced riverbed topography. Although strongly determined the top and a pointed tip on the bottom. Close to the bottom end
by the weirs, the discharge regime is influenced by the growth of a slotted hole holds the temperature sensor, a thermistor (Davis
macrophytes in summer (Bal and Meire, 2009). Discharges and Instruments Model 7817; Hayward, CA, USA). The sensor is con-
flow velocities therefore are higher in winter. The combined effects nected to an electric multimeter (ABB Metrawatt); the value of
lead to relatively constant water levels throughout the year, dis- the electric resistance of the sensor is converted to a temperature
rupted only by occasional peak discharges. value via a calibration function. The position of the temperature
The riverbed is composed of fine sand containing a varying frac- sensor in the riverbed is determined by reading scaled marks with
tion of organic matter, especially at the inner banks of meanders. respect to the surface water level.
The sediments are usually compacted along the banks; while in Five summer and six winter measurement campaigns, each
the middle of the river the sediments are loose, showing some pool consisting of two consecutive days have been conducted at the
C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104 97

consist of measurements at five depths: 0.0, 0.10, 0.25, 0.50 m and


a maximum depth. We measured the temperature profile starting
from the interface between river and riverbed sediment indicated
as 0.0 m depth up to a maximum depth generally between 0.40–
1.20 m (average 0.84 m). Deeper measurements were either pre-
vented by a compacted sandy layer or by the limited length of
the instrument. The measurements started usually between 9
and 11 a.m. and ended between 3 and 6 p.m.; around 15 min were
required to measure a temperature profile with the T-stick
instrument.
During the second day of a campaign measurements were taken
at the left bank (L), center (C) and right bank (R) of six transects at
the locations 1, 4, 7, 11, 13 and 14 (Fig. 3). The points L and R are
located a quarter of the width of the river away from respectively
the left and right bank. All measurement campaigns were per-
formed in a similar pattern.
Furthermore, at location 4, 7 and 13 piezometer nests were in-
stalled in the riverbed. Data loggers recorded groundwater levels at
three different depths (0.1–0.15 m, 0.5–0.6 m and 0.9–1.2 m) with
hourly frequency. Slug tests were performed to determine hydrau-
lic conductivity values.
Fig. 4. Longitudinal cross-section along the Aa River between weir 3 (right) and For the simulations presented in this work, the thermal conduc-
weir 4 (left) showing the geology (DOV, 2010). tivity is assumed to be 1.8 J s1 m1 K1; taking into account the
organic content of the uppermost hyporheic sediments. The other
model parameters are given in Table 1. At a depth below 4 m the
groundwater shows an almost constant temperature the year
round, about 1–2 °C above the mean annual surface temperature
(Anderson, 2005). The average temperature of about 650 tempera-
ture profiles (the complete set collected at the Aa River site) was
12.2 °C. This value corresponds well to the average temperature
measured in a piezometer on the southern edge of the flood plain.
Hence, the value is used as a constant for the lower boundary at
5 m depth, whereas the upper boundary of the model domain is
defined at 0.1 m depth and uses the value of the respective temper-
ature profile (Schmidt et al., 2006).

4. Results and discussion

The temperature–depth variation in the river is indicated in


Fig. 6, which shows the average temperatures of 13 profiles along
the center line of the river for 11 measurement campaigns. The
maximum temperature recorded at the interface (0.0 m depth)
was 23.6 °C on 3 July 2006 and the absolute minimum was 3.2 °C
on 26 January 2005. At the maximum depth (0.84 m) the temper-
Fig. 5. Scheme for measuring temperature profiles in the riverbed of the Aa with
the T-stick instrument.
ature varied between 17.2 °C on 25 August 2005 and 7.6 °C on 9
February 2006. The temperature differences between the interface
and the deepest measurement point vary between 1.6 and 7.3 °C,
Aa River field site between August 2004 and February 2007. During with an average of 2.9 °C for all measurement campaigns. Hence,
each campaign 32 temperature profiles were measured at 26 spa- this temperature difference was found both in winter and summer
tially distributed points (Fig. 3), six of the points were investigated seasons.
on both days. We assumed, based on Anibas et al. (2009) that the The relative large dataset of about 850 simulations performed
time between 10 July and 10 September and between 10 January with the Arriaga and Leap (2006) and with STRIVE allowed statis-
and 10 March, referred to as ‘summer’ and ‘winter’ respectively,
are most favorable for a thermal steady state investigation. In total Table 1
around 250 temperatures profiles fulfilling the requirement for a Physical properties and boundary conditions used for the steady state STRIVE model
thermal steady state assumption were used for analysis. and the modeling procedure of Arriaga and Leap (2006).
Fourteen points along the centerline of the river were measured Parameter Value Unit
during the first day of a campaign. Point 1 is located around 30 m
Thermal conductivity, k 1.8 J s1 m1 K1
from weir 4, point 14 around 40 m from weir 3, resulting in a dis- Density of water, qw 1000 kg m3
tance of about 100 m between each point (Fig. 3). However, mea- Specific heat capacity of water, cw 4180 J kg1 K1
surement point 12 is not used for most of the analysis, as it was Upper boundary condition, T0a Variable °C
not possible to reach the center of the river. The spatial position Lower boundary condition, TLb 12.2 °C

of the measurement points was established with a Garmin etrex a


Located at the interface between surface water–riverbed or 0.1 m below the
GPS receiver resulting in a spatial error of about 1.5 m during the interface.
b
different campaigns. As indicated in Fig. 5, the temperature profiles Located 5.0 m below the interface between surface water–riverbed.
98 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

et al. (2006). By applying a Wilcox on Matched Pairs (N = 202)


and the Sign test (N = 195) using different upper boundaries,
including or excluding the uppermost measurement point at the
interface (0.0 m depth; Fig. 1) a significant diurnal influence was
detected in the uppermost 0.1 m of the sediment.

4.1. Spatial patterns

Figs. 7a and 8a show the temperature distribution in the river-


bed interpolated along a vertical, longitudinal cross-section. The
bar graphs of Figs. 7b and 8b suggest that the examined river reach
can be divided into two zones; a zone where higher fluxes are ob-
served in the upstream part of the river (points 9, 10 and 13), and a
zone of lower fluxes in the downstream part (points 5–8). This dif-
ference is confirmed by a Kolmogorov–Smirnov two sample
(N = 127) and a Mann–Whitney U test (N = 127) tests and coincides
with the change in geology as indicated in Figs. 3 and 4. At the cen-
ter line the variability of the fluxes between the measurement
campaigns is relatively similar along the whole river stretch. The
winter data however shows a relative higher variation compared
to the summer.
Fig. 6. Measured vertical profiles of riverbed temperatures averaged for 13
longitudinal points and presented for 11 measurement campaigns, showing
Measurements at six river cross-sections were performed; the
different temperature gradients between summer and winter seasons. median fluxes for each measurement point for (a) winter and (b)
summer seasons are plotted in Fig. 9 versus their position along
the river. Characteristic is that the average fluxes of 79 and
tical analysis of the results (level of significance p = 0.05). First, we 53 mm d1 and their ranges at the left bank and right bank
tested if the distribution of the results is normal (Kolmogorov– respectively are higher than at the center (32 mm d1). A Krus-
Smirnov and Shapiro–Wilk test; N = 202). Since the thermal steady kal–Wallis multiple comparison p values (2-tailed) test (N = 145)
state analysis is limited to negative fluxes both tests indicated that showed significant differences between the center and the banks.
the simulated exchange fluxes are not normally distributed, hence As the composition of the riverbed sediments is relatively uniform,
non-parametric statistical tests were applied for further analysis. the difference in flux can be explained by different hydraulic gradi-
Since six points were measured on both days of the two-day ents and heterogeneity of the underlying geologic layers (Conant,
campaigns, we examined with a Wilcox on matched pairs test 2004; Kalbus et al., 2009), in this case the Formations of Diest
(N = 42) and the Sign test (N = 38) if a significant difference exists and Kasterlee. As indicated in Fig. 4, the upstream section of the
between the estimated fluxes for the center locations. The tests river is dominated by the higher conductive Kasterlee Formation
show no significant differences and hence all thermal profiles are causing a more local flow system compared to the downstream
treated as a single data set. part. There, the Quaternary is underlain directly by the thick Diest
By comparing results of the STRIVE model simulations with Formation, creating deeper, more regional discharges.
Arriaga and Leap (2006) (Kolmogorov–Smirnov two sample test; The analysis of the transect measurements across the river indi-
N = 109) no significant differences were found. This verifies STRIVE cate that the bank areas of the river not only contribute relatively
with the approach of Arriaga and Leap (2006), while it also shows more groundwater discharge compared to the center of the river,
that both approaches are applicable for shallow thermal profiles but that also the flow mechanism at the banks is different. The
measured in a riverbed. temporal variability of the fluxes at the banks is higher, indicating
The measurement setup described in Fig. 5 has an estimated shallow transport processes. Fig. 9 shows that the left bank gener-
uncertainty for the depth determination of ±0.01 m for a single ally shows higher fluxes than the right bank. In summer higher
measurement resulting in a possible maximum cumulative error fluxes occur at the left bank in the upstream part whereas in the
for the measured domain of 0.10 m. A Kruskal–Wallis ANOVA test downstream part higher fluxes are observed at the right bank.
by ranks and median (N = 192) with Arriaga and Leap (2006) for The differences in flux between the left and the right banks as well
domains of ±0.10 m showed no significant difference in resulting as the differences between upstream and downstream can be ex-
fluxes. This proves that the T-stick is sufficient accurate for collect- plained by hydromorphological patterns (Woessner, 2000). Due
ing temperature–depth profiles. to topography (Fig. 3) higher gradients are apparent upstream.
For steady state boundary conditions the error between the Moreover, a meander forces the groundwater flow to converge to
modeled and measured T-profile is dependent on the thickness the left bank, while the flow pattern towards the right bank is
of the modeled domain (Schmidt et al., 2006). This was confirmed diverging. In the downstream part, the valley is widening and the
by a Kolmogorov–Smirnov two sample test (N = 108), which com- groundwater head gradients are lower leading to smaller differ-
pared simulations using a model domain equal to the depth of the ences between right and left bank.
measured profile (Fig. 5) with a model domain equal to 5.0 m. In the upper reach however, points 11 and 14 show dissimilar,
Hence, the lower boundary should be defined at a depth where relatively low fluxes compared to the neighboring points. At point
the temperature is not changing between the different simulated 11, this can be explained by the uneven bathymetry of the river
profiles, thus where a constant groundwater temperature can be cross-section due to the meander at this location. The right and
assumed. center measurements are located in the shallow inner bend of
An important question is the influence of diurnal transient ther- the meander, while the measurement at the left bank is located
mal signals on the applicability of the steady state assumption. The in the deeper part and shows higher fluxes (Fig. 9).
Aa River shows diurnal temperature changes of about 2–3 °C in Piezometer nests (Fig. 3) were used to verify the two heat trans-
winter and 2–5 °C in summer, which are lower than the variations port models. The observed head differences indicate an upward
measured in the small stream in Germany examined by Schmidt gradient during most time of the year. The horizontal hydraulic
C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104 99

Fig. 7. Temperature distribution and vertical groundwater discharge in winter. (a) Two dimensional, longitudinal cross-section (vertical axis exaggerated) along the river
centerline indicating the measured temperature distribution in the riverbed; crosses show measurement locations. (b) Bar graph indicating the measurement points and the
estimated vertical groundwater flux averaged for all winter seasons as simulated with STRIVE.

Fig. 8. Temperature distribution and vertical groundwater discharge in summer. (a) Two dimensional, longitudinal cross-section (vertical axis exaggerated) along the river
centerline indicating the measured temperature distribution in the riverbed; crosses show measurement locations. (b) Bar graph indicating the measurement points and the
estimated vertical groundwater flux averaged for all summer seasons as simulated with STRIVE.

conductivity, Kh, of the streambed was calculated from rising and wards location 4. The temporal trend also corresponds with esti-
falling head slug tests performed in piezometer nest 13. The aver- mated fluxes from the heat transport; significantly higher
age estimated Kh of the riverbed was 2.7 105 m s1. Anisotropy ra- discharges are observed in winter than in summer. We conclude
tios of horizontal over vertical hydraulic conductivity, Kh/Kv, of 3– that the flux estimated with thermal and hydraulic head methods
10 were tested to calculate vertical groundwater fluxes. A ratio of show similar temporal and spatial patterns and acceptable abso-
nine resulted in a lowest absolute difference of 22.4 mm d1 be- lute differences.
tween the estimated fluxes based on the hydraulic head gradients
and the heat transport simulation. With a comparable approach 4.2. Seasonal differences
Schmidt et al. (2006) obtained a similar difference. The exchange
fluxes based on the measured hydraulic heads follow the spatial Seasonal and diurnal transient influences can be taken into ac-
trend observed from the heat transport investigation, indicating count prior to the measurement or simulation by measuring at cer-
highest discharge at location 13, followed by a decreasing trend to- tain times in the year or by excluding the area close to the interface
100 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

Fig. 9. Bar graph of the vertical groundwater fluxes of measurement points along the Aa River (1 is downstream; 14 is upstream) as derived from the STRIVE model; (a) shows
the median of the winter results; green, blue and red bars indicate the measurements at the left bank, centerline and right bank respectively; (b) presents the results for the
summer season. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

from the simulation. However, this is usually not possible with 4.3. Net flows
irregular, hence meteorological events such as phases of relative
cold or warm weather. The effect of the irregular influences on Spatially distributed plots of the groundwater–surface water
the profiles becomes clearly visible when temperature profiles of interaction along the examined river reach were produced by
different winter seasons are compared. Two groups, profiles with interpolating the median flux values of all 26 points on the river
a steep gradient (e.g. 26 January 2005 or 03 July 2006) and with surface area of 20,400 m2. For the interpolation a multilog radial
a relative low gradient (e.g. 14 January 2007 or 08 September bases function (Surfer 8.04, Golden Software, 2003) was used with
2006) can be distinguished in Fig. 6 by their different shapes and R2 = 1800, anisotropy ratio 2.5 and an angle of 20°, representing the
gradients. When these profiles are fitted with the steady state heat morphology and orientation of the river course (Fig. 10). GIS tech-
transport model higher estimated fluxes are obtained for the steep niques were used to determine the net fluxes between groundwa-
profiles as confirmed by a Mann–Whitney U test (N = 127) and a ter and surface water body (Table 2).
Kolmogorow–Smirnow test (N = 127). The spatial interpolation (Fig. 10) resulted in a median flux of
The strong thermal variations in deeper layers as shown in Fig. 6 44 and 55 mm d1 for summer and winter respectively (Table 2,
is a phenomenon not explained by flux changes unless a significant No. 1), i.e. a seasonal difference of 23%. This is equivalent to net
infiltration is assumed. Infiltration is however not supported by groundwater discharges of 10.5 L s1 in summer and 12.9 L s1 in
piezometer measurements. If we assume that irregular meteoro- winter (Table 2, No. 2). These groundwater discharges are respec-
logical influences are responsible for the different temperature dis- tively 0.7% and 0.4% of the average summer and winter discharge
tributions within the seasons, our analysis shows an error of the at weir 4. Hence, the groundwater discharge has a relatively higher
flux estimation of around ±50%. This is exemplified by the median contribution during summer than in winter. Extrapolating these
results of the summer campaigns, which vary between 21 and discharges over the full length of the Aa River (about 37 km) and
38 mm d1, and the results for winter between 23 and comparing with long-term average total river discharge we esti-
49 mm d1. mate a groundwater contribution to the main channel of about
Since a thermal steady state is approached both in summer and 15%. Baseflow separation for the Aa River resulted in a baseflow
winter (Anibas et al., 2009), we hypothesize that seasonal differ- percentage of 75% (Batelaan and De Smedt, 2007). Hence, it is clear
ences in groundwater–surface water interaction can be evaluated that although groundwater discharge through the riverbed is sig-
with the presented methodology. Kolmogorov–Smirnov two sam- nificant (15%), the major component (60%) of baseflow is drainage
ple (N = 135) and Mann–Whitney U tests (N = 135) confirm that from tributaries and ditches in the Aa River catchment, which is
the Aa River has a different hydrological flow regime in winter obvious from presence of the dense network (Fig. 3).
and summer. The median flux of all winter profiles along the center Beside minimum and maximum fluxes (Table 2, Nos. 3–4), Ta-
of the river is 36 mm d1 while the summer flux is 27 mm d1. ble 2 (Nos. 5 and 8) also shows discharge values of the 90th and
A transient thermal simulation with STRIVE over several 10th percentiles of the interpolated fluxes. The 90th percentile is
months in 2006 (Anibas et al., 2009) results in an average flux of not changing significantly between the seasons (Table 2, No. 5),
24 mm d1 at measurement point 7, showing that the magnitude whereas the 10th percentile shows an increase of 52% from sum-
of flux can be reproduced well with the steady state simulation mer to winter (Table 2, No. 8). The maximum fluxes in the river
(Figs. 7 and 8). thus do not change significantly over the seasons (Table 2, No. 3),
C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104 101

Fig. 10. Spatial interpolation of the median groundwater flux along the Aa River of (a) six winter and (b) five summer campaigns 2004–2007. The width of the river is five
times exaggerated.

Table 2 Table 3
Summary of estimates of flux, net discharge and discharge area as derived from Advantages and disadvantages of the steady-state thermal method for estimating
spatially interpolated groundwater–surface water interaction along the Aa River for groundwater–surface water interaction.
summer and winter.
Advantages
Parameter Unit Summer Winter Cost-effective
Simple field measurement
Median net flux mm d1 44 55
Allows spatially distributed data collection
Net dischargea L s1 10.5 12.9
Small range of parameter values
Maximum flux mm d1 75 77
Minimal modeling effort
Minimum flux mm d1 22 36
Short simulation time
90th percentile flux mm d1 73 75
Eludes limitations of other methods
Discharge area 90th percentile %b 9.6 13.4
Net discharge 90th percentile %c 16.2 18.6 Disadvantages
10th percentile flux mm d1 25 38 Indirect method
Discharge area 10th percentile %b 7.9 22.6 Point estimate
Net discharge 10th percentile %c 4.8 15.5 Purely vertical discharge estimation
Only exfiltrating conditions
a
Surface area of the river stretch 20,400 m2. Limited temporal resolution
b
Percentage of discharge area of total river surface area of 20,400 m2 with dis- Measurement only useful in summer and winter
charge higher than 90th (or lower than 10th) percentile. Significant but unknown error band
c
Percentage of the net discharge.

whereas the minimum fluxes alter between winter and summer Smedt, 2001, 2007) was used to simulate groundwater recharge
(Table 2, No. 4). Consequently, there are seasonal differences in to condition this steady state MODFLOW model (Harbaugh and
the discharge areas, as follows from the 90th and 10th percentiles, McDonald, 1996) with a 50 m grid cell resolution. Drain and river
which occupy respectively 10% and 8% of the river surface area in leakages were extracted for the Aa River section from this model,
summer, while in winter these areas comprise respectively 13% resulting in a spatial average of the leakage flux of 15.3 mm d1.
and 23% (Table 2, Nos. 6 and 9). Thus the increased total ground- This value is significantly lower as estimated with the heat trans-
water discharge (Table 2, No. 2) in winter is mostly caused by an port model (Table 2, No. 1). The spatial distribution of the fluxes
increase of low net discharges (Table 2, No. 10) and its correspond- however is similar, showing high fluxes upstream and low fluxes
ing discharge area (Table 2, No. 9). The areas indicating high net downstream. The lateral extent of the groundwater model is signif-
discharges show however a gentle increase (Table 2, No. 7). icantly larger and the horizontal and vertical resolution of the
The spatially interpolated fluxes can be used to verify ground- model is much lower than the spatial scales of the roaming T-stick
water models or vice versa. Woldeamlak et al. (2007) and Dams surveys. It is therefore likely that the differences in estimated
et al. (2008) developed a groundwater model for the catchment fluxes can be attributed to the differences in resolution and the
in which the Aa River is situated. WetSpass (Batelaan and De steady state character of the groundwater model.
102 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

5. Conclusions of strong changes in weather conditions (e.g. at a drop in temper-


ature with an incoming rain front) is therefore not recommended.
The presented methodology indirectly detects the movement of Preferably, the temperature measurements should be performed
water in the subsurface by using heat as a tracer. The thermal steady during stable weather conditions of three or more days. These in-
state investigation results in an acceptable flux approximation. creases on one hand the accuracy of the method, but also reduces
Moreover, measurements of thermal profiles are easy to perform its flexibility. Temperature variations between entire winter and
(e.g. with the T-stick), very robust, cost-effective and have proven summer seasons of different years (Fig. 6) cannot be avoided what-
their reliability in different applications. A major advantage is cer- soever. We regard these influences on the simulated flux results as
tainly the fact that spatial differences can be detected by performing acceptable with respect to the limitations of the thermal method as
‘roaming surveys’, collecting data from multiple locations. Such a well as in comparison to uncertainties of other methods used for
spatial distribution cannot be achieved with temperature measure- estimation of groundwater–surface water exchange.
ments in boreholes or piezometers, or with most other field methods. For relatively shallow rivers of up to several kilometer length,
Furthermore, practical problems of temperature measurements in we suggest a hierarchical setup of measurement campaigns, where
piezometers can be overcome with a T-stick, including thermal an initial survey using the thermal method can detect places of
instabilities of the water column or problems related to heat trans- special interest (McClain et al., 2003). Consequently, as presented
port along and across the piezometer pipe. Opposite to the high sen- in this article, the survey can be extended by successive investiga-
sitivity of hydraulic gradient measurements, the thermal method tions, allowing detection of seasonal differences or spatial varia-
shows a high robustness regarding vertical misplacement of the tem- tions. Results of the initial survey are guiding continued research
perature sensor. This robustness is explained by the fact that the efforts and help in establishing measurement locations for gauges,
maximum measurement error is small with respect to the whole piezometers, seepage meters, etc. This hierarchical setup assures
model domain. The measurement errors have the largest impact capturing the significant features of groundwater–surface water
when they occur close to the upper boundary. The cumulative mea- interaction and increases the efficiency of the employed
surement error close to the interface however is still small in case a T- instrumentation.
stick like instrument is used. Despite the uncertainties of the thermal From the spatial interpolation for the River Aa, and supported
steady state methodology it can be used as a ‘stand alone’ technique by statistical analysis it is concluded that in winter the absolute
to estimate the groundwater–surface water interaction at first hand. flux of groundwater–surface water exchange is higher than in
If lower uncertainty is required, we recommend thermal modeling in summer. This difference is attributed to the in general higher
transient mode and/or in combination with other methods. The used groundwater head gradients in winter. But, as the surface water
models have small computational requirements, while their input discharge changes significantly with the seasons, the relative con-
parameters have a limited range, which can be assumed from litera- tribution of exfiltration to the river discharge (i.e. the baseflow) is
ture. Table 3 briefly summarizes the advantages and disadvantages higher during the summer season. The bank areas contribute rela-
of the presented methodology. tively more to the exchange than the center part of the river
The steady state approach of Arriaga and Leap (2006) proved to stretch. They also show stronger variability of fluxes. This indicates
be very efficient as the data preparation, model setup and calibra- the importance of the banks in groundwater–surface water inter-
tion was less time consuming than for the STRIVE model. However, action. Local shallow groundwater flow systems and likely higher
STRIVE can also be used for transient simulation (Anibas et al., hydraulic conductivities of the bank zones are thought to be
2009) if continuous data sets are available. Beside the discussed responsible for these phenomena.
use, temperature is an important ecological parameter and moder- The upper reach of the examined stretch of the Aa River showed
ates biological and chemical processes in groundwater and surface in general a higher flux and more heterogeneous flow patterns
water systems; hence the measured temperatures can be used as than the lower reach. Different mechanisms can be detected, the
input parameter for various other models examining biogeochem- hydrogeology of the subsurface of the riverbed (Fig. 4), shallow
ical transport, reaction and retention processes. The thermal steady groundwater flow mechanisms at the banks, deep flow at the cen-
state model performed best with a domain of several meters (e.g. ter of the river channel, the morphology of the river with respect to
5.0 m) thickness with a constant groundwater temperature at the the surrounding alluvium, the topography (Fig. 3) and the bathym-
lower boundary. The fluxes are integrated over the model domain, etry of the riverbed. It is concluded that these difference between
which can explain the difference between the estimated fluxes and the upper and the lower part are the expression of a combination
results of other models or data from piezometer nests, as they have of causes of which topography and hydrogeology are dominant.
different horizontal, vertical and temporal scales. The presented Spatial interpolation techniques can be used to calculate net
models however show similar spatial and temporal trends. discharge fluxes from the riverbed towards the river, on a scale
Since a thermal and hydrological steady state has to be assumed were other methods like differential gauging of surface water dis-
the applicability of the methodology is limited to summer and win- charge are below their detection limit. The estimated fluxes indi-
ter and to locations where no sudden changes of the hydrological cate that the riverbed of the main channel of the whole Aa River
conditions are expected. Diurnal thermal transient influences have account for about 15% of the total river discharge at its outlet. It
an effect at the upper model boundary, regardless the size of the is concluded that the remaining about 60% of the baseflow of the
river system investigated. We confirm Schmidt et al. (2006) find- Aa is produced all over the catchment by diffuse discharge in small
ings that the temperatures of the uppermost section of the river- tributaries, ditches, drains, swamps, etc. The spatial interpolation
bed should not be used for a steady-state thermal analysis. This indicates that the significant increase in net flux between summer
leads to a loss of measured information and limits the model do- and winter is due to an increase in low fluxes as well as an enlarge-
main that can be investigated with an instrument like the T-stick. ment of the zones with these relatively low fluxes. On the other
To maximize the trade-off between temperature information and hand regions with high fluxes increased moderately in extent
diurnal transient influences we propose to use a temperature mea- and flux rates.
sured between 0.10 and 0.20 m below the interface as upper model Eventually, the flux results of the thermal method are useful as
boundary. input for calibration or validation of distributed hydrological or
Corrections for diurnal and seasonal thermal influences can be groundwater flow models. For reducing model uncertainty these
considered before the analysis. This is however not the case for fluxes are complementary to the traditional head calibration data.
the effects of irregular meteorological events. Measuring at times The presented methodology for estimation of fluxes by thermal
C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104 103

modeling proves to be a simple and reliable method, improves the European Commission, 2000. Directive 2000/60/EC of the European Parliament and
of the Council establishing a framework for Community action in the field of
understanding of groundwater–surface water interaction.
water policy. Official Journal of the European Community L327, 1–72.
Fairley, J.P., Nicholson, K.N., 2005. Imaging lateral groundwater flow in the shallow
subsurface using stochastic temperature fields. Journal of Hydrology 321, 276–
Acknowledgements
285. doi:10.1016/j.jhydrol.2005.08.017.
Fleckenstein, J., Anderson, M., Fogg, G., Mount, J., 2004. Managing surface water–
Financial support from the Research Foundation-Flanders groundwater to restore fall flows in the Consumnes River. Journal of Water
(FWO) for the work on the Aa River as part of the project ‘A funda- Resources Planning and Management – ASCE 130 (4), 301–310. doi:10.1016/
(ASCE)0733-9496(2004)130:4(301).
mental study on exchange processes in river ecosystems’ Freeze, R.A., Cherry, J.A., 1979. GroundWater. Prentice-Hall Inc., Englewood Cliffs,
(G.0306.04) is greatly appreciated. Further support was received NJ.
from the Belgian Science Policy Office in the framework of the MA- Harbaugh, A.W., McDonald, M.G., 1996. User’s documentation for MODFLOW-96, an
update to the US Geological Survey modular finite-difference ground-water
MUD (SR/00/105) and FRAC-WECO (SD/TE/02) projects. flow model. US Geological Survey Open-file Report 56, 96–485.
We would also like to thank the numerous assistants for their Hatch, C.E., Fisher, A.T., Revenaugh, J.S., Constantz, J., Ruehl, C., 2006. Quantifying
invaluable help in the field. surface water–groundwater interactions using time series analysis of
streambed thermal records: method development. Water Resources Research
42 (10), 1–5. doi:10.1029/2005WR004787.
References Hayashi, M., Rosenberry, D.O., 2002. Effects of groundwater exchange on the
hydrology and ecology of surface water. Ground Water 40 (3), 309–316.
Hoehn, E., Cirpka, O.A., 2006. Assessing residence times of hyporheic ground water
Anderson, M.P., 2005. Heat as a groundwater tracer. Ground Water 43 (6), 951–968.
in two alluvial flood plains of the Southern Alps using water temperature and
Anibas, C., Fleckenstein, J., Volze, N., Buis, K., Verhoeven, R., Meire, P., Batelaan, O.,
tracers. Hydrology and Earth System Sciences 10, 553–563.
2009. Transient or steady-state? Using vertical temperature profiles to quantify
Kalbus, E., Reinstorf, F., Schirmer, M., 2006. Measuring methods for groundwater–
groundwater–surface water exchange. Hydrological Processes 23, 2165–2177.
surface water interactions: a review. Hydrology and Earth System Sciences 10,
doi:10.1002/hyp.7289.
873–887.
Arriaga, M.A., Leap, D.I., 2006. Using solver to determine vertical groundwater
Kalbus, E., Schmidt, C., Reinstorf, F., Krieg, R., Schirmer, M., 2008. How streambed
velocities by temperature variations, Purdue University, Indiana, USA.
temperatures can contribute to the determination of aquifer heterogeneity.
Hydrogeology Journal 14 (1–2), 253–263. doi:10.1007/s10040-004-0381-x.
Grundwasser 13 (2), 91–100. doi:10.1007/s00767-008-0066-9.
Bal, K., Meire, P., 2009. The influence of macrophyte cutting on the hydraulic
Kalbus, E., Schmidt, C., Molson, J.W., Reinstorf, F., Schirmer, M., 2009. Influence of
resistance of Lowland Rivers. Journal of Aquatic Plant Management 47, 65–68.
aquifer and streambed heterogeneity on the distribution of groundwater
Batelaan, O., De Smedt, F., 2001. WetSpass: a flexible, GIS based, distributed
discharge. Hydrology and Earth System Sciences 13 (1), 69–77.
recharge methodology for regional groundwater modeling. In: Gehrels, H.,
Keery, J., Binley, A., 2007. Temperature Measurements for Determining
Peters, J., Hoehn, E., Jensen, K., Leibundgut, C., Griffioen, J., Webb, B.,
Groundwater–Surface Water Fluxes. Environment Agency-science Report,
Zaadnoordijk, W.J. (Eds.) Impact of Human Activity on Groundwater
SC030155/SR9, Bristol, United Kingdom.
Dynamics. IAHS Publ. 269, IAHS, Wallingford, UK, pp. 11–17.
Keery, J., Binley, A., Crook, N., Smith, J.W.N., 2007. Temporal and spatial variability of
Batelaan, O., De Smedt, F., 2007. GIS-based recharge estimation by coupling
groundwater–surface water fluxes: development and application of an
surface–subsurface water balances. Journal of Hydrology 337 (3–4), 337–355.
analytical method using temperature time series. Journal of Hydrology 336
doi:10.1016/j.jhydrol.2007.02.001.
(1–2), 1–16. doi:10.1016/j.jhydrol.2006.12.003.
Baxter, C., Hauer, F.R., Woessner, W.W., 2003. Measuring groundwater–stream
Lapham, W.M., 1989. Use of Temperature Profiles Beneath Streams to Determine
water exchange: new techniques for installing minipiezometers and estimating
Rates of Vertical Ground-water Flow and Vertical Hydraulic Conductivity.
hydraulic conductivity. Transactions of the American Fisheries 132 (3), 493–
Water-Supply Paper 2337, US Geological Survey.
502.
Lee, D.R., 1977. Device for measuring seepage flux in lakes and estuaries. Limnology
Becker, M.W., 2006. Potential for satellite remote sensing of groundwater. Ground
and Oceanography 22 (1), 140–147.
Water 44 (2), 306–318. doi:10.1111/j.1745-6584.2005.00123.x.
Loheide, S.P., Gorelick, S.M., 2006. Quantifying stream–aquifer interactions through
Becker, M.W., Georgian, T., Ambrose, H., Siniscalchi, J., Fredrick, K., 2004. Estimating
the analysis of remotely sensed thermographic profiles and in situ temperature
flow and flux of groundwater discharge using water temperature and velocity.
histories. Environmental Science and Technology 40 (10), 3336–3341.
Journal of Hydrology 296 (1–4), 221–233. doi:10.1016/j.jhydrol.2004.03.025.
doi:10.1021/es0522074.
Boulton, A.J., Findlay, S., Marmonier, P., Stanley, E.H., Valett, H.M., 1998. The
Lowry, C.S., Walker, J.F., Hunt, R.J., Anderson, M.P., 2007. Identifying spatial
functional significance of the hyporheic zone in streams and rivers. Annual
variability of groundwater discharge in a wetland stream using a distributed
Review of Ecology and Systematics 29, 59–81.
temperature sensor. Water Resources Research 43 (10), W10408. doi:10.1029/
Bredehoeft, J.D., Papadopulos, I.S., 1965. Rates of vertical groundwater movement
2007WR006145.
estimated form the Earth’s thermal profile. Water Resources Research 1 (2),
Marechal, R., Laga, P. (Eds.), 1988. Voorstel Lithostratigrafische Indeling van het
325–328.
Paleogeen. Nationale Commissie Voor Stratigrafie, Commissie Tertiair, p. 208.
Brunke, M., Gonser, T., 1997. The ecological significance of exchange processes
McClain, M.E., Boyer, E.W., Dent, C.L., Gergel, S.E., Grimm, N.B., Groffmann, P.M.,
between rivers and groundwater. Freshwater Biology 37 (1), 1–33.
Hart, S.C., Harvey, J.W., Johnston, C.A., Mayorga, E., McDowell, W.H., Pinay, G.,
Cardenas, M.B., Zlotnik, V.A., 2003. Three-dimensional model of modern channel
2003. Biogeochemical hot spots and hot moments at the interface of terrestrial
bend deposits. Water Resources Research 39 (6), 1141. doi:10.1029/
and aquatic ecosystems. Ecosystems 6 (4), 301–312.
2002WR001383.
Murdoch, L.C., Kelly, S.E., 2003. Factors affecting the performance of conventional
Carey, S.K., Quinton, W.L., 2005. Evaluating runoff generation during summer using
seepage meters. Water Resources Research 39 (6), 1163. doi:10.1029/
hydrometric, stable isotope and hydrochemical methods in a discontinuous
2002WR001347.
permafrost alpine catchment. Hydrological Processes 19 (1), 95–114.
National Research Council (NRC), 2004. Groundwater Fluxes Across Interfaces.
Cey, E.E., Rudolph, D.L., Parkin, G.W., Aravena, R., 1998. Quantifying groundwater
National Academic Press, Washington, DC, p. 85.
discharge to a small perennial stream in southern Ontario, Canada. Journal of
Rosenberry, D.O., 2008. A seepagemeter designed for the use in flowing water.
Hydrology 210, 21–37.
Journal of Hydrology 359 (1–2), 118–130. doi:10.1016/j.jhydrol.2008.06.029.
Chen, X.H., 2000. Measurement of streambed hydraulic conductivity and its
Saenger, N., Kitanidis, P.K., Street, R.L., 2005. A numerical study of surface–
anisotropy. Environmental Geology 39 (12), 1317–1324.
subsurface exchange processes at a riffle-pool pair in the Lahn River, Germany.
Commission, European., 2006. Directive 2006/118/EC of the European Parliament
Water Resources research 41, W12424. doi:10.1029/2004WR003875.
and of the Council on the protection of groundwater against pollution and
Schmidt, C., Bayer-Raich, M., Schirmer, M., 2006. Characterization of spatial
deterioration. Official Journal of the European Community L372, 19–31.
heterogeneity of groundwater–stream water interactions using multiple
Conant Jr., B., 2004. Delineating and quantifying ground water discharge zones
depth streambed temperature measurements at the reach scale. Hydrology
using streambed temperatures. Ground Water 42 (2), 243–257.
and Earth System Sciences 10, 849–859.
Constantz, J., 2008. Heat as a tracer to determine streambed water exchanges.
Schmidt, C., Conant Jr., B., Bayer-Raich, M., Schirmer, M., 2007. Evaluation and field-
Water Resources Research 44, W00D10. doi:10.1029/2008WR006996.
scale application of an analytical method to quantify groundwater discharge
Constantz, J., Niswonger, R.G., Steward, A.E., 2008. Analysis of temperature
using mapped streambed temperatures. Journal of Hydrology 347, 292–307.
gradients to determine stream exchanges with ground water. In: Rosenberry,
doi:10.1016/j.jhydrol.2007.08.022.
D.O., LaBaugh, J.W. (Eds.), Field Techniques for Estimating Water Fluxes
Schmidt, C., Kalbus, E., Krieg, R., Bayer-Raich, M., Leschik, S., Reinstorf, F.,
Between Surface Water and Ground Water. Techniques and Methods, US
Martienssen, M., Schirmer, M., 2008. Schadstoffmassenströme zwischen
Geological Survey Report (4-D2).
Grundwasser, Flussbettsedimenten und Oberflächenwasser am regional
Dams, J., Woldeamlak, S.T., Batelaan, O., 2008. Predicting land-use change and its
kontaminierten Standort Bitterfeld. Grundwasser 13, 133–146. doi:10.1007/
impact on the groundwater system of the Kleine Nete catchment, Belgium.
s00767-008-0076-7.
Hydrology and Earth System Sciences 12, 1369–1385.
Schön, J.H., 1998. Physical Properties of Rocks, Fundamentals and Principles of
Domenico, P.A., Schwartz, F.W., 1998. Physical and Chemical Hydrogeology, second
Petrophysics, second ed. Pergamon.
ed. John Wiley and Sons Inc., New York.
Shaw, R., Prepas, E., 1990. Groundwater–lake interactions: I. Accuracy of seepage
DOV, 2010. Databank Ondergrond Vlaanderen. Database Subsurface of Flanders.
meter estimates of lake seepage. Journal of Hydrology 119, 105–120.
<https://dov.vlaanderen.be/>.
104 C. Anibas et al. / Journal of Hydrology 397 (2011) 93–104

Smith, J.W.N., 2005. Groundwater–Surface Water Interactions in the Hyporheic Stonestrom, D.A., Constantz, J., 2003. Heat as a Tool for Studying the Movement of
Zone. Environment Agency-science Report SC030155/SR1, Bristol, United Groundwater near Streams. USGS Circular 1260, Reston, Virginia.
Kingdom. Suzuki, S., 1960. Percolation measurements based on heat flow through soil with
Soetaert, K., De Clippele, V., Herman, P., 2002. FEMME, a flexible environment for special reference to paddy fields. Journal of Geophysical Research 65 (9), 2883–
mathematically modeling the environment. Ecological Modelling 151 (2–3), 2885.
177–193. Woessner, W.W., 2000. Stream and fluvial plain ground water interactions:
Stallman, S., 1965. Steady one-dimensional fluid flow in the semi-infinite porous rescaling hydrogeologic thought. Ground Water 38 (3), 423–429.
medium with sinusoidal surface temperature. Journal of Geophysical Research Woldeamlak, S., 2007. Spatio-temporal Impacts of Climate and Land-use Changes
70 (12), 2821–2827. on the Groundwater and Surface Water Resources of a Lowland Catchment. PhD
Stanford, J.A., 1998. Rivers in the landscape: introduction to the special issue on Thesis, Vrije Universiteit Brussel, Brussels, Belgium.
riparian and groundwater ecology. Freshwater Biology 40, 402–406. Woldeamlak, S., Batelaan, O., De Smedt, F., 2007. Effects of climate change on the
Stonestrom, D.A., Blasch, K.W., 2003. Determining Temperature and Thermal groundwater system in the Grote-Nete catchment, Belgium. Hydrogeology
Properties for Heat-based Studies of Surface-water Ground-water Journal 15 (5), 891–901. doi:10.1007/s10040-006-0145-x.
Interactions. USGS Circular 1260, Reston, Virginia, pp. 73–80 (Appendix A).

You might also like