You are on page 1of 23

Surv Geophys (2008) 29:247269

DOI 10.1007/s10712-008-9051-1
ORIGINAL PAPER

Monitoring Continental Surface Waters by Satellite


Altimetry
Stephane Calmant Frederique Seyler Jean Francois Cretaux

Received: 30 March 2008 / Accepted: 9 December 2008 / Published online: 24 January 2009
Springer Science+Business Media B.V. 2009

Abstract The monitoring of continental water stages is a requirement for meeting human
needs and assessing ongoing climatic changes. However, regular gauging networks fail to
provide the information needed for spatial coverage and timely delivery. Although the
space missions discussed here were not primarily dedicated to hydrology, 18 years of
satellite altimetry have furnished complementary data that can be used to create hydrological products, such as time series of stages, estimated discharges of rivers or volume
change of lakes, river altitude profiles or leveling of in situ stations. Raw data still suffer
uncertainties of one to several decimeters. These require specific reprocessing such as
waveform retracking or geophysical correction editing; much work still remains to be
done. Besides, measuring the flow velocity appears feasible owing to SAR interferometer
techniques. Inundated surfaces, and the time variations of their extent, are currently almost
routinely computed using satellite imagery. Thus, the compilation of the continuous efforts
of the scientific community in these various investigative directions, such as recording
from space the discharges of rivers or the change in water volume stored in lakes, can be
foreseen in the near future.
Keywords

Satellite altimetry  Hydrology  Rivers and lakes

1 Introduction
The radar altimeters employed on altimetry mission satellites transmit a short microwave
pulse in the nadir direction, and the echo reflected by the surface is examined. The time for
the pulse to be reflected back to the altimeter corresponds to the distance (or range)
travelled by the electromagnetic pulse between the satellite and the Earths surface,
S. Calmant (&)  J. F. Cretaux
LEGOS, UMR CNRS/UT3/IRD/CNES, Universite Paul Sabatier, UT3, 31400 Toulouse, France
e-mail: stephane.calmant@ird.fr
F. Seyler
LMTG, UMR CNRS/UT3/IRD, Universite Paul Sabatier, UT3, 31400 Toulouse, France

123

248

Surv Geophys (2008) 29:247269

assuming that the pulse is propagating at the speed of light. The height H of the reflecting
surface is determined by the difference between the satellite orbit (Alt) and the altimeter
range R measurement, with the addition of various corrections that take into account the
time delays related to the propagation of the pulse through the atmosphere and the ionosphere and, lastly, correction for solid tidal effects on the Earth:
H Alt  R DTC WTC IC Ts 

where DTC is the Dry Tropospheric Correction, WTC the Wet Tropospheric Correction,
IC the Ionospheric Correction and Ts the solid tide correction. In Eq. (1), H is considered
with respect to a mathematical reference ellipsoid, similar to that used for satellite orbits. A
full discussion of the derivation of altimetric heights and their associated errors can be
found in Fu and Cazenave (2001). Hydrology often requires that elevation be expressed in
terms of altitudes instead of ellipsoidal height. In that case, H must be corrected for a geoid
model.
So far, not a single satellite altimetric mission has been designed primarily for monitoring continental waters. However, the radar altimetric missions TOPEX/Poseidon (T/P
hereafter), ERS 1 and 2, Jason, GFO and ENVISAT have registered data over the lakes and
the great river basins of the world for about 18 years. How these data could be useful to the
hydrological community, what their actual drawbacks are, how they can be resolved in
time, and how to evaluate their values for different kinds of applications, are all still open
questions. Clearly, the hydrological products derived from altimetry data can be improved,
but it may also be necessary to investigate the usual hydrological practices to make these
products more adequate for hydrology purposes.
The issues at stake concerning the continental water monitoring archive are particularly important and will be reviewed in the first section of this article. From the first report
published by NASA on radar altimetry, in 1969, to the current studies involving radar
altimetry for river and wetland storage and various flow studies, the relentless efforts of
the scientific community will also be reviewed in the second section. In order to provide
the hydrologists community with adequate products, teams of geodesists processed such
data and developed databases accessible via the Internet; these are also presented.
Technical issues such as re-tracking, correction algorithms, multi-mission temporal series
building, and error and uncertainty estimations are also presented. Applications of interest
are discussed in the last section of this article. This article also presents the major
advantages and drawbacks of hydrometric data used in hydrology today, either in situ or
spaceborne.

2 The Issues at Stake


Water resources on the continental surface are limited, and the spatial and temporal distribution of this resource does not always meet the most crucial needs. Water resources readily
available for human consumption and for ecosystems are found in lakes and rivers, correspond to only 0.27% of the fresh water and about 0.007% of the total amount of water in the
world (Agencia National das Aguas, http://hidroweb.ana.gov.br/doc/WRMB/part1.htm).
The quantity and quality of the fresh water supplies will be a major problem in the coming
decades. The world consumption of fresh water reached 54% of the total stock in 1995. It is
likely to equal the total fresh water resource available in North Africa and South Asia by
2025, when Asia will be using ten times more water than the rest of the world (World
Resources Institute 2000http://www.wri.org/wri/wr).

123

Surv Geophys (2008) 29:247269

249

One billion people lack sufficient water for domestic consumption today, and it is
estimated that, in 30 years, 5.5 billion people will be living in areas with moderate to
serious water shortages (Population Reference Bureau 1997). In the 1970s, the emergence
of environmentalism, followed by the adoption of the sustainability principle, in combination with the threat of a worldwide water shortage in the years to come, has led a number
of countries to completely revise their strategies and governance tools applied to the
integrated management of water resources. The monitoring of continental surface water
stored in rivers, lakes, and wetlands has turned into the scientific hydrologist communitys
primary objective and a major concern for a majority of state organizations.
Unfortunately, our knowledge of the global dynamics of terrestrial surface waters is still
very poor (SWOT, Satellite Water Ocean Topography, http://www.geology.ohio-state.edu/
water/). Yet, studies of the integrated, global nature of the hydrological cycle are essential
to our understanding of natural climate variability and to predict a climatic response to
anthropogenic forcing (Koster et al. 1999). Most international water management groups
underline the need for core hydrological data. Nowadays, monitoring of the continental
water resource (temporal variability of river stages and river discharges) is provided via
hydrologic networks. These networks are organized on a national basis. The challenges
common to most regions include inadequate monitoring networks, gaps in records, a
general decline in the number of stations, chronic under-funding, differences in processing
and quality control, and differences in data policies (WMO 2003). Thus, major issues in the
poorer regions of the world include poor status or outright lack of monitoring networks,
support infrastructure and data quality. For example, the WMO first identified those
National Hydrological Networks in Africa that still maintain a hydrological data archive on
paper. Replies to a questionnaire sent to Hydrological Advisors in 39 countries showed that
82% use paper for archiving their data (WMO 2004), which means that they are not
available to the scientific community or the decision makers.
Maintaining a hydrological network in remote basins is a costly task. To give the
example of the Brazilian part of the Amazon basin network, managed by ANA (Agencia
Nacional de Aguas) and other federal or state entities, there is on average one limnimetric
station for each 7,200 km2, About 270 limnimetric stations are managed by ANA in the
Amazon basin, from which only 210 have been regularly gauged, and only 61% are still
operating today. The cost of this network operation is about 3 million dollars per year, a
huge expense for the country.
Moreover, the question of water resources must be analyzed in the scope of the
changing climate. River and wetland water stages are integrated measurements of the basin
response to climatic stress. The availability of raw data is a key issue here. Presently, these
measurements are much more reliable than the spatially averaged estimation of rain,
evapo-transpiration and infiltration. For example, Roads et al. (2003) compared climate
model outputs with ground-truth data over the continental United States. Using predictions
from various climate models, they found that the runoff predictions are often in error by
50%, and even mismatches with observations as high as 100% were not uncommon. Coe
(2000) found similar results for many of the worlds large river basins.
Thus, these models are now becoming limited as a result of the decline in observations
of water discharge and water storage (Alsdorf et al. 2003). As rivers, lakes and wetlands
form the main fresh water resource, there is a need for a global, homogeneous, continuing
(over many years) monitoring system which delivers fast-access data on continental water
stages and, wherever possible, on water volumes and river discharges too. We now
examine how satellite altimetry can or could bring some responsesalthough partialto
the problem of monitoring water stages worldwide.

123

250

Surv Geophys (2008) 29:247269

3 Pioneering Works
In 1969, NASA established the long-term objectives of radar spatial altimetry from space,
reporting the activities of the Williamstown group on solid Earth and ocean physics (Kaula
1969). In a report on SKYLAB altimeter results, Brown (1977) analyzed the backscattering
and typical waveforms of the Salt Lake of Utah. Miller (1979) showed applications for
determining the water stages of continental lakes from measurements collected by the
GEOS-3 radar altimeter. Pioneering studies used the data collected by Seasat in 1978
(Brooks 1982) and Geosat from 1986 to 1988 (Morris and Gill 1994a), demonstrating the
feasibility of using data from ocean-designed missions for land waters. Mason et al. (1985)
conducted a study of the ability of ERS1 data to measure water stage variations in relation
to climate change.
On continental surfaces, radar altimetry left the domain of prospective studies with the
first studies dedicated to the Great Lakes in the United States, using SEASAT data (Brooks
1982), Geosat (Morris and Gill 1994a), and then T/P (Morris and Gill 1994b; Birkett
1995a). The African Great Lakes have also been the subject of various studies (Birkett
1995b; Cazenave et al. 1997; Ponchaut and Cazenave 1998; Birkett et al. 1999; Mercier
2001; Mercier et al. 2002). The first studies of great river basins began with Koblinsky
et al. (1993) who searched for specular waveforms in the GEOSAT data to estimate the
water levels on four sites on the Amazon. They estimated a 70 cm rms discrepancy
between satellite and in situ measurements, partly attributed to uncertainties in the in situ
record, but mostly due to poor modeling of the orbit. Orbit errors are decreasing with the
new generations of satellite. From the aforementioned 50 cm error estimated for GEOSAT
by Koblinsky et al. (1993), the uncertainty in the radial component of satellite orbits is now
estimated to be 15 cm for ERS-1 and 3 cm for T/P (Le Traon et al. 1995). Now, data
currently used to establish the time series of water stages over the worlds rivers and lakes
originate from the T/PJason series (10-day repeat period, respectively since 1992, 2002
and 2008), ENVISAT (following the ERS series started in 1991 with a 35-day repeat
period) missions and, to a lesser extent, the GFO mission. Time series of river stages may
cover a decade (Birkett 1995b, 1998; De Oliveira Campos et al. 2001; Maheu et al. 2003).
Satellite altimetry has also been used to determine the height profiles of rivers. Cudlip
et al. (1992) used SEASAT data to establish an Amazon water profile. They used 32
crossings of the river by the satellite tracks over a 18-day period in July, 1978. These
authors estimated the water level accuracy to be between 10 and 20 cm. More recently,
Birkett et al. (2002) published time series of stage fluctuations over the Amazon basin
using T/P measurements between 1992 and 1999. It was concluded that accuracy might
range from tens of centimeters to several meters (1.1 m average rms). These authors found
that the water-surface gradient of the main stem varies both spatially and temporally, with
values ranging from 1.5 cm/km downstream to 4.0 cm/km for more upstream reaches.
Overall, they demonstrated that altimetric data from the T/P mission could monitor the
transient flood waves of this continental-scale river basin.

4 Available Database for Hydrology


At present, various databases enable the retrieval of time series of water stages of the great
basins of the world. The first database came from the processing of ERS and ENVISAT at
de Montfort University, UK, on behalf of ESA. Two types of product are currently
available: the River Lake Hydrology product (RLH), designed for hydrologists with no

123

Surv Geophys (2008) 29:247269

251

prior knowledge of radar altimetry and organized in accordance with river/lake crossing
points, with each product corresponding to one crossing point, and the River Lake
Altimetry (RLA) product, for radar altimetry experts. It is structured around orbit or by 35day cycle and provides all crossing points for a predefined region (European Space
Agency, http://earth.esa.int/riverandlake/description.htm). Another database is Hydroweb
(http://www.legos.obs-mip.fr/soa/hydrologie/hydroweb/), created in LEGOS (Laboratoire
dEtudes en Oceanographie et Geodesie Spatiale), Toulouse, France. It collects river
observations (*200 sites mainly using T/P data) and *150 lakes and reservoirs in the
world in a multi-mission processing of the data from ENVISAT, GFO, T/P and Jason. The
third one is the Global Reservoir and Lake Elevation Database. It is maintained by the US
Department of Agriculture (http://www.pecad.fas.usda.gov/cropexplorer/global_reservoir/).
It mainly consists of time series over lakes and reservoirs using the T/P-JASON suite.
Owing to the use of Interim data, e.g., data not fully checked and delivered by the Agencies
before the final product is available, this site is able to propose Near Real Time (NRT)
values. Yet, because it is restricted to the T/P-JASON suite, this database suffers from a
lack of measurements for many sites given the low performance of JASON-1 over continental waters. The CASH project (Contribution of Spatial Altimetry to Hydrology)
funded by the French Ministry of Research and Technology has produced a retracked T/P
database (http://www.hydrospace-cash.fr/). This product consists of a water level time
series for each river crossing for eight great basins in the world, expressed with respect to
the geoid (EGM96), with the frequency of one mean point at every overflight (i.e., every
10 days at best for Topex/Poseidon). The radar ranges used to compute water stages for
this database had their echoes being re-processed using various algorithms, called trackers.
These trackers were those already used routinely by ESA to process the ENVISAT radar
measurements.

5 Issues Affecting the Quality of Altimetry Data for Continental Waters


and Error Budgets
The altimetry datasets used by land water investigators were primarily intended and processed for other scientific purposes, namely to measure heights of the ocean surface (T/P,
Jason, GFO, part of ERS 1 and 2, and ENVISAT) or ice caps (ERS 1 and 2, and ENVISAT). Using these data turns out to be problematic, in particular for rivers and to a lesser
extent for lakes. Here, we now consider the major points and methods followed by
investigators to overcome them.
The basic measurement of nadir altimetry is the two-way travel time of a pulse emitted
by the antenna onboard the satellite and bounced back by the Earths surface. The precision
of the measurement will then strongly depend on the capability to pick up (within the time
window) the radar echo that corresponds to the actual height of the reflector at the nadir of
the antenna. This time is to be found within the spread of the energy received at the
satellite antenna (Fig. 1). For all the current missions, the energy returns of the individual
echoes emitted at about 12 kHz are averaged between 10 and 18 Hz (*100 measurement
packets) to form the so-called waveforms that are distributed to the scientific community as
SDRs (Sensor Data Records). The shape of the return waveforms over rivers is highly
variable, different from the consistent return waveforms over oceans (Fig. 2). Smith (1997)
stated that the major difficulty in retrieving ranges over continental waters results from the
variability in shape of the return waveforms when onboard trackers are designed for ocean
waveforms.

123

252

Surv Geophys (2008) 29:247269

Fig. 1 Position of the pulse relative to the reflecting surface and energy received by the satellite. The
vertical axis is positioned at the time of emission of the pulse. The dotted line stands for the theoretical time
giving the correct two-way travel time to be used for the range (half amplitude of the peak)

Regular ocean-type trackers expect long tail shapes of energy distribution (view A in
Fig. 3) whereas the echoes bouncing off rivers are often specular (Guzkowska et al. 1990)
or a combination of specular echoes (view B in Fig. 3). Thus, in the best case, energy is
received but the range estimate is erroneous or not estimated (views C and D in Fig. 3); in
the worst case, the altimeter loses tracking and subsequent echoes are lost. The antennareflector range is determined by fitting the waveform with a predefined analytical function
(called waveform tracking). If the analytical function is not well suited to the waveform
shape, this tracking leads to wrong estimates of the height value (or even no estimate at
all). It is worth noting that the radar altimetry data collected by the ongoing ENVISAT
mission are nominally retracked with four algorithms. In turn, retracking the radar
waveforms collected by the ERS 1 & 2, T/P, Jason, and GFO missions requires that some
retracking procedures be conducted. Berry et al. (2005) showed that, by retracking multimission altimeter data with tuned filters, high quality height data can be obtained from the
vast majority of lakes with surface areas greater than 500 km2. Results from the Amazon
Basin also show the great improvement in both quantity and quality of data obtained. The
retracking of the T/P data with ENVISAT algorithms that has been performed by CLS, a
subsidiary of the French Space agency CNES (Centre National dEtudes Spatiales), for the
CASH project has provided promising results in terms of accuracy improvement and
recovery of data that are missing in the Topex/Poseidon data distributed to users as Merged
Geophysical Data Records (MGDRs, Mercier and Zanife 2006). More in-depth information about this key point of tracking algorithms can be found in Calmant and Seyler (2006)
and the references given there.
The second issue is the problem of off-nadir reflections, in particular the hooking effect.
This effect occurs when the ground area illuminated by the radar beam at the nadir of the

123

Surv Geophys (2008) 29:247269

253

Fig. 2 Samples of Topex waveforms over rivers in the Amazon basin. a Ocean-like waveform processed
by the onboard tracker, e.g., the algorithm implemented onboard the satellite to provide a range estimate.
b Multi-peak waveform rejected by the tracker. c Specular waveform processed by the tracker. d Specular
waveform, similar to (c), rejected by the tracker

antenna is bouncing back little energy in comparison to the energy bounced back by a
water surface at the edge of the radar ground footprint. Thus, the latter is dominant in the
echo returned although lessened by the antenna pattern. The tracking procedure will then
determine the range by fitting this return of energy that traversed a skewed path. Because
the range estimate assumes that the target is at the satellite nadir, this leads to an overestimated range, i.e., to an underestimate of the height of the reflecting water surface. This
is not uncommon in continental waters and has to be carefully taken into account because
these measurements can be erroneously interpreted (Frappart et al. 2006a, b). The
geometry of such a mis-measurement is depicted in Fig. 4 and an example of a real case is
given in Fig. 5.
The third geometrical effect impacting accuracy is the slope effect. Indeed, within a
radar footprint a few kilometers wide, the river height can change significantly. This
changes the waveform shape and affects range determination. Also, in braided reaches,
branches may have different heights, which will spread the return time of energy and
contribute to an erroneous range determination.
The width of the radar beam of ocean-oriented missions is several kilometers since a
surface average is needed over oceans to reduce as much as possible height measurement
corruption due to wind driven waves. The drawback for hydrology studies is that echoes
over rivers whose width is less than, say, 1 kmi.e., most of themare polluted by energy

123

254

Surv Geophys (2008) 29:247269

Fig. 3 Waveforms over different water bodies: an example of waveforms over oceanic surfacesshowing
a high similarity along the trackis presented in the left-hand figure. An example of waveforms collected
over the Amazon basin, at the confluence of the Solimoes (waters in blue in the false color Landsat image)
and Negro river (waters in black in the Landsat image) is presented in the right-hand figure. This example
highlights the high variability in the shape of the waveforms making necessary that echo retracking is
performed on a case basis. Redrawn after Mercier and Zanife (2006)

Fig. 4 Geometry of the hooking effect. When the satellite is at location A on its orbit, i.e., not yet flying
over the water body, radar energy reaches the water body on the borders of the footprint. If the energy
bounced back by this water body dominates the total energy returned to the satellite antenna (i.e., when the
ground at the nadir is poorly reflecting), the retracking procedure will estimate the two-way travel time by
fitting this peak of energy, estimating a slant distance (h0 ) instead of a nadir one (h). The true height, i.e., the
one at the nadir will be obtained only when the satellite is a location B on its orbit. The resulting along-track
heights exhibit a characteristic parabola pattern similar to the one presented in Fig. 5

reflected by riverbanks or islets. The heterogeneity of the reflecting surface is estimated to


be the first cause of error in terms of magnitude (Rosmorduc et al. 2006).
Lastly, the air density, the amount of water vapor and the electrons in the ionosphere
modify the travel time of radar waves throughout the atmosphere. The electron content in
the ionosphere and air pressure are given by independent datasets. However, the amount of
water vapor is estimated using microwave radiometers employed onboard together with the
radar altimeters. Current microwave radiometers fail to estimate the atmospheric content of
water vapor over continents because the signature of the atmospheric water vapor is mixed
with that of ground wetness. Thus, this effect, ranging from a few centimeters to some tens
of centimeters, cannot be accurately corrected for over land waters and corrections can
only be estimated from large scale global datasets such as ECMWF or NCEP. Yet, these

123

Surv Geophys (2008) 29:247269

255

Fig. 5 Example of the hooking effect in the altimetry data (ENVISAT) over a small river in the Amazon
basin (Igarape Nelson Pinheiro, Rio Negro sub-basin). Left view (a) Map of the crossing between
N. Pinheiro river and ENVISAT track (blue dots). The green polygon stands for the window from which
measurements were extracted to be plotted in the right figure. The river appears in white in the JERS mosaic,
i.e., highly energetic echoes, since the water echoes are coherently double bounced off of the tree trunks
surrounding the river. Right view (b) cross section displaying where the ENVISAT passes (each black line)
over the river and adjacent banks. Note that the capture of echoes bounced by the river surface before and
after the satellite passes over the river itself (green polygon) turns into a parabolic apparent reflector under
the banks on both sides of the river. The data within the green polygon are marked as red dots in the lefthand view

model-based correctionsalthough satisfactory in most casescan suffer from large


errors, in particular in the case of mountain lakes. Mercier and Zanife (2006) have shown
that the errors due to changes in the altitude of the reflecting surface (and thus the thickness
of the atmosphere column) are not taken into account. They also demonstrated that a
computation of the atmospheric corrections such as the DTC based on the use of a Digital
Elevation Model to estimate the altitude of the reflecting surface was not adequate. They
proposed a method where the altitude of the reflecting surface is deduced from the altimetric measurement itself. Cretaux et al. (2008) have shown in the case of Lake Issykulkul that the Dry Tropospheric Corrections delivered by the T/P, ENVISAT, JASON and
GFO Geophysical Data Records (GDRs) were wrong (Fig. 6), either because the altitude of
the lake was not accounted for in the computation of the reference air pressure or because
this altitude varies erroneously within the Lake surface in the DTM used for the computation of this reference air pressure, producing an artificial variation of the correction along
the tracks crossing the lake, and hence erroneous height variations over the lake surface. In
this case, the largest difference between corrections on a given day exceeds 60 cm. Such an
error is a major issue when merging water stages derived from different altimetry missions
in order to increase the sampling frequency.
From Eq. (1), it is clear that the quality of the orbit directly enters into the error budget
of the water height. Yet, orbit modeling has dramatically improved over the last decade
(Fig. 7). Thus, in view of the aforementioned error sources, this error can be considered of
less importance for continental waters. This fact is used by web services such as
Cropexplorer to propose NRT preliminary values of water height using interim orbits,
given that the water height finally computed with a more refined orbit will not differ
significantly in regard to the overall uncertainties.

123

256

Surv Geophys (2008) 29:247269

Fig. 6 Model-derived dry tropospheric correction at lake Issyk-kul (After Cretaux et al. 2008). The pink
band stands for the range of DTC expected from the records of a meteorological station on the lake shore

Fig. 7 Improvement in orbit error from GEOS-3 in 1975 to JASON and ENVISAT in 2002

Several studies have dealt with the estimation of the error budget in the altimetric series
of continental water stages. In particular, Birkett et al. (2002) compared T/P altimetric
series with in situ series. Discrepancies up to several meters have been reported in this
study. However, it should be pointed out that such comparisons suffer severe limitations.
On the one hand, the satellite track usually fails to cross the river directly over a gauging
station. Comparing both series assumes that the water stage varies identically in both
places, but this is often not true. Indeed, water stage variations along a reach are affected
by changes in the cross section geometry. Also, flow routing should be performed to
account for the delay from the location of the gauge and that of the satellite track when the
distance between both locations is significant with respect to flow velocity and sampling
rate. Typically, the delay is 1 daye.g., the sampling rate of gauging stationsfor a flow
to cover 85 km with a 1 m/s wave. Given that water stages tend to vary by several

123

Surv Geophys (2008) 29:247269

257

centimeters or decimeters within a day, the error induced when the delay is not taken into
account can contribute significantly to the discrepancy between both series. On the other
hand, when the gauge is not leveled, scatter between altimetric and in situ series is
computed with reducedzero mean series, and the resulting estimate of the discrepancy
does not take into account possible biases, such as the tracker-dependent bias. Roux et al.
(2008) investigated ways of estimating errors in virtual stations through the use of interpolated series, i.e., using more than one in situ gauging station and taking into account the
propagation time between the in situ and virtual gauges. They found that the random error
is roughly at the decimeter level except for low water, where the error can be *1 m or
more. This result is also consistent with the results presented by Bercher et al. (2006).
Alternatively, the accuracy of altimetric measurements can be assessed by means of a
comparison with GPS ground truth height measurements right under the satellite track.
Although this technique is widely used to assess the accuracy of altimetric data over
oceanic surfaces (Bonnefond et al. 2003, among others), this kind of fieldwork has seldom
been conducted over rivers and the results published to date, as in Frappart et al. (2006a,
b), are too limited to allow definite conclusions to be drawn as to the determination of the
error budget in altimetric series using this approach. Although intrinsic system errors of
radar measurements are the same over land waters and oceans, e.g., centimetric accuracy,
the overall uncertainty of altimetric measurements over continental waters is now a couple
of decimeters (Birkett et al. 2002; Frappart et al. 2006a, b).

6 Temporal Sampling
The time step of the altimetric series is given by the orbit repeat period. For the current
radar altimetric missions, this period ranges from 10 days for T/P and Jason to 35 days for
ERS-2 and ENVISAT. GFO has an intermediate repeat period of 17 days. In terms of
sampling rate, this is much lower than the time step of in situ measurements usually
collected once or twice a day, or even much more often, every 15 min in automated
networks of developed countries. Bercher et al. (2006) evaluated the amount of information lost due to this under-sampling.
To improve this time step, multi-mission series can be constructed. This is often possible over large lakes that are frequently crossed by several tracks and/or several missions.
For example, JASON, ENVISAT and GFO altogether have overflown Lake Victoria about
20 times each month since 2002 (Fig. 8). Moreover, stage variations of most lake levels do
not vary significantly on a daily basis. Thus, taking into account all the available passes
should lead to a highly satisfactory time sampling of some dozens of the worlds largest
lakes. As far as rivers are concerned, the situation is more critical. In some fortunate
places where satellite tracks intersect over a river reach, a multi-track virtual station can
be created. An example of such a series is given in Fig. 9. These combined series raise the
issue of the type of errors in the different data series. Indeed, the error budget of each
mission includes biases that must be accounted for prior to combining the data from
different missions.
Another method (Fig. 10) consists of interpolating satellite-based altimetry based on a
linear model exploiting data at a limited number of in situ limnimetric stations in order to
obtain time-series with a 1-day sampling period (Roux et al. 2008). The precision of
interpolation has been investigated in terms of methodology, sensitivity to model rapid
stage variations, robustness with regard to missing values and the effect of random noise.
An optimisation method based on a multi-objective criterion (OPT) furnishes the best

123

258

Surv Geophys (2008) 29:247269

Fig. 8 a Satellite tracks crosscutting the surface of Lake Victoria, including GFO, T/P and JASON,
ENVISAT and ICEsat. The color-coded surface in the background are the ellipsoidal height variation due to
gravity used to separate the geographical and temporal variations in the altimetry measurements along the
satellite tracks. b Temporal variations. The color coding of the dots is as follows: GFO in blue, T/P in pink,
JASON-1 in black, ENVISAT in red, ICEsat in yellow. The green line stands for in situ measurements

Fig. 9 Example of multi-satellite time series at a crossover formed by the tracks of the ERS and T/P
missions over the Rio Negro. Note that the three independent series have been adjusted for their relative
biases in order to produce a consistent series

absolute results, as it is the method that is least sensitive to missing values and random
noise, two factors that systematically affect radar altimetry data. Results also show that
taking into account more than one in situ reference station significantly decreases the RMS
errors in the predicted stages. Taking into account time shifts between stations improves
the results too.

123

Surv Geophys (2008) 29:247269

259

Fig. 10 Example of time series at an ENVISAT virtual station. The daily values are predicted using both
the ENVISAT series and readings at remote stations

7 Applications
A range of applications derived from the altimetric measurements of continental water
levels has been performed or is currently underway. A brief review of these applications is
now given.
7.1 Levelling of Hydrological Network Gauge Stations
A number of great river basins are located in remote geographical areas. These basins
include hydrological gauges, where topographic leveling has not been carried out or, in
some cases, leveling uncertainty is too high due to the difficulties in conducting the
conventional processes of terrestrialspiritleveling. By way of an example, for the
Amazon basin, most hydrological gauges from the ANA network (Brazilian National
Agency for Water) are located outside the topographic leveling routes of the IBGE
(Brazilian Institute for Geography and Statistics). Additionally, gauge stations within the
hydrographic basin of the Amazon and in the neighboring countries of Brazil are all
unleveled, apart from the Iquitos station in Peru. This is a major drawback for hydrodynamical modeling of the basin, since these kinds of models require that the hydrographic
parameters of the river, such as bed slope, be entered into the model using a common
altitudinal reference. Cauhope (2004) leveled gauges in the Curuai Varzea and the adjacent
Amazon reach using ENVISAT time series. For the Tapajos River, a tributary of the
Amazon River, Calmant and Seyler (2004) showed that ICESat measurements are particularly well-suited for gauge leveling. Kosuth et al. (2006) established altimetric
levelling from Topex Poseidon for 97 hydrometric stations along 27,740 km of the

123

260

Surv Geophys (2008) 29:247269

Amazon hydrographic network. Validation has been undertaken for 23 stations, comparing
altimetric data with leveling values obtained from bi-frequency GPS positioning. However,
leveling of other gauge networks worldwide still has to be conducted.
7.2 Estimation of Discharge from Stage Altimetric Measurements
Dense stage and discharge estimations have many uses. For example, discharge values are
essential for water management, extreme flow prediction, and hydrological modeling. Rain
gauges within remote watersheds are often sparsely distributed, although the spatially
averaged estimation of rain is an important parameter for Global Climate Models. As
previously pointed out, a better distribution of the integrated response of the basin to the
incoming rain would allow a better validation of these models. Better-distributed discharge
values could also be used to constrain models of weathering processes and carbon flux
estimations.
Several works have examined the ability of spatial data such as altimetry or imagery to
retrieve river discharge. Recently, Bjerklie et al. (2005) has estimated in-bank river discharge on the basis of hydraulic relationships constrained with remotely sensed width
information and channel slope obtained from topographic maps. Coe and Birkett (2004)
estimated the mean monthly discharge of the Chari River at NDjamena, Chad. They used
T/P surface water stages upstream from the gauging station calibrated with the groundbased gauge height and discharge data using simple empirical regression techniques.
Kouraev et al. (2004) estimated discharge for the Ob river (Siberia) along two T/P tracks
crosscutting the river in the vicinity of the Salekhard gauging station. T/P measurements
were found to provide reliable water level (H) time series that could then be used to
estimate water discharge (Q) from the rating curve between H and Q at Salekhard, located
6570 km away from the T/P tracks. Zakharova et al. (2006) used the same method to
derive discharge at T/P virtual stations of the Amazon main stream. These studies suggest
that remotely sensed river hydraulic data could be used to estimate the discharge at a
specific location directly, if nearby ground-based discharge measurements are used to
develop discharge ratings in conjunction with the remotely observed variable(s). As
pointed out by Bjerklie et al. (2005), discharge ratings developed from ground-based flow
measurements and remotely sensed hydraulic information are site specific. Leon et al.
(2008) developed a model based on a diffusion-cum-dynamic wave propagation assumption, using in situ discharges and radar altimetry data to estimate rating curves at the
satellite track crossings in the Negro River Basin, Amazon. The calibration phase led to
differences of less than 4% between measured and estimated outflows and validation has
yielded less than 10% errors.
By estimating discharges by combining altimetric water stage and remote in situ discharges, denser stage-discharge rating curves within a basin can be obtained. For example,
out of the 571 gauges listed by ANA, 46 are located in the Negro River sub-basin and 25
have complete records covering the last 20 years. Along the T/P tracks, water level time
series were built for 88 T/P crossings with river and floodplains (Frappart et al. 2005).
Based on altimetric heights, 3.5 times more measurement points are available in the Negro
River basin. These points are evenly distributed within the basin, making possible a
regionalization of the water fluxes. Yet, it is worth noting that these rating curves share
with the in situ curves the need for regular updates to account for possible changes in the
hydrological characteristics of the stem.
In addition, SAR interferometry appears to be the most promising technique to retrieve
surface flow velocities (Goldstein and Zebker 1987; Romeiser and Runge 2007; Romeiser

123

Surv Geophys (2008) 29:247269

261

et al. 2007 among others). Yet, this technique still suffers from technical limitations, such
as being able to estimate the velocity component only in the line-of-sight direction. Thus,
under the assumption that the velocity vector is mostly parallel to the river channel, and
that a global coverage implies that the spacecraft orbits the Earth at large inclination, the
flow velocity in many EastWest trending river segments is likely to be poorly resolved.
7.3 Estimation of Spatial and Temporal Variations of Water Storage: Rivers
and Wetlands
As pointed out by Alsdorf et al. (2003), for the past 100 years our understanding of the
hydraulic characteristics and hydrological mass-balances of surface water runoff have
largely been derived from discharge measurements at in-channel gauging stations. Measurement of in-channel discharge unfortunately does not provide the information necessary
for understanding flow and storage in off-river-channel environments, such as wetlands,
floodplains, and anabranches (e.g., braided channels); these environments are increasingly
recognized for their importance in the biogeochemical cycling of waterborne constituents.
Using radar interferometry, Alsdorf et al. (2001) estimated water height changes in an
Amazon lake to be about 12 2.4 cm, and a volume change of 280 106 m3 during the
44 days between the two JERS images used. These authors reviewed a T/P crossing of the
lake and estimated the change in water stage to be of 20 10 cm during the same period.
The T/P nadir measurements might appear less accurate than interferometry but, so far,
only these measurements offer 10-day periodic estimates over more than a decade. In an
extensive study of T/P measurements over the Amazon basin, Birkett et al. (2002) have
successfully distinguished rivers from floodplains in a number of cases. A small phase
offset of a few days in stage variations between river and nearby floodplain has occasionally been observed.
Frappart et al. (2005), determined spatio-temporal variations of water volume over the
main stream and floodplain located in the Negro River basin, using area variation estimates
for a seasonal cycle captured by the Synthetic Aperture Radar (SAR) onboard the Japanese
Earth Resources Satellite (JERS-1), and changes in water level from the T/P altimetry,
combined with in situ hydrographic stations. Similarly, Frappart et al. (2006a, b) monitored the flood propagation along the Mekong River by combining satellite altimetry data
and imagery. A volume variation of 331 km3 was estimated for the whole Negro sub-basin,
enhancing the complex relationship between the volume potentially stored, the inundated
area and the volume flow during the same period. Altimetry data is useful to study the
hydrological cycle of rivers. Also using the T/P data, Maheu et al. (2003) mapped the flood
propagation along strike the La Plata basin, from the Pantanal wetland to the mouth in the
South Atlantic Ocean.
7.4 Water Profiles and Geodynamical Implications
River free surface slope is an important parameter in floodwave propagation models and
sediment transport calculations. For example, slope values of a few centimeters per kilometer were evaluated for the Amazon main stem using barometric estimates of elevation
performed at some gauging stations (Salati and Marques 1984; Sioli 1984; Nordin and
Meade 1986; Meade et al. 1991). Guzkowska et al. (1990) and Cudlip et al. (1992) used 15
of the 32 crossings of the Amazon River by the SEASAT altimeter to provide an estimate
of the elevation profile of the Amazon, whereas Mertes et al. (1996) and Dunne et al.
(1998) used the SEASAT falling stage measurements to calculate 14 gradient values for

123

262

Surv Geophys (2008) 29:247269

Fig. 11 Example of height profile determined by satellite altimetry for a non-levelled river (Godavari,
India)

the Amazon main stem. Birkett et al. (2002) used T/P measurements along the main stem
of the Amazon to estimate the spatial and temporal variation of the gradient values.
The slope of the riverbed is an important parameter for modeling river hydrodynamics.
Recently, Leon et al. (2006a) have proposed a methodology to derive stream profiles from
the riverbeds height and slope. This method takes advantage of the fact that altimetry data
all have a common reference, whereas in situ measurements are only referred to a local
origin, united from one gauge to the other. The height of the river bed at virtual stations is
determined as that height at which discharge vanishes in rating curves established by
combining times series of water height by satellite altimetry with discharges predicted by
routing the flow recorded at remote gauges. To give an example of the potential of
altimetric data to provide river slopes, even for rivers located in semi-arid regions, the
slope of the Godavari river is presented in Fig. 11. The Godavari river is a sacred river that
runs through India. It originates at 1,620 m above mean sea level, runs eastwards for about
1,500 km and empties into the Bay of Bengal. It runs through the Godavari graben, within
steep banks in the upper part of its course. With a mean annual discharge of 3,200 m3/s
(Singh and Swamy 2006) that ranges far below the Amazon basin with its 209,000 m3/s
mean annual discharge, the Godavari river is also an example of the capabilities of satellite
altimetry in retrieving heights for medium-size rivers, at least partly.
7.5 Hydrologic Regime of Ungauged or Poorly Gauged Basins
A very promising application of radar altimetry concerns the characterization of hydrologic
regime of poorly gauged tropical rivers. Leon et al. (2006b) have studied the 200,000 km2
basin of Rio Caqueta, the most important river basin within Colombian Amazonia. Over 32
gauging stations installed in the Rio Caqueta basin, only six stations are active today. This
point is mainly due to political problems in the FARC controlled region. Only one station,
namely Vila Betancourt, near the border of Brazil, has a full history of discharge estimations in the period matching that of the altimetric data. Along the 1,270 km length, 13
virtual stations have been determined, 12 for ENVISAT crossings and only one with T/P.
Together with the river stages along the river, the discharge at the virtual station has been
estimated by hydrodynamic modelling. Hence, stage discharge relationships have been
computed, leading also to estimates of the river depth at the virtual station and to the slope
of the free surface river between virtual stations. The Caqueta river is characterized by a
hydrological regime highly variable in space and time, and an average width of about

123

Surv Geophys (2008) 29:247269

263

1.1 km at high stage. Despite these factors that impact the quality of the altimetric data,
discharges have been estimated with a mean error of only 15% and the river depth with a
mean difference of less than 1.1 m. Moreover, a 15% mean error in discharge estimates is
interesting since discharge computed using in situ measurements often has a worse
accuracy since it is extrapolated from rating curves.
7.6 Monitoring Lakes for Climate Change Assessment and the Impact of Human
Activity
The monitoring of levels by satellite altimetry is much better for lakes than it is for rivers.
As stated previously, lake surfaces can be cross cut by several missions and, if large
enough, by several tracks of some missions. Merging these measurements together enables
one to produce high sampling rate time series of the level of the Lake, as displayed in
Fig. 8 for the case of Lake Victoria. Such a procedure requires that the curvature of the
lake surface be computed beforehand in order to have a common height reference all over
the surface of the lake (Fig. 8). In the present case of Lake Victoria, the rms difference
between the altimetry series and the in situ series is 8 cm, after the mean of each series has
been removed.
The assessment of the lake water balance could provide improved knowledge of
regional and global climate change and a quantification of the human stress on water
resources across all continents, as studied by Birkett (2000) in the case of Lake Chad and
adjacent wetlands or by Mercier et al. (2001) and Birkett et al. (1999) to highlight the links
between the levels in the African Great Lakes and oceanic climatic indices. Hostetler
(1995) noted that deep and steep sided lakes are good proxies for high amplitude-low
frequency changes, while shallow water basins are better targets for rapid-low amplitude
changes. Central Asia is a place where many lakes of both types, e.g., large bodies in flat
areas such as the Aral sea and narrow bodies in steep mountainous areas such as the
Karakul and Togtogul lakes, are encountered. Very few of these lakes are monitored with
in situ data. However, many of them can be monitored by satellite altimetry (Fig. 12).
Shallow lakes are also extremely sensitive for revealing decreased water inputpossibly of human originand rising evaporation. The Aral sea is typical of such a case and it
has been extensively studied, including using altimetry (Cretaux et al. 2005). The Aral sea
is located in an arid zone characterised by marked differences between summer and winter
temperatures and has low precipitation all year round. Evaporation is approximately ten
times greater than precipitation and the sea being maintained at equilibrium by the
inflowing waters of the Amu Darya and Syr Darya. Around 1960, a decision was made to
develop an intensive cotton and rice economy in its vicinity. In such an arid zone, irrigation
provided the means to reach the planned agricultural objectives of the Soviet Union
government. Large-scale development of ground infrastructure (irrigation channels, reservoirs) started in the 1960s and the volume of water utilised for irrigation increased to
around 100 km3/year, overtaking the annual Amu Daria and Syr Daria inflows. As a result,
the level of the Aral Sea dropped by 13 m between 1960 and 1989 (Fig. 13). This led to the
split of the Aral sea into so-called Small Aral in the north and Big Aral in the south.
Satellite altimetry tracks crosscut both lakes and have measured the level variations since
1992. During that period, Big Aral shrank at a rate of 6080 cm/year (Cretaux et al. 2005).
The corresponding decrease in surface extent and volume was 67,000 km2 and 1,083 km3
in 1960 (Bortnik 1999) to 16,000 km2 and 100 km3 in 2004 (Cretaux et al. 2005). The
difference between evaporation and precipitation for Big Aral represents an average loss of
2530 km3/year over the last decade.

123

264

Surv Geophys (2008) 29:247269

Fig. 12 Central Asian lakes and reservoirs monitored by satellite altimetry (http://www.legos.
obs-mip.fr/soa/hydrologie/hydroweb/). The image is from a mosaic of Landsat images taken in 1990

Cretaux et al. (2005) also showed that the reduction of lake volume compared to the
hydrological budget deduced from gauges implied an underground water inflow of
5 3 km3/year to the Big Aral. During the last 2 years, the level and volume of the Big
Aral has continuously shrunk with a rate as high as about 10 km3/year in water loss and
1 m/year in decrease of the sea level (Fig. 13). Since 1989, the Small Aral experienced a
different situation since it has continued to be fed by the Syr Darya River and therefore has

Fig. 13 Big Aral sea volume variations taken from in situ measurements and from altimetry data (after
Cretaux et al. 2005)

123

Surv Geophys (2008) 29:247269

265

Fig. 14 Variation of volume of small Aral taken from satellite altimetry (black stars) and from hydrological
in situ measurements (orange triangles)

dried up less than the Big Aral (Fig. 14). There are two main explanations: first, the area of
the basin is much smaller and the effect of evaporation is reduced, and, secondly, during
the years 1992 to 1999 a dam was built in the Bergs strait in order to separate both lakes
and stop the loss of water from the Syr Darya into the desert. This dam was destroyed (and
rebuilt) three times during this period. Aladin et al. (2005) demonstrated that during the
period of 1993 to 1999 the existence of the dam contributed to the restoration of the Small
Aral. In August 2005, a solid dam, funded by World Bank, has been built in the Bergs
Strait separating the Small and the Big Aral. It induced a 2-m increase in the sea level of
Small Aral since then and, every spring, the release of some water through open gate in the
dam has allowed the regulation the level of the Small Aral to an approximate level of 42 m
(Fig. 13).

8 Conclusions
Altimetric data offer many possibilities for the monitoring of continental waters. Among
them is the ability of tracking either in-channel fluxes or flow and storage in off-riverchannel environments, such as wetlands, floodplains, and anabranches. Also, that all the
series are naturally leveled is a great benefit of satellite altimetry with respect to in situ
gauges. These are critical for surface water balance. The development of methods to
estimate the river discharge using remotely sensed data would provide the means to
increase the streamflow measurement network globally. Typically, in situ data collection
and management activities are undertaken at the national level, where there is a need for
regionally coordinated systems and actions. Worldwide coverage and near-real time
availability of measurement is definitely another major benefit of satellite altimetry with
respect to local measurements.

123

266

Surv Geophys (2008) 29:247269

The major drawback in the use of altimetric height for water stage monitoring is the
temporal sampling rate. Clearly, the 10-day period of T/P and Jason and the 35-day period
for ENVISAT cannot compete with observations made daily or twice a day at most gauges
around the world. In some applications like flood events, even more frequent data, every
few minutes, are required. On the other hand, over lakes that can be overflown by several
missions and several tracks of each mission, the time sampling can be much better, up to
three to four times per week and then compare with the frequency of in situ sampling.
Another limitation is measurement uncertainty that has been seen to vary from a few
centimeters to some meters or more in the worst cases. With respect to radar altimetry,
accuracy can to a certain extent be improved by adapting the processing of echo waveforms to the continental case. This uncertainty is due to the ground point target size which
ranges from kilometers for T/P to a few hundred meters for ENVISAT and 70 m for
ICESat. Roughly speaking, uncertainty decreases with the size of the footprint, as the
combination of water and vegetation or the merging of different water bodies in a single
footprint becomes less likely. When it comes to identifying and separating peaks of energy
reflected by small water bodies, the sensor and echo processing capability becomes a major
issue.
Higher resolution is foreseen for the future missions. It should reach a few hundred
meters for AltiKa, since the footprint in the Ka band radar is smaller than it is in the Ku band
and Cryosat-2 owing to footprint slicing by virtue of interferometry. Lastly, again owing to
interferometry techniques, a full coverage of the continental domain, a resolution of a few
tens of meters and a slope accuracy of 10 lrad along river courses are expected with Surface
Water Ocean Topography (SWOT), the first mission specifically dedicated to the monitoring of continental waters (SWOT Homepage: www.geology.ohio-state.edu/water).
The prospects offered by future missions will definitely enhance the ability of spatial
data to be included in models and change the management and monitoring of water
resources. Clearly, satellite altimetry is not likely to replace in situ measurements in the
near future, but a combination of both systems will certainly enhance our ability to monitor
the cycle of surface water from the regional scale to the global scale.
Acknowledgments The authors thank P. Bates, an anonymous reviewer and the Editor for their in-depth
reviews that greatly helped to improve this article.

References
Agencia National das Aguas. http://hidroweb.ana.gov.br/doc/WRMB/part1.htm. Accessed Nov 2008
Aladin NV, Cretaux J-F, Plotnikov IS, Kouraev AV, Smurov AO, Cazenave A, Egorov AN, Papa F (2005)
Modern hydro-biological state of the small Aral sea. Environmetric 16:118. doi:10.1002/env.709
Alsdorf DE, Birkett CM, Dunne T, Melack J, Hess L (2001) Water level changes in a large Amazon lake
measured with spaceborne radar interferometry and altimetry. Geophys Res Lett 28(14):26712674
Alsdorf D, Lettenmaier D, Vorosmarty C et al (2003) The need for global, satellite-based observations of
terrestrial surface waters. EOS Trans 84(29):269276
Bercher N, Kosuth P, Bruniquel J (2006) Quality of river water level time series issued from satellite radar
altimetry: influence of river hydrology and satellite measurement accuracy and frequency. Presented at
EGU General Assembly, Vienna, April 2006
Berry PAM, Garlick JD, Freeman JA, Mathers EL (2005) Global inland water monitoring from multimission altimetry. Geophys Res Lett 32:L16401. doi:10.1029/2005GL022814
Birkett CM (1995a) The contribution of Topex/Poseidon to the global monitoring of climatically sensitive
lakes. J Geophys Res 100(C12):2517925204
Birkett CM (1995b) The global remote sensing of lakes, wetlands and rivers for hydrological and climate
research. Geoscience and Remote Sensing Symposium, IGARSS 95, Quantitative Remote Sensing for
Science and Applications, vol 3. pp 19791981

123

Surv Geophys (2008) 29:247269

267

Birkett CM (1998) Contribution of the Topex NASA radar altimeter to the global monitoring of large rivers
and wetlands. Water Resour Res 34(5):12231239
Birkett CM (2000) Synergistic remote sensing of Lake Chad: variability of basin inundation. Remote Sens
Environ 72:218236
Birkett CM, Murtugudde R, Allan T (1999) Indian Ocean climate event brings floods to east Africas lakes
and the Sudd marsh. Geophys Res Lett 26:10311034
Birkett CM, Mertes LAK, Dunne T, Costa M, Jasinski J (2002) Altimetric remote sensing of the Amazon:
application of satellite radar altimetry. J Geophys Res 107(D20):8059. doi:10.1029/2001JD000609
Bjerklie DM, Moller D, Smith LC, Dingman SL (2005) Estimating discharge in rivers using remotely sensed
hydraulic information. J Hydrol 309:191209
Bonnefond P, Exertier P, Laurain O, Menard Y, Orsoni A, Jan G, Jansou E (2003) Absolute calibration of
Jason-1 and TOPEX/Poseidon altimeters in Corsica. Mar Geod 26:261284
Bortnik VN (1999) Alteration of water level and salinity of the Aral sea, creeping environmental problems
and sustainable development in the Aral sea basin. Cambridge University Press, Cambridge, UK, pp
4765
Brooks RL (1982) Lake elevation from satellite radar altimetry from a validation area in Canada. Report,
Geoscience Research Corporation, Salibury, Maryland, USA
Brown GS (1977) Skylab S-193 radar experiment analysis and results. NASA CR-2763, February 1977
Calmant S, Seyler F (2004) Tapajos hydraulic slope at the confluence with the Amazon from combined
satellite altimetric data. EGU, Nice, April 2004
Calmant S, Seyler F (2006) Continental surface waters from satellite altimetry. C R Geosciences 338:1113
1122
Cauhope M (2004) Hauteurs deau dune plaine dinondation amazonienne par altimetrie spatiale. Rapport
de stage de DEA Sciences de la Terre et lEnvironnement. IMFT, Toulouse, France, p 30 (in French)
Cazenave A, Bonnefond P, DoMinh K (1997) Caspian sea level from Topex/Poseidon altimetry: level now
falling. Geophys Res Lett 24:881884
Coe MT (2000) Modeling terrestrial hydrological systems at the continental scale: testing the accuracy of an
atmospheric GCM. J Climatol 13:686704
Coe MT, Birkett CM (2004) Calculation of river discharge and prediction of lake height from satellite radar
altimetry: example for the Lake Chad basin. Water Resour Res 40:W10205. doi:10.1029/2003WR002543
Cretaux J-F, Kouraev AV, Papa F, Berge Nguyen M, Cazenave A, Aladin NV, Plotnikov IS (2005) Water
balance of the big Aral sea from satellite remote sensing and in situ observations. J Great Lakes Res
31(4):520534
Cretaux J-F, Calmant S, Romanovski V, Shibuyin A, Lyard F, Berge-Nguyen M, Cazenave A, Hernandez F
(2008) Implementation of a new absolute calibration site for radar altimeter in the continental area:
lake Issykkul in Central Asia. J Geod (in press)
Cudlip W, Ridley JK, Rapley CG (1992) The use of satellite radar altimetry for monitoring wetlands. In:
Remote sensing and global change. Proceedings of the 16th annual conference of Remote Sensing
Society, London, UK, pp 207216
De Oliveira Campos I, Mercier F, Maheu C, Cochonneau G, Kosuth P, Blitzkow D, Cazenave A (2001)
Temporal variations of river basin waters from Topex/Poseidon satellite altimetry. Application to the
Amazon basin. CR Acad Sci Paris 333:633643
Dunne T, Mertes LAK, Meade RH, Richey JE, Forsberg BR (1998) Exchanges of sediment between the
floodplain and channel of the Amazon river in Brazil. GSA Bull 110(4):450467
Frappart F, Martinez JM, Seyler F, Leon JG, Cazenave A (2005) Determination of the water volume
variation in the Negro river sub-basin by combination of remote sensing and in-situ data. Remote Sens
Environ 99:387399
Frappart F, Calmant S, Cauhope M, Seyler F, Cazenave A (2006a) Results of ENVISAT RA-2 derived
levels, validation over the Amazon basin. Remote Sens Environ 100:252264
Frappart F, Dominh K, Lhermitte J, Ramilllien G, Cazenave A, LeToan T (2006b) Water volume change in
the lower MEKONG basin from satellite altimetry and other remote sensing data. Geophys J Int
167:570584. doi:10.1111/j.1365-246X.2006.03184.x
Fu L, Cazenave A (2001) Satellite altimetry and earth sciences: a handbook of techniques and applications.
Academic Press, London (UK), 464 p
Goldstein RM, Zebker HA (1987) Interferometric radar measurement of ocean surface currents. Nature
328(6132):707709
Guzkowska MAJ, Rapley CG, Rideley JK, Cudlip W, Birkett CM, Scott RF (1990) Developments in inland
water and land altimetry. ESA contract report 78391881FIFL
Hostetler SW (1995) Hydrological and thermal response of lakes to climate: description and modeling. In:
Physics and chemistry of lakes. Springer-Verlag, Berlin, Germany

123

268

Surv Geophys (2008) 29:247269

Kaula WM (1969) The terrestrial environment: solid earth and ocean physics. NASA report study at
Williamstown, MA, NASA CR-1579, August 1969
Koblinsky CJ, Clarke RT, Brenner AC, Frey H (1993) Measurements of river level variations with satellite
altimetry. Water Resour Res 29(6):18391848
Koster RD, Houser PR, Engman ET, Kustas WP (1999) Remote sensing may provide unprecedented
hydrological data. American Geophysical Union. http://www.agu.org/eos_elec
Kosuth P, Blitzkow D, Cochonneau G (2006) Establishment of an altimetric reference network over the
Amazon basin using satellite radar altimetry (Topex Poseidon). Proceedings of the symposium on
15 years of progress in radar altimetry, Venice, Italy, 1318 March 2006
Kouraev A, Zakharova E, Samain O, Mognard N, Cazenave A (2004) Ob river discharge from TOPEX/
Poseidon satellite altimetry (19922002). Remote Sens Environ 93:238245
Leon JG, Calmant S, Seyler F, Bonnet M-P, Cauhope M, Frappart F, Filizola N (2006a) Rating curves and
estimation of average water depth at the upper Rio Negro river based on satellite altimeter data and
modeled discharges. J Hydrol 328:481496
Leon JG, Seyler F, Calmant S, Bonnet M, Cauhope M (2006b) Hydrological parameter estimation for
ungauged basin based on satellite altimeter data and discharge modeling. A simulation for the Caqueta
river (Amazonian basin, Colombia). Hydrology and Earth System Sciences. SRef-ID: 1812-2116/
hessd/2006-3-3023
Leon JG, Bonnet MP, Seyler F, Calmant S, Cauhope M (2008) Distributed water flow estimates of the upper
Negro river by a Muskingum-Cunge routing model using altimetric spatial data. J Hydrol
Le Traon PY, Gaspar P, Bouyssel F, Makhmara H (1995) Using Topex/Poseidon data to enhance ERS1 data.
J Atmos Ocean Technol 12(1):161170
Maheu C, Cazenave A, Mechoso CR (2003) Water level fluctuations in the Plata basin (South America)
from Topex/Poseidon satellite altimetry. Geophys Res Lett 30(3):11431146
Mason IM, Rapley CG, Street-Perrott FA, Guzkowska M (1985) ERS-1 observations of lakes for climate
research. Proceedings of the EARSeL/ESA symposium on European remote sensing opportunities,
Strasbourg, 31 March-3 April 1985
Meade RH, Rayol JM, da Conceicao SC, Natividade JRG (1991) Backwater effects in the Amazon river
basin of Brazil. Environ Geol Water Sci 18(2):105114
Mercier F (2001) Altimetrie spatiale sur les eaux continentales: apport des missions Topex/Poseidon et
ERS1&2 a` letude des lacs, mers interieures et bassins fluviaux. The`se Univ. Toulouse III-Paul
Sabatier, 9/11/2001, p 190
Mercier F, Zanife O-Z (2006) Improvement of the Topex/Poseidon altimetric data processing for hydrological purposes (CASH project). Proceedings of the symposium on 15 years of progress in radar
altimetry, Venice, Italy, 1318 March 2006
Mercier F, Cazenave A, Maheu C (2002) Interannual lake level fluctuations (19931999) in Africa from
Topex/Poseidon: connections with oceanatmosphere interactions over the Indian ocean. Glob Planet
Changes 32:141163
Mertes LAK, Dunne T, Martinelli LA (1996) Channel-floodplain geomorphology along the SolimoesAmazon river, Brazil. GSA Bull 108(9):10891107
Miller LS (1979) Topographic and backscatter characteristics of GEOS 3 overland data. J Geophys Res 84B8:40454054
Morris CS, Gill SK (1994a) Variation of Great Lakes waters from geosat altimetry. Water Resour Res
30:10091017
Morris CS, Gill SK (1994b) Evaluation of the Topex/Poseidon altimeter system over the Great Lakes.
J Geophys Res 99(C12):2452724539
Nordin CF Jr, Meade RH (1986) The Amazon and the Orinoco in McGraw Hill Yearbook of Sciences and
Technology. McGraw Hill, New York, USA, pp 385390
Ponchaut F, Cazenave A (1998) Continental lake level variations from Topex/Poseidon (19931996). C R
Acad Sci Paris 326:1320
Population Reference Bureau (1997) http://www.prb.org/Content/NavigationMenu/Other_reports/19971999/WorldPopMoreThanNos_Eng.pdf)
Roads J, Lawford R, Bainto E, Berbery E, Chen S, Fekete B, Gallo K, Grundstein A, Higgins W, Kanamitsu
M, Krajewski W, Lakshmi V, Leathers D, Lettenmaier D, Luo L, Maurer E, Meyers T, Miller D,
Mirchell K, Mote T, Pinker R, Reichler T, Robinson D, Robock A, Smith J, Srinivasan G, Verdin K,
Vinnikov K, Vonder Haar T, Vorosmarty C, Williams S, Yarosh E (2003) GCIP water and energy
budget synthesis (WEBS). J Geophys Res 108(D16):8609. doi:10.1029/2002JD002583
Romeiser R, Runge H (2007) Theoretical evaluation of several possible along-track inSAR modes of
TerraSAR-X for ocean current measurements. IEEE Trans Geosci Remote Sens 45(1):2135

123

Surv Geophys (2008) 29:247269

269

Romeiser R, Runge H, Suchandt S, Sprenger J, Weibeer H, Sohrmann A, Stammer D (2007) Current


measurements in rivers by spaceborne along-track InSAR. IEEE Trans Geosci Remote Sens 45(12):
40194031
Rosmorduc V, Benveniste J, Lauret O, Milagro M, Picot N (2006) In: Benveniste J, Picot N (ed) Radar
altimetry tutorial. http://www.altimetry.info
Roux E, Cauhope M, Bonnet M-P, Calmant Seyler F (2008) Daily water stage estimated from satellite
altimetric data for large river basin monitoring. Hydrol Sci JJournal des Sciences Hydrologiques 531:8199
Salati E, Marques J (1984) Climatology of the Amazon region. In: Sioli H (ed) The Amazon, limnology and
landscape ecology of a mighty tropical river and its basin. Monographs in biology, vol 56. Kluwer
Academics, Norwell, MA, pp 85126
Singh B, Swamy ASR (2006) Delta sedimentation: east coast of India. Technology Publication, Dehradun,
India, iv, p 400
Sioli H (1984) The Amazon and its main affluents. In: Sioli H (ed) The Amazon, limnology and landscape
ecology of a mighty tropical river and its basin. Monographs in biology, vol 56. Kluwer Academics,
Norwell, MA, pp 127165
Smith LC (1997) Satellite remote sensing of river inundation area, stage, and discharge: a review. Hydrol
Process 11:14271439
WMO (2003) Report GTOS32. HWRP/GCOS/GTOS Expert meeting on hydrological data for global
studies. Report, Toronto, Canada, 1820 Nov 2002, GCOS 84, GTOS 32, WMO/TDN1156
WMO (2004) WMO statement on the status of the global climate. WMO Tech Rep 966, 22p
World Resources Institute (2000) http://www.wri.org/wri/wr
Zakharova EA, Kouraev AV, Cazenave A, Seyler F (2006) Amazon river discharge estimated from TOPEX/
Poseidon altimetry. Geosciences Comptes Rendus (French Academy of Sciences) 338(3):188196

123

You might also like