[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (10)

Search Parameters:
Keywords = global navigation satellite reflectometers

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 9904 KiB  
Article
Improved GNSS-R Altimetry Methods: Theory and Experimental Demonstration Using Airborne Dual Frequency Data from the Microwave Interferometric Reflectometer (MIR)
by Oriol Cervelló i Nogués, Joan Francesc Munoz-Martin, Hyuk Park, Adriano Camps, Raul Onrubia, Daniel Pascual, Christoph Rüdiger, Jeffrey P. Walker and Alessandra Monerris
Remote Sens. 2021, 13(20), 4186; https://doi.org/10.3390/rs13204186 - 19 Oct 2021
Cited by 4 | Viewed by 3572
Abstract
Altimetric performance of Global Navigation Satellite System - Reflectometry (GNSS-R) instruments depends on receiver’s bandwidth and signal-to-noise ratio (SNR). The altimetric delay is usually computed from the time difference between the peak of the direct signal waveform and the maximum of the derivative [...] Read more.
Altimetric performance of Global Navigation Satellite System - Reflectometry (GNSS-R) instruments depends on receiver’s bandwidth and signal-to-noise ratio (SNR). The altimetric delay is usually computed from the time difference between the peak of the direct signal waveform and the maximum of the derivative of the reflected signal waveform. Dual-frequency data gathered by the airborne Microwave Interferometric Reflectometer (MIR) in the Bass Strait, between Australia and Tasmania, suggest that this approach is only valid for flat surfaces and large bandwidth receivers. This work analyses different methods to compute the altimetric observables using GNSS-R. A proposed novel method, the Peak-to-Minimum of the 3rd Derivative (P-Min3D) for narrow-band codes (e.g., L1 C/A), and the Peak-to-Half Power (P-HP) for large bandwidth codes (e.g., L5 or E5a codes) show improved performance when using real data. Both methods are also compared to the Peak-to-Peak (P-P) and Peak-to-Maximum of the 1st Derivative (P-Max1D) methods. The key difference between these methods is the determination of the delay position in the reflected signal waveform in order to compute the altimetric observable. Airborne experimental results comparing the different methods, bands and GNSS-R processing techniques show that centimeter level accuracy can be achieved. Full article
(This article belongs to the Special Issue Applications of GNSS Reflectometry for Earth Observation II)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) MIR instrument, and (<b>b</b>) 19-element up-looking antenna array mounted in a fixed-wing aircraft ready for an airborne campaign (Melbourne, Australia) [<a href="#B52-remotesensing-13-04186" class="html-bibr">52</a>].</p>
Full article ">Figure 2
<p>Map of the location of the track from the MIR transect (yellow), CryoSat-2 (red), and the specular reflections position of the GPS PRN 3 (green) and Galileo E08 (blue).</p>
Full article ">Figure 3
<p>CryoSat-2 altimetry data used as reference ground-truth in this study and geoid (EGM96), both with respect to WGS84. CryoSat-2 flight direction was opposite to the one from the MIR transect analyzed.</p>
Full article ">Figure 4
<p>Diagram representing the different magnitudes involved in Equations (<a href="#FD4-remotesensing-13-04186" class="html-disp-formula">4</a>)–(<a href="#FD6-remotesensing-13-04186" class="html-disp-formula">6</a>).</p>
Full article ">Figure 5
<p>Diagram illustrating different methods to compute the time difference between the direct and the reflected correlations for the altimetry observable. P-P (purple), P-Max1D (green), and P-HP (yellow).</p>
Full article ">Figure 6
<p>(<b>a</b>) Simulated ideal L1 C/A waveform, with infinite bandwidth, and its first, second, and third derivatives. (<b>b</b>) Simulated reflected power function as a function of the delay. (<b>c</b>) Simulated L1 C/A reflected waveform (convolution of <b>a</b> and <b>b</b>). The maximum of the first derivative, and minimum of the third derivative correctly track the delay 0.</p>
Full article ">Figure 7
<p>(<b>a</b>) Simulated L1 C/A waveform low-pass filtered (bandwidth ≈ 5.115 MHz). (<b>b</b>) Simulated reflected L1 C/A waveform low-pass filtered (convolution of <a href="#remotesensing-13-04186-f007" class="html-fig">Figure 7</a>a and <a href="#remotesensing-13-04186-f006" class="html-fig">Figure 6</a>b). The maximum of the first derivative shifts towards the negative delays, while the minimum of the third derivative remains closer to the zero delay.</p>
Full article ">Figure 8
<p>Delay position tracking of the maximum of the first derivative and minimum of the third derivative as a function of the bandwidth. The first derivative maximum shifts towards negative delays for finite bandwidths, while the third derivative minimum remains closer to the zero delay for a wider range of bandwidths (larger than ∼5 MHz).</p>
Full article ">Figure 9
<p>(<b>a</b>) Simulated reflected power function as a function of the delay with a leading edge slope of 86.2 u/(C/A chips) (SWH ≈ 1.7 m). (<b>b</b>) Comparison between the reflected power function with a leading edge slope of 86.2 u/(C/A chips) compared to the one from Case 1 (<a href="#remotesensing-13-04186-f006" class="html-fig">Figure 6</a>b) with infinite slope. (<b>c</b>) Simulated low-pass filtered version of a reflected waveform for a rough sea (SWH ≈ 1.7 m, bandwidth ≈ 5.115 MHz).</p>
Full article ">Figure 10
<p>Delay position tracking of the maximum of the first derivative and minimum of the third derivative as a function of the bandwidth (considering leading-edge of (<b>a</b>) slope 86.2 u/(C/A chips), SWH ≈ 1.7 m and (<b>b</b>) a slope 60 u/(C/A chips), SWH ≈ 2.4 m). The minimum of the third derivative tracks the mid-rise position of the leading-edge of the reflected power function (marked with a dashed green line, while the maximum of the triangle is marked with a dashed red line). The maximum of the first derivative tracks a point near the zero delay for very large bandwidths, and then it starts to shift.</p>
Full article ">Figure 11
<p>Sample of altimetry estimates computed using cGNSS-R at L1 with 3 different methods: P-P, P-Max1D, and P-Min3D, showing (<b>a</b>) the error introduced in the time series, and (<b>b</b>) the scatter plot of the unbiased SSH obtained by each method and CryoSat-2 SSH values. Computed using the cGNSS-R technique with <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 100 ms and <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 100 s. For this case, P-HP would present similar behaviorthan P-Max1D but with an even lower slope as shown in <a href="#remotesensing-13-04186-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 12
<p>Sample of a normalized (<b>a</b>) L1, and (<b>b</b>) L5 cGNSS-R reflected waveform acquired by MIR, and its first and third derivatives. Computed with <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 1 ms and <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 100 ms.</p>
Full article ">Figure 13
<p>Examples of normalized reflected waveforms shapes acquired by MIR, showing (<b>a</b>) L1 waveform, (<b>b</b>) L5 waveform and (<b>c</b>) E5a waveform, all computed with cGNSS-R technique and with <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 1 ms and <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </semantics></math> = 100 ms.</p>
Full article ">Figure 14
<p>Allan variance computed for (<b>a</b>) L1 and (<b>b</b>) L5 and E5a for the computed altimetric methods and coherent integration times.</p>
Full article ">Figure 15
<p>Unbiased root mean squared error (ubRMSE), with respect to CryoSat-2, for (<b>a</b>) L1 and (<b>b</b>) L5 and E5a for the computed altimetric methods and coherent integration times.</p>
Full article ">
11 pages, 32927 KiB  
Communication
A Coastal Experiment for GNSS-R Code-Level Altimetry Using BDS-3 New Civil Signals
by Fan Gao, Tianhe Xu, Xinyue Meng, Nazi Wang, Yunqiao He and Baojiao Ning
Remote Sens. 2021, 13(7), 1378; https://doi.org/10.3390/rs13071378 - 3 Apr 2021
Cited by 19 | Viewed by 3118
Abstract
High temporal and spatial resolutions are the key advantages of the global navigation satellites system-reflectometry (GNSS-R) technique, while low precision and instabilities constrain its development. Compared with conventional Ku/C band nadir-looking radar altimetry, the precision of GNSS-R code-level altimetry is restricted by the [...] Read more.
High temporal and spatial resolutions are the key advantages of the global navigation satellites system-reflectometry (GNSS-R) technique, while low precision and instabilities constrain its development. Compared with conventional Ku/C band nadir-looking radar altimetry, the precision of GNSS-R code-level altimetry is restricted by the smaller bandwidth and the lower transmitted power of the signals. Fortunately, modernized GNSS broadcast new open-available ranging codes with wider bandwidth. The Chinese BDS-3 system was built on 31 July 2020; its inclined geostationary orbit and medium circular orbit satellites provide B1C and B2a public navigation service signals in the two frequency bands of B1 and B2. In order to investigate their performance on GNSS-R code-level altimetry, a coastal experiment was conducted on 5 November 2020 at a trestle of Weihai in the Shandong province of China. The raw intermediate frequency data with a 62 MHz sampling rate were collected and post-processed to solve the sea surface height every second continuously for over eight hours. The precisions were evaluated using the measurements from a 26 GHz radar altimeter mounted on the same trestle near our GNSS-R setup. The results show that a centimeter-level accuracy of GNSS-R altimetry—based on B1C code after the application of the moving average—can be achieved, while for B2a code, the accuracy is about 10 to 20 cm. Full article
(This article belongs to the Special Issue GNSS-R Earth Remote Sensing from SmallSats)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Basic concept of deriving code-level path delay from waveforms of direct and reflected signals using the data component (red) and the pilot component (purple).</p>
Full article ">Figure 2
<p>Photos of the geodetic GNSS chock-ring antenna (<b>a</b>), electronic total station (<b>b</b>), and the monostatic radar altimeter (<b>c</b>).</p>
Full article ">Figure 3
<p>The configuration of antennas for our GNSS-R altimetry experiment (<b>a</b>), the photograph of the antennas’ arrangement (<b>b</b>) and top view of the trestle bridge (<b>c</b>) during the experiment.</p>
Full article ">Figure 4
<p>SSHs derived from B1C code-level delay and radar altimeter measurements for more than eight hours.</p>
Full article ">Figure 5
<p>SSHs derived from B2a code-level delay and radar altimeter measurements for more than eight hours.</p>
Full article ">Figure 6
<p>Differences between measured SSHs using monostatic radar and B1C signals at different elevations.</p>
Full article ">Figure 7
<p>Differences between measured SSHs using monostatic radar and B2a signals at different elevations.</p>
Full article ">Figure 8
<p>Cross-correlated waveforms of B1C code for direct (blue) and reflected (red) signals.</p>
Full article ">Figure 9
<p>Cross-correlated waveforms of B2a code for direct (blue) and reflected (red) signals.</p>
Full article ">Figure 10
<p>SSHs derived from B1C code-level delay measurements with moving average and radar altimeter for more than eight hours.</p>
Full article ">Figure 11
<p>SSHs derived from B2a code-level delay measurements with moving average and radar altimeter for more than eight hours.</p>
Full article ">Figure 12
<p>Differences between measured SSHs using monostatic radar and smoothed solutions from B1C signals.</p>
Full article ">Figure 13
<p>Differences between measured SSHs using monostatic radar and smoothed solutions from B2a signals.</p>
Full article ">
23 pages, 4575 KiB  
Article
Soil Moisture Estimation Synergy Using GNSS-R and L-Band Microwave Radiometry Data from FSSCat/FMPL-2
by Joan Francesc Munoz-Martin, David Llaveria, Christoph Herbert, Miriam Pablos, Hyuk Park and Adriano Camps
Remote Sens. 2021, 13(5), 994; https://doi.org/10.3390/rs13050994 - 5 Mar 2021
Cited by 25 | Viewed by 4318
Abstract
The Federated Satellite System mission (FSSCat) was the winner of the 2017 Copernicus Masters Competition and the first Copernicus third-party mission based on CubeSats. One of FSSCat’s objectives is to provide coarse Soil Moisture (SM) estimations by means of passive microwave measurements collected [...] Read more.
The Federated Satellite System mission (FSSCat) was the winner of the 2017 Copernicus Masters Competition and the first Copernicus third-party mission based on CubeSats. One of FSSCat’s objectives is to provide coarse Soil Moisture (SM) estimations by means of passive microwave measurements collected by Flexible Microwave Payload-2 (FMPL-2). This payload is a novel CubeSat based instrument combining an L1/E1 Global Navigation Satellite Systems-Reflectometer (GNSS-R) and an L-band Microwave Radiometer (MWR) using software-defined radio. This work presents the first results over land of the first two months of operations after the commissioning phase, from 1 October to 4 December 2020. Four neural network algorithms are implemented and analyzed in terms of different sets of input features to yield maps of SM content over the Northern Hemisphere (latitudes above 45° N). The first algorithm uses the surface skin temperature from the European Centre of Medium-Range Weather Forecast (ECMWF) in conjunction with the 16 day averaged Normalized Difference Vegetation Index (NDVI) from the Moderate Resolution Imaging Spectroradiometer (MODIS) to estimate SM and to use it as a comparison dataset for evaluating the additional models. A second approach is implemented to retrieve SM, which complements the first model using FMPL-2 L-band MWR antenna temperature measurements, showing a better performance than in the first case. The error standard deviation of this model referred to the Soil Moisture and Ocean Salinity (SMOS) SM product gridded at 36 km is 0.074 m3/m3. The third algorithm proposes a new approach to retrieve SM using FMPL-2 GNSS-R data. The mean and standard deviation of the GNSS-R reflectivity are obtained by averaging consecutive observations based on a sliding window and are further included as additional input features to the network. The model output shows an accurate SM estimation compared to a 9 km SMOS SM product, with an error of 0.087 m3/m3. Finally, a fourth model combines MWR and GNSS-R data and outperforms the previous approaches, with an error of just 0.063 m3/m3. These results demonstrate the capabilities of FMPL-2 to provide SM estimates over land with a good agreement with respect to SMOS SM. Full article
(This article belongs to the Special Issue Applications of GNSS Reflectometry for Earth Observation II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Soil Moisture and Ocean Salinity (SMOS)-derived Soil Moisture (SM) from [<a href="#B40-remotesensing-13-00994" class="html-bibr">40</a>] projected on a 36 km Equal-Area Scalable Earth (EASE) grid on (<b>a</b>) 15 October 2020 and (<b>b</b>) 5 November 2020, respectively, and and on a 9 km grid on (<b>c</b>) 15 October 2020 and (<b>d</b>) 5 November 2020, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) Sixteen day averaged NDVI from MODIS [<a href="#B49-remotesensing-13-00994" class="html-bibr">49</a>] on 22 October 2020 and (<b>b</b>) skin temperature from ECMWF [<a href="#B50-remotesensing-13-00994" class="html-bibr">50</a>] on 15 October 2020.</p>
Full article ">Figure 3
<p>Flexible Microwave Payload-2 (FMPL-2) Microwave Radiometer (MWR) measurements geo-located using the Nearest Neighbor Interpolation (NNI) algorithm comprising five days of measurements with and centered on (<b>a</b>) 7 October 2020, (<b>b</b>) 21 October 2020, (<b>c</b>) 3 November 2020, and (<b>d</b>) 10 November 2020.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>d</b>) GNSS-R reflectivity calibrated as shown in [<a href="#B37-remotesensing-13-00994" class="html-bibr">37</a>], (<b>b</b>,<b>e</b>) the SMOS SM (9 km) estimation interpolated over the GNSS-R specular reflection position, and (<b>c</b>,<b>f</b>) the incidence angle of the reflection. (<b>a</b>–<b>c</b>) from 1–31 October 2020 and (<b>d</b>,<b>e</b>) from 1 November to 4 December 2020.</p>
Full article ">Figure 5
<p>(<b>a</b>) Scatter-density plot between the ANN output using NDVI and skin temperature as input data and the SMOS SM product as the network reference value; and (<b>b</b>) the error histogram between the ANN output and the SMOS reference.</p>
Full article ">Figure 6
<p>(<b>a</b>) Low resolution skin temperature (°C), and (<b>b</b>) sample FMPL-2 standard deviation (°K) in the along-track direction after applying the NNI algorithm. Note that land-water transitions in the northern part of Siberia are highlighted in red.</p>
Full article ">Figure 7
<p>(<b>a</b>) Scatter-density plot between the ANN output using FMPL-2 data and the SMOS reference value; and (<b>b</b>) the error histogram between the ANN output and the SMOS reference.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) Scatter-density plot between the ANN output using FMPL-2 data and the SMOS reference value; and (<b>b</b>) the error histogram between the ANN output and the SMOS reference.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>) FMPL-2/MWR down-scaled SM estimations corresponding to (<b>a</b>) 1–5 October 2020, (<b>b</b>) 10–15 October 2020, (<b>c</b>) 25–30 October 2020, and (<b>d</b>) 9–14 November 2020. (<b>e</b>–<b>h</b>) Errors with respect to the SMOS SM product, for the same date periods specified in (<b>a</b>–<b>d</b>). Note that, SMOS GT data are not available in regions that are blank (e.g., northern part of Russia).</p>
Full article ">Figure 9
<p>(left) Scatter-density plot and (right) error histogram of the GNSS-R based ANN output with respect to the collocated SMOS SM product from Barcelona Expert Center on Remote Sensing (BEC) [<a href="#B40-remotesensing-13-00994" class="html-bibr">40</a>], showing the four different values of <span class="html-italic">N</span> used to compute the movmean and movstd inputs for the network.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>c</b>) GNSS-R-derived SM estimations corresponding to <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>. (<b>b</b>–<b>d</b>) Error with respect to the collocated SMOS SM product. The map presents the measurements collected by FMPL-2 from (<b>a</b>,<b>b</b>) 1–31 October 2020 and from 1 November to 4 December 2020. Only reflections collocated with SMOS SM data are presented.</p>
Full article ">Figure 11
<p>(left) Scatter-density plot and (right) the error histogram of the combined MWR and GNSS-R ANN output with respect to the collocated SMOS SM product from BEC [<a href="#B40-remotesensing-13-00994" class="html-bibr">40</a>], for the four different values of <span class="html-italic">N</span> used to compute movmean and movstd, used as inputs for the ANN.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>c</b>) Combined MWR and GNSS-R SM estimations corresponding to <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>. (<b>b</b>–<b>d</b>) Error with respect to the collocated SMOS SM product. The map presents the measurements collected by FMPL-2 from (<b>a</b>,<b>b</b>) 1–31 October 2020 and from 1 November to 4 December 2020. Only reflections collocated with SMOS SM data are presented.</p>
Full article ">Figure 13
<p>Std(Err) evolution as a function of the number of consecutive Fresnel zones used to derive the SM estimation. Comparison between the airborne case (Microwave Interferometric Reflectometer instrument [<a href="#B62-remotesensing-13-00994" class="html-bibr">62</a>]) and the spaceborne case (FMPL-2). Note that the X-axis is normalized by the number of Fresnel zones used to derive the SM estimation (i.e., the <span class="html-italic">N</span> parameter used in <a href="#sec3dot3-remotesensing-13-00994" class="html-sec">Section 3.3</a>). A regression curve is fit to the Std(Err) of both error curves to ease the comparison.</p>
Full article ">
20 pages, 32122 KiB  
Article
Single-Pass Soil Moisture Retrieval Using GNSS-R at L1 and L5 Bands: Results from Airborne Experiment
by Joan Francesc Munoz-Martin, Raul Onrubia, Daniel Pascual, Hyuk Park, Miriam Pablos, Adriano Camps, Christoph Rüdiger, Jeffrey Walker and Alessandra Monerris
Remote Sens. 2021, 13(4), 797; https://doi.org/10.3390/rs13040797 - 22 Feb 2021
Cited by 26 | Viewed by 3597
Abstract
Global Navigation Satellite System—Reflectometry (GNSS-R) has already proven its potential for retrieving a number of geophysical parameters, including soil moisture. However, single-pass GNSS-R soil moisture retrieval is still a challenge. This study presents a comparison of two different data sets acquired with the [...] Read more.
Global Navigation Satellite System—Reflectometry (GNSS-R) has already proven its potential for retrieving a number of geophysical parameters, including soil moisture. However, single-pass GNSS-R soil moisture retrieval is still a challenge. This study presents a comparison of two different data sets acquired with the Microwave Interferometer Reflectometer (MIR), an airborne-based dual-band (L1/E1 and L5/E5a), multiconstellation (GPS and Galileo) GNSS-R instrument with two 19-element antenna arrays with four electronically steered beams each. The instrument was flown twice over the OzNet soil moisture monitoring network in southern New South Wales (Australia): the first flight was performed after a long period without rain, and the second one just after a rain event. In this work, the impact of surface roughness and vegetation attenuation in the reflectivity of the GNSS-R signal is assessed at both L1 and L5 bands. The work analyzes the reflectivity at different integration times, and finally, an artificial neural network is used to retrieve soil moisture from the reflectivity values. The algorithm is trained and compared to a 20-m resolution downscaled soil moisture estimate derived from SMOS soil moisture, Sentinel-2 normalized difference vegetation index (NDVI) data, and ECMWF Land Surface Temperature. Full article
(This article belongs to the Special Issue Applications of GNSS Reflectometry for Earth Observation)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Highlighted in black is the NSW area of Australia where the flight was conducted, and (<b>b</b>) definition of the Yanco areas A (34<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>43<math display="inline"> <semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics> </math> S, 146<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>05<math display="inline"> <semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics> </math> E) and B (34<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>59<math display="inline"> <semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics> </math> S, 146<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>18<math display="inline"> <semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics> </math> E).</p>
Full article ">Figure 2
<p>The tracks denoted by the <span class="html-italic">Received Power</span> color scale represent the received power (in arbitrary units) by the Microwave Interferometer Reflectometer (MIR) instrument, and the coloured squares are the in-situ OzNet soil moisture (m<math display="inline"> <semantics> <mrow> <msup> <mrow/> <mn>3</mn> </msup> <mo>/</mo> </mrow> </semantics> </math>m<math display="inline"> <semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics> </math>) sensors at 5 cm depth. (<b>a</b>) Shows the received power and in-situ measurement during <span class="html-italic">Dry</span> flight at (<b>a</b>) Site A, and (<b>b</b>) Site B; and during <span class="html-italic">Wet</span> flight at (<b>c</b>) Site A, and (<b>d</b>) Site B. Note: anomalous low reflectivity values for large banking angles during the turns are not used in the study.</p>
Full article ">Figure 3
<p>Normalized difference vegetation index (NDVI) retrieved by Sentinel-2. The selected data sets contain Copernicus Sentinel data corresponding to 2018-05-01 (Dry) and 2018-06-16 (Wet) from the Sentinel Hub [<a href="#B51-remotesensing-13-00797" class="html-bibr">51</a>]. <span class="html-italic">Dry</span> flight at (<b>a</b>) Site A, and (<b>b</b>) Site B; and during <span class="html-italic">Wet</span> flight at (<b>c</b>) Site A, and (<b>d</b>) Site B. Note that negative NDVI values in (<b>a</b>,<b>c</b>) correspond to the Coleambally Canal.</p>
Full article ">Figure 4
<p>SM retrieved by the combination of SMOS SM and downscaled using Sentinel-2 NDVI. <span class="html-italic">Dry</span> flight at (<b>a</b>) Site A, and (<b>b</b>) Site B; and during <span class="html-italic">Wet</span> flight at (<b>c</b>) Site A, and (<b>d</b>) Site B. Credits: Barcelona Expert Center [<a href="#B52-remotesensing-13-00797" class="html-bibr">52</a>]. Note that the color scale is different in each plot to maximize the contrast and ease its visualization.</p>
Full article ">Figure 5
<p>Histograms of the collocated Sentinel-2 NDVI (<b>a</b>) and downscaled SM (<b>b</b>) values over the MIR specular points for both the Dry and Wet flights.</p>
Full article ">Figure 6
<p>MIR reflectivity for both (<b>a</b>) Dry and (<b>b</b>) Wet flights (dB). Average Reflectivity over the selected water bodies is ∼−2.1 dB at an incidence angle ∼20<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>. The resultant reflectivity has been evaluated against a saline water model [<a href="#B57-remotesensing-13-00797" class="html-bibr">57</a>], assuming a very low salinity (psu = 1 ppm) and a temperature of 15 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C.</p>
Full article ">Figure 7
<p>MIR reflectivity Probability Density Function (PDF) for both (<b>a</b>,<b>b</b>) Dry and (<b>c</b>,<b>d</b>) Wet flights. The Y-axis is the normalized counts values of the PDF, and the average reflectivity for the four beams at the <span class="html-italic">Dry</span> flight are: −17.6 dB, −16.8 dB, −16.3 dB, and −16.7 dB, with a standard deviation of ∼4.7 dB; and for the <span class="html-italic">Wet</span> flight are: −11.2 dB, −10.8 dB, −10.5 dB, and −10.4 dB, with a standard deviation of ∼4.8 dB. Note that, the tracking algorithm selects those satellites whose reflection has a very small incidence, and the <span class="html-italic">average</span> incidence angle for the four beams is below ∼30<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>.</p>
Full article ">Figure 8
<p>Normalized histogram PDFs of the MIR reflectivities retrieved at different effective integration times at (<b>a</b>) L1 and (<b>b</b>) L5.</p>
Full article ">Figure 9
<p>Geolocated MIR reflectivities in Yanco site A during <span class="html-italic">Wet</span> flight at different <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math>: (<b>a</b>) 20 ms and (<b>b</b>) 1000 ms.</p>
Full article ">Figure 10
<p>Estimated Surface Roughness normalized PDF in Site B during <span class="html-italic">Wet</span> flight seen by L1 and L5 signals, assuming a flat reflectivity <math display="inline"> <semantics> <mrow> <mo>Γ</mo> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics> </math> dB, and using (<b>a</b>) <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math> = 20 ms, and (<b>b</b>) at <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math> = 1000 ms and 5000 ms only for the L1 case.</p>
Full article ">Figure 11
<p>Geolocated <math display="inline"> <semantics> <mo>Γ</mo> </semantics> </math>, movstd(<math display="inline"> <semantics> <mo>Γ</mo> </semantics> </math>), and Noise-to-signal ratio (NSR), defined as movstd(<math display="inline"> <semantics> <mo>Γ</mo> </semantics> </math>) −<math display="inline"> <semantics> <mo>Γ</mo> </semantics> </math>, at <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics> </math> 5000 ms at L1 C/A for the <span class="html-italic">Dry</span> flight. Black boxes identify areas with reflectivity drops due to vegetated areas and an increase of the surface roughness.</p>
Full article ">Figure 12
<p>Same as <a href="#remotesensing-13-00797-f011" class="html-fig">Figure 11</a> but for the <span class="html-italic">Wet</span> flight.</p>
Full article ">Figure 13
<p>Scatter density plot of the Noise-to-Signal Ratio (NSR) computed at <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math> = 5000 ms with respect to the reflectivity values at <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math> = 20 ms at L1 for (<b>a</b>) <span class="html-italic">Dry</span> and (<b>b</b>) <span class="html-italic">Wet</span> flights, and at L5 for (<b>c</b>) <span class="html-italic">Dry</span> and (<b>d</b>) <span class="html-italic">Wet</span> flights.</p>
Full article ">Figure 14
<p>ANN estimated SM vs. SMOS/Sentinel-2 downscaled SM at L1. Columns from left to right increasing <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics> </math> for 0.1, 1, 2, and 5 s. Row from top to bottom, ANN cases 1 to 4.</p>
Full article ">Figure 15
<p>As for <a href="#remotesensing-13-00797-f014" class="html-fig">Figure 14</a> but for L5.</p>
Full article ">Figure 16
<p>Standard deviation of the error of the ANN with respect to Sentinel-2/SMOS downscaled soil moisture for the four different cases for different integration times, at (<b>a</b>) L1 and (<b>b</b>) L5.</p>
Full article ">
17 pages, 1624 KiB  
Article
Untangling the Incoherent and Coherent Scattering Components in GNSS-R and Novel Applications
by Joan Francesc Munoz-Martin, Raul Onrubia, Daniel Pascual, Hyuk Park, Adriano Camps, Christoph Rüdiger, Jeffrey Walker and Alessandra Monerris
Remote Sens. 2020, 12(7), 1208; https://doi.org/10.3390/rs12071208 - 9 Apr 2020
Cited by 19 | Viewed by 4154
Abstract
As opposed to monostatic radars where incoherent backscattering dominates, in bistatic radars, such as Global Navigation Satellite Systems Reflectometry (GNSS-R), the forward scattered signals exhibit both an incoherent and a coherent component. Current models assume that either one or the other are dominant, [...] Read more.
As opposed to monostatic radars where incoherent backscattering dominates, in bistatic radars, such as Global Navigation Satellite Systems Reflectometry (GNSS-R), the forward scattered signals exhibit both an incoherent and a coherent component. Current models assume that either one or the other are dominant, and the calibration and geophysical parameter retrieval (e.g., wind speed, soil moisture, etc.) are developed accordingly. Even the presence of the coherent component of a GNSS reflected signal itself has been a matter of discussion in the last years. In this work, a method developed to separate the leakage of the direct signal in the reflected one is applied to a data set of GNSS-R signals collected over the ocean by the Microwave Interferometer Reflectometer (MIR) instrument, an airborne dual-band (L1/E1 and L5/E5a), multi-constellation (GPS and Galileo) GNSS-R instrument with two 19-elements antenna arrays with 4 beam-steered each. The presented results demonstrate the feasibility of the proposed technique to untangle the coherent and incoherent components from the total power waveform in GNSS reflected signals. This technique allows the processing of these components separately, which increases the calibration accuracy (as today both are mixed and processed together), allowing higher resolution applications since the spatial resolution of the coherent component is determined by the size of the first Fresnel zone (300–500 meters from a LEO satellite), and not by the size of the glistening zone (25 km from a LEO satellite). The identification of the coherent component enhances also the location of the specular reflection point by determining the peak maximum from this coherent component rather than the point of maximum derivative of the incoherent one, which is normally noisy and it is blurred by all the glistening zone contributions. Full article
(This article belongs to the Special Issue Applications of GNSS Reflectometry for Earth Observation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flight path on 6 June 2018 in which the data used for the coherency analysis were acquired.</p>
Full article ">Figure 2
<p>Cross-correlation process of the Microwave Interferometer Reflectometer (MIR) data based on the parallel code-phase search (PCPS) algorithm.</p>
Full article ">Figure 3
<p>Total power waveform and coherent component processing algorithm of MIR data, including the phase retrieval of the peak.</p>
Full article ">Figure 4
<p>Variance calculus algorithm in the presence of bit transitions for both direct and reflected waveforms.</p>
Full article ">Figure 5
<p>Coherency of a GPS L1CA direct signal (top figure) within 40 ms of integration with a two bit transitions, coherency of the same GPS signal once reflected over the sea (middle figure), and phase evolution of the peak for each integration sample (bottom figure). All figures with <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>32.768</mn> </mrow> </semantics> </math> MHz, 1 sample = 30 nanoseconds.</p>
Full article ">Figure 6
<p>Coherency of a GPS L1CA direct signal (top figure) within 20 ms of integration with a bit transition in the middle, coherency of the same GPS signal once reflected over the sea (middle figure), and phase evolution of the peak for each integration sample (bottom figure). All figures with <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>32.768</mn> </mrow> </semantics> </math> MHz, 1 sample = 30 nanoseconds.</p>
Full article ">Figure 7
<p>Coherency of a GPS L5Q without secondary codes direct signal (top figure) within 40 ms of integration with a two bit transitions, coherency of the same GPS signal once reflected over the sea (middle figure), and phase evolution of the peak for each integration sample (bottom figure). All figures with <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>32.768</mn> </mrow> </semantics> </math> MHz, 1 sample = 30 nanoseconds</p>
Full article ">Figure 8
<p>L5QCoherency of a GPS L5Q with secondary codes direct signal (top figure) within 40 ms of integration with a two bit transitions, coherency of the same GNSS signal once reflected over the sea (middle figure), and phase evolution of the peak for each integration sample (bottom figure). All figures with <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>32.768</mn> </mrow> </semantics> </math> MHz, 1 sample = 30 nanoseconds</p>
Full article ">Figure 9
<p>Left column: Total power waveform and coherent component for different integration times (from 5 to 40 ms). Right column: phase evolution of GPS L1 direct (black) and reflected (red) signals for the 1 ms coherently integrated waveforms used for different integration times (from 5 to 40 ms).</p>
Full article ">Figure 10
<p>Left column: Total power waveform and coherent component for different integration times (from 5 to 40 ms). Right column: phase evolution of GPS L1 direct (black) and reflected (red) signals for the 1-ms coherently integrated waveforms used for different integration times (from 5 to 40 ms).</p>
Full article ">Figure 11
<p>Left column: Total power waveform and coherent component for different integration times (from 5 to 40 ms). Right column: phase evolution of GPS L5Q direct (black) and reflected (red) signals for the 20 ms coherently integrated waveforms used for different integration times (from 40 to 200 ms).</p>
Full article ">Figure 12
<p>Left column: Total power waveform and coherent component for different integration times (from 5 to 40 ms). Right column: phase evolution of GPS L5Q direct (black) and reflected (red) signals for the 20 ms coherently integrated waveforms used for different integration times (from 40 to 200 ms).</p>
Full article ">Figure 13
<p>Peak position estimation depending on the integration time (40 to 100 ms) from the coherent component at L5Q with secondary codes, <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>262.144</mn> </mrow> </semantics> </math> MHz, 1 sample = 3.81 nanoseconds. Note all functions are normalized for the sake of clarity.</p>
Full article ">Figure 14
<p>Swell period in the flight path the 6th of June, 2018. Downloaded from Ventusky [<a href="#B27-remotesensing-12-01208" class="html-bibr">27</a>].</p>
Full article ">
11 pages, 9243 KiB  
Article
Satellite Cross-Talk Impact Analysis in Airborne Interferometric Global Navigation Satellite System-Reflectometry with the Microwave Interferometric Reflectometer
by Raul Onrubia, Daniel Pascual, Hyuk Park, Adriano Camps, Christoph Rüdiger, Jeffrey P. Walker and Alessandra Monerris
Remote Sens. 2019, 11(9), 1120; https://doi.org/10.3390/rs11091120 - 10 May 2019
Cited by 16 | Viewed by 4006
Abstract
This work analyzes the satellite cross-talk observed by the microwave interferometric reflectometer (MIR), a new global navigation satellite system (GNSS) reflectometer, during an airborne field campaign in Victoria and New South Wales, Australia. MIR is a GNSS reflectometer with two 19-element, dual-band arrays, [...] Read more.
This work analyzes the satellite cross-talk observed by the microwave interferometric reflectometer (MIR), a new global navigation satellite system (GNSS) reflectometer, during an airborne field campaign in Victoria and New South Wales, Australia. MIR is a GNSS reflectometer with two 19-element, dual-band arrays, each of them having four steerable beams. The data collected during the experiment, the characterization of the arrays, and the global positioning system (GPS) and Galileo ephemeris were used to compute the expected delays and power levels of all incoming signals, and the probability of cross-talk was then evaluated. Despite the MIR highly directive arrays, the largest ever for a GNSS-R instrument, one of the flights was found to be contaminated by cross-talk almost half of the time at the L1/E1 frequency band, and all four flights were contaminated ∼5–10% of the time at the L5/E5a frequency band. The cross-talk introduces an error of up to 40 cm of standard deviation for altimetric applications and about 0.24 dB for scatterometric applications. Full article
(This article belongs to the Special Issue Radio Frequency Interference (RFI) in Microwave Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Microwave interferometric reflectometer (MIR) instrument and up-looking array mounted inside the airplane, and (<b>b</b>) down-looking array covered with a radome hanging from the airplane’s fuselage (with part of the fairing removed).</p>
Full article ">Figure 2
<p>Ground track of sea flight over the Bass Strait. The color scale represents the flight height above the sea level.</p>
Full article ">Figure 3
<p>Retrieved L5/E5a interferometric global navigation satellite system-reflectometry (iGNSS-R) waveform (blue), Galileo E5a tracked conventional global navigation satellite system-reflectometry (cGNSS-R) waveform (red), global positioning system (GPS) L5 interfering cGNSS-R waveform (yellow), and reconstructed contaminated waveform (purple). The data was collected during the flight over the Bass Strait.</p>
Full article ">Figure 4
<p>Skyplot of the tracked satellites in the (<b>a</b>) flight over the entrance to Port Philip Bay, (<b>b</b>) sea flight over the Bass Strait, and flights over Yanco under dry (<b>c</b>) and wet (<b>d</b>) soil conditions. The color scale indicates the angular distance to the closest satellite. Red lines highlight cross-talk from other satellites.</p>
Full article ">Figure 5
<p>Cumulative histogram of the probability of having cross-talk in the (<b>a</b>) L1/E1 and, (<b>b</b>) L5/E5a frequency bands for the different MIR flights.</p>
Full article ">Figure 6
<p>Probability density function (PDF) of the estimation error of the (<b>a</b>) maximum derivative position (DER), (<b>b</b>) half-power position (HALF), (<b>c</b>) peak position (MAX), and (<b>d</b>) peak amplitude. The “XT” PDFs show the error estimation when the cross-correlation peaks are close enough to interfere with each other, while the “NOXT” show the error estimation when the cross-correlation peaks are far enough not to interfere. The analysis distinguishes between the interfering cross-correlation peak arriving later than the tracked satellite cross-correlation peak (“first peak”), or arriving sooner (“second peak”).</p>
Full article ">
30 pages, 56449 KiB  
Article
The Global Navigation Satellite Systems Reflectometry (GNSS-R) Microwave Interferometric Reflectometer: Hardware, Calibration, and Validation Experiments
by Raul Onrubia, Daniel Pascual, Jorge Querol, Hyuk Park and Adriano Camps
Sensors 2019, 19(5), 1019; https://doi.org/10.3390/s19051019 - 27 Feb 2019
Cited by 11 | Viewed by 4392
Abstract
This manuscript describes the Microwave Interferometric Reflectometer (MIR) instrument, a multi-beam dual-band GNSS-Reflectometer with beam-steering capabilities built to assess the performance of a PAssive Reflectrometry and Interferometry System—In Orbit Demonstrator (PARIS-IoD) like instrument and to compare the performance of different GNSS-R techniques and [...] Read more.
This manuscript describes the Microwave Interferometric Reflectometer (MIR) instrument, a multi-beam dual-band GNSS-Reflectometer with beam-steering capabilities built to assess the performance of a PAssive Reflectrometry and Interferometry System—In Orbit Demonstrator (PARIS-IoD) like instrument and to compare the performance of different GNSS-R techniques and signals. The instrument is capable of tracking up to 4 different GNSS satellites, two at L1/E1 band, and two at L5/E5 band. The calibration procedure of the up- and down-looking arrays is presented, the calibration performance is evaluated, and the results of the validation experiments carried out before the field experiments are shown in this paper. Full article
(This article belongs to the Section Remote Sensors)
Show Figures

Figure 1

Figure 1
<p>Block diagram of the MIR instrument. The red, blue, orange, and purple lines correspond to the radio frequency (RF) signals of each of the four beams from both the up- and the down-looking array. The light green is a 1 Pulse Per Second (PPS) signal. The grey line is a 10 MHz reference signal. The cyan lines correspond to serial protocols of communication, such as Inter-Integrated Circuit (I2C), Serial Peripheral Interface (SPI) or Universal Asynchronous Receiver-Transmitter (UART). The dark green lines are Ethernet lines. The black lines are other radio frequency (RF) signals, such as the calibration signals, or the GNSS signals for the GNSS receiver. SDR stands for Software Defined Radio. IMU stands for Inertial Measurement Unit. GPS-DO stands for GPS—Disciplined Oscillator.</p>
Full article ">Figure 2
<p>3D model of the antenna elements (<b>a</b>), and photography of one of the manufactured antennas (<b>b</b>). The antenna measures 95 × 95 × 8 mm.</p>
Full article ">Figure 3
<p>Measured antenna radiation pattern of one of the manufactured antennas at L1/E1 (<b>a</b>), and at L5/E5 (<b>b</b>).</p>
Full article ">Figure 4
<p>Measured cross-polarization discrimination patterns at L1/E1 (<b>a</b>), and L5/E5a (<b>b</b>) frequency bands. The colorscale was limited to ±20 dB to ease the viewing.</p>
Full article ">Figure 5
<p>Top view of the up-looking array (<b>a</b>), and back view (uncovered) of the down-looking array (<b>b</b>). The up-looking array image includes the baseline used to define each antenna element by a <math display="inline"> <semantics> <mrow> <mo>{</mo> <msub> <mi>m</mi> <mi>i</mi> </msub> <mo>,</mo> <msub> <mi>n</mi> <mi>i</mi> </msub> <mo>}</mo> </mrow> </semantics> </math> pair of values. The azimuthal pointing angle <math display="inline"> <semantics> <mi>φ</mi> </semantics> </math> is defined anti-clockwise from vector <math display="inline"> <semantics> <mover accent="true"> <mi>m</mi> <mo stretchy="false">→</mo> </mover> </semantics> </math> to <math display="inline"> <semantics> <mover accent="true"> <mi>n</mi> <mo stretchy="false">→</mo> </mover> </semantics> </math>. The off-boresight pointing angle <math display="inline"> <semantics> <mi>θ</mi> </semantics> </math> is defined from vector <math display="inline"> <semantics> <mover accent="true"> <mi>z</mi> <mo stretchy="false">→</mo> </mover> </semantics> </math> to the antenna array ground plane.</p>
Full article ">Figure 6
<p>Block diagram of one of the beamformers (<b>a</b>), close view of one of the phase shifters manufactured (<b>b</b>), and front picture of one of the assembled beamformers (<b>c</b>). Note in the assembled beamformer 19 inputs (red), and the 4 outputs (blue).</p>
Full article ">Figure 7
<p>Used model for a whole RF chain.</p>
Full article ">Figure 8
<p>Ideal gain circle (dashed line), and modeled gain ellipsoid (solid line) in the IQ gain plane.</p>
Full article ">Figure 9
<p>Developed selection algorithm to choose the best satellites available [<a href="#B13-sensors-19-01019" class="html-bibr">13</a>].</p>
Full article ">Figure 10
<p>Block diagram of the calibration system. <math display="inline"> <semantics> <msub> <mi>f</mi> <mrow> <mi>L</mi> <mi>O</mi> </mrow> </msub> </semantics> </math> is the frequency of the local oscillator (LO), <math display="inline"> <semantics> <msub> <mi>B</mi> <mi>w</mi> </msub> </semantics> </math> is the −3 dB bandwidth, <math display="inline"> <semantics> <msub> <mi>f</mi> <mi>c</mi> </msub> </semantics> </math> is the −3 dB cut-off frequency, <span class="html-italic">L</span> stands for attenuation, and <span class="html-italic">G</span> stands for gain.</p>
Full article ">Figure 11
<p>Measured spectra of the transmitted calibration signal at L1/E1 frequency band (<b>right</b>), and at L5/E5a frequency band (<b>left</b>). In green, the region of the spectrum that will be used for calibration. The aliases of the up-conversion are highlighted in red. The local oscillator leakages are highlighted in yellow.</p>
Full article ">Figure 12
<p>MIR rack. The top floor of the rack has the embedded system (red), the power supplies (white), the calibration system (blue), and the GNSS receiver (green). The middle floor has the beamformers (orange). The lower floor has the Octoclock (magenta), the SDR (purple), the gigabit ethernet swithc (cyan), and the extra amplifiers for the down-looking RF chain (yellow).</p>
Full article ">Figure 13
<p>Received and calibration signal’s paths to the receiver in one of the bands and beams.</p>
Full article ">Figure 14
<p>Estimated phase (<b>a</b>) and gain (<b>b</b>) of the different <math display="inline"> <semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> </mrow> </msub> </semantics> </math> intervals. The time averaged phase response estimation is plot in yellow, and the linear approximation in cyan. A value of <math display="inline"> <semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>500</mn> </mrow> </semantics> </math> was used. The time geometrically averaged amplitude of the transfer function is plot in yellow, and the geometric average gain is plot as a dot in cyan.</p>
Full article ">Figure 15
<p>Estimated phase between the direct and reflected RF channels for the four different USRPs during three sets of experiments. The lines show the drift between the estimations, and the markers are error bars.</p>
Full article ">Figure 16
<p>Estimated group delay between the direct and reflected RF channels for the four different USRPs during three sets of experiments. The lines show the drift between the estimations, and the markers are error bars.</p>
Full article ">Figure 17
<p>Estimated gain between the direct and reflected RF channels for the four different USRPs during three sets of experiments. The lines show the drift between the estimations, and the markers are error bars. The mean and the standard deviation have been computed as geometric mean and geometric standard deviation.</p>
Full article ">Figure 18
<p>MIR down-looking array being calibrated and measured at the Universitat Politècnica de Catalunya (UPC) anechoic chamber. The 19 elements MIR array is at the back left, while the transmitting reference horn antenna is at the front right. The transmitting antenna has linear polarization and can rotate along the boresight axis.</p>
Full article ">Figure 19
<p>IQ gain circles before (<b>a</b>) and after calibration (<b>b</b>). The radii lines mark where the circle starts to “rotate”. A red circle with average radius has been added to ease the viewing the effect of the amplitude unbalance.</p>
Full article ">Figure 20
<p>Effect of the calibration in the radiation patterns. Planar cuts in the <math display="inline"> <semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics> </math> plane at L1/E1 (<b>a</b>) and L5/E5 (<b>b</b>).</p>
Full article ">Figure 21
<p>Planar cuts of the radiation pattern along the <math display="inline"> <semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics> </math> plane when tilting the beam in the same plane for different off-boresight pointing angles, at the L1/E1 (<b>a</b>) and L5/E5 (<b>b</b>) frequency bands.</p>
Full article ">Figure 22
<p>Directivity loss as a function of the pointing direction at L1/E1 (<b>a</b>), and at L5/E5 (<b>b</b>), for the down-looking array.</p>
Full article ">Figure 23
<p>Azimuthal cuts of the full radiation pattern pointing to the boresight of the antenna, at the L1/E1 (<b>a</b>) and L5/E5 (<b>b</b>) frequency bands.</p>
Full article ">Figure 24
<p>Frontal hemisphere of the measured radiation pattern of the down-looking array, at the L1/E1 (<b>a</b>) and at L5/E5 (<b>b</b>) frequency bands.</p>
Full article ">Figure 25
<p>Side lobes for different azimuthal cuts of the antenna pattern when pointing to an off-boresight angle of 35<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math> at the L1/E1 (<b>a</b>) and L5/E5 (<b>b</b>) frequency bands, respectively.</p>
Full article ">Figure 26
<p>MIR up-looking array and rack during the rooftop experiments (<b>a</b>), and retrieved waveforms from GPS and Galileo satellites in both L1/E1 and L5/E5 frequency bands (<b>b</b>) during the rooftop experiments [<a href="#B23-sensors-19-01019" class="html-bibr">23</a>].</p>
Full article ">Figure 27
<p>Tracked C/N0 in the L1-1 beam (<b>a</b>) and L1-2 beam (<b>b</b>), both red circled, during the experiment in the laboratory rooftop [<a href="#B24-sensors-19-01019" class="html-bibr">24</a>].</p>
Full article ">Figure 27 Cont.
<p>Tracked C/N0 in the L1-1 beam (<b>a</b>) and L1-2 beam (<b>b</b>), both red circled, during the experiment in the laboratory rooftop [<a href="#B24-sensors-19-01019" class="html-bibr">24</a>].</p>
Full article ">
1242 KiB  
Article
On the Synergy of Airborne GNSS-R and Landsat 8 for Soil Moisture Estimation
by Nilda Sánchez, Alberto Alonso-Arroyo, José Martínez-Fernández, María Piles, Ángel González-Zamora, Adriano Camps and Mercè Vall-llosera
Remote Sens. 2015, 7(8), 9954-9974; https://doi.org/10.3390/rs70809954 - 5 Aug 2015
Cited by 42 | Viewed by 7148
Abstract
While the synergy between thermal, optical, and passive microwave observations is well known for the estimation of soil moisture and vegetation parameters, the use of remote sensing sources based on the Global Navigation Satellite Systems (GNSS) remains unexplored. During an airborne campaign performed [...] Read more.
While the synergy between thermal, optical, and passive microwave observations is well known for the estimation of soil moisture and vegetation parameters, the use of remote sensing sources based on the Global Navigation Satellite Systems (GNSS) remains unexplored. During an airborne campaign performed in August 2014, over an agricultural area in the Duero basin (Spain), an innovative sensor developed by the Universitat Politècnica de Catalunya-Barcelona Tech based on GNSS Reflectometry (GNSS-R) was tested for soil moisture estimation. The objective was to evaluate the combined use of GNSS-R observations with a time-collocated Landsat 8 image for soil moisture retrieval under semi-arid climate conditions. As a ground reference dataset, an intensive field campaign was carried out. The Light Airborne Reflectometer for GNSS-R Observations (LARGO) observations, together with optical, infrared, and thermal bands from Landsat 8, were linked through a semi-empirical model to field soil moisture. Different combinations of vegetation and water indices with LARGO subsets were tested and compared to the in situ measurements. Results showed that the joint use of GNSS-R reflectivity, water/vegetation indices and thermal maps from Landsat 8 not only allows capturing soil moisture spatial gradients under very dry soil conditions, but also holds great promise for accurate soil moisture estimation (correlation coefficients greater than 0.5 were obtained from comparison with in situ data). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area with the main land uses and the sampling layout.</p>
Full article ">Figure 2
<p>(<b>left</b>) Orthophoto of the study area and (<b>right</b>) preliminary reflectivity (dB) map from the UPC LARGO GNSS-R instrument.</p>
Full article ">Figure 3
<p>Results of the pre-treatment of the datasets: (<b>a</b>) Landsat 8 true-color composite, (<b>b</b>) Landsat 8 temperature, (<b>c</b>) LARGO observations binned between 40° and 60° incidence angles, (<b>d</b>) LARGO observations binned between 60° and 80° incidence angles, (<b>e</b>) reflectivity map from LARGO interpolation at 40°–60° incidence angles, (<b>f</b>) reflectivity map from LARGO interpolation at 60°–80°.</p>
Full article ">Figure 4
<p>Field-based soil moisture map interpolated to 30 m.</p>
Full article ">Figure 5
<p>Landsat 8 bands reflectance (<b>a</b>) and indices (<b>b</b>), both of them averaged for the main land covers in the study area. In (<b>c</b>), several examples of these land covers over a 5-4-3 composite.</p>
Full article ">Figure 6
<p>Relationship between the indices the best correlated with soil moisture and soil moisture.</p>
Full article ">Figure 7
<p>(<b>a</b>) Ground observations distribution. (<b>b</b> to <b>h</b>) Estimated soil moisture (% vol.) maps following the seven alternatives considered.</p>
Full article ">Figure 8
<p>Relationship between the estimated soil moisture resulting from the linking model and the <span class="html-italic">in situ</span> soil moisture measurements using GVRI, NDWI2-red, and NDWI2-NIR.</p>
Full article ">
2910 KiB  
Article
New Passive Instruments Developed for Ocean Monitoring at the Remote Sensing Lab—Universitat Politècnica de Catalunya
by Adriano Camps, Xavier Bosch-Lluis, Isaac Ramos-Perez, Juan F. Marchán-Hernández, Nereida Rodríguez, Enric Valencia, Jose M. Tarongi, Albert Aguasca and René Acevo
Sensors 2009, 9(12), 10171-10189; https://doi.org/10.3390/s91210171 - 14 Dec 2009
Cited by 15 | Viewed by 16942
Abstract
Lack of frequent and global observations from space is currently a limiting factor in many Earth Observation (EO) missions. Two potential techniques that have been proposed nowadays are: (1) the use of satellite constellations, and (2) the use of Global Navigation Satellite Signals [...] Read more.
Lack of frequent and global observations from space is currently a limiting factor in many Earth Observation (EO) missions. Two potential techniques that have been proposed nowadays are: (1) the use of satellite constellations, and (2) the use of Global Navigation Satellite Signals (GNSS) as signals of opportunity (no transmitter required). Reflectometry using GNSS opportunity signals (GNSS-R) was originally proposed in 1993 by Martin-Neira (ESA-ESTEC) for altimetry applications, but later its use for wind speed determination has been proposed, and more recently to perform the sea state correction required in sea surface salinity retrievals by means of L-band microwave radiometry (TB). At present, two EO space-borne missions are currently planned to be launched in the near future: (1) ESA’s SMOS mission, using a Y-shaped synthetic aperture radiometer, launch date November 2nd, 2009, and (2) NASA-CONAE AQUARIUS/SAC-D mission, using a three beam push-broom radiometer. In the SMOS mission, the multi-angle observation capabilities allow to simultaneously retrieve not only the surface salinity, but also the surface temperature and an “effective” wind speed that minimizes the differences between observations and models. In AQUARIUS, an L-band scatterometer measuring the radar backscatter (σ0) will be used to perform the necessary sea state corrections. However, none of these approaches are fully satisfactory, since the effective wind speed captures some sea surface roughness effects, at the expense of introducing another variable to be retrieved, and on the other hand the plots (TB0) present a large scattering. In 2003, the Passive Advance Unit for ocean monitoring (PAU) project was proposed to the European Science Foundation in the frame of the EUropean Young Investigator Awards (EURYI) to test the feasibility of GNSS-R over the sea surface to make sea state measurements and perform the correction of the L-band brightness temperature. This paper: (1) provides an overview of the Physics of the L-band radiometric and GNSS reflectometric observations over the ocean, (2) describes the instrumentation that has been (is being) developed in the frame of the EURYI-funded PAU project, (3) the ground-based measurements carried out so far, and their interpretation in view of placing a GNSS-reflectometer as secondary payload in future SMOS follow-on missions. Full article
(This article belongs to the Special Issue State-of-the-Art Sensors Technology in Spain)
Show Figures


<p>Derivation of the brightness temperature sensitivity to wind speed. Scatter plots of the brightness temperature change at (a) horizontal (ΔT<sub>B h</sub>) and (b) vertical (ΔT<sub>B v</sub>) polarizations <span class="html-italic">vs.</span> the 10-m height wind speed (U<sub>10</sub>). Solid line: linear fit, and dashed lines: percentile 50% as a function of U<sub>10</sub> for incidence angles from 25° to 65°. (c) Derived T<sub>B h,v</sub> sensitivity to wind speed as a function of (solid line) polarization and incidence angle, associated ±1 σ error bars, and (dashed lines) linear fit. All data points used (<a href="#f6-sensors-09-10171" class="html-fig">Figure 6</a> from [<a href="#b3-sensors-09-10171" class="html-bibr">3</a>]).</p>
Full article ">
<p>Sun glint over the sea under: (a) calm and (b) windy conditions.</p>
Full article ">
<p>(left) GNSS-R geometry of observation from Garraf coast (South of Barcelona). Red: isorange annuli (same delay), Yellow: isoDoppler hyperbolae (same Doppler), Green dots with same delay and Doppler. (right) measured DDM.</p>
Full article ">
<p>Measured (not normalized) 1 s incoherently integrated DDM with the threshold applied to compute its volume. Noise is well below the threshold [<a href="#b9-sensors-09-10171" class="html-bibr">9</a>].</p>
Full article ">
<p>Block diagram of the <span class="html-italic">PAU/RAD</span> concept with just one receiver.</p>
Full article ">
<p>Schematic of the <span class="html-italic">PAU-Real Aperture</span> with 16 elements [<a href="#b4-sensors-09-10171" class="html-bibr">4</a>].</p>
Full article ">
<p>(a) <span class="html-italic">PAU-Real Aperture</span> without radome in UPC anechoic chamber for testing beamforming performance, (b) front view of the 4 × 4 element array, and (c) inside of the instrument showing the distribution of the dual-polarization receivers, FPGAs for data processing, and the temperature control.</p>
Full article ">
<p>(a) <span class="html-italic">PAU</span>-Synthetic Aperture's topology, (b) View of the whole instrument with one arm opened, and (c) Detail of the hub, including digital correlator unit (right hand side, bottom), base-band clock distribution (right-hand side, top), calibration unit, including 2-level noise source and Pseudo-Random Noise (PRN) generator (left-hand side, bottom), power supply (left-hand side, up).</p>
Full article ">
<p><span class="html-italic">griPAU</span> instrument mounted on a 6 m tower with automatic antenna positioner, during the ALBATROSS 2009 field experiment in Gran Canaria (Canary Islands, Spain).</p>
Full article ">
8026 KiB  
Article
PAU/GNSS-R: Implementation, Performance and First Results of a Real-Time Delay-Doppler Map Reflectometer Using Global Navigation Satellite System Signals
by Juan Fernando Marchan-Hernandez, Adriano Camps, Nereida Rodriguez-Alvarez, Xavier Bosch-Lluis, Isaac Ramos-Perez and Enric Valencia
Sensors 2008, 8(5), 3005-3019; https://doi.org/10.3390/s8053005 - 6 May 2008
Cited by 14 | Viewed by 14010
Abstract
Signals from Global Navigation Satellite Systems (GNSS) were originally conceived for position and speed determination, but they can be used as signals of opportunity as well. The reflection process over a given surface modifies the properties of the scattered signal, and therefore, by [...] Read more.
Signals from Global Navigation Satellite Systems (GNSS) were originally conceived for position and speed determination, but they can be used as signals of opportunity as well. The reflection process over a given surface modifies the properties of the scattered signal, and therefore, by processing the reflected signal, relevant geophysical data regarding the surface under study (land, sea, ice…) can be retrieved. In essence, a GNSS-R receiver is a multi-channel GNSS receiver that computes the received power from a given satellite at a number of different delay and Doppler bins of the incoming signal. The first approaches to build such a receiver consisted of sampling and storing the scattered signal for later post-processing. However, a real-time approach to the problem is desirable to obtain immediately useful geophysical variables and reduce the amount of data. The use of FPGA technology makes this possible, while at the same time the system can be easily reconfigured. The signal tracking and processing constraints made necessary to fully design several new blocks. The uniqueness of the implemented system described in this work is the capability to compute in real-time Delay-Doppler maps (DDMs) either for four simultaneous satellites or just one, but with a larger number of bins. The first tests have been conducted from a cliff over the sea and demonstrate the successful performance of the instrument to compute DDMs in real-time from the measured reflected GNSS/R signals. The processing of these measurements shall yield quantitative relationships between the sea state (mainly driven by the surface wind and the swell) and the overall DDM shape. The ultimate goal is to use the DDM shape to correct the sea state influence on the L-band brightness temperature to improve the retrieval of the sea surface salinity (SSS). Full article
(This article belongs to the Special Issue Remote Sensing of Natural Resources and the Environment)
Show Figures


<p>Data paths in the implemented reflectometer.</p>
Full article ">
<p>Delay Doppler Map (DDM) generator architecture. The ‘Amps to 2C’ block performs the two's complement encoding, whereas the ‘AC’ block accumulates the inputs.</p>
Full article ">
<p>Detail of the main cores that compose the DDM generator: (a) ‘Sign and amplitudes’ block in <a href="#f2-sensors-08-03005" class="html-fig">Fig. 2</a> and (b) ‘Oscillator block’ in <a href="#f3-sensors-08-03005" class="html-fig">Fig. 3a</a>.</p>
Full article ">
<p>Phase MSBs role on the wave generation.</p>
Full article ">
<p>Sketch of the <span class="html-italic">C/A generator</span> block. The <span class="html-italic">Resampler</span> component drives the Linear Feedback Shift Registers (LFSR) inside the block <span class="html-italic">G1 and G2,</span> out of where the C/A code is obtained. The values of M1, M2 and nint_value determine the code offset and the satellite ID.</p>
Full article ">
<p>Architecture of the delay offset block.</p>
Full article ">
<p>Correlation of the 5745-sample C/A code using standard FFT blocks with 2<sup>13</sup> = 8192 input samples. (a) Two correlation maxima instead of a single one. (b) Relation of the actual code offset and the retrieved one.</p>
Full article ">
<p>Interrupts and their associated Interrupt Service Routine (ISR) dataflow: (a) Buffer ISR and (b) UART ISR</p>
Full article ">
<p>Implemented interface for receiving, displaying and storing the DDMs. On the left the 4-satellite operational mode is seen, whereas on the right the 1-satellite operational mode is shown. These videos correspond to actual measurements on the Garraf Cliffs (click to play them).</p>
Full article ">
Back to TopTop