OBSERVATION OF HURRICANE WINDS USING SYNTHETIC APERTURE RADAR

Similar documents
Development of SAR-Derived Ocean Surface Winds at NOAA/NESDIS

Wind Direction Analysis over the Ocean using SAR Imagery

Geophysical Model Functions for the Retrieval of Ocean Surface Winds

PRELIMINARY STUDY ON DEVELOPING AN L-BAND WIND RETRIEVAL MODEL FUNCTION USING ALOS/PALSAR

EVALUATION OF ENVISAT ASAR WAVE MODE RETRIEVAL ALGORITHMS FOR SEA-STATE FORECASTING AND WAVE CLIMATE ASSESSMENT

IMPROVED BAYESIAN WIND VECTOR RETRIEVAL SCHEME USING ENVISAT ASAR DATA: PRINCIPLES AND VALIDATION RESULTS

CHANGE OF THE BRIGHTNESS TEMPERATURE IN THE MICROWAVE REGION DUE TO THE RELATIVE WIND DIRECTION

Global Wind Speed Retrieval From SAR

Reprocessed QuikSCAT (V04) Wind Vectors with Ku-2011 Geophysical Model Function

Using several data sources for offshore wind resource assessment

OCEAN vector winds from the SeaWinds instrument have

THE QUALITY OF THE ASCAT 12.5 KM WIND PRODUCT

Azimuthal variations of X-band medium grazing angle sea clutter

ON THE USE OF DOPPLER SHIFT FOR SAR WIND RETRIEVAL

Synthetic Aperture Radar imaging of Polar Lows

RZGM Wind Atlas of Aegean Sea with SAR data

High resolution wind retrieval for SeaWinds

HIGH RESOLUTION WIND FIELDS OVER THE BLACK SEA DERIVED FROM ENVISAT ASAR DATA USING AN ADVANCED WIND RETRIEVAL ALGORITHM

GLOBAL VALIDATION AND ASSIMILATION OF ENVISAT ASAR WAVE MODE SPECTRA

ADVANCES ON WIND ENERGY RESOURCE MAPPING FROM SAR

Application of a new algorithm using Doppler information to retrieve complex wind fields over the Black Sea from ENVISAT SAR images

ERS-1/2 Scatterometer new products: mission reprocessing and data quality improvement

Surface Wave Parameters Retrieval in Coastal Seas from Spaceborne SAR Image Mode Data

THE POLARIMETRIC CHARACTERISTICS OF BOTTOM TOPOGRAPHY RELATED FEATURES ON SAR IMAGES

STUDY OF LOCAL WINDS IN MOUNTAINOUS COASTAL AREAS BY MULTI- SENSOR SATELLITE DATA

ENVISAT WIND AND WAVE PRODUCTS: MONITORING, VALIDATION AND ASSIMILATION

DANISH METEOROLOGICAL INSTITUTE

Comparison of Geophysical Model Functions for SAR Wind Speed Retrieval in Japanese Coastal Waters

ERS WAVE MISSION REPROCESSING- QC SUPPORT ENVISAT MISSION EXTENSION SUPPORT

High resolution wind fields over the Black Sea derived from Envisat ASAR data using an advanced wind retrieval algorithm

Ocean Wave Parameters Retrieval from TerraSAR-X Images Validated against Buoy Measurements and Model Results

THE EFFECT OF RAIN ON ASCAT OBSERVATIONS OF THE SEA SURFACE RADAR CROSS SECTION USING SIMULTANEOUS 3-D NEXRAD RAIN MEASUREMENTS

Satellite information on ocean vector wind from Scatterometer data. Giovanna De Chiara

Global Ocean Internal Wave Database

Offshore wind resource mapping in Europe from satellites

Analyses of Scatterometer and SAR Data at the University of Hamburg

USING SATELLITE SAR IN OFFSHORE WIND RESOURCE ASSESSMENT

Wavelet Analysis for Wind Fields Estimation

Sentinel-1A Ocean Level-2 Products Validation Strategy

Air-Sea Interaction Spar Buoy Systems

Combining wind and rain in spaceborne scatterometer observations: modeling the splash effects in the sea surface backscattering coefficient

IMPROVED OIL SLICK IDENTIFICATION USING CMOD5 MODEL FOR WIND SPEED EVALUATION ON SAR IMAGES

Comparison of data and model predictions of current, wave and radar cross-section modulation by seabed sand waves

WIND ENERGY MAPPING USING SYNTHETIC APERTURE RADAR. C. B. Hasager

SENSOR SYNERGY OF ACTIVE AND PASSIVE MICROWAVE INSTRUMENTS FOR OBSERVATIONS OF MARINE SURFACE WINDS

Towards an Optimal Inversion Method. for SAR Wind Retrieval 1

RIVET Satellite Remote Sensing and Small Scale Wave Process Analysis

Assessing the quality of Synthetic Aperture Radar (SAR) wind retrieval in coastal zones using multiple Lidars

HIGH RESOLUTION WIND AND WAVE MEASUREMENTS FROM TerraSAR-X IN COMPARISON TO MARINE FORECAST

SIMON YUEH, WENQING TANG, ALEXANDER FORE, AND JULIAN CHAUBELL JPL-CALTECH, PASADENA, CA, USA GARY LAGERLOEF EARTH AND SPACE RESEARCH, SEATTLE, WA, US

RapidScat wind validation report

The Ice Contamination Ratio Method: Accurately Retrieving Ocean Winds Closer to the Sea Ice Edge While Eliminating Ice Winds

Cross-Calibrating OSCAT Land Sigma-0 to Extend the QuikSCAT Land Sigma-0 Climate Record

Super-parameterization of boundary layer roll vortices in tropical cyclone models

Introduction EU-Norsewind

HIGH RESOLUTION WIND RETRIEVAL FOR SEAWINDS ON QUIKSCAT. Jeremy B. Luke. A thesis submitted to the faculty of. Brigham Young University

Review of Equivalent Neutral Winds and Stress

Offshore wind mapping Mediterranean area using SAR

THE SEAWINDS scatterometer was flown twice, once on

OPERATIONAL AMV PRODUCTS DERIVED WITH METEOSAT-6 RAPID SCAN DATA. Arthur de Smet. EUMETSAT, Am Kavalleriesand 31, D Darmstadt, Germany ABSTRACT

Marine wind speed retrieval from RADARSAT-2 dual-polarization imagery

SPATIAL AND TEMPORAL VARIATIONS OF INTERNAL WAVES IN THE NORTHERN SOUTH CHINA SEA

WaMoS II Wave Monitoring System

Does direct impact of SST on short wind waves matter for scatterometry?

The RSS WindSat Version 7 All-Weather Wind Vector Product

Wind Stress Working Group 2015 IOVWST Meeting Portland, OR

SATELLITE REMOTE SENSING APPLIED TO OFF-SHORE WIND ENERGY

Scatterometer Data Interpretation: Measurement Space and Inversion

830 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 48, NO. 2, FEBRUARY /$ IEEE

Determination Of Nearshore Wave Conditions And Bathymetry From X-Band Radar Systems

Flow separation and lee-waves in the marine atmosphere

Observation of Wind Direction Change on the Sea Surface Temperature Front Using High-Resolution Full Polarimetric SAR Data

An Atlas of Oceanic Internal Solitary Waves (May 2002) by Global Ocean Associates Prepared for the Office of Naval Research - Code 322PO

Analyzing Surface Wind Fields Near Lower Cook Inlet And Kodiak Waters Using SAR

Legendre et al Appendices and Supplements, p. 1

Refined Source Terms in WAVEWATCH III with Wave Breaking and Sea Spray Forecasts

High-Frequency Scattering from the Sea Surface and Multiple Scattering from Bubbles

High-Resolution Measurement-Based Phase-Resolved Prediction of Ocean Wavefields

An Atlas of Oceanic Internal Solitary Waves (February 2004) by Global Ocean Associates Prepared for Office of Naval Research Code 322 PO

On Measurement of the Sea Surface Wind Vector by the Airborne Weather Radars Having a Various Scanning Sector

OBSERVATIONS OF ATMOSPHERIC GRAVITY WAVES OVER THE CHINESE SEAS BY SPACEBORNE SYNTHETIC APERTURE RADAR

Validation of 12.5 km Resolution Coastal Winds. Barry Vanhoff, COAS/OSU Funding by NASA/NOAA

Beach Wizard: Development of an Operational Nowcast, Short-Term Forecast System for Nearshore Hydrodynamics and Bathymetric Evolution

First studies with the high-resolution coupled wave current model CWAM and other aspects of the project Sea State Monitor

Prediction of Nearshore Waves and Currents: Model Sensitivity, Confidence and Assimilation

IMPROVEMENTS IN THE USE OF SCATTEROMETER WINDS IN THE OPERATIONAL NWP SYSTEM AT METEO-FRANCE

Determination of Nearshore Wave Conditions and Bathymetry from X-Band Radar Systems

High resolution mapping of Bora winds in the northern Adriatic Sea using synthetic aperture radar

NONLINEAR INTERNAL WAVES IN THE SOUTH CHINA SEA

SeaWinds wind Climate Data Record validation report

A. Bentamy 1, S. A. Grodsky2, D.C. Fillon1, J.F. Piollé1 (1) Laboratoire d Océanographie Spatiale / IFREMER (2) Univ. Of Maryland

Observed near-surface currents under high wind speeds

DEVELOPMENT AND VALIDATION OF A SAR WIND EMULATOR

ERGS are large expanses of sand in the desert. Aeolian

QuikScat/Seawinds Sigma-0 Radiometric and Location Accuracy Requirements for Land/Ice Applications

Influence of rounding corners on unsteady flow and heat transfer around a square cylinder

Synthetic aperture radar observations of sea surface signatures of atmospheric gravity waves over Southeast Asian coastal waters

Monitoring Conditions Offshore with Satellites

Impact of Typhoons on the Western Pacific: Temporal and horizontal variability of SST cooling Annual Report, 2011 James F. Price

PROPAGATION OF LONG-PERIOD WAVES INTO AN ESTUARY THROUGH A NARROW INLET

Transcription:

OBSERVATION OF HURRICANE WINDS USING SYNTHETIC APERTURE RADAR Jochen Horstmann 1, Wolfgang Koch 1,DonaldR.Thompson 2, and Hans C. Graber 3 1 Institute for Coastal Research, GKSS Research Center, Geesthacht, Germany 2 Applied Physics Lab, Johns Hopkins University, Laurel, Maryland, United States 3 Center for Southeastern Tropical Advanced Remote Sensing, University of Miami, Miami, Florida, United States ABSTRACT In the last five years several synthetic aperture radar (SAR) images of hurricanes have been acquired by the Canadian satellite RADARSAT as well as the European satellite ENVISAT. Several of these SAR images have captured hurricanes of category 4 and 5. These SAR data give the unique opportunity to investigate the possibilities of SAR data for observation and forecast of hurricanes. In this study wind fields under hurricane conditions are retrieved from the SAR utilizing WiSAR. WiSAR is an algorithm, which has shown to give good results under low and moderate wind conditions. The algorithm extracts wind directions from wind induced streaks imaged by the SAR at scales above 200m. Wind speeds are extracted from the SAR measured normalized radar cross section (NRCS) utilizing the C-band model CMOD5, which describes the dependency of the NRCS on wind. It will be shown that the algorithm enables to measure wind directions as well as wind speeds of over 50 ms 1. The SARretrieved wind fields are compared to results of a high resolution numerical hurricane model. 1. INTRODUCTION The ability of space borne microwave radars to measure the wind vector near the ocean surface relies on the fact that the near-surface wind field generates small scale roughness on the surface that increases with increasing wind speed. For radar backscatter at moderate incident angles, (roughly 20 to 60 ), the normalized radar cross section (NRCS) is proportional to the spectral density of the surface roughness on the scales comparable to the radar wavelength. For this mid-incidence regime and wind speeds below 25 ms 1 or so, the NRCS is typically largest when the wind blows directly toward the radar (upwind) and decreases to a minimum when the wind direction is orthogonal to the radar look direction (cross wind). Another smaller maximum in NRCS occurs when the wind blows directly away from the radar (down wind). The relation between the near-surface wind vector and NRCS can be described by an equation of the form σ 0 = au γ [1 + b cos Φ + c cos 2Φ] (1) where σ 0 represents NRCS, u represents wind speed, Φ represents the relative angle between the radar look and wind direction, and θ is the nadir incidence angle. The quantities a, γ, b and c are empirical parameters determined by measured data and are usually dependent on incidence angle and wind speed. Proceedings of SEASAR 2006, 23-26 January 2006, Frascati, Italy, (ESA SP-613, April 2006) Equation (1) shows that σ 0 is an exponential function of wind speed and a harmonic function of its direction. It is important to note that though the model function can relate wind speed and direction to NRCS, a specific NRCS value cannot be associated with a unique wind speed and direction pair. This significantly complicates the inversion from NRCS to wind speed. If wind direction is known a priori, it is possible to use Eq.( 1) and the measured NRCS to estimate wind speed. An approach convenient for operational use is to use predicted wind directions from operational meteorological models (Monaldo et al. 2004). Another method is to estimate the wind direction directly using linear features in the SAR image itself (Horstmann et al. 2002). Both approaches have proven successful for moderate wind speed regimes (below 25 ms 1 ). At high wind speeds, especially hurricane winds, the surface roughness structure grows more complicated not only because of the addition of more wave breaking and surface foam, but also because of the more complex structure of the long wave field generated at previous times when the hurricane was at a different location. It is, therefore, likely that conventional empirical geophysical model functions (GMFs) developed using data for moderate wind conditions may also begin to break down. In the following sections, we briefly describe some of the salient aspects of SAR wind retrieval schemes and the results of applying them in the high wind regimes observed in Hurricane Ivan (Horstmann et al. 2005) and Katrina. We will show that the newly-developed Cmod5 GMF (Hersbach 2003) outperforms the commonly used Cmod4 (Stoffelen & Anderson 1997) in this high wind speed case studies. In particular, we will show using NRCS measurements from two SAR images collected over Hurricanes that the dependence of NRCS on the relative angle between the radar look direction and the local wind direction weakens, and perhaps disappears altogether, at very high wind speeds near the hurricane eye. We will further show that such dependence is better modelled by the Cmod5 GMF. 2. SAR WIND RETRIEVAL 2.1. Wind Direction Estimation The most popular method for SAR wind direction retrieval is based on the imaging of linear features aligned along the mean surface wind direction (Gerling 1986). Most of these features are associated with wind streaks (Drobinski & Foster 2003) and marine

2 atmospheric boundary layer (MABL) rolls (Etling & Brown 1993) that are visible in SAR images at scales larger than 200 m or so. To retrieve the orientation of these linear features, three methods have been developed: the Fast Fourier Transformation Method (FFT-Method) (Gerling 1986; Lehner et al. 1998) the Local Gradient Method (LG-Method) (Horstmann et al. 2002; Koch 2004) and the Projection Method (Wackerman et al. 2004). Comparisons among these methods have shown that the LG-Method yields the smallest residuals, with a typical error of 18 (Horstmann & Koch 2005). Investigations of the orientation of wind streaks at scales larger than 50m with real aperture radar (RAR) imagery have confirmed that wind streaks are very well aligned with the mean surface wind direction (Dankert et al. 2003). These results also suggest that wind direction retrievals are likely to be most accurate when linear features at the smallest possible scales visible from spaceborne SARs are used. These scales are typically 200 m, and are limited by the spatial resolution of the SAR system. Wind direction retrieval based on features whose scales are > 3km can often represent MABL rolls, whose orientations are more likely to significantly differ from the mean surface wind direction (Etling & Brown 1993). For the present investigation, the LG-Method is utilized to retrieve the SAR wind direction. In this approach, a SAR image is sequentially smoothed and reduced to resolutions of 100, 200 and 400 m. The resulting three SAR images retain spatial scales greater than 200, 400 and 800 m, respectively. From each of these images, local directions defined by the normal to the local gradient (to within a 180 ambiguity), are computed. In the next step, pixels associated with non wind induced phenomena, e.g. land, surface slicks, and sea ice, are masked and excluded from further analysis. From all of these retrieved wind directions, only the most frequent directions in a predefined grid cell are selected. These resulting wind directions from the 100, 200 and 400 m resolution SAR images typically vary by only a few degrees, except for cases where additional features are present in the image; e.g., ocean surface waves, internal waves, and artifacts from the original SAR image processing. The 180 directional ambiguity can be removed if wind shadowing is present. This is often the case in the lee of coastlines. If such features are not present in the image other a priori information, e.g., weather charts, atmospheric models or in situ measurements, are used to remove ambiguities. 2.2. Wind Speed Retrieval With the wind direction determined, we can use the NRCS measured by the SAR to retrieve the wind speed using GMFs of the form given in Eq. (1). For C-band, VV-polarization NRCS, there are a number of popular model functions for which the coefficients and exact model form have been determined empirically using scatterometer data acquired by the European satellites ERS and ERS-2 and co-located winds from the European Center for Medium-Range Weather Forecast. The most commonly used C- band model function is Cmod4 (Stoffelen & Anderson 1997), and the most recently developed GMF is Cmod5 (Hersbach 2003). Both of these model functions are directly applicable for wind speed retrieval from C-band VV polarized SAR images (e.g. Vachon & Dobson 1996; Lehner et al. 1998; Horstmann et al. 2003). For wind speed retrieval from C-band SAR images acquired at HHpolarization (i.e., the configuration for RADARSAT- 1 images), no similar well-developed GMF exists. To meet this deficiency a hybrid model function is used that consists of one of the prior mentioned VVpolarization empirical models and a C-band polarization ratio (Horstmann et al. 2000; Thompson & Beal 2000; Vachon & Dobson 2000). The polarization ratio (PR) is defined as the ratio of HH-polarization NRCS to VV-polarization NRCS. The nature of the PR is an active area of research, and several different PRs have been proposed in literature (Thompson et al. 1998; Mouche et al. 2005). The PR proposed by Thompson et al. (1998) neglects wind speed and wind direction dependence and has the form PR = (1 + α tan2 θ) 2 (1 + 2 tan 2 θ), (2) 2 where α is an empirical constant set to 0.6 inorder to yield consistency with the measurements of Unal et al. (1991). Several other values for α have been suggested in literature for RADARSAT- 1 and ENVISAT SAR data. These vary between 0.4 and 1.2 (Horstmann et al. 2000; Vachon & Dobson 2000; Monaldo et al. 2002; Horstmann & Koch 2005). Comparisons of RADARSAT SAR imagery produced from different SAR processing facilities showed that the different estimates of α may be due to the different calibrations of RADARSAT SAR data at different facilities. Recently, Mouche et al. (2005) have proposed a C-band PR, constructed using airborne RAR data acquired at C-band with both VV- and HH-polarization, that is also dependent on the wind direction. Comparisons of C-band SAR retrieved wind speeds, using the Cmod4 at low to moderate wind speeds (up to 20 ms 1 ) resulted in errors of 2ms 1 (Monaldo et al. 2002; Horstmann et al. 2003; Monaldo et al. 2004; Horstmann & Koch 2005). It is also well known that the Cmod4 underestimates the wind speeds at high winds (> 20 ms 1 ) when applied to scatterometer and SAR data (Donnelly et al. 1999; Katsaros et al. 2002; Horstmann et al. 2003). The new Cmod5 algorithm was specifically designed to provide better estimates of the NRCS at higher wind speeds (Hersbach 2003). It was constructed primarily using co-locations between ERS-2 scatterometer backscatter triplets and first-guess model winds. For extremely high wind conditions, results from recent aircraft campaigns were also included (Donnelly et al. 1999). Differences between Cmod4 and Cmod5 for low to moderate wind speeds are relatively minor. At high wind speeds (25 to 30 ms 1 ) however, the differences become quite significant. In partic-

3 ular, the NRCS from Cmod4 continues to increases monotonically with wind speed for all incident angles, while that predicted by Cmod5 increases much more slowly with wind speed and for lower incident angles (25 to 30 depending on wind speed), can even reach a maximum value and decrease with further increase in wind speed. Furthermore, dependence of the NRCS on wind direction from Cmod5 becomes much weaker in the high wind regime compared to that of Cmod4. 3. HIGH WIND SPEED DEPENDENCE OF THE NRCS A RADARSAT SAR image of Hurricane Ivan, acquired on September 10, 2004 at 2307 UTC is shown in Fig. 1. This image is ideally suited to investigate the dependence of the C-band NRCS on wind speed and direction under extreme wind conditions. At the time of acquisition, Hurricane Ivan was situated about 80 km south of Kingston, Jamaica, and was classified by the US National Hurricane Center as a category 4 hurricane with maximum wind speeds on the order of 65 ms 1. The color scale in Fig. 1 represents NRCS values from the calibrated SAR image in the range 6 db to 9 db. The roughly circular hurricane eye is easily seen in the center of the figure where the NRCS is quite low. The variation of the NRCS with distance from the eye can be seen using the radii of the concentric circles shown in the figure. The NRCS increases dramatically at a distance of about 12 km from the eye, and remains roughly independent of the angular location around the eye for distances out to about 18 km. One can also see in Fig. 1, evidence of counter-clockwise circulation from the wind streaks as well as the effect of rain 0 13 km 23 km 43 km 63 km Figure 1. SAR image acquired by the Canadian RADARSAT satellite on September 10, 2004 at 2307 UTC showing the surface signature of Hurricane Ivan. The plot beneath the image shows the incident angle as a function of distance. 13 km 23 km 43 km 63 km Figure 2. NRCS versus direction around the concentric circles in the SAR image of Fig. 1. bands (in a distance of 50 km). To quantitatively investigate the wind direction dependence of the NRCS at various distances from the hurricane eye, we show in Fig. 2 the NRCS plotted as a function of azimuth angle in a clockwise direction around each of the concentric circles in Fig. 1 starting from 0 as shown at the top of that figure. Because of the cyclonic nature of the hurricane, these circular transects span all wind directions with respect to the antenna look direction (toward the right side of the image in Fig. 1). To reduce the effect of speckle noise in the transects, the NRCS was averaged over an area of 1 km. Note that the azimuth directions relative to the antenna look direction in the abscissa of the plot in Fig. 1 correspond roughly to the counter-clockwise wind directions expected in a (northern-hemisphere) hurricane. Thus, the directions 90 and 270 are approximately cross wind looks, while 0 and 180 are up and down wind, respectively. Note also that the different radii of the plots in Fig. 1 were selected carefully in order to avoid areas in the SAR image which are significantly affected by non-wind induced artifacts, e.g., rain bands. Fig. 2 shows that for the plots with radii of 13 and 23 km, the NRCS is fairly constant around the entire circle. For these radii, the maximum change in NRCS over the entire range is less than 1 db or so. For the circle of 43 km radius, the wind direction dependence of the NRCS plot begins to become apparent. In particular, this curve shows a wind direction dependence more similar to the second harmonic behavior seen in standard model functions such as Cmod4 Stoffelen & Anderson (1997). The 63 km plot, on the other hand, shows asymmetric dependence on wind direction. This asymmetry is partially due to the fact that the difference in incident angle between the near- and far-range is becoming significant for this case. (As shown in the plot at the bottom of Fig. 1, the difference in incident angle between the direction values of 90 and 270 is 7 or so.) It is important to note from the Hurricane Ivan image of Fig. 1 that the backscatter NRCS in the high wind regime near the hurricane eye wall appears to

4 Figure 3. SAR wind map from the Cmod5 (left panel), and Cmod4 (center panel) inversion, and the differences Cmod5 - Cmod4 (right panel). a) c) b) 13 km 23 km 43 km 63 km 50 ms 45 ms 35 ms 25 ms 50 ms 45 ms 35 ms 25 ms Figure 4. NRCS values versus wind direction from the Cmod5 (left panel) and Cmod4 (center panel) for the constant wind speeds shown in the legend. For both cases, Eq. (2) with α = 0.6 was used for the polarization ratio. The right-hand panel shows the variation in incident angle with wind direction for each case. The curves in this figure complement the data for the concentric circles in Fig. 1 as shown by the plots in Fig. 2. be independent of wind direction. We show in the following section that this apparent loss of direction sensitivity, which is not well represented by the Cmod4 model function, can have a large impact on the retrieved wind speeds in this regime. 4. RETRIEVED WIND FIELDS FROM HURRICANE IVAN As discussed in Section 2.2 above, differences between the Cmod4 and Cmod5 C-band GMFs can be significant for hurricane force winds. We have just seen in Section 3 that the SAR NRCS measured over Hurricane Ivan may in fact be more characteristic of behavior of Cmod5 than of Cmod4. To further investigate this issue, we have inverted the NRCS from the Ivan SAR image of Fig. 1 to produce two wind maps based on the Cmod5 and Cmod4 model functions and shown in left and center panels of Fig. 3, respectively. For both inversions, we have used the LG-Method to determine the wind directions and the polarization ratio given by Eq. (2) with α = 0.6. From an examination of the first two panels in Fig. 3, differences between the two GMFs are immediately clear. Perhaps most obvious is the fact that Cmod5 (left-hand panel) yields significantly higher winds than does Cmod4 (center panel); 50 to 55 ms 1 versus 25 to 30 ms 1, respectively. As mentioned earlier, the maximum wind speed reported by the National Hurricane Center for Ivan at the overpass time was around 65 ms 1. We see from Fig. 3 that the predicted winds from Cmod5 are fairly close to this value, while those from Cmod4 are significantly lower. The right-hand panel of this figure shows the difference image (Cmod5 - Cmod4). Notice the large differences especially near the northeast hurricane eye wall where the maximum winds are expected. The small circular dots shown at the wind vectors on the Cmod5 and Cmod4 wind maps (left and center panels in Fig. 3) represent wind speeds predicted by an interactive objective kinematic analysis (IOKA) model for tropical cyclones (Cox et al. 1992). These dots are color coded using the same wind speed scale as the SAR map so that they become less visible in regions where the agreement between the model and the extracted SAR wind speed is good. One can see that the agreement is generally good for the Cmod5 map, especially in the higher wind region near the eye wall. For the Cmod4 wind map, the agreement is significantly worse in this region. It is also apparent from Fig. 3 that the angular dependence of the predicted wind speeds associated with Hurricane Ivan is closer to the expected behavior. In particular, the Cmod5 wind speeds near the hurricane eye are not only larger than those from Cmod4, but they are much more uniform around the full angular extent of the eye. To see this effect more clearly, we show in Fig. 4 plots of the predicted NRCS versus wind direction for Cmod5 (4a) and Cmod4 (4b) that would occur around the concentric circles in Fig. 1. These curves should be

5 compared with the corresponding curves in Fig. 2 extracted directly from the NRCS data from Hurricane Ivan shown in Fig. 1. Note that the curves computed from Cmod5 compare reasonably well with the corresponding curves in Fig. 1. The Cmod4 curves in Fig. 4b, on the other hand, show more sensitivity to wind direction at the higher wind speeds, and are generally rather different from those computed using Cmod5. We remind the reader that the asymmetry about 180 apparent in the NRCS predictions of both GMFs Fig. 4 is due to the fact that the incident angle varies around the concentric circles in Fig. 1. As mentioned already in Section 3, this variation of incident angle (as well as the fact that wind may not be constant) also affects the angular dependence of the circular transects in Fig. 2. The extent of this variation is shown in Fig. 4c. The results shown in Fig. 4 further confirm our conclusions based on the wind maps in Fig. 3 that the Cmod5 GMF provides a better representation of the Hurricane Ivan wind field. 5. RETRIEVED WIND FIELD FROM HURRICANE KATRINA Concerning the Hurricane Ivan investigations, a polarization ratio must be applied to account for the HH-polarized NRCS of RADARSAT. The EN- VISAT ASAR image recorded on August 28, 2005 was acquired with VV-polarization and therefore enables to apply the GMF Cmod5 directly to retrieve the wind speed. Fig. 5 shows the wind directions as retrieved from the LG-Method and the wind speeds resulting from the GMF Cmod5. Similar to the results found for Hurricane Ivan at HH-pol, maximum wind speeds of over 55 ms 1 were extracted near the eye of Katrina from the VV-pol ENVISAT ASAR data. The NRCS near the eye of both hurricanes was fairly constant with no significant dependence on the wind direction in this region. At larger distances from the hurricane eye more significant dependence on wind direction was observed. As already mentioned above, there remains uncertainty about the proper GMF for HH-pol wind retrievals. As part of a future effort to further analyze our data base of hurricane imagery, we plan to examine various models for the C-band polarization ratio (e.g. Mouche, et al) to see how they affect the extracted SAR wind vector, especially in the sensitive transition region in the vicinity of the hurricane eye wall. 6. CONCLUSION AND OUTLOOK We attempted to extract quantitative wind speed estimates under hurricane conditions from SAR imagery using the recently developed Cmod5 geophysical model function. At least for the two shown examples, we conclude that Cmod5 provides much better wind speed estimates for hurricane force winds than the more commonly used Cmod4. The major difference between these two GMFs is that for Cmod5, the NRCS (at a particular incident angle) becomes rather insensitive to the local wind vector (speed and direction) for winds in excess of 35 ms 1 or so. The high-wind dependence is quite different from that of Cmod4, and as we have attempted to demonstrate, can have a major impact on the extracted wind field for the extreme conditions such as those of Hurricane Ivan and Katrina. Although the Cmod5 GMF was developed empirically (Hersbach 2003), its loss of wind sensitivity at high wind speeds has a more fundamental basis. Recent papers also indicate that the aerodynamic roughness (i.e. the drag coefficient) over the ocean approaches a limiting value in high wind conditions (Powell et al. 2003; Donelan et al. 2004). This behavior is significantly different from its behavior at intermediate wind speeds where the drag coefficient increases with wind speed. Qualitatively, the wind speed dependence of the NRCS predicted by Cmod5 mirrors that observed for the drag coefficient. Much work remains to be done to further refine wind estimation of hurricane force winds. It is clear that more cases are needed to substantiate the findings discussed here. Also, the issue of the proper C-band GMF for HH-polarization, discussed in Section 2.2, must be resolved. The use of high-resolution SAR wind mapping under extreme wind conditions is just beginning to be recognized as a viable tool for hurricane tracking and prediction. Although much work remains, we believe that the potential payoff is well worth continued effort. ACKNOWLEDGMENTS Figure 5. High-resolution wind map of Hurricane Katrina as retrieved from the ENVISAT ASAR image acquired on 28. August 2005 at 15:50 UTC. The authors thank MDA s Geospatial Services for their collaboration with CSTARS in the Hurricane Watch program to receive RADARSAT data. The ENVISAT ASAR data were kindly provided

6 by the European Space Agency. The National Oceanographic Partnership Program (NOPP) provided high-resolution wind fields from numerical models. REFERENCES Cox A., Greenwood J., Cardone V., Swail V., 1992, An interactive objective kinematic analysis system, In: Fourth Int. Workshop on Wave Hindcasting. Dankert H., Horstmann J., Rosenthal W., 2003, Ocean Wind Fields Retrieved from Radar-Image Sequences, J. Geophys. Res., vol. 108(C11), no. C11, pp. 3352, doi: 10.1029/2003JC002056. Donelan M., Haus B.K., Reul N., et al., 2004, On the limiting aerodynamic roughness of the ocean in very strong winds, Geophys. Res. Lett, vol. 31, 18306,doi:10.1029/2004GL019460. Donnelly W., Carswell J., McIntosh R., Chang P., Wilkerson J., 1999, Revised ocean backscatter models at C and Ku band under high wind conditions, J. Geophys. Res., vol. 104, pp. 11 485 11 498. Drobinski P., Foster R., 2003, On the Origin of Near- Surface Streaks in the Neutrally-Stratified Planetary Boundary Layer, Bound.-Layer Meteorol., vol. 108, no. 2, pp. 247 256. Etling D., Brown R., 1993, Roll vortices in the planetary boundary layer: A review, Boundary-Layer Meteorol., vol. 18, pp. 215 248. Gerling T., 1986, Structure of the surface wind field from Seasat SAR, J. Geophys. Res., vol. 91, pp. 2 308 2 320. Hersbach H., 2003, CMOD5 an improved geophysical model function for ERS C-band scatterometry, Internal report, European Centre for Medium-Range Weather Forecast. Horstmann J., Koch W., 2005, Measurement of Ocean Surface Winds Using Synthetic Aperture Radars, J. Oceanic Eng., vol. 30, no. 3, pp. 508 515. Horstmann J., Koch W., Lehner S., Tonboe R., 2000, Wind retrieval over the ocean using synthetic aperture radar with C-band HH polarization, IEEE Trans. Geosci. Remote Sens., vol. 38, no. 5, pp. 2 122 2 131. Horstmann J., Koch W., Lehner S., Tonboe R., 2002, Ocean winds from RADARSAT ScanSAR, Can. J. Remote Sens., vol. 28, no. 3, pp. 524 533. Horstmann J., Schiller H., Schulz-Stellenfleth J., Lehner S., 2003, Global Wind Speed Retrieval from SAR, IEEE Trans. Geosci. Remote Sens., vol. 41, no. 10, pp. 2277 2286. Horstmann J., Thompson D., Monaldo F., Iris S., Graber H., 2005, Can synthetic aperture radars be used to estimate hurricane force winds?, Geophys. Res. Lett., vol. 32, no. L22801, pp. doi:10.1029/2005gl023992. Katsaros K., Vachon P., Lio W., Black P., 2002, Microwave Remote Sensing of Tropical Cyclones from Space, J. of Oceanography, vol. 58, pp. 137 151. Koch W., 2004, Directional analysis of SAR images aiming at wind direction, IEEE Trans. Geosci. Remote Sens., vol. 42, no. 4, pp. 702 710. Lehner S., Horstmann J., Koch W., Rosenthal W., 1998, Mesoscale wind measurements using recalibrated ERS SAR images, J. Geophys. Res., vol. 103, pp. 7 847 7 856. Monaldo F., Thompson D., Beal R., Pichel W., Clemente-Colon P., 2002, Comparison of SARderived wind speed with model predictions and ocean buoy measurements, IEEE Trans. Geosci. Remote Sens., vol. 39, no. 12, pp. 2587 2600. Monaldo F., Thompson D., Pichel W., Clemente- Colon P., 2004, A Systematic Comparison of QuikSCAT and SAR Ocean Surface Wind Speeds, IEEE Trans. Geosci. Remote Sens., vol. 42, no. 2, pp. 283 291. Mouche A., Hauser D., Daloze J.F., Gurin C., 2005, Dual Polarization Measurements at C-Band over the Ocean: Results from Airborne Radar Observations and Comparison with ENVISAT ASAR data, IEEE Trans. Geosci. Remote Sens., vol. 43, no. 4, pp. 10.1109/TGRS.2005.843951, 753 769. Powell M., Vickery P., Reinhold T., 2003, Reduced drag coefficient for high wind speeds in tropical cyclones, Nature, vol. 422, pp. doi: 10.1038/nature01481, 53 64. Stoffelen A., Anderson D., 1997, Scatterometer data interpretation: Estimation and validation of the transfer function CMOD4, J. Geophys. Res., vol. 102, pp. 5 767 5 780. Thompson D., Beal R., 2000, Mapping of mesoscale and submesocale wind fields using synthetic aperture radar, John Hopkins APL Tech. Dig., vol. 21, no. 1, pp. 58 67. Thompson D., Elfouhaily T., Chapron B., 1998, Polarization ratio for microwave backscattering from the ocean surface at low to moderate incidence angles, In: Proc. Int. Geosci. Remote Sens. Symp. 1998, Seattle, USA. Unal C., Snoeij P., Swart P., 1991, The polarizationdependent relation between radar backscatter from the ocean surface and surface wind vectors at frequencies between 1 and 18 GHz, IEEE Trans. Geosci. Remote Sens., vol. 29, pp. 621 626. Vachon P., Dobson F., 2000, Wind retrieval from RADARSAT SAR images: Selection of a suitable C-band HH polarization wind retrieval model, Can. J. Remote Sens., vol. 26, no. 4, pp. 306 313. Vachon P.W., Dobson F., 1996, Validation of wind vector retrieval from ERS SAR images over the ocean, Global Atmos. Ocean Syst., vol. 5, pp. 177 187. Wackerman C., Pichel W., Clemente-Colón P., 2004, A projection method for automatic estimation of wind vectors with RADARSAT SAR imagery, ESA SP-Series, vol. SP-565, pp. 55 60.