08.02.2015 Views

Th`ese Marouan BOUALI - Sites personnels de TELECOM ParisTech

Th`ese Marouan BOUALI - Sites personnels de TELECOM ParisTech

Th`ese Marouan BOUALI - Sites personnels de TELECOM ParisTech

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

École Doctorale<br />

d’Informatique,<br />

Télécommunications<br />

et Électronique <strong>de</strong> Paris<br />

Thèse<br />

présentée pour obtenir le gra<strong>de</strong> <strong>de</strong> docteur<br />

<strong>de</strong> l’Ecole Nationale Supérieure <strong>de</strong>s Télécommunications<br />

Spécialité : Signal et Images<br />

<strong>Marouan</strong> <strong>BOUALI</strong><br />

Destriping data from multi<strong>de</strong>tector imaging<br />

spectrometers : a study on the MODIS<br />

instrument<br />

Soutenue le 36 avril 2040 <strong>de</strong>vant le jury composé <strong>de</strong><br />

Bidule<br />

Truc Muche<br />

Machin<br />

Chose<br />

Tartampion<br />

Patrice Henry<br />

Nozha Boujemaa<br />

Saïd Ladjal<br />

Prési<strong>de</strong>nt<br />

Rapporteurs<br />

Examinateurs<br />

Directeurs <strong>de</strong> thèse


3<br />

« Rajouter une citation ici. »<br />

Auteur – Oeuvre


5<br />

Résumé<br />

Nou abordons dans cette thèse le problème <strong>de</strong> ...<br />

Abstract<br />

In this thesis, we tackle the issue of ...


7<br />

Table <strong>de</strong>s matières<br />

1 Introduction 11<br />

2 Remote Sensing with MODIS 15<br />

2.1 The evolution of remote sensing . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.2 Applications of satellite imagery . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.2.1 Monitoring land changes . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.2.2 Studying the atmosphere dynamics . . . . . . . . . . . . . . . . . . . 17<br />

2.2.3 Cryosphere and climate change . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2.4 Recent exemples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2.4.1 The Eyjafjallajökull . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.2.4.2 Deepwater Horizon oil spill . . . . . . . . . . . . . . . . . . 20<br />

2.3 The importance of oceans in the Earth system . . . . . . . . . . . . . . . . 21<br />

2.3.1 The ocean’s carbon cyle . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.3.2 Phytoplancton and ocean color . . . . . . . . . . . . . . . . . . . . . 22<br />

2.4 Constraints in remote sensing . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.4.1 Sensor Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.4.2 Atmospheric correction . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.4.3 Sun glint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.4.4 Cloud coverage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.5 The Mo<strong>de</strong>rate Resolution Imaging Spectroradiometer (MODIS) . . . . . . . 28<br />

2.5.1 Context and objectives . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.5.2 Technical specifications . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.5.3 Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.5.4 MODIS products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.5.5 Stripe noise on MODIS . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3 Standard <strong>de</strong>striping techniques and application to MODIS 39<br />

3.1 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.2 Moment Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

3.3 Histogram Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.4 Overlapping Field-of-View Method . . . . . . . . . . . . . . . . . . . . . . . 48


8 TABLE DES MATIÈRES<br />

3.5 Frequency filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.5.1 Spectral Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.5.2 Band-pass filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.6 Haralick Facet filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.7 Multiresolution approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.7.1 Limitations of fourier transform . . . . . . . . . . . . . . . . . . . . . 57<br />

3.7.2 Multiresolution analysis . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.7.3 Wavelet basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.7.4 Filter banks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.7.5 2D wavelet basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.7.6 Destriping with wavelet coefficient thresholding . . . . . . . . . . . . 62<br />

3.8 Assessing <strong>de</strong>striping quality . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.8.1 Noise Reduction Ratio and Image Distortion . . . . . . . . . . . . . 65<br />

3.8.2 Radiometric Improvement Factors . . . . . . . . . . . . . . . . . . . 66<br />

3.8.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4 A Variational approach for the <strong>de</strong>striping issue 73<br />

4.1 PDEs and variational methods in image processing . . . . . . . . . . . . . . 73<br />

4.2 Rudin, Osher and Fatemi Mo<strong>de</strong>l . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.3 Striping as a texture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

4.3.1 Yves Meyer’s mo<strong>de</strong>l for oscillatory functions . . . . . . . . . . . . . . 83<br />

4.3.2 Vese-Osher’s Mo<strong>de</strong>l . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

4.3.3 Osher-Solé-Vese’s Mo<strong>de</strong>l . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

4.3.4 Other u + v mo<strong>de</strong>ls . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.3.5 Experimental results and discussion . . . . . . . . . . . . . . . . . . 88<br />

4.4 Destriping via gradient field integration . . . . . . . . . . . . . . . . . . . . 91<br />

4.5 A unidirectional variational <strong>de</strong>striping mo<strong>de</strong>l . . . . . . . . . . . . . . . . . 94<br />

4.6 Optimal regularization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

4.6.1 Tadmor-Nezzar-Vese (TNV) hierarchical <strong>de</strong>composition . . . . . . . 99<br />

4.6.2 Osher et al. iterative regularization method . . . . . . . . . . . . . . 100<br />

4.6.3 Stopping criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

4.6.4 Experimental results and discussion . . . . . . . . . . . . . . . . . . 103<br />

5 Application : Restoration of Aqua MODIS Band 6 111<br />

5.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

5.2 Existing restoration techniques . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.2.1 Global interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.2.2 Local interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

5.3 Proposed approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

5.3.1 Spectral similarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

5.3.1.1 Spectral Correlation Measure . . . . . . . . . . . . . . . . . 115<br />

5.3.1.2 Spectral Angle Measure . . . . . . . . . . . . . . . . . . . . 116


9<br />

5.3.1.3 Euclidian Distance Measure . . . . . . . . . . . . . . . . . . 116<br />

5.3.1.4 Spectral Information Divergence Measure . . . . . . . . . . 116<br />

5.3.2 Spectral inpainting . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

5.4 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

5.4.1 Validation of spectral inpainting using NDSI measurements . . . . . 119<br />

5.4.2 Assessing the impact of stripe noise . . . . . . . . . . . . . . . . . . 122<br />

Conclusion 126<br />

Bibliographie 135


10 TABLE DES MATIÈRES


11<br />

Chapitre 1<br />

Introduction<br />

Over the years, climate change and its impact on life have become an alarming question<br />

for science to answer. Historical weather records show evi<strong>de</strong>nce of cyclic brutal climate<br />

fluctuations occuring since the formation of Earth and due to volcanic eruptions, meteorite<br />

impacts, solar activity, plate tectonic movements and <strong>de</strong>viations in the Earth rotational<br />

axis. Today, human industrial activities constitues an extra variable that interfers with<br />

land, ocean and atmosphere natural processes.<br />

To un<strong>de</strong>rstand, quantifie and mo<strong>de</strong>l the dynamic interactions between the earth system<br />

components, scientists have to relie on information available over the entire globe with a<br />

satisfactory time frequency. These requirements can only be satisfied by the use of satellite<br />

sensors orbiting the earth at a high altitu<strong>de</strong>s. Multiyear to multi<strong>de</strong>cadal data sets collected<br />

in the ultraviolet, visible, infrared and microwave portions of the energy spectrum, provi<strong>de</strong><br />

valuable information for researchers to interconnect the Earth’s main geophysical variables.<br />

Nevertheless, the complexity of physical processes involved in global change imposes<br />

data quality standards that satellite instruments fail to meet, <strong>de</strong>spite the continuous technological<br />

innovations in imaging sensors <strong>de</strong>sign. In<strong>de</strong>ed, a given geophysical variable (chlorophyll<br />

concentration for exemple) is <strong>de</strong>termined from raw satellite measurements subsequently<br />

used as input for geolocation, radiometric calibration, atmospheric correction, sun<br />

glint correction, bio-optical mo<strong>de</strong>ls, vicarious calibration, temporal bining...and the list is<br />

certainly not exaustive. This long processing chain makes the estimation of any geophysical<br />

parameter extremely sensitive to initial quantity acquired by satellite instruments.<br />

In simple words, small errors in the instrument response will irreversibly compromise the<br />

accuracy of high level products.<br />

Given the crucial role of oceans in global climate change and the stringent requirements<br />

associated with ocean color remote sensing, space agencies have given high priority<br />

to instrument <strong>de</strong>sign and pre-launch/on-orbit calibration procedures. Despite this effort,<br />

many issues related to image quality can still be observed as visual artifacts and require<br />

further processing. These typically inclu<strong>de</strong> straylight, optical leaks, gaussian noise, speckle<br />

noise (for active sensors), blurring, <strong>de</strong>ad pixels, line dropouts..


12 1. Introduction<br />

A common issue for Earth observing instruments, known as striping effect have persisted<br />

for more then 30 years, i.e, since the launch of Landsat 1. It was since reported on<br />

many imaging spectrometers :<br />

- Landsat Multi Spectral Scanner (MSS) and Thematic Mapper (TM)<br />

- Geostationary Operational Environmental Satellites (GOES)<br />

- Advanced Very High Resolution Radiometer (AVHRR)<br />

- Mo<strong>de</strong>rate Resolution Imaging Spectrometer (MODIS)<br />

- Compact High Resolution Imaging Spectrometer (CHRIS)<br />

- HYPERION<br />

- MEdium Resolution Imaging Spectrometer (MERIS)<br />

- Compact Reconnaissance Imaging Spectrometer for Mars (CRISM)<br />

- GLobal Imager (GLI)<br />

- Advanced Land Imager (ALI)<br />

The previous list inlu<strong>de</strong>s both whiskbroom and pushbroom sensors which are the two<br />

main techniques used for Earth observation.<br />

Whiskbroom instruments, also known as cross-track scanners acquire a series of lines in<br />

the direction perpendicular to the satellite orbital motion. A scan sweep from one si<strong>de</strong> of<br />

the swath to the other is ensured by a continuously rotating mirror. A limited set of <strong>de</strong>tectors,<br />

sensitive to spectific wavelenghts, then captures the radiation emitted or reflected<br />

by the Earth before its convertion into digital numbers.<br />

Pushbroom sensors (along-track scanners) also exploit the orbital motion of the platform<br />

to generate the second dimension of the acquired signal. They differ from the whiskbroom<br />

<strong>de</strong>sign in that the rotating mirror is replaced by a linear array of numerous <strong>de</strong>tectors that<br />

capture simultaneously each pixel of a single scan line.<br />

In both acquisition principles, images are formed by enterlacing or concatanating scan<br />

lines, acquired separately by different <strong>de</strong>tectors. Consequently, the imperfect calibration<br />

of individual <strong>de</strong>tectors induces a sharp pattern across or along the scanning direction that<br />

compromises both visual interpretation and quantitative analysis. More specifically, striping<br />

is disturbingly visible over low-radiance homogeneous regions, which often coinci<strong>de</strong><br />

with oceanographic areas. In fact, the impact of stripe noise on satellite <strong>de</strong>rived geophysical<br />

variables, including ocean color products, is increasingly attracting the interest of<br />

remote sensing research groups.<br />

Although an extensive litterature has tackled the striping issue, existing techniques are<br />

often not able to satisfie the requirements imposed by several remote sensing applications<br />

namely, complete stripe removal without signal distortion or additional post-processing<br />

artifacts.<br />

The goal of this thesis is to analyse the limitations of standard <strong>de</strong>striping methods<br />

and to explore the issue of stripe noise removal using variational mo<strong>de</strong>ls. We illustrate the


13<br />

results of our work using data from NASA’s MODIS instrument.<br />

Our focus on the MODIS sensor is justified by 1) the complexity and amplitu<strong>de</strong> of<br />

stripe noise visible on its emissive bands and 2) the importance of MODIS data in variety<br />

of remote sensing disciplines<br />

This thesis is organised as follows :<br />

Chapter 1 briefly exposes main aspects and constraints of ocean color remote sensing.<br />

This chapter also <strong>de</strong>scribes the Mo<strong>de</strong>rate Resolution Imaging Spectrometer (MODIS) and<br />

characterises the striping effect on MODIS products.<br />

Chapter 2 constitutes a non exaustive state-of-the-art of the <strong>de</strong>striping litterature. Several<br />

approaches are <strong>de</strong>scribed and illustrated on MODIS data. Experimental results and<br />

comparative study are presented at the end of the chapter. The limitations of existing<br />

techniques are analysed and used as a basis to <strong>de</strong>fine the requirements of an optimal<br />

<strong>de</strong>striping.<br />

Chapter 3 is the main contribution of this work. After an overview of variational methods<br />

for image processing applications, we explore Rudin, Osher and Fatemi total variation<br />

mo<strong>de</strong>l. Yves Meyers variational mo<strong>de</strong>ls for oscillatory paterns, are used in an attempt to<br />

extract stripe noise as a texture component. Our exploration of the striping issue from a<br />

variational angle eventually comes down to a unidirectional variation based regularizing<br />

mo<strong>de</strong>l able to efficiently remove all the stripe noise without signal distortion.<br />

Chapter 4 present an application that beneficits from the <strong>de</strong>veloped <strong>de</strong>striping technique.<br />

We propose a new approache to restore Aqua MODIS band 6 based on non-local filters<br />

and spectral similarity.


14 1. Introduction


15<br />

Chapitre 2<br />

Remote Sensing with MODIS<br />

2.1 The evolution of remote sensing<br />

Look far and wi<strong>de</strong>... This could well have been one of the un<strong>de</strong>rlying rules used by<br />

evolution to produce life as we can see it all over the Earth. From the multiple eyes of<br />

arachnophibia species to the mobile stereoscopic eyes of the chameleon, nature offers a<br />

variety of elegant examples. The very origin of hominid bipedism have been investigated<br />

by many evolutionnary scientists and attributed to a survival need. The ability to stand<br />

on two feet in the hostile environment of prehistory, enabled early Homo-Sapiens to look<br />

beyond savanna’s high grasses and thus avoid any uncomfortable encounters.<br />

Jumping back to comtemporary times, the numerous satellites orbiting the earth can<br />

be also perceived as the result of a long-term survival instinct. Remote sensing aims at<br />

measuring information and <strong>de</strong>termining specific properties of an object or phenomenon<br />

without direct physical contact. Today, this term commonly refers to the set of techniques<br />

used to <strong>de</strong>rive geophysical variables of the Earth (or other planets, stars, galaxies...).<br />

The <strong>de</strong>velopment of remote sensing is a direct consequence of the technological advances<br />

ma<strong>de</strong> in aviation from the begining of twentieth century. The first and second world war<br />

triggered an interest in aviation strong enough to attract heavy investments from most of<br />

the countries involved in war. Planes were not only used as a fast transportation vehicule,<br />

but were also consi<strong>de</strong>red a strategic advantage for their ability to monitor simultaneously<br />

wi<strong>de</strong> spatial areas. The invasion of Normandie is a good illustration of how planes were<br />

put to contribution. Aerial photographes were taken and analysed to establish appropriate<br />

landing surfaces and <strong>de</strong>termine the <strong>de</strong>pth and state of coastal waters. More significantly,<br />

infrared filters were already inclu<strong>de</strong>d in the primitive imaging <strong>de</strong>vices to distinguish natural<br />

vegetation from enemy camouflages. The following steps towards mo<strong>de</strong>rn remote sensing<br />

were achieved as consecutive technological intimidation attempts between USA and USSR<br />

during the cold war, leading to a brutal race to space ; Spoutnik and Explorer, quickly<br />

followed by the creation of NASA, the first men on the moon and the first habited spatial<br />

station (Saliout-1).


16 2. Remote Sensing with MODIS<br />

In 1970, inspired by spatial photographs taken during Mercury and Gemini’s missions,<br />

and motivated by the possibility to monitor agriculture crop yelds, NASA launched<br />

Landsat 1, the first satellite aiming at observing the earth surface and providing data<br />

for scdientists from different disciplines. Following the success of the Landsat series ( 7<br />

sensors since 1972), today’s intense research in remote sensing technologies is motivated<br />

by <strong>de</strong>eper questions. Climate change and its potential disastrous impact on human life<br />

urges an improved un<strong>de</strong>rstanding of earth sciences, ranging from carbon cycle mo<strong>de</strong>lling<br />

to weather forecasting. Such achievements are stricly limited by the quality of data collected<br />

by satellite /aerial sensors and require a continous increase of spatial, temporal and<br />

spectral resolution.<br />

2.2 Applications of satellite imagery<br />

Satellite and airborne sensors provi<strong>de</strong> a continuous flow of data exploited in fields as<br />

diverse as telecommunications, agriculture, <strong>de</strong>forestation, geology, hydrology, oceanography,<br />

urban management and fire monitoring to cite a few. For its abilities to capture wi<strong>de</strong><br />

portions of the globe, satellite <strong>de</strong>rived data offers a high potential for the surveillance of<br />

both local and global changes occuring in the Earth main systems.<br />

2.2.1 Monitoring land changes<br />

Among many human-induced activites, <strong>de</strong>forestation is highly responsible for both<br />

weather and climate change. Forests play a crucial role in the climate stability as they<br />

absorb and trap large quantities of carbon dioxi<strong>de</strong> (CO2) present in the atmopshere. The<br />

massive <strong>de</strong>struction of forests for urbanization and agricultural <strong>de</strong>velopment, inevitably<br />

releases in the atmopshere all the CO2 and other greenhouse gazes stored in trees for<br />

<strong>de</strong>ca<strong>de</strong>s. Assessing the amount and rate of carbon emissions from <strong>de</strong>forestation or other<br />

activities, remotely sensed data can be used to monitor primary variables of the Earth<br />

land surface such as the spectral reflectance, albedo and temperature. In addition, higher<br />

or<strong>de</strong>r variables in the form of vegetation indices can provi<strong>de</strong> valuable information on land<br />

change. A well-known variable is the Normalized Difference Vegetation In<strong>de</strong>x (NDVI)<br />

[Rouse et al., 1973], <strong>de</strong>fined as the ratio between the difference and the sum of reflectances<br />

in the visible (red) and near-infrared spectral bands :<br />

NDVI = ρ NIR − ρ Red<br />

ρ NIR + ρ Red<br />

(2.1)<br />

A major disadvantage of he NDVI in<strong>de</strong>x is that it tends to saturate over <strong>de</strong>nse vegetation<br />

and is highly <strong>de</strong>pen<strong>de</strong>nt on the background soil composition. In the presence of high<br />

biomass, the Enhanced Vegetation In<strong>de</strong>x (EVI) [Huete, 2002] provi<strong>de</strong>s an optimized vegetation<br />

signal able to capture high variability. and minimizes the influence of atmospheric


17<br />

Figure 2.1 – Deforestation in Brazil, in the region of Mato Grosso, monitored using the<br />

Normalized Difference Vegetation In<strong>de</strong>x (NDVI) (TL) RGB image from Terra MODIS<br />

captured on june 17, 2002 (TR) NDVI in<strong>de</strong>x (BL) RGB image from Terra MODIS<br />

captured in the same region on june 28 2006 (BR) NDVI in<strong>de</strong>x<br />

effects. The EVI is <strong>de</strong>fined as :<br />

EVI =<br />

2.5ρ NIR − ρ Red<br />

L + ρ NIR + C 1 ρ Red − C 2 ρ Blue<br />

(2.2)<br />

where the coefficient L corrects for canopy background, while C 1 and C 2 reduce the effects<br />

of aerosol. The values selected for the MODIS-EVI product are L = 1, C1 = 6 and<br />

C 2 =7.5.<br />

2.2.2 Studying the atmosphere dynamics<br />

The atmosphere <strong>de</strong>termines the amount of solar radiance actually reaching the Earth<br />

surface and therefore acts as a regulator of local and global climate processes. Depending


18 2. Remote Sensing with MODIS<br />

Figure 2.2 – A dust storm over the Red Sea, monitored using the Normalized Difference<br />

Dust In<strong>de</strong>x (NDDI) (Left) RGB image acquired by Aqua MODIS on May 13, 2005<br />

(Right) NDDI*<br />

on the nature and concentration of its components, the atmosphere can scatter and absorb<br />

a given percentage of the energy coming from the sun. In addition, the solar radiation<br />

reaching the Earth’s surface and reflected back to the atmosphere can be re-scatter or reabsorbed<br />

by the atmposhere. The equilibrium between incoming and outgoing energy is<br />

highly <strong>de</strong>pen<strong>de</strong>nt on the atmospheric characteritics in terms of clouds, water vapor and<br />

aerosols content. Clouds for exemple, tend to cool the Earth during the day by reflecting<br />

away incoming solar energy. At night, the presence of clouds blocks the outgoing energy<br />

and consequently warms the Earth. In addition to cloud <strong>de</strong>tection, satellite-based imagery<br />

is vital to measure the type and concentration of aerosols present in the atmosphere. Aerosols<br />

are microscopic particules abundant in <strong>de</strong>sert dust, volcanic ashes, fire smoke, sea<br />

salt and pollution <strong>de</strong>rived gazes. Their impact on the Earth’s radiation budget through albedo<br />

variations results directly from the scattering and absorption of sun light. Indirectly,<br />

aerosols intervene in the formation of clouds as they provi<strong>de</strong> a surface where water vapor<br />

emanating from the Earth surface can con<strong>de</strong>nsate and form liquid droplets. Furthermore,<br />

the size of aerosol particules influences greatly the optical properties of clouds. Clouds<br />

containing high concentrations of aerosols emitted from industrial activies, are composed<br />

of smaller liquid water drops than those containing aerosols <strong>de</strong>rived from natural processes.<br />

The resulting high brightness of these clouds increases the portion of solar radiation reflected<br />

back to space and tend to reduce the Earth surface temperature.<br />

The study of the atmosphere through the analysis of multispectral/hyperspectral satellite<br />

imagery is an important step in the un<strong>de</strong>rstanding of ocean-land-atmosphere physical interactions.<br />

Furthermore, the atmospheric layer between the Earth surface and space acts<br />

as a filter and attenuates the signal measured by satellite intruments. Estimation of the<br />

atmosphere optical characteristics is necessary to compute efficient atmosperic correction<br />

and <strong>de</strong>rive accurately primary variables used in global climate mo<strong>de</strong>ls.


19<br />

Figure 2.3 – Monitoring snow cover with the Normalized Difference Snow In<strong>de</strong>x (NDSI)<br />

(Left) RGB image acquired by Terra MODIS on May 5, 2001 in Alaska (Right) NDSI,<br />

notice how the in<strong>de</strong>x erroneously indicates snow cover over the ocean<br />

2.2.3 Cryosphere and climate change<br />

The cryosphere <strong>de</strong>signates the Earth system where water can be found in solid form. It<br />

is composed of snow, lake and sea ice, glaciers, ice caps and frozen ground and represents<br />

during the winter seasons, 17 % of the Earth surface. For instance, the Greenland and<br />

Antarctic Ice sheets together are equivalent to 70 m of sea level rise. Given its mass and<br />

latent heat capacity, the cryosphere plays a role in climate change almost as important as<br />

the oceans. First, the high surface reflectivity of snow covered regions have a significant<br />

impact on the Earth albedoand even small variations in snow or ice fractional cover can lead<br />

to short-term climate changes. In addition, the quantity of fresh water stored on cryosphere<br />

can interfer with ocean currents. The variations in sea level, temperature and salinity of<br />

ocean waters due to seasonal snowmelt can modify the interactions between oceans and the<br />

atmosphere. Furthermore, global warming in the polar region induces permafrost melting<br />

which in turn releases large quantities of methane stored in frozen soils. Un<strong>de</strong>rstanding<br />

the interactions between cryosphere and other parts of the Earth system, requires short<br />

and long-term quantitative measurements provi<strong>de</strong>d by satellite instruments.<br />

2.2.4 Recent exemples<br />

The year 2010 was subject to few unfortunate hazards. The political and economical<br />

implications of these events have emphasized the crucial role played by satellite sensors in<br />

the management of such disasters.


20 2. Remote Sensing with MODIS<br />

Figure 2.4 – The ash plume resulting from the Eyjafjallajokull eruption, monitored by<br />

two sensors on April 17, 2010 (Left) Image acquired by NASA’s Aqua MODIS on April 17,<br />

2010 showing a diffuse cloud of volcanic ash and a column plume rising at higher altitu<strong>de</strong>s<br />

(Right) Image from CNES/NASA CALIOP, lidar instrument aboard CALIPSO <strong>de</strong>dicated<br />

to the study of the atmosphere vertical profile of aerosol.<br />

2.2.4.1 The Eyjafjallajökull<br />

Around 20 Mars 2010, Eyjafjallajökull, a volcano located in the south of Iceland began<br />

its first erupting phase. A more explosive phase started on 14 april 2010. The volcano being<br />

located beneath glacial ice, the melting water flowing back in the volcano vent resulted in<br />

the formation of silica particules and ash, explosively rejected in the atmosphere at heights<br />

reaching 10 km. In addition, the south-easterly path of the Jet Stream, unusually stable<br />

at that specific period, progressively dispersed the ash plume across Europe’s airspace. Although<br />

not specifically quantified by manufacturers, the sensitivity of aircraft engines to<br />

ash particules, forced the Fe<strong>de</strong>ral Aviation Administration (FAA) to shut down air-traffic<br />

in most european countries. The financial losses caused by airspace closure to the airline<br />

industry were estimated approximatively at 150 ME per day from 15 to 23 April. The<br />

quantity of rejected ash and the trajectory of the resulting plumes were eventually monitored<br />

with numerous satellite-based instruments including CALIPSO, VEGETATION<br />

and MODIS (figure 2.4). As the eruptive phase of the volcano persisted, satellite images<br />

of the ash cloud were used to <strong>de</strong>termine its trajectory and establish low-risk air corridors.<br />

2.2.4.2 Deepwater Horizon oil spill<br />

In April 20, 2010, an offshore drilling platform, Deepwater Horizon, located southeast<br />

of the Louisiana coast explo<strong>de</strong>d and sank in the ocean. The explosion was followed by the<br />

release of 8500 m 3 of cru<strong>de</strong> oil per day and is consi<strong>de</strong>red today as the largest oil spill in the<br />

history of petroleum industry. Data collected from several spatial instruments were used<br />

to analyse the extend of the oil spill, and evaluate its spreading evolution with respect<br />

to wind direction. The reflection of sun light off the ocean surface (sun glint) highlights


21<br />

Figure 2.5 – RGB images from MODIS showing the evolution of the oil spill after the<br />

Deepwater Horizon platform explosion in 2010 (TL) April 29 (TR) May 4 (BL) May 9<br />

(BR) In 24 May, the oil spill reach the shores of Blind Bay and Redfish Bay at the eastern<br />

edge of the Mississippi River <strong>de</strong>lta. Credit : NASA/MODIS Rapid Response Team<br />

oil contaminated waters, which appear in satellite imagery as bright silvery stains. High<br />

concentrations of oil in the ocean surface tend to reduce the magnitu<strong>de</strong> of waves and the<br />

formation of whitecaps. This change in the ocean surface optical properties can increase<br />

the reflection of light in the sensor viewing direction (see section 2.4.3) specially if the<br />

oil-covered waters are located near the specular reflection.<br />

2.3 The importance of oceans in the Earth system<br />

2.3.1 The ocean’s carbon cyle<br />

Oceans occupie 71% of the earth and constitute a large reservoir of carbon present<br />

in both organic and non-organic form. Their impact on the regulation of our climatic<br />

system flows directly from the fundamental role they play in the global carbon cycle and<br />

justifies the importance given by space agencies to the study of physical and biological<br />

processes along the ocean-atmosphere interface. The carbon cycle <strong>de</strong>signates the process


22 2. Remote Sensing with MODIS<br />

Figure 2.6 – The ocean carbon cycle. Credit : NASA Robert Simmon<br />

of carbon exchange fluxes between the earth’s dynamic systems, namely biosphere, lithosphere,<br />

hydrosphere and atmosphere. Oceans are the earth’s largest active carbon reservoir<br />

and contain approximatively 4.10 19 gC. They hold a double role in the global carbon cycle<br />

refered to as solubility and biological pumps. The exchange of carbon dioxi<strong>de</strong> between<br />

ocean and atmosphere occurs in the air-sea interface and its flux F can be mo<strong>de</strong>lled as :<br />

(<br />

)<br />

F = K. PCO Ocean<br />

2<br />

− P Atmosphere<br />

CO 2<br />

(2.3)<br />

where PCO Ocean<br />

2<br />

(respectively P Atmosphere<br />

CO 2<br />

) is the partial pressure of CO 2 in the ocean (resp.<br />

atmosphere) and K a coefficient known as piston velocity. The term PCO Ocean<br />

2<br />

is <strong>de</strong>pen<strong>de</strong>nt<br />

on other thermodynamic parameters such as the water temperature and salinity, and thus<br />

dictates the directional exchange of CO 2 while its rate is related to K and to wind driven<br />

turbulence. The <strong>de</strong>positing of atmospheric CO 2 in the <strong>de</strong>ep layers f the ocean is ensured<br />

by a continuous mechanism known as solubility pump ; Once trapped in surface waters,<br />

the CO 2 is transfered to <strong>de</strong>eper ocean layers. This process occurs in high latitu<strong>de</strong>s areas,<br />

where overturning circulation is generated by the formation and sinking of <strong>de</strong>nse waters<br />

(figure 2.6).<br />

2.3.2 Phytoplancton and ocean color<br />

The intense biological activity witin the ocean and the resulting photosynthetic production<br />

(also known as primary production) constitutes the fastest exchange of CO 2 between<br />

oceans and atmosphere. In eutophic areas (upper layer of the oceans located between 0-<br />

200m), algal micro-organisms such as phytoplanton, absorb the energy <strong>de</strong>rived from solar


23<br />

radiance and other mineral compounds to convert CO 2 into organic matter in the form of<br />

particulate or dissolved organic carbon. This process is refered to as photosynthesis :<br />

nCO 2 + nH 2 O → (CH 2 O) n + nO 2 (2.4)<br />

Phytoplancton is a major regulator of the glocal carbon cycle. These microscopic algae<br />

are in fact at the base of the marine food chain and ensure the survival of most species<br />

including other micro-organisms such as zooplancton, responsible for the exportation of<br />

CO 2 to the <strong>de</strong>ep ocean. Phytoplancton organisms are composed of pigments with specific<br />

ligh absorption spectra. Chlorophyll-a for example, is an ubiquitous pigment that attributes<br />

the green color to most marine and continental plants. Despite their microscopic<br />

size, the concentration of phytoplancton in the water column <strong>de</strong>termines the ocean color.<br />

From oligotrophic to eutrophic waters, the increase of phytoplancton biomass and its<br />

primary production shifts the color of the ocean from <strong>de</strong>ep blue to green. Consequently,<br />

satellite-<strong>de</strong>rived measurements of the water-leaving radiance can be used to evaluate the<br />

phytoplancton biomass and quantify the ocean carbon fluxes. The spectral analysis of<br />

the radiance emanating from the oceans can be coupled to bio-optical mo<strong>de</strong>ls in or<strong>de</strong>r<br />

to retrived the ocean biomass constituents. Chlorophyll-a pigment concentration, a major<br />

geophysical variable in oceanography, can be expressed as :<br />

[Chla] = 10 (c 0+c 1 .ρ+c 2 .ρ 2 +c 3 .ρ 3 ) (2.5)<br />

where c 0 , c 1 , c 2 , c 3 are empirically-<strong>de</strong>rived constants and ρ = ρ 488 /ρ 551 is the ratio of<br />

remote sensing reflectances at wavelenght 488 nm and 551 nm [Aiken et al., 1995].<br />

Another parameter used to study the optical properties of ocean waters is the diffuse<br />

attenuation coefficient K. This coefficient <strong>de</strong>termines the attenuation of light intensity<br />

within the water column. At a wavelenght of 490 nm, K is a direct indicator of water<br />

turbidity and is <strong>de</strong>fined as :<br />

K(490) = 0.016 + 0.156<br />

where L w <strong>de</strong>signates the water leaving radiance.<br />

( )<br />

Lw (490) −1.54<br />

(2.6)<br />

L w (555)<br />

2.4 Constraints in remote sensing<br />

2.4.1 Sensor Calibration<br />

In or<strong>de</strong>r to provi<strong>de</strong> reliable information, the response of an instrument needs to be<br />

characteriszed precisely with respect to a lage data set of controlled input signals. This<br />

process known as calibration, occurs during the pre-launch and in-flight stages of the sensor<br />

mission as data quality highly <strong>de</strong>pends on it. The raw counts measured by a sensor<br />

can not be used directly for quantitative studies because they are not associated with


24 2. Remote Sensing with MODIS<br />

any unit. Radiometric calibration converts instrument digital numbers (DN) to physical<br />

radiance values (W.sr −1 .m −2 ). Since CZCS (Coastal Zone Color Scanner), most space<br />

instruments relie on on-board, in-flight internal calibration systems that inclu<strong>de</strong> a solar<br />

diffuser plate oriented towards the sun. The solar irradiance reaching the diffuser is then<br />

used as a radiance reference to adjust the sensor absolute radiometric calibration. Furthermore,<br />

pre-flight calibrations are not necessarely optimal in the actual space environment<br />

of the sensors and might require further adjustments. On many instruments, the continous<br />

exposure of the diffuser plate to solar radiations and bombardment by space particules,<br />

induces a progressive <strong>de</strong>gradation that needs to be accounted for in the overall system<br />

(see section ). This can be achieved using an additional diffuser to monitor the primary<br />

diffuser <strong>de</strong>gradation (MERIS) or by comparing lunar observations to the light scattered<br />

by the solar diffuser (MODIS and SeaWiFS).<br />

In addition to absolute and relative radiometric calibration, vicarious calibration techniques<br />

are used to minimize errors between satellite-<strong>de</strong>rived data and stable ground targets<br />

with known radiance values. Typical calibration sites inclu<strong>de</strong> <strong>de</strong>sertic regions, ice sheets,<br />

clouds and ocean targets contaminated with sun glint. Assuming the availability of an<br />

extensive in situ data set and a reliable atmospheric correction, the sensor response can<br />

be adjusted so that the values of a geophysical variable estimated from satellite radiances,<br />

Figure 2.7 – The ocean color from satellite imagery is <strong>de</strong>termined by the organic constituents<br />

of the ocean upper surface. These images acquired by MODIS illustrate phytoplancton<br />

blooms in the South Atlantic Ocean, off of the cost of Argentina. Phytoplancton<br />

blooms offer a wi<strong>de</strong> variety of colors ranging from turquoise blue to dark green. The strong<br />

color variations are due to the pigment composition of each phytoplancton specie and their<br />

<strong>de</strong>pth in the eutophic layer.


25<br />

coinci<strong>de</strong> with those <strong>de</strong>rived from ground measurements.<br />

2.4.2 Atmospheric correction<br />

Prior and after its reflection on the Earth surface, solar radiation is subject to molecular<br />

and aerosol scattering. The cumulation of these processes, combined with the high altitu<strong>de</strong><br />

of satellites, reduces the magnitu<strong>de</strong> of the radiance reaching the sensor. The received signal,<br />

known as Top-Of-Atmosphere (TOA) can be expressed as a sum of radiance contributions<br />

related to several processes. Atmospheric Correction then estimates the radiative portion<br />

of these effect prior to the generation of surface reflectances. Atmospheric correction is<br />

particularly complex in the case of ocean color remote sensing where the TOA radiance<br />

can be written in a simplified form as :<br />

L T OA = L Rayleigh + L Aerosol + L Rayleigh−Aerosol + TL Glint + tL W ater (2.7)<br />

In the previous equation,<br />

- T and t are the direct and diffuse atmospheric transmittances between the ocean surface<br />

and the instrument. The portion of energy that reaches the sensor is related to the optical<br />

thickness of aerosols contained in the atmopshere<br />

- L Glint is the specular reflection of sun light on the oceanic surface towards the sensor<br />

viewing direction<br />

- L Rayleigh and L Aerosol both translate scattering effects of light and <strong>de</strong>pend on atmospheric<br />

pressure, temperature and polarization.<br />

- L Rayleigh−Aerosol accounts for coupling processes between Rayleigh and Mie scaterring.<br />

- L W ater is the signal to be retrieved after atmopheric correction and corresponds to the<br />

water-leaving radiance<br />

The L W ater term represents a weak portion (less than 10%) of the TOA signal and its<br />

extraction from the combined surface/atmosphere system highly <strong>de</strong>pends on the accuracy<br />

of the atmopheric correction. In fact, a 1% error on the measured L T OA leads to a 10%<br />

error on the estimated L W ater . In addition to scattering effects, many gazes present in the<br />

atmosphere such as ozone, oxygen and water vapor can absord solar radiation in specific<br />

regions of the electromagnetic spectrum.<br />

The effective removal of atmospheric contribution from the TOA signal is a major requirement<br />

for the generation of operational land and ocean colour products.<br />

2.4.3 Sun glint<br />

The greatest obstacle for the generation of ocean color products is the presence of sun<br />

glint. Sun glint is the specular reflection of sunlight off the ocean surface and into the<br />

satellite sensor. This optical phenomena occurs when the ocean surface directs the solar<br />

radiation in the exact viewing direction of the sensor and as such, <strong>de</strong>pends on the sea<br />

surface state, the sun position and the satellite viewing geometry. The high intensity of


26 2. Remote Sensing with MODIS<br />

Figure 2.8 – Sun glint, a major obstacle for the generation of ocean color products appears<br />

on MODIS images as a wi<strong>de</strong> and bright vertical stripe. It results from the reflection of sun<br />

light off the ocean surface.<br />

sun glint radiance L Glint (often close to the sensor saturation) compromises the estimation<br />

of water-leaving radiance and all the <strong>de</strong>rived oceanic geophysical variables. In addition,<br />

atmospheric correction over ocean targets often fails to distinguish sun glint from high<br />

concentrations of white aerosols. Many techniques have been <strong>de</strong>vised to remove the sun<br />

glint contribution from the TOA signal. The most common approach for medium resolution<br />

instruments relie on the well-known Cox and Munk statistical mo<strong>de</strong>l of sea surface<br />

roughness [Cox and W.Munk, 1954]. For a given sensor viewing geometry, the amount of<br />

sun glint radiance can be expressed as a function of the probability <strong>de</strong>nsity function of sea<br />

surface slopes which in turn, <strong>de</strong>pend on the wind speed and direction. The reflectance due<br />

to sun glint can be predicted from the sensor viewing geometry and the wind speed with :<br />

ρ Glint (λ, θ s ,θ v ,φ s ,φ v ,W)= P (θ s,θ v ,φ s ,φ v ,W)f(w, λ)<br />

4cos 4 βcosθ v cosθ s<br />

(2.8)<br />

where :<br />

- θ s and φ s are the zenith and azimuth angles of the sun<br />

- θ v and φ v are the zenith and azimuth angles of the sensor<br />

- f(w, λ) is the Fresnel reflectance at the ocean surface for an angle of inci<strong>de</strong>nce of w<br />

- W is the wind speed<br />

- P (θ s ,θ v ,φ s ,φ v ,W) is the probability distribution function of Cox/Munk mo<strong>de</strong>l corresponding<br />

to a 4 th or<strong>de</strong>r Gramm-Charlier expansion and often approximated with a Gaussian<br />

distribution.<br />

Although implemented in numerous ocean colour sensors, Cox and Munk based correcting<br />

schemes [Montagner et al., 2003], [Wang and Bailey, 2001], [Fukushima et al., 2007], [Ottaviani<br />

et al., 2008] can only process mo<strong>de</strong>rate sun glint. Also, the results are limited by<br />

the accuracy and resolution of available wind data. Motivated by these limitations, Steinmetz,<br />

Deschamps and Ramon, [Steinmetz et al., 2008] recenly introduced a new approach,


27<br />

POLYMER based on neural networks. The TOA reflectance is mo<strong>de</strong>lled as :<br />

ρ T OA (λ) =c 0 + c 1 λ −1 + c 2 λ −4 + tρ W ater (λ) (2.9)<br />

where c 0 inclu<strong>de</strong>s spectrally flat components such as sun glint, c 1 λ −1 accounts for aerosol<br />

effects and c 2 λ −4 represents coupling processes between sun glint and aerosols. The water<br />

reflectance ρ W ater is assumed to be a function of chlorophyll concentration, <strong>de</strong>rived from<br />

bio-optical mo<strong>de</strong>ls. Rather different techniques are employed to process sun glint effects<br />

on high resolution imagery (1-20m)[Hochberg et al., 2003], [Hedley et al., 2005], [Lyzenga<br />

et al., 2006], [Kutser et al., 2009]. They are mostly based on the assumption of neglectable<br />

water-leaving radiance at Near-Infrared (NIR) bands and the linear relationship of reflectances<br />

at visible and NIR bands. An aternative option for sun glint correction is sun glint<br />

avoidance. Several instruments are <strong>de</strong>signed with the ability to tilt the sensors viewing<br />

direction (SeaWiFS) or to inclu<strong>de</strong> a multi-angle viewing mo<strong>de</strong> (POLDER). The presence<br />

of such mechanisms can reduce consi<strong>de</strong>rably the loss of data related to strong sun glint<br />

contamination and increase the spatial coverage of ocean products.<br />

2.4.4 Cloud coverage<br />

Persisting cloud coverage in satelite imagery disturbs the study of many geological and<br />

oceanographic processes because it reduces the availability of exploitable data. More specifically,<br />

it disables the <strong>de</strong>tection of changes occuring at temporal frequencies higher than<br />

the persistence of clouds and, therefore poses a serious challenge for applications related to<br />

disaster management (see section 2.2.4.2). As a reponse to the limited set of measurements<br />

available in high level geophysical variables, many studies have focused on possible techniques<br />

to enhance the daily spatial coverage. A common approach in ocean color remote<br />

sensing is to exploit data measured by many sensors at slightly different times. Un<strong>de</strong>r the<br />

assumption of stationnay oceanic processes, measurements acquired within a reasonable<br />

time frame can be merged to fill in the gaps related to clouds. In the context of ocean color<br />

remote sensing, combination of data from multiple instruments was explored for SST in<br />

[Reynolds and Smith, 1994] and have since been <strong>de</strong>eply investigated by the SIMBIOS project<br />

of NASA [M. E. and Gregg, 2003], [Fargion and McClain], [Kwiatkowska and Fargion,<br />

2002]. Common techniques for the merging of chlorophyll concentration products inclu<strong>de</strong> :<br />

- Weighted averaging : estimation of missing values is <strong>de</strong>termined from different sensors<br />

using a weighted average where the weights are <strong>de</strong>pen<strong>de</strong>nt on the accuracy of observations.<br />

- Statistical objective analysis, also known as optimal interpolation, consists in filling<br />

gaps through spatial and temporal interpolation.<br />

More recently, the Short-term Prediction and Research Transition (SPoRT) program of<br />

NASA has <strong>de</strong>velopped a composite product for MODIS SST [Haines et al., 2007] to improve


28 2. Remote Sensing with MODIS<br />

Figure 2.9 – The issue of limited spatial coverage illustrated on Aqua MODIS Level 3<br />

nightime 11 µm Sea Surface Temperature (SST) (Left) Daily product with missing measurements<br />

due to clouds, sun glint and distance between swaths (Right) 8 day composite<br />

product showing a substantial increase in spatial coverage.<br />

regional weather prediction mo<strong>de</strong>ls. The methodology tackles the issue of cloud contamination<br />

using a temporal compositing approach, able to estimate SST missing values while<br />

ensuring global spatial continuity of SST gradients.<br />

2.5 The Mo<strong>de</strong>rate Resolution Imaging Spectroradiometer<br />

(MODIS)<br />

2.5.1 Context and objectives<br />

In 1978, the Coastal Zone Color Scanner (CZCS) was launched aboard the spacecraft<br />

Nimbus-7. Designed as a proof-of-concept instrument with a mission duration of one year,<br />

CZCS primary goal was to <strong>de</strong>termine the potential of satellite imagery for the quantification<br />

of ocean chlorophyll concentration and other dissoved and suspen<strong>de</strong>d organic<br />

matter. Despite many limitations related to its prelaunch characterisation and on-orbit<br />

calibration, CZCS imposed the basis of mo<strong>de</strong>rn ocean color satellite sensors. The CZCS<br />

operational period (1978-1986) was followed by the launch of several instruments <strong>de</strong>dicated<br />

to the study of oceans, MOS, OCTS, POLDER and SeaWiFS, each including specific<br />

technological innovations with increased dynamic range, signal-to-noise ratio and number<br />

of spectral bands. Motivated by the <strong>de</strong>manding requirements of remote sensing scientists,<br />

the improvement in satellite sensors <strong>de</strong>sign reached its climax with the Mo<strong>de</strong>rate Resolution<br />

Imaging Spectrometer (MODIS). As a multipurpose mission, MODIS primary goals<br />

are :<br />

- Ensure the continuity of data collection provi<strong>de</strong>d by heritage sensors such as AVHRR,<br />

CZCS, SeaWiFS and HIRS<br />

- Deliver products in a variety of disciplines with improved radiometric quality<br />

- Monitor changes that occur at short time-scales combining Terra and Aqua MODIS


29<br />

morning and afternoon observations<br />

- Provi<strong>de</strong> highly consistent time series of observations used to improve our un<strong>de</strong>rstanding<br />

of climate change at seasonal-to-<strong>de</strong>cadal time scales<br />

The first prototype of MODIS was launched on December 18, 1999 aboard the Terra<br />

EOS-AM-1 platform. The MODIS on the Aqua EOS-PM1 spacecraft was launched on<br />

May 4, 2002.<br />

2.5.2 Technical specifications<br />

MODIS was initially conceptualized as a double instrument MODIS-N (Nadir) and<br />

MODIS-T (Tilt). Aiming improved ocean colour capabilities, MODIS-T was <strong>de</strong>signed with<br />

the ability to tilt away from specular reflection directions and avoid sun glint effects.<br />

Despite its success on SeaWiFS, a tilting mechanism was not retained for the MODIS<br />

project because the combination of data collected from Terra and Aqua MODIS was proven<br />

to provi<strong>de</strong> almost similar spatial coverage [Gregg, 1992], [Gregg and Woodward, 2007],<br />

with the additional advantage of both morning and afternoon observations. MODIS is<br />

composed of 36 spectral bands ranging from the visible (0.4µm) to the far infrared (14µm)<br />

and centered at wavelenghts <strong>de</strong>dicated to three major applications. Bands 1-7 are <strong>de</strong>voted<br />

to the study of land remote sensing, cloud <strong>de</strong>tection and aerosol estimation. These bands<br />

are centered at wavelenghts similar to Landsat TM and measure data at spatial resolutions<br />

of 250 m for bands 1-2 and 500 m for bands 3-7. Stringent requirements associated with<br />

ocean color monitoring and studies conducted on CZCS and SeaWiFS instruments lead to<br />

nine spectral bands on MODIS (8-16). Compared to SeaWiFS, MODIS ocean color bands<br />

are narrower (average of 10 nm width compared to 20 nm on SeaWiFS), and, therefore<br />

allow more reliable atmospheric correction with higher signal-to-noise ratio values. Most<br />

of the remaining bands 17-26 were spectrally positioned with respect to HIRS, AVHRR<br />

and ATRS. To provi<strong>de</strong> accurate Sea Surface Temperatures (SST), two split-windows at<br />

mid-wave infrared (MWIR) (bands 23-24) and long-wave infrared (LWIR) (bands 31 and<br />

32) were inclu<strong>de</strong>d. The split-window composed of bands 31-32 enables the <strong>de</strong>rivation of<br />

day time SST measurements, because channels 23 and 24 are contaminated with sun glint<br />

effects, still persisting in the MWIR portion of the electromagnetic spectrum. MODIS<br />

was <strong>de</strong>signed as a whiskbroom sensor ; it uses the obital motion of the satellite to acquire<br />

successive lines using a scanning mirror that rotates at ± 55˚. MODIS swath reaches 2330<br />

km and allows a global coverage of the entire earth every one to two days.<br />

2.5.3 Components<br />

Compared to its pre<strong>de</strong>cessors, MODIS inclu<strong>de</strong>s many components (figure 2.10), each<br />

playing a specific role in the acquisition process. As we shall see, emphasis was given to the<br />

calibration of the instrument. We present here a brief <strong>de</strong>scription of the main subsytems.


30 2. Remote Sensing with MODIS<br />

Main application Band Spectral Center Spectral Width<br />

Land/Cloud/Aerosols Boundaries<br />

1 645 nm 25 nm<br />

2 856 nm 15 nm<br />

3 469 nm 20 nm<br />

4 555 nm 20 nm<br />

Land/Cloud/Aerosols Properties 5 1240 nm 20 nm<br />

6 1640 nm 24 nm<br />

7 2130 nm 50 nm<br />

8 412 nm 15 nm<br />

9 443 nm 10 nm<br />

10 493 nm 10 nm<br />

11 531 nm 10 nm<br />

Ocean Color/Phytoplankton 12 551 nm 10 nm<br />

13 667 nm 10 nm<br />

14 678 nm 10 nm<br />

15 748 nm 10 nm<br />

16 869 nm 15 nm<br />

17 905 nm 30 nm<br />

Atmospheric Water Vapor 18 936 nm 10 nm<br />

19 940 nm 50 nm<br />

20 3.750 µm 0.180 µm<br />

Surface/Cloud Temperature<br />

21 3.959 µm 0.030 µm<br />

22 3.959 µm 0.030 µm<br />

23 4.050 µm 0.060 µm<br />

Atmospheric Temperature<br />

24 4.465 µm 0.065 µm<br />

25 4.515 µm 0.067 µm<br />

26 1.375 µm 0.030 µm<br />

Cirrus Clouds Water Vapor 27 6.715 µm 0.360 µm<br />

Cloud Properties 29 8.550 µm 0.300 µm<br />

Ozone 30 9.730 µm 0.300 µm<br />

Surface/Cloud Temperature<br />

31 11.030 µm 0.500 µm<br />

32 12.020 µm 0.500 µm<br />

33 13.335 µm 0.300 µm<br />

Cloud Top Altitu<strong>de</strong><br />

34 13.635 µm 0.300 µm<br />

35 13.935 µm 0.300 µm<br />

36 14.235 µm 0.300 µm<br />

Table 2.1 – MODIS spetral bands<br />

More <strong>de</strong>tails are provi<strong>de</strong>d in (Barnes et al., 1998).


31<br />

Figure 2.10 – Main Components of the MODIS instrument<br />

Scan mirror assembly The energy from Earth radiation is reflected into the focal<br />

plane assembly via a double sid<strong>de</strong>d scan mirror, composed of nickel-plated Beryllium. The<br />

rotation of the mirror at 20.3 rpm is ensured by a 2-phase brushless DC motor and the<br />

stability of its speed is monitored by a 14-bit optical enco<strong>de</strong>r with an accuracy of 11-<br />

microradian . The system mirror-motor-enco<strong>de</strong>r has been limited to a weight of 4.3 kg<br />

and only requires a power of 2.9 W.<br />

Focal plane assembly Dichroic beamsplitters are used to separate the scene radiation<br />

into 4 Focal Plane Assemblies (FPA) associated with each spectral region (VIS, NIR,<br />

SWIR-MWIR, LWIR). Each FPA contains <strong>de</strong>tector arrays of pixels which size ranges<br />

from 135 to 540 µm. Bands with spatial resolution of 1 km have 10 <strong>de</strong>tectors array, while<br />

bands at 250m and 500m, have 20 and 40 <strong>de</strong>tectors. This is illustrated in figure 2.11.<br />

The FPAs are highly responsible for the presence of stripe noise in the images as striping<br />

originates partly from the miscalibration and non-linear responses of the pixels contained<br />

in the <strong>de</strong>tector arrays.<br />

The onboard calibration system inclu<strong>de</strong>s four complementary components :<br />

Solar Diffuser The solar diffuser (SD) is a pressed plate located in the forward part<br />

of the instrument and used for the calibration of the reflective bands. Periodically (once<br />

per orbit, at the north and south pole to avoid loss of data), the SD provi<strong>de</strong>s diffuse<br />

solar illumination to the scan mirror. Assuming a good characterization of the SD surface<br />

reflectance and precise estimation of the sun position, the solar diffuser radiance provi<strong>de</strong>s<br />

a stable source for absolute radiometric calibration.


32 2. Remote Sensing with MODIS<br />

Figure 2.11 – Four Focal Plane Assemblies composed of <strong>de</strong>tector arrays<br />

Solar Diffuser Stability Monitor Due to exten<strong>de</strong>d exposure to sun radiation during<br />

the mission, the solar diffuser can be subject to a progressive <strong>de</strong>terioration that translates<br />

as slight variations in its bidirectional reflectance distribution function (BRDF). Such<br />

<strong>de</strong>gradation can impact the radiometric calibration of the sensor and is monitored in<strong>de</strong>pen<strong>de</strong>ntly<br />

with the solar diffuser stability monitor (SDSM). Deviations of the SD response<br />

can be <strong>de</strong>duced by comparing measurements from its solar-illumated surface with measurements<br />

obtained directly from the sun. This is done with a spherical integrating source<br />

(SIS) composed of nine filtered <strong>de</strong>tectors that succcessively view the SD, a dark scene<br />

(space view) and the sun.<br />

Spectral Radiometric Calibration Assembly The Spectral Radiometric Calibration<br />

Assembly (SRCA) provi<strong>de</strong>s on orbit radiometric, spectral and spatial calibration without<br />

interfering with the sensor current acquisition. An integrating sphere with four lamps<br />

directs light on a toroidal relay mirror that reflects it towards a Czerny-Turner monochromator.<br />

A grating/mirror assembly then points the monocromatic light into the VIS, NIR<br />

and SWIR bands to evaluate spectral response <strong>de</strong>viations. Radiometric response is evaluated<br />

when the entrance and exit aperture of the SRCA are open and a mirror replaces<br />

the grating. The SRCA integrating sphere provi<strong>de</strong>s six-level radiometric sources used for<br />

the radiometric calibration of the reflective bands. Band-to-Band spatial registration is<br />

achieved with reticule patterns placed at the exit of the monochromator and projected<br />

into MODIS optical system.


33<br />

Blackbody : The calibration of MODIS emissive bands is achieved with the Blackbody,<br />

a component with zero reflectivity and an effective emissivity above 0.992. The temperature<br />

of the Blackbody is <strong>de</strong>termined with a precision of ±0.1 K and is measured with twelve<br />

thermistors. For every scan line (1.47s), a two-point calibration of the thermal bands is<br />

achieved with the scan mirror successively viewing space and the blackbody.<br />

2.5.4 MODIS products<br />

The generation of MODIS products from instrument raw data is done at NASA’s<br />

DAAC (Distributed Active Archive Center). The distribution of the products is then<br />

ensured by NASA’s Goddard Space Flight Center, the United States Geological Survey’s<br />

(USGS) EROS Data Center and the NOOA’s National Snow and Ice Data Center<br />

(NSIDC). MODIS products are organized in three levels.<br />

Level 1A products are composed of digital counts and data related to spacecraft and<br />

instrument telemetry. L1A data is processed at the EOS Data and Operations Systems<br />

(EDOS), then stored in 5-min granules containing all necessary information for further<br />

geolocation and calibration.<br />

Level 1B products are <strong>de</strong>duced from L1A and correspond to Top of the Atmosphere<br />

(TOA) calibrated radiances.<br />

Level 2 products result from the application of atmospheric correction and bio-optical<br />

algorithms to L1B data. L2 data represents geolocated geophysical variables.<br />

Level 3 products correspond to spatially binned L2 data accumulated during a time<br />

period of one day, 3 days, 8 days, a mounth or an entire year. L3 data are mapped in an<br />

Equidistant Cylindrical Projection grid of the earth and available at resolutions of 4 km<br />

and 9 km.<br />

MODIS geophysical products belong to four major disciplines, Ocean, Land, Atmosphere<br />

and Cryosphere (Table 2.2).<br />

2.5.5 Stripe noise on MODIS<br />

The analysis of MODIS level 1B data <strong>de</strong>rived from reflective and emmissive bands reveals<br />

the presence of striping with slightly different aspects. [Gumley, 2002] was the first to<br />

point out that three types of stripes can actually be seen on MODIS (figure 2.12). Mirror<br />

si<strong>de</strong> stripes (mirror banding) affects most reflective bands and some emissive bands located<br />

below the MWIR range. However, they are visible only over bright targets. Mirror banding<br />

appears to be the result of a quasi constant offset between forward and backward scans<br />

(we recall that MODIS scanning mirror is double si<strong>de</strong>d) and is systematically located in<br />

regions displaying radiance values high enough to bring the sensor close to its saturation


34 2. Remote Sensing with MODIS<br />

Figure 2.12 – Three types of stripe noise on Terra MODIS Level 1B geolocated calibrated<br />

radiances (Left) Detector-to-<strong>de</strong>tector stripes (Band 27) (Center) Mirror si<strong>de</strong> stripes<br />

(Band 9) (Right) Random stripes (Band 33). All the images are 200×200 pixels in size<br />

with a resolution of 1km<br />

mo<strong>de</strong>. Typical cases of mirror banding can be seen in homogeneous areas contaminated<br />

with sun glint or high concentrations of aerosol. The analysis of oceanographic data indicates<br />

that mirror banding reaches a maximal amplitu<strong>de</strong> in the specular direction, (highest<br />

level of sun glint) and tends to fa<strong>de</strong> away as the sun glint intensity <strong>de</strong>creases. This shows<br />

that mirror si<strong>de</strong> stripes are <strong>de</strong>pen<strong>de</strong>nt on the signal level and can be correlated with the<br />

scan angle.<br />

Detector-to-<strong>de</strong>tector stripes take the form of a periodic pattern of stripes and unlike mirror<br />

banding they cover entire MODIS swaths. Studies related to other sensors Horn and<br />

Woodham [1979], attribute the presence of these periodic stripes to a poor gain/offset<br />

calibration of the indivual <strong>de</strong>tectors composing the sensor. Furthermore, <strong>de</strong>tectors responses<br />

display strong non linear effects in the low radiance range (figure 3.4).<br />

The third type of stripes are random and appear clearly on Terra MODIS band 33 as<br />

black and white stripes with a limited lenght over a given scan line. Noisy stripes are<br />

presumably due to errors in the internal system and random noise.<br />

The processing of level 1B data to level 2 geophysical products does not inclu<strong>de</strong> a correcting<br />

algorithm for striping. The most recent version of MODIS data (collection 5),<br />

inclu<strong>de</strong>s a <strong>de</strong>striping procedure [] limited only to land surface reflectances. As atmospheric<br />

correction and bio-optical algorithms combine the information from multiple spectral<br />

bands to generate level 2 data, striping tends to be emphazised in the <strong>de</strong>rived geophysical<br />

products. Mirror banding clearly affects ocean colour products such as normalized<br />

water leaving radiance and chlorophyll concentration (figure 2.13). Level 2 daytime Sea<br />

Surface Temperature (SST) is computed from bands 31 and 32, and shows evi<strong>de</strong>nce of<br />

<strong>de</strong>tector-to-<strong>de</strong>tector stripes. Going further in MODIS processing chain, striping appears<br />

to persist even in level 3 products, altough this effect migh also be originating from the<br />

poor resolution of level 3 data (4km). It is clear, that stripe noise on level 1B data impacts<br />

the quality of higher level products and needs to be corrected efficiently before the gene-


35<br />

ration of level 2 and level 3 products. Because the stripe noise does not affect similarly<br />

all spectral bands, the <strong>de</strong>rivation of geophysical variables often results in a stripe noise<br />

amplification. For instance, striping can hardly be seen on bands 31 and 32. Yet, SWIR<br />

and LWIR SST products display stripe noise. MODIS SST is obtained with an NLSST<br />

(Non linear SST) algorithm that can be expressed as :<br />

SST = a 0 + a 1 .BT 11 + a 2 .(BT 11 − BT 12 )+a 3 .(BT 11 − BT 12 ).( 1 − 1) (2.10)<br />

µ<br />

where a 0 , a 1 , a 2 , a 3 are time <strong>de</strong>pen<strong>de</strong>nt coefficients <strong>de</strong>termined by the Rosentiel School of<br />

Marine and Atmospheric Science from in-situ measurements. BT 11 and BT 12 are brightness<br />

temperatures of bands 31 and 32, and µ is the cosine of the sensor zenith angle. The<br />

difference between bands 31 and 32 is used to correct atmospheric effects due to waper<br />

vapor absorption and is responsible for the introduction of striping in the SST. The amplification<br />

of striping can be stronger on products generated from algorithms that inclu<strong>de</strong><br />

multiplicative operations between spectral bands, which is the case of many bio-optical<br />

mo<strong>de</strong>ls used to estimate oceanographic variables.


36 2. Remote Sensing with MODIS<br />

Figure 2.13 – Stripe noise on Aqua MODIS Level 2 ocean products (TL) Remote sensing<br />

reflectance at 412 nm (TR) Chlorophyll concentration (BL) Diffuse attenuation coefficient<br />

at 490 nm (BR) Sea surface temperature at 11 µm


37<br />

Discipline Product ID Product name<br />

MOD01 Level-1A Radiance Counts<br />

– MOD02 Level-1B Calibrated Geolocation Radiances<br />

MOD03 Geolocation Data Set<br />

MOD04 Aerosol Product<br />

MOD05 Total Precipitable Water<br />

Atmosphere<br />

MOD06 Cloud Product<br />

MOD07 Atmospheric Profiles<br />

MOD08 Grid<strong>de</strong>d Atmospheric Product<br />

MOD035 Cloud Mask<br />

Cryosphere<br />

MOD010 Snow Cover<br />

MOD029 Sea Ice Cover<br />

MOD09 Surface Reflectance (Atmospheric Correction)<br />

MOD011 Land Surface Temperature and Emissivity<br />

MOD012 Land Cover/Land Cover Change<br />

MOD013 Grid<strong>de</strong>d Vegetation Indices (NDVI & EVI)<br />

MOD014 Thermal Anomalies - Fires and Biomass Burning<br />

Land MOD015 Leaf Area In<strong>de</strong>x (LAI) and FPAR<br />

MOD016 Evapotranspiration<br />

MOD017 Vegetation Production, Net Primary Productivity (NPP)<br />

MOD040 Grid<strong>de</strong>d Thermal Anomalies<br />

MOD043 Surface Reflectance BRDF/Albedo Parameter<br />

MOD044 Vegetation Cover Conversion<br />

MOD018 Normalized Water-leaving Radiance<br />

MOD019 Pigment Concentration<br />

MOD020 Chlorophyll Fluorescence<br />

MOD021 Chlorophylla Pigment Concentration<br />

MOD022 Photosynthetically Available Radiation (PAR)<br />

MOD023 Suspen<strong>de</strong>d-Solids Concentration<br />

MOD024 Organic Matter Concentration<br />

MOD025 Coccolith Concentration<br />

Ocean MOD026 Ocean Water Attenuation Coefficient<br />

MOD027 Ocean Primary Productivity<br />

MOD028 Sea Surface Temperature<br />

MOD029 Sea Ice Cover<br />

MOD031 Phycoerythrin Concentration<br />

MOD032 Processing Framework and Match-up Database<br />

MOD036 Total Absorption Coefficient<br />

MOD037 Ocean Aerosol Properties<br />

MOD039 Clean Water Epsilon<br />

Table 2.2 – MODIS products


38 2. Remote Sensing with MODIS


39<br />

Chapitre 3<br />

Standard <strong>de</strong>striping techniques<br />

and application to MODIS<br />

3.1 Data<br />

The characteristics of stripe noise on the MODIS instrument are more complex than<br />

those observed on other imaging spectrometers. We shall see that early <strong>de</strong>striping techniques<br />

<strong>de</strong>velopped to improve the quality of Landsat MSS/TM images, provi<strong>de</strong> unsatisfactory<br />

results for MODIS data. As illustrated in figure 2.12, images extracted from MODIS<br />

are contaminated with three different types of stripes, ubiquitous on few spectral bands.<br />

Although this increases the difficulties associated with <strong>de</strong>striping, it constitutes - in addition<br />

to the importance of MODIS in earth science research - a complementary motivation<br />

to explore new techniques.<br />

Despite a growing number of <strong>de</strong>striping methodologies, very few take into account particular<br />

cases of stripe noise. In fact, a major concern on MODIS is the presence of random<br />

stripes, so far discussed only in [Rakwatin et al., 2007]. In this chapter, we explore several<br />

techniques and we analyse the results obtained on images extracted from specific emissive<br />

bands of MODIS.<br />

The algorithms <strong>de</strong>scribed in each section are applied on both Terra and Aqua level 1B<br />

TOA calibrated radiances. Given the consi<strong>de</strong>rable size of MODIS swaths (2330 km cross<br />

track ), we illustrate most of the results on 512 × 512 sub-images to allow accurate visual<br />

examination of small scale features. We specifically selected bands 27, 30 and 33 for Terra<br />

MODIS and bands 27, 30 and 36 for Aqua MODIS as they are representative of an extreme<br />

case due to severe striping with strong non-linear effects. All these bands display periodic<br />

stripes (mainly <strong>de</strong>tector-to-<strong>de</strong>tector), except band 33 of Terra MODIS which contains<br />

mostly random stripes. The data set used for this study is composed of three images for<br />

Terra MODIS and three images for Aqua MODIS, each extracted from a different spectral<br />

band (see figure 3.1). The scene captured by Terra MODIS was acquired on July 1 st , 2009<br />

in the Mediterannean Sea. The images from Aqua MODIS were acquired on November


40 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.1 – MODIS data set used for the analysis of <strong>de</strong>striping algorithms and composed<br />

of level 1B TOA calibrated radiances (TL) Terra band 27 (TC) Terra band 30 (TR) Terra<br />

band 33 (BL) Aqua band 27 (BC) Aqua band 30 (BR) Aqua band 36<br />

10, 2009, in the Pacific Ocean on the west coast of USA.<br />

3.2 Moment Matching<br />

The need for post-processing algorithms and the <strong>de</strong>velopment of <strong>de</strong>striping methods<br />

quickly followed the distribution of data collected by Landsat MSS. The MSS incorporates<br />

a rotating scan mirror that redirects the scene energy to six separate <strong>de</strong>tectors.<br />

After completion of an entire scan sweep, the orbital motion of the satellite ensures the<br />

acquisition of six additional lines and so forth. The six individual <strong>de</strong>tectors of MSS are<br />

not perfectly i<strong>de</strong>ntical and the corresponding transfer functions are subject to slight <strong>de</strong>viations<br />

from each other. During the acquisition process, the radiometric drift between<br />

<strong>de</strong>tectors produces an un<strong>de</strong>sirable striping effect in the image(*add*). The removal of<br />

this artifact requires an accurate characterization of the <strong>de</strong>tectors response in term of<br />

<strong>de</strong>viations from nominal operational mo<strong>de</strong>. Assuming that input/output functions of the<br />

<strong>de</strong>tectors established during the pre-launch calibration stage remain the same during onorbit<br />

operations, a simple inversion of these functions could provi<strong>de</strong> a reliable correction


41<br />

for striping. The drift observed on MSS <strong>de</strong>tectors response was attributed in [Horn and<br />

Woodham, 1979] to the <strong>de</strong>gradation of photomultipliers with respect to exposure time.<br />

Despite the calibration system on-board LANDSAT, MSS images still contain stripes and<br />

additional radiometric correction were <strong>de</strong>velopped by [Strome and Vishnubhatla, 1973]<br />

and [Sloan and Orth, 1977]. The resulting method is referred to as moment matching and<br />

is based on the assumption of both linear and time invariant behavior of the <strong>de</strong>tectors.<br />

Stripe noise is assumed to be the result of relative uncorrected gain and offset coefficients.<br />

Hereafter, we <strong>de</strong>note d k the signal acquired by <strong>de</strong>tector number k. In the case of MODIS,<br />

k =1..10 (for FPAs with 1 km resolution), images are formed by interlacing signals from<br />

the 10 <strong>de</strong>tectors. Un<strong>de</strong>r the assumption of linear responses, the signal from <strong>de</strong>tector k can<br />

be written as :<br />

d k = G k .d c k + O k (3.1)<br />

where G k and O k are gain and offset parameters, d k is the noisy signal and d c k<br />

is the true<br />

signal for <strong>de</strong>tector k. Once G k and O k are <strong>de</strong>termined, values of <strong>de</strong>tector k are corrected<br />

as :<br />

d c k = d k − O k<br />

(3.2)<br />

G k<br />

Unless calibration targets with known radiance values are used, absolute values for G k and<br />

O k remain un<strong>de</strong>termined. Nevertheless, <strong>de</strong>striping only requires relative gain/offset coefficients.<br />

These can be estimated statistically from the entire image. In fact, the acquisition<br />

process of whiskbroom systems makes it reasonnable to assume that signals acquired by<br />

each <strong>de</strong>tector share similar statistical characteristics. The validity of this hypothesis holds<br />

for images with high dimensions. Relative gain/offsets can then be <strong>de</strong>termined from the<br />

mean value and standard <strong>de</strong>viation of a signal d ref acquired by a pre<strong>de</strong>termined reference<br />

<strong>de</strong>tector. Let us <strong>de</strong>note µ k and σ k , - respectively µ ref et σ ref - the mean value and standard<br />

<strong>de</strong>viation of d k , - resp. d ref -. Moment matching adjusts the response of a noisy <strong>de</strong>tector<br />

by forcing its mean value and standard <strong>de</strong>viation to coinci<strong>de</strong> with those <strong>de</strong>rived from the<br />

reference <strong>de</strong>tector. Using equation (3.1), this translates to :<br />

and gain and offset are obtained as :<br />

µ k = G k .µ ref + O k<br />

σ k = G k .σ ref<br />

(3.3)<br />

G k =<br />

σ k<br />

σ ref<br />

O k = µ k − µ ref . σ k<br />

σ ref<br />

(3.4)<br />

Replacing equations (3.4) in (3.2), the signal from <strong>de</strong>tector k is corrected with :<br />

ˆd c k = σ ref .(d k − µ k )<br />

σ k<br />

− µ ref (3.5)


42 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Detector Terra Band 27 Terra Band 30 Terra Band 33<br />

– Mean value Std <strong>de</strong>viation Mean value Std <strong>de</strong>viation Mean value Std <strong>de</strong>viation<br />

d 1 12982 1043.6 16933 2441.9 23072 1660.9<br />

d 2 12982 1043.6 16933 2441.4 23072 1660.9<br />

d 3 12982 1043.7 16934 2442.5 23072 1661.1<br />

d 4 12981 1043.7 16934 2442.1 23072 1661.3<br />

d 5 12982 1043.9 16934 2442.1 23072 1661.6<br />

d 6 12982 1043.9 16934 2442.1 23072 1661.0<br />

d 7 12984 1044.1 16934 2442.0 23072 1661.0<br />

d 8 12981 1044.0 16934 2442.0 23072 1660.9<br />

d 9 12980 1044.0 16934 2442.0 23072 1660.9<br />

d 10 12980 1043.9 16934 2442.0 23072 1660.6<br />

Table 3.1 – Mean/standard <strong>de</strong>viation values (DN) of 10 <strong>de</strong>tectors for Terra<br />

MODIS bands 27, 30 and 33. The small statistical <strong>de</strong>viation between <strong>de</strong>tectors<br />

discredits the application of moment matching on MODIS data<br />

The moment matching technique only rectifies first/second or<strong>de</strong>r statistics and as such<br />

requires a careful selection of the reference signal. If a <strong>de</strong>tector with strong statistical<br />

<strong>de</strong>viations from others (a <strong>de</strong>tector responsible for random stripes for exemple) is used<br />

as a reference, moment matching can result in very poor <strong>de</strong>striping. Many strategies based<br />

on the analysis of inter-<strong>de</strong>tector response statistics can be used (see section 3.4). A<br />

common option is to rely on mean/standard <strong>de</strong>viation values computed over the entire<br />

scene. Altough the simplicity of its implementation makes it a popular technique, moment<br />

matching suffers from many limitations <strong>de</strong>scribed in [Horn and Woodham, 1979]. A major<br />

issue pointed out on MSS is the impact of non-linearities in photomultipliers responses,<br />

presumably abscent in more sophisticated imaging <strong>de</strong>vices. Mean/standard <strong>de</strong>viation values<br />

reported in tables 3.1, 3.2 and experiments conducted on both Terra and Aqua MODIS<br />

data, illustrate how moment matching only results in partial <strong>de</strong>striping (figure 3.3). In<strong>de</strong>ed,<br />

the gain/offset mo<strong>de</strong>l only modifies the affine response of the <strong>de</strong>tectors and fails to<br />

take into account non-linear effects, highly responsible for residual stripes. Local analysis<br />

of MODIS swaths reveals a <strong>de</strong>pen<strong>de</strong>ncy of <strong>de</strong>tectors linear response to the signal intensity.<br />

This can be verified experimentally by estimating gains/offsets from swaths covering<br />

different levels of radiances (oceans and clouds). Furthermore, first or<strong>de</strong>r statistics make<br />

the method very sensitive to the geophysical content of the images. In many cases, small<br />

clouds or other highly reflective targets, are only visible by a limited set of <strong>de</strong>tectors and<br />

the resulting statistical bias is not accounted for in the moment matching procedure.<br />

3.3 Histogram Matching<br />

The hypothesis of linear and stationnary response exploited by the moment matching<br />

method is too strong to provi<strong>de</strong> reliable results on MODIS data. Investigation of<br />

tables 3.1, 3.2 shows that on Terra/Aqua bands severely contaminated with stripes, <strong>de</strong>tectors<br />

have very similar mean/standard <strong>de</strong>viation values. The visual examination of su-


Figure 3.2 – Sub-images acquired by each of the 10 <strong>de</strong>tectors of Terra MODIS in band<br />

27. From Top to bottom and left to right, signals from <strong>de</strong>tectors d 1 to d 10 . Visual analysis<br />

indicates that stripes are not only due to differences between mean values and variances<br />

of adjacent <strong>de</strong>tectors. Images <strong>de</strong>rived from some individual <strong>de</strong>tectors also display strong<br />

striping. This is clearly visible for <strong>de</strong>tectors 1 and 6<br />

43


44 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.3 – (Left) Noisy image from Terra MODIS band 27 (Right) Image <strong>de</strong>striped<br />

with moment matching<br />

Detector Aqua Band 27 Aqua Band 30 Aqua Band 36<br />

– Mean value Std <strong>de</strong>viation Mean value Std <strong>de</strong>viation Mean value Std <strong>de</strong>viation<br />

d 1 14026 3508.2 14684 2872.0 22497 1420.3<br />

d 2 14026 3508.1 14685 2871.7 22497 1420.4<br />

d 3 14027 3507.8 14685 2871.2 22498 1420.3<br />

d 4 14027 3508.3 14685 2871.0 22498 1420.5<br />

d 5 14027 3508.4 14685 2870.9 22497 1420.7<br />

d 6 14028 3508.3 14685 2870.9 22498 1420.8<br />

d 7 14028 3508.2 14686 2870.6 22498 1420.8<br />

d 8 14027 3507.2 14685 2870.6 22498 1420.6<br />

d 9 14029 3508.3 14686 2870.3 22498 1420.6<br />

d 10 14030 3508.5 14687 2870.2 22499 1420.6<br />

Table 3.2 – Mean/standard <strong>de</strong>viation values of 10 <strong>de</strong>tectors for Aqua MO-<br />

DIS bands 27, 30 and 36<br />

bimages extracted from individual <strong>de</strong>tectors (figure 3.2) shows evi<strong>de</strong>nce of striping. This<br />

means that suddle <strong>de</strong>viations occur within a single <strong>de</strong>tector between two consecutive scan<br />

sweeps. Given the short time period between two scans, 1.477 seconds, these intra-<strong>de</strong>tector<br />

<strong>de</strong>viations cannot be related to a potential physical <strong>de</strong>gradation but are very likely to be<br />

induced by the on-board calibration procedure itself ; For instance, the blackbody (section<br />

) provi<strong>de</strong>s a two-point calibration for all emmissive bands every 1.477s.<br />

The inefficiencies of moment matching due to non linear effects on MSS, gave rise to an<br />

improved method. To overcome the limitations of first or<strong>de</strong>r statistics, [Horn and Woodham,<br />

1979] introduced a refined approach where the probability distribution function of<br />

scene radiances captured by each <strong>de</strong>tector is assumed to be i<strong>de</strong>ntical. In fact, radiations<br />

emmitted or reflected by the earth reach all the <strong>de</strong>tectors with approximatively the same


45<br />

magnitu<strong>de</strong>. The <strong>de</strong>rived technique, histogram matching, then consists in adjusting the empirical<br />

cumulative distribution function (ECDF) of each <strong>de</strong>tector to an ECDF selected as<br />

reference. Let us consi<strong>de</strong>r the signal measured by <strong>de</strong>tector k as a discrete random variable<br />

X k , with a probability distribution p k (x). The ECDF of X k is <strong>de</strong>fined as :<br />

and computed as :<br />

P k (x) =P (X k ≤ x) (3.6)<br />

P k (x) =<br />

x∑<br />

p k (i) (3.7)<br />

We consi<strong>de</strong>r P ref to be the ECDF of a reference <strong>de</strong>tector, x is a value measured by <strong>de</strong>tector<br />

k and x ′ its corresponding corrected value. Forcing <strong>de</strong>tector k to have the same ECDF as<br />

the reference <strong>de</strong>tector, the following relation holds :<br />

i=0<br />

P ref (x ′ )=P k (x) (3.8)<br />

The ECDF P ref is a non increasing function and can be inversed to <strong>de</strong>termine an estimate<br />

value of x ′ as :<br />

x ′ = P −1<br />

ref (P k(x)) (3.9)<br />

When applied to the set of acquired values, equation (3.9) provi<strong>de</strong>s a normalization lookup<br />

table that associates to every signal value x, its corrected value x ′ (figures 3.4 and 3.5).<br />

Analogously to moment matching, the reference <strong>de</strong>tector can be <strong>de</strong>termined experimentally<br />

from the analysis of each <strong>de</strong>tector or by selecting the ECDF of the entire swath.<br />

Following the results obtained on Landsat MSS, the histogram matching method was<br />

later used in 1985 by [Poros and Peterson, 1985] for the <strong>de</strong>striping of Landsat TM. The limitations<br />

of histogram matching were first discussed in [Wegener, 1990] and a modification<br />

of the original approach was introduced to reinforce the assumption of similar <strong>de</strong>tectors<br />

ECDFs. The consi<strong>de</strong>rable size of captured swaths systematically garantees the acquisition<br />

of geophysical data diverse enough in terms of radiances to contradict the hypothesis of<br />

statistically i<strong>de</strong>ntical ECDFs. Wegener suggested to constrain the estimation of <strong>de</strong>tectors<br />

statistics to the only homogeneous regions of the swath. In his approach, a noisy image<br />

is <strong>de</strong>composed into subimages which size is a multiple of the number of <strong>de</strong>tectors in the<br />

instrument. In the case of Landsat MSS, images are fragmented into blocs of 12 × 12<br />

pixels (MSS has 6 <strong>de</strong>tectors), and used for the computation of ECDFs if they satisfy a<br />

homogeneous criteria established by Bienaymé-Tchebychev inequality :<br />

P (|x − µ| > kσ) ≤ k −2 (3.10)<br />

where µ and σ are mean/standard <strong>de</strong>viation values of a subimage, and k ≤ 1 a real number<br />

that <strong>de</strong>termines the percentage of rejected sub-images.<br />

Another interesting implementation of histogram matching was introduced by [Weinreb<br />

et al., 1989], for the <strong>de</strong>striping of meteorological data acquired by GOES (Geostationary


46 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d1<br />

d7<br />

d8<br />

d9<br />

d10<br />

9000 10500 12000 13500 15000 16500<br />

Radiance values<br />

Corrected Radiances<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

d1<br />

d7<br />

d8<br />

d9<br />

d10<br />

6000<br />

6000 8000 10000 12000 14000 16000<br />

Measured Radiances<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d1<br />

d3<br />

d5<br />

d6<br />

d8<br />

12000 14000 16000 18000 20000 22000<br />

Radiance values<br />

Corrected Radiances<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

d1<br />

d3<br />

d5<br />

d6<br />

d8<br />

6000<br />

6000 8000 10000 12000 14000 16000<br />

Measured Radiances<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d1<br />

d2<br />

d5<br />

d7<br />

d9<br />

20000 21500 23000 24500 26000<br />

Radiance values<br />

Corrected Radiances<br />

25000<br />

22000<br />

19000<br />

16000<br />

13000<br />

d1<br />

d2<br />

d5<br />

d7<br />

d9<br />

10000<br />

10000 13000 16000 19000 22000 25000<br />

Measured Radiances<br />

Figure 3.4 – (Left) From top to bottom, Empirical Cumulative Distribution Functions<br />

for noisy <strong>de</strong>tectors of Terra MODIS bands 27, 30 and 33. (Right) From top to bottom,<br />

Normalization Look-up table for the same <strong>de</strong>tectors obtained with the histogram matching<br />

technique (IMAPP) and used to correct striping. Nonlinear effects are highly present in<br />

the low-radiance range<br />

Operational Environmental Satellites). The methodology was <strong>de</strong>dicated to the pre-launch<br />

calibration of the 8 <strong>de</strong>tectors composing GOES I-M (launched by NOAA in 1990). While<br />

the original method of [Horn and Woodham, 1979] <strong>de</strong>rives and applies a normalization


47<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d3<br />

d6<br />

d8<br />

d9<br />

d10<br />

6000 9000 12000 15000 18000 21000<br />

Radiance values<br />

Corrected Radiances<br />

18000<br />

15000<br />

12000<br />

9000<br />

6000<br />

d3<br />

d6<br />

d8<br />

d9<br />

d10<br />

3000<br />

3000 6000 9000 12000 15000 18000<br />

Measured Radiances<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d3<br />

d4<br />

d8<br />

d9<br />

d10<br />

7000 10000 13000 16000 19000 22000<br />

Radiance values<br />

Corrected Radiances<br />

17000<br />

14000<br />

11000<br />

8000<br />

d3<br />

d4<br />

d8<br />

d9<br />

d10<br />

5000<br />

5000 8000 11000 14000 17000<br />

Measured Radiances<br />

ECDF<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

d1<br />

d2<br />

d3<br />

d4<br />

d8<br />

18000 19500 21000 22500 24000 25500<br />

Radiance values<br />

Corrected Radiances<br />

24000<br />

22000<br />

20000<br />

18000<br />

d1<br />

d2<br />

d3<br />

d4<br />

d8<br />

16000<br />

16000 18000 20000 22000 24000<br />

Measured Radiances<br />

Figure 3.5 – (Left) From top to bottom, Empirical Cumulative Distribution Functions<br />

for noisy <strong>de</strong>tectors of Aqua MODIS bands 27, 30 and 36. (Right) From top to bottom,<br />

Normalization Look-up table for the same <strong>de</strong>tectors obatined with the histogram matching<br />

technique (IMAPP). Comparison with Terra normalization look-up table indicates<br />

a improved pre-launch calibration<br />

look-up table to a single image, Weinreb suggests using the same table to process in<strong>de</strong>pen<strong>de</strong>nt<br />

acquisitions. The normalization table is generated from a specific sample covering<br />

a wi<strong>de</strong> dynamic range of radiances. An image acquired by GOES-7 on May 18, 1988 was


48 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.6 – (Left) Noisy image from Terra MODIS band 27 (Right) Image <strong>de</strong>striped<br />

with histogram matching (IMAPP)<br />

used to establish the 8 <strong>de</strong>tectors normalization curves. The look-up table was then successfully<br />

applied to <strong>de</strong>stripe an in<strong>de</strong>pen<strong>de</strong>nt image collected two weeks later on June 1,<br />

1988.<br />

More recently, this approach was adopted by [Xiaoxiang et al., 2007] for the <strong>de</strong>striping of<br />

CMODIS (SZ-3 Chinese MODIS) data.<br />

It is currently implemented in the NASA IMAPP (International MODIS/AIRS Processing<br />

Package) software and distributed by the university of Madison-Wisconsin for MODIS<br />

users (http ://cimss.ssec.wisc.edu/imapp/). Additional features specific to MODIS data<br />

have been inclu<strong>de</strong>d in the software. Two sets of look-up tables were <strong>de</strong>veloped for both<br />

Terra and Aqua MODIS, and their application <strong>de</strong>pends on the acquisition date of images to<br />

be <strong>de</strong>triped. The comparative study of section 3.8 uses the histogram matching technique<br />

implemented in the NASA IMAPP software.<br />

3.4 Overlapping Field-of-View Method<br />

The Overlapping Field-of-View method (OFOV) was specifically <strong>de</strong>signed for MODIS.<br />

Its basic principle was introduced by [Antonelli et al., 2004] and recently examined by<br />

[di Bisceglie et al., 2009]. This technique relies on another artifact of MODIS, the bowtie<br />

effect. When the scan angle of the mirror increases, the actual resolution of pixels also<br />

increases due to the earth curvature. The nominal resolution of a pixel is 1 × 1 km at<br />

Nadir and reaches 4.8 × 2km at the begining and ending of every scan, which correspond<br />

to scan angles of ±55˚. As a result of pixel resolution growth, a swath will cover an<br />

exten<strong>de</strong>d area of 20km along track at scan extremums, causing overlaps with previous<br />

and next swaths (figure 3.7a). Areas affected with the bowtie effect will display i<strong>de</strong>ntical


49<br />

Figure 3.7 – (Left) Bowtie effect due to the growth of pixel size in the cross track<br />

direction of a swath (Right) Image from Terra MODIS band 17, extracted from an area<br />

with a scan angle raging from 47˚ to 55˚. Bowtie effect appears as overlapping fields of<br />

view.<br />

geometrical features in successive scans. This is clearly visible along land/ocean transitions<br />

(figure 3.7b). The OFOV method then exploits the information redundancy to equalize<br />

the <strong>de</strong>tectors responses. In the bowtie region (areas corresponding to scan angles higher<br />

than ±25˚), some of MODIS <strong>de</strong>tectors are viewing exactly the same scene and, therefore<br />

thus <strong>de</strong>viations in the <strong>de</strong>tectors reponses are only due to stripe noise. The OFOV algorithm<br />

is divi<strong>de</strong>d in two steps. Following the classification of the <strong>de</strong>tectors as in-family and outof<br />

family <strong>de</strong>tectors, a reference <strong>de</strong>tector is selected and equalization functions are <strong>de</strong>rived<br />

using only bow-tie measurements. The equalization curves are then applied to the <strong>de</strong>tectors<br />

over the entire swath. To assess the accuracy of relative inter-<strong>de</strong>tectors calibration, a Bowtie<br />

Based Detector Distance (BTBDD) between <strong>de</strong>tectors i and j is <strong>de</strong>fined as :<br />

d i,j =<br />

∑ Ni,j<br />

n=1 |I i(n) − I j (n)| .W (n)<br />

∑ Ni,j<br />

n=1 W (n) (3.11)<br />

where N i,j is the number of OFOVs of <strong>de</strong>tectors i and j, with an overlapping percentage<br />

above η, a value fixed to 65%. I i (n) (respectively I j (n)) is the radiance measured by<br />

<strong>de</strong>tector i (respectively j) on the n th OFOV and W (n) is the overlapping percentage. In<br />

addition to the BTBDD distances, the similarity between <strong>de</strong>tectors ECDFs is computed<br />

using the Kolmogorov-Smirnov distance :<br />

D i,j = 1<br />

N obs<br />

N obs<br />

∑<br />

|P i (r(n)) − P j (r(n))| (3.12)<br />

n


50 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Table 3.3 – Mean Kolmokorov-Smirnov <strong>de</strong>tectors distance for bands 27, 30<br />

and 33 for Terra MODIS and 27, 30 and 36 for Aqua MODIS<br />

– Terra Aqua<br />

Detector Band 27 Band 30 Band 33 Band 27 Band 30 Band 36<br />

d 1 0.3544 0.1108 0.0413 0.0265 0.0114 0.0189<br />

d 2 0.1637 0.0253 0.0150 0.0253 0.0113 0.0142<br />

d 3 0.1624 0.0261 0.0149 0.0338 0.0120 0.0192<br />

d 4 0.1604 0.0253 0.0140 0.0246 0.0130 0.0139<br />

d 5 0.1562 0.0312 0.0181 0.0245 0.0111 0.0114<br />

d 6 0.1579 0.0273 0.0130 0.0278 0.0111 0.0106<br />

d 7 0.1973 0.0259 0.0154 0.0255 0.0106 0.0118<br />

d 8 0.2188 0.0429 0.0142 0.0958 0.0125 0.0135<br />

d 9 0.3195 0.0260 0.0218 0.0353 0.0134 0.0134<br />

d 1 0 0.2587 0.0248 0.0152 0.0514 0.0379 0.0140<br />

where P i and P j are the ECDFs of <strong>de</strong>tectors i and j, and r(n) is the set of radiances<br />

observed by both <strong>de</strong>tectors. The Kolmogorov-Smirnov distance measures the distributional<br />

distance for every couple of <strong>de</strong>tectors and can be averaged for a single <strong>de</strong>tector i ≠ j as :<br />

D i =<br />

1<br />

card(M i ) − 1<br />

∑<br />

D i,j<br />

(3.13)<br />

where M i is the set of <strong>de</strong>tectors. [Antonelli et al., 2004] and [di Bisceglie et al., 2009],<br />

specify that the classification of <strong>de</strong>tectors and the selection of a reference one is <strong>de</strong>termined<br />

from both BTBDD and Kolmogorov-Smirnov distances. However, it is not mentioned how<br />

this is done. As an alternative, we will rely only on the Kolmogorov-Smirnov distances ;<br />

the weakest value D M i<br />

i<br />

<strong>de</strong>termines the reference <strong>de</strong>tector to be used for the radiometric<br />

equalization. Let us <strong>de</strong>note by B r (n) and B i (n) the set of radiances measured in the<br />

bowtie region by a reference <strong>de</strong>tector r and a noisy <strong>de</strong>tector i. The equalization function<br />

is obtained with a polynomial regression between B r (n) and B i (n) :<br />

j∈M i<br />

j≠i<br />

B r (n) =p 0 + p 1 .B i (n)+p 2 .B 2 i (n)+... + p N .B N i (n) (3.14)<br />

where p 0 , p 1 ...p N are the coefficients of the best fitting polynome. The radiometric equalisation<br />

of the signal I i (n) measured by the <strong>de</strong>tector i over the entire swath is then corrected<br />

as :<br />

Î i (n) =p 0 + p 1 .I i (n)+p 2 .I 2 i (n)+... + p N .I N i (n) (3.15)<br />

All the other <strong>de</strong>tectors are corrected with the same procedure. Destriping results obtained<br />

with the OFOV technique are illustrated in figure 3.7. When the equalization is based on<br />

polynomial functions of or<strong>de</strong>r 1, the OFOV method might be comparable to the moment<br />

matching technique because only the affine response of the <strong>de</strong>tectors is modified. However,<br />

while moment matching is based on statistical assumptions satisfied only over homogeneous<br />

areas, the OFOV method relies entirely on the bowtie effect to equalize the <strong>de</strong>tectors


51<br />

Figure 3.8 – (Left) Noisy image from Terra MODIS band 27 (Right) Image <strong>de</strong>striped<br />

with the IFOV method<br />

response. To this extent, the OFOV method can be viewed as a radiometric calibration<br />

procedure. It is worth mentionning that only the range of radiances contained in the bowtie<br />

region is corrected and additional processing is required to cover the entire dynamic<br />

range of the acquired signal. In our implementation, the equalization functions were <strong>de</strong>rived<br />

from measurements with scan angles higher than ±25˚. Linear and quadratic fitting<br />

computed on the selected images showed little differences and only first or<strong>de</strong>r polynomes<br />

were used. Radiance values not contained in the bowtie region were not equalized.<br />

3.5 Frequency filtering<br />

Frequency filtering methods has been wi<strong>de</strong>ly used for the elimination of stripes on<br />

satellite images. [Srinivasan et al., 1988] proposed a new approach for Landsat TM and<br />

MSS based on a spectral analysis of the stripe noise. Despite extensive research in this<br />

direction [Crippen, 1989], [Simpson et al., 1995], [Simpson et al., 1998], [Chen et al., 2003]<br />

and good visual results on GOES and Landsat sensors, frequency filtering can only be<br />

viewed as a cosmetic improvement. The smoothed results obtained with low-pass filtering<br />

can nonetheless be used as a reference to compare the performances of other <strong>de</strong>triping<br />

techniques. In this subsection, we briefly explore the application of frequency filtering on<br />

MODIS data.<br />

3.5.1 Spectral Analysis<br />

In most cases, striping can be consi<strong>de</strong>red as a periodic noise. The application of a lowpass<br />

filter (an averaging filter in the spatial domain for exemple) reduces the visual impact


52 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.9 – (Left) Noisy image from Terra MODIS band 27 (Right) Module of its<br />

fourier transform, showing lobes in the vertical axis of the fourier domain<br />

of stripes but inevitably removes fine structures of the signal and, therefore compromisies<br />

further quantitative analysis. To minimize the loss of high-frequency information, the<br />

filtering procedure shall take into account the periodic nature of stripes prior to the <strong>de</strong>sign<br />

of a band-pass filter. Stripe-related frequencies are <strong>de</strong>termined from the spectrum of the<br />

noisy image I s , which we consi<strong>de</strong>r as a two dimensional vector of size M × N. Its discrete<br />

fourier transform is given by :<br />

I F s (u, v) =<br />

M∑<br />

N∑<br />

i=1 k=1<br />

( ( jui<br />

I s (i, k)exp −<br />

2πM + jvk ))<br />

2πN<br />

(3.16)<br />

where j is the imaginary unity (j = √ −1). Periodic stripes on images translate in the<br />

fourier spectrum as a signature taking the form of multiple lobes, concentrated along the<br />

vertical axis (figure 3.9b). An accurate estimation of periodic stripe frequencies requires a<br />

distinction from the frequencies related to the true signal structures. For instance, the frequency<br />

power spectrum estimated on the striped signal values, barely reveals the presence<br />

of periodic noise (figure 3.10a). The spectral components of stripes can be highlighted<br />

using the averaging method of<br />

3.5.2 Band-pass filtering<br />

The spectral analysis of MODIS data <strong>de</strong>scribed above, i<strong>de</strong>ntifies the frequencies to be<br />

reduced. [Simpson et al., 1995] proposed several possible implementations of Finite Impulsional<br />

Response (FIR) filters. The approach we retain here is based on a two-dimensional<br />

filter in fourier space composed of wells centered at stripe frequencies. The band-pass filter


53<br />

12<br />

11<br />

10<br />

10<br />

Power spectrum<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Power spectrum<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

−2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

3<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

Figure 3.10 – (Left) Power spectrum computed directly on the noisy image, frequncies<br />

of striping are not visible (Right) Power spectrum computed as an average of the columns<br />

periodograms ; striping frequencies appear as distinct peaks<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

u<br />

u<br />

v<br />

v<br />

Figure 3.11 – Frequency response of the filter H 1 . As the value of σ increases, the wells<br />

(represented here as lobes for visual clarity) overlap. When frequencies located near the<br />

center of fourier domain are reached by the lobes, the band-pass filter H 1 becomes a high<br />

pass filter<br />

takes the form :<br />

H 1 (u, v) =1− ∑<br />

(u s,v s)<br />

exp<br />

(− (u − u s) 2 +(v − v s ) 2 )<br />

σ 2<br />

(3.17)<br />

where u s and v s are the center coordinates of the wells in the fourier domain and σ<br />

controls the sharpness of the wells. The <strong>de</strong>sign of the FIR filter H 1 in the fourier 2D<br />

domain, implicitly takes into account the unidirectional spatial property of stripes noise<br />

by placing the wells along the vertical axis, centered at the same coordinates as the stripe<br />

lobes observed in figure 3.8. With increasing values of σ, the wells start to overlap with<br />

each other and eventually affect low-frequencies (figure 3.11b). For high values of σ, the<br />

filter H 1 shifts from a band-pass filter to a high-pass filter that removes all the low-pass


54 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.12 – (Left) Original image from Terra MODIS band 27 (Right) Image <strong>de</strong>striped<br />

with the filter H 1<br />

components. This results in high-pass filtered images containing mostly stripe noise. In<br />

addition, the mean value of filtered images differs from the original signal. To overcome<br />

this issue, we consi<strong>de</strong>r the following low-pass filter :<br />

H B (u, v) =exp<br />

(− u2 + v 2 )<br />

σ 2 B<br />

(3.18)<br />

where σ B is small enough so that the filter H B mainly retrieves the mean value of the<br />

noisy image. The <strong>de</strong>striped image is then obtained with the filter :<br />

H 2 (u, v) =H 1 (u, v)+H B (u, v) (3.19)<br />

An alternative implementation of the FIR filter can be done in the spatial domain by<br />

convolving each column with a unidimensional spatial FIR filter kernel <strong>de</strong>rived from the<br />

Parks-McClellan (PM) algorithm [Parks and Mcclellan, 1972]. Destriping in the spatial<br />

domain overcomes many limitations compared to the 2D fourier filtering. It removes the<br />

constraints related to images dimensions being a power of two and slightly reduces horizontal<br />

edge effects.<br />

Another FIR filtering approach was presented in [Srinivasan et al., 1988] for the <strong>de</strong>striping<br />

of Landsat. The fourier response of the proposed filter is composed of concentric<br />

rings centered at stripe frequencies. This method however, smoothes data equally in both<br />

directions and introduces blurring and ringing artefacts along the discontinuities of the<br />

images.


55<br />

Image distortion ID<br />

1<br />

0.98<br />

0.96<br />

0.94<br />

0.92<br />

0.9<br />

0.88<br />

0.86<br />

0.84<br />

0<br />

0 10 20 30 40 50<br />

σ<br />

ID<br />

NR<br />

2000<br />

1600<br />

1200<br />

800<br />

400<br />

Noise reduction NR<br />

Image distortion ID<br />

1<br />

0.98<br />

0.96<br />

0.94<br />

0.92<br />

0.9<br />

0.88<br />

0.86<br />

0.84<br />

ID<br />

NR<br />

30<br />

25<br />

20<br />

15<br />

10<br />

0<br />

0 10 20 30 40 50<br />

σ<br />

5<br />

Noise reduction NR<br />

Figure 3.13 – ID and NR in<strong>de</strong>xes as a function of σ in the filter H 2 for images of (Left)<br />

Terra band 27 and (Right) Terra band 30. The value of σ B is fixed to 5. Images from band<br />

27 are representative of smooth atmospheric effects and therefore, small scale structures<br />

in the images are mostly due to striping. In (Left) for σ > 21, the filter H 2 behaves<br />

as a high-pass filter that retains only high-frequency components, coinci<strong>de</strong>nt with stripe<br />

noise, hence the <strong>de</strong>crease in NR. This effect is not visible on images from band 30 where<br />

high-frequencies are composed of land-ocean-clouds discontinuities.<br />

3.6 Haralick Facet filtering<br />

Destriping techniques <strong>de</strong>scribed this far are only <strong>de</strong>dicated to only periodic stripes<br />

and fail to process random stripes. Application on images extracted from Terra MODIS<br />

band 33 do not display any improvement (and therefore are not illustrated). The study<br />

presented in [Rakwatin et al., 2007] was the first to tackle the issue of random stripes on<br />

MODIS. Rakwatin proposed an hybrid approach that combines histogram matching for<br />

the removal of <strong>de</strong>tector-to-<strong>de</strong>tector stripes and mirror banding, with an iterated weighted<br />

least squares facet filtering for random stripes. The facet mo<strong>de</strong>l introduced in [Haralick and<br />

Watson, 1981] <strong>de</strong>composes a given image into connected facets. A given pixel is contained<br />

in K 2 different blocks each composed of K × K pixels.<br />

For a pixel i, we <strong>de</strong>note W i,k a resolution cellcomposed of k = {1, .., K × K} neighbooring<br />

pixels. A given pixel is contained in several resolution cells as they overlap with each other.<br />

Let us <strong>de</strong>note by J ik (r, c) the signal value at row r and column c in the resolution cell<br />

W ik . Haralick’s sloped facet mo<strong>de</strong>l assumes that J ik (r, c) can be expressed as :<br />

J ik (r, c) =α ik r + β ik c + γ ik + n ik (r, c) (3.20)<br />

where α ik , β ik and γ ik represent the slope plan coefficients of the facet mo<strong>de</strong>l in the<br />

cell W ik and n ik is the noise. Denoting (−L, −L) (resp. (L, L)) the relative row-column<br />

coordinates of a cell upper-left corner (resp. lower-right corner), the values of a resolution<br />

cell slope plane coefficients can be estimated by minimizing the following quadratic energy


56 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

functional :<br />

E(α ik ,β ik ,γ ik )=<br />

L∑<br />

L∑<br />

r=−L c=−L<br />

(α ik r + β ik c + γ ik − J ik (r, c)) 2 (3.21)<br />

Least square estimates of the facet mo<strong>de</strong>l coefficients are obtained by consi<strong>de</strong>ring the<br />

partial <strong>de</strong>rivatives of the functional E with respect to α ik , β ik and γ ik and lead to the<br />

following :<br />

α ik =<br />

β ik =<br />

γ ik =<br />

3<br />

L(L + 1)(2L + 1) 2<br />

3<br />

L(L + 1)(2L + 1) 2<br />

1 ∑<br />

L<br />

(2L + 1) 2<br />

L∑<br />

r=−L c=−L<br />

L ∑<br />

r=−L<br />

L ∑<br />

c=−L<br />

r<br />

c<br />

L∑<br />

c=−L<br />

L∑<br />

r=−L<br />

J ik (r, c)<br />

J ik (r, c)<br />

J ik (r, c)<br />

(3.22)<br />

Let us consi<strong>de</strong>r the particular case where L = 1, which corresponds to blocks composed<br />

of 3 × 3 pixels. Using the least square estimates of α ik , β ik and γ ik from equation (3.22)<br />

with L = 1 , Ĵ ik (r, c) can be written as :<br />

Ĵ ik (r, c) = 1 18 [(−3r − 3c + 2)J ik(−1, −1) + (−3r + 2)J ik (−1, 0) + (−3r +3c + 2)J ik (−1, 1)<br />

+(−3c + 2)J ik (0, −1) + 2J ik (0, 0) + (3c + 2)J ik (0, 1)<br />

+(3r − 3c + 2)J ik (1, −1) + (3r + 2)J ik (1, 0) + (3r +3c + 2)J ik (1, 1)]<br />

(3.23)<br />

Each 3 × 3 pixels block is represented by a set of 9 coefficients and is associated with an<br />

estimate error <strong>de</strong>fined as :<br />

L<br />

ɛ 2 ik (r, c) =<br />

∑<br />

L∑<br />

r=−L c=−L<br />

) 2 (Ĵik (r, c) − J ik (r, c)<br />

(3.24)<br />

In the original approach introduced in [], the value of a pixel is <strong>de</strong>termined using the slope<br />

coefficients of the block which induces minimal error ɛ ik . Another alternative proposed in<br />

Li and Tam [2000] is to estimate the value at pixel i with an iterative procedure :<br />

Ĵ (n+1)<br />

ik<br />

=<br />

L∑<br />

L∑<br />

r=−L c=−L<br />

w ik (r, c)Ĵ (n)<br />

ik<br />

(3.25)<br />

The weighting coefficients w ik <strong>de</strong>pend on the estimate errors obtained for each bloack and<br />

are computed as :<br />

(<br />

)<br />

L∑ L∑<br />

w ik (r, c) =1/ ɛ ik (r, c) ɛ −1<br />

ik (r, c) (3.26)<br />

r=−L c=−L


57<br />

Figure 3.14 – (Left) Image from Terra MODIS band 33 affected mostly with random<br />

stripes (Right) Destriped result obtained after histogram matching and Haralick’s sloped<br />

facet mo<strong>de</strong>l filtering.<br />

The application of Haralick’s mo<strong>de</strong>l is not <strong>de</strong>voted to the correction of periodic stripes and<br />

Rakwatin et al. suggest using the iterative facet filtering procedure only on specific pixels.<br />

Detector-to-<strong>de</strong>tector stripes are initially corrected via the histogram matching technique.<br />

Lines acquired by noisy <strong>de</strong>tectors are then visually <strong>de</strong>tected and processed with the sloped<br />

facet mo<strong>de</strong>l. In pratice, we selected a facet window of size 3 × 3 pixels. As the number of<br />

iterations of the weighted facet filtering procedure increases (3.25), the visual impact of<br />

random stripes <strong>de</strong>creases. For the image from Terra MODIS band 33, 10 iterations were<br />

required to achieve a cosmetic improvement (see figure 3.14).<br />

3.7 Multiresolution approach<br />

3.7.1 Limitations of fourier transform<br />

The Fourier transform is a remarquable tool in signal processing. Shifting to the frequency<br />

domain offers the possibility to extract information that would otherwise be unperceptible<br />

in the spatial or temporal domain. Nevertheless, the fourier transform has a<br />

major drawback. The shift in fourier domain is inevitably followed by a loss of temporal/spatial<br />

information ; The frequency of a given event can only be known at the expense<br />

of it occuring times. A compromise can be achieved with the short-term fourier transform<br />

(STFT). It consists in limiting the computation of fourier transform to local portions of<br />

the signal, using a fixed size sliding analysing window. The Heiseinberg principle then<br />

highlights the limitations of the STFT ; For small sized windows, a good temporal localisation<br />

is achieved with approximative frequency localisation. For increasing windows size,


58 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

one eventually converges to the original fourier transform where temporal information is<br />

completely lost.<br />

3.7.2 Multiresolution analysis<br />

Multiresolution analysis, and more precisely wavelet transform constitutes a powerful<br />

alternative to fourier transform, exploited in numerous signal processing applications.<br />

Multiresolution analysis allows the <strong>de</strong>composition of a function f ∈ L 2 (R 2 ), as a sum of<br />

approximations associated with different resolution levels. In [Mallat, 2000], Mallat <strong>de</strong>fines<br />

an multiresolution approximation as a set of closed vector sub-spaces {V j } j∈Z that verify<br />

several properties among which :<br />

∀j ∈ Z, ⊂ V j+1 ⊂ V j ⊂ L 2 (R) (3.27)<br />

∀j ∈ Z,f(t) ∈ V j ⇔ f( t 2 ) ∈ V j+1 (3.28)<br />

The embedding property (3.27) of vector spaces {Vj} ensures that the approximation of<br />

the function f at resolution 2 j is obtained from an approximation at a higher resolution<br />

2 j+1 . The property (3.28) garantees that the projection of f on the space V j contains<br />

twice as much <strong>de</strong>tails as the projection on the space V j+1 . Denoting φ, a function ∈ L 2 (R)<br />

which translations {φ(t − k)} k∈Z form an orthonormal basis of the space V 0 , it is shown<br />

that the family of functions {φ j,k } k∈Z obtained as dilatations and translations of φ is an<br />

orthonormal basis of V j where :<br />

φ j,k = 1 √<br />

2 j φ ( t<br />

2 j − k )<br />

(3.29)<br />

The function φ, known as scale function or father wavelet, is dilated and translated to<br />

form an orthonormal basis of the space V j , where the orthorgonal projection of a function<br />

f <strong>de</strong>fines its approximation at the resolution 2 −j .<br />

3.7.3 Wavelet basis<br />

Going from a resolution 2 j to a lower resolution 2 j+1 , leads to a loss of information.<br />

Details visible in the approximation of resolution 2 j are lost at the resolution 2 j+1 . The<br />

embedding property of multiresolution approximations and the inclusion of vector space<br />

V j in V j−1 can be used to <strong>de</strong>fine the space of <strong>de</strong>tails W j as the orthogonal complementary<br />

of V j in V j−1 :<br />

V j−1 = V j ⊕ W j (3.30)<br />

Similarly to the vector space of approximations V 0 , it is shown that the space of <strong>de</strong>tails W 0<br />

is generated from an orthonormal basis composed of translated version {ψ(t − k)} k∈Z of a


59<br />

function ψ, the mother wavelet. The family of functions {ψ j,k } k∈Z obtained as dilatations<br />

and translations of ψ is an orthonormal basis of W j and is expressed as :<br />

ψ j,k = √ 1 ( ) t<br />

ψ<br />

2 j 2 j − k (3.31)<br />

The wavelet transform of a function f then provi<strong>de</strong>s the approximation coefficients <strong>de</strong>noted<br />

a j [k] and the <strong>de</strong>tails coefficients d j [k]. These are obtained by projecting f on the vector<br />

spaces V j and W j :<br />

a j [k] =< f, φ j,k > (3.32)<br />

d j [k] =< f, ψ j,k > (3.33)<br />

where the symbol < ., . > <strong>de</strong>notes the scalar product in L 2 (R).<br />

3.7.4 Filter banks<br />

The property (3.27) implies that the vector space V j is inclu<strong>de</strong>d in V j−1 . Then, any<br />

function in V j−1 can be written as a linear combination of a function in V j . If we consi<strong>de</strong>r<br />

a given sequence h[k], the function √ 1<br />

2<br />

φ ( t<br />

2)<br />

can be expressed with respect to the family<br />

of functions {φ(t − k)} k∈Z as :<br />

(<br />

1 t<br />

∞∑<br />

√ φ = h[k]φ(t − k) (3.34)<br />

2 2)<br />

where :<br />

k=−∞<br />

h[k] =< √ 1 φ( t ),φ(t − k) > (3.35)<br />

2 2<br />

W j being also a sub-space of V j−1 , the function √ 1<br />

2<br />

ψ ( t<br />

2)<br />

can be expressed with respect to<br />

the family of functions {φ(t − k)} k∈Z and a different sequence g[k] :<br />

where :<br />

1<br />

√<br />

2<br />

ψ<br />

( t<br />

=<br />

2)<br />

∞∑<br />

k=−∞<br />

g[k]φ(t − k) (3.36)<br />

g[k] =< √ 1 ψ( t ),φ(t − k) > (3.37)<br />

2 2<br />

Equations (3.34) and (3.36) are known as the two scale equations and are used to expresss<br />

the inter-scale relationship between the scaling function and the mother wavelet<br />

with respect to their translations and the coefficients of filters h[k] and g[k]. Through the<br />

combination of equations (3.32), (3.33), (3.34) and (3.36) it is shown that wavelet <strong>de</strong>composition<br />

and reconstruction are computed as a series of discrete convolutions with filters h<br />

and g. In the <strong>de</strong>composition stage, approximations and <strong>de</strong>tails coefficients are given with :<br />

∞∑<br />

a j+1 [k] = h[n − 2k]a j [n] =a j ⋆ h[2k] (3.38)<br />

n=−∞


60 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

d j+1 [k] =<br />

∞∑<br />

n=−∞<br />

g[n − 2k]d j [n] =d j ⋆ g[2k] (3.39)<br />

where h[k] =h[−k] and ⋆ is the convolution symbol. The approximation coefficients a j+1<br />

result from the convolution of approximations a j with the low-pass filter h, followed by<br />

a sub-sampling of factor 2. The sub-sampling operation consists in preserving only one<br />

coefficient out of two. Details coefficients d j+1 are <strong>de</strong>duced from the convolution of d j<br />

with the high-pass filter g. The reconstruction of the signal is obtained with the inverse<br />

wavelet transform :<br />

∞∑<br />

∞∑<br />

a j [k] = h[k − 2n]a j+1 [n]+ g[k − 2n]d j+1 [n]<br />

(3.40)<br />

n=−∞<br />

=ă j+1 ⋆h + ˘d j+1 ⋆g<br />

n=−∞<br />

The successive approximations a j are obtained with a convolution of signals ă j+1 and ˘d j+1<br />

with the filters h and g. ă results from an up-sampling of factor 2 of a, where zeros are<br />

inserted between succesive samples as :<br />

ă[2n] =a[n]<br />

ă[2n + 1] = 0<br />

(3.41)<br />

The <strong>de</strong>composition and illustration in filter banks is illustrated in figure . Filters ¯h, ḡ,<br />

h and g are quadrature mirror filters due to the orthogonality relation between h and g :<br />

3.7.5 2D wavelet basis<br />

g[k] = (−1) 1−n h[1 − k] (3.42)<br />

The wavelet <strong>de</strong>composition of a bidimensional signal f ∈ L 2 (R 2 ) can be seperately<br />

computed along its dimensions, using separable wavelet basis generated from the tensor<br />

product of unidimensional orthogonal wavelet basis. In the 2D case, the wavelet <strong>de</strong>composition<br />

is computed first on the lines and then on the columns (or inversely). The sequence<br />

{Vj 2}<br />

j∈Z <strong>de</strong>fined as Vj 2 = V j ⊗ V j is a separable multiresolution approximation of L 2 (R 2 ).<br />

Similarly to the unidimensional case, the space of <strong>de</strong>tails Wj<br />

2 is <strong>de</strong>fined as the orthogonal<br />

complementary of Vj 2 in Vj−1 2 : Vj−1 2 = Vj 2 ⊕ Wj 2 (3.43)<br />

An orthonormal wavelet basis of L 2 (R 2 ) can be constructed from the scaling functions<br />

φ and the mother wavelet ψ. To this prupose, let us <strong>de</strong>fine ∀(x, y) ∈ R 2 , the following<br />

wavelets :<br />

ψ 1 (x, y) =φ(x)φ(y)<br />

ψ 2 (x, y) =ψ(x)φ(y)<br />

ψ 3 (x, y) =ψ(x)ψ(y)<br />

(3.44)


61<br />

Figure 3.15 – (Left) Noisy image from Terra MODIS band 30 (Right) Wavelet <strong>de</strong>composition<br />

at two resolution levels showing that (1) stripe noise is isolated in the only<br />

horizontal <strong>de</strong>tails d 1 j (2) Wavelet coefficients associated with stripes have a magnitu<strong>de</strong> of<br />

the same or<strong>de</strong>r than those related to the image edges<br />

We then consi<strong>de</strong>r the dilated and translated versions of these wavelets <strong>de</strong>fined for k =1, 2, 3<br />

as :<br />

ψj,m,l k = 1 ( x − 2 j m<br />

2 j ψk 2 j , y − )<br />

2j l<br />

2 j (3.45)<br />

The wavelet family {ψj,m,l 1 ,ψ2 j,m,l ,ψ3 j,m,l } (m,l)∈Z 2 is an orthonormal basis of the <strong>de</strong>tails<br />

vector space Wj 2 and the family {ψ1 j,m,l ,ψ2 j,m,l ,ψ3 j,m,l } (j,m,l)∈Z3 is an orthonormal basis of<br />

L 2 (R 2 ).<br />

The separability of bidimensional wavelet basis is an interesting property for the multiresolution<br />

analysis of images. In fact, the wavelet family {ψj,m,l 1 ,ψ2 j,m,l ,ψ3 j,m,l } (m,l)∈Z 2 allows<br />

the extraction of <strong>de</strong>tails in the horizontal, vertical and diagonal directions. This particular<br />

feature is in<strong>de</strong>ed the main motivation behind the use of wavelet analysis for the striping issue.<br />

Wavelet <strong>de</strong>composition and reconstruction of a function f ∈ L 2 (R 2 ) is computed with<br />

separable bidimensional convolutions. Using the conjugate mirror filters h and g of the<br />

mother wavelet ψ, wavelet coefficients at level 2 j+1 are obtained from the approximation<br />

at level 2 j with :<br />

a j+1 [m, l] =a j ⋆ ¯h[m]¯h[l]<br />

d 1 j+1[m, l] =a j ⋆ ¯h[m]ḡ[l]<br />

d 2 j+1[m, l] =a j ⋆ ḡ[m]¯h[l]<br />

d 3 j+1[m, l] =a j ⋆ ḡ[m]ḡ[l]<br />

(3.46)<br />

The bidimensional wavelet <strong>de</strong>composition of an image contaminated with stripe noise is<br />

illustrated in figure 3.15.


62 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

3.7.6 Destriping with wavelet coefficient thresholding<br />

As exposed in the previous section, the discrete dyadic wavelet transfom offers a sparse<br />

representation of a signal, well suited for many image processing applications. In the case<br />

of <strong>de</strong>noising, a major limitation not mentionned this far is the translation invariance condition,<br />

not satisfied by the discrete dyadic wavelet transfom. This drawback does not affect<br />

the wavelet <strong>de</strong>composition in itself, however, if the wavelet coefficients are modified, the<br />

wavelet reconstruction introduces artifacts along the image discontinuities. These artifacts,<br />

visible as pseudo-Gibbs oscillations come from the <strong>de</strong>cimation of wavelet coefficients between<br />

successive <strong>de</strong>composition levels. To avoid these effects, we will rely on the stationary<br />

wavelet transform [Nason and Silverman, 1995]. The multiscale <strong>de</strong>composition of an image<br />

provi<strong>de</strong>s wavelet coefficients with a magnitu<strong>de</strong> <strong>de</strong>pen<strong>de</strong>nt on the local regularity of the<br />

signal in a given resolution. Assuming a mo<strong>de</strong>rate amount of noise in the observed signal,<br />

strong wavelet coefficients will be distributed along the edges of an image while weak coefficients<br />

will be located in homogenenous areas. The noise can then be reduced from the<br />

image by consi<strong>de</strong>ring only specific wavelet coefficients in the reconstruction process. The<br />

seminal work of [Donoho and Johnstone, 1994] <strong>de</strong>scribes two <strong>de</strong>noising techniques based<br />

on the thresholding of wavelet coefficients. A first intuitive approach consists in setting<br />

to zero all wavelet coefficients with an amplitu<strong>de</strong> lower than a fixed threshold λ. This<br />

operation is known as hard thresholding and is represented by a function S hard that takes<br />

the following form :<br />

{ 0 if d<br />

Sλ hard (d k k<br />

j )=<br />

j ≤ λ<br />

d k j if d k j >λ (3.47)<br />

where {d k j } k=1,2,3 are <strong>de</strong>tail coefficients in the horizontal (k = 1), diagonal (k = 2) and<br />

vertical (k = 3) directions, obtained at the resolution 2 j .<br />

An alternative approach in the selection of wavelet coefficients is based on soft thresholding<br />

or wavelet shrinkage ; Coefficients that exceed a threshold are attenuated by the value of<br />

the threshold. The resulting thresholding function S soft<br />

λ<br />

is given by :<br />

{<br />

S soft<br />

0 if d<br />

λ<br />

(d k k<br />

j )=<br />

j ≤ λ<br />

d k j − sign(dk j )λ if dk j >λ (3.48)<br />

The quality of the reconstructed signal highly <strong>de</strong>pends on the choice of the threshold which<br />

can be the same for all resolutions or vary from one resolution to the other. A common<br />

methodology relies on the universal thresholding of [Donoho and Johnstone, 1994] where<br />

the value of the threshold is fixed as :<br />

λ =ˆσ √ 2 log(M × N) (3.49)<br />

where M × N is the number of pixels in the image and ˆσ is an estimation of the noise<br />

variance, <strong>de</strong>termined from the <strong>de</strong>tail coefficients of the highest resolution with :<br />

λ<br />

ˆσ = median{dk 0 } k=1,2,3<br />

0.6745<br />

(3.50)


63<br />

The highest resolution is used for the estimation of σ because its wavelet coefficients are<br />

mostly related to the noisy component of the signal.<br />

Going back to the striping issue, let us un<strong>de</strong>rscore an important point. Most <strong>de</strong>noising<br />

technique presented in the litterature and based on wavelet coefficients thresholding are<br />

<strong>de</strong>votd to the elimination of isotropic noise (gaussian, poissonian, speckle noise). Although<br />

the basic principle of wavelet thresholding remains valid for our study, specific aspects such<br />

as the choice of the threshold have to be revised and adapted for the case of stripe noise.<br />

Destriping via wavelet thresholding have already been used on MODIS in [Yang et al.,<br />

2003]. However, the proposed methodology is based on a soft thresholding applied to <strong>de</strong>tail<br />

coefficients of all directions. A refinement of this approach was introduced in [Torres and<br />

Infante, 2001] for the <strong>de</strong>striping of Landsat MSS images. This technique, which we apply<br />

here to MODIS data, exploits the unidirectional signature of striping and its impact on the<br />

multiresolution <strong>de</strong>composition. As illustrated in figure 3.15, the presence of stripe noise<br />

affects only the horizontal component of the image. It is then reasonnable to restrain the<br />

manipulation of wavelet coefficients to the horizontal <strong>de</strong>tails d 1 j . Figure 3.15 also indicates<br />

that unlike white gaussian noise, striping translates as wavelet coefficients with very high<br />

intensity. The amplitu<strong>de</strong> of wavelet coefficients associated with the image edges is actually<br />

dominated by stripe noise and application of classic thresholding strategies cannot be<br />

consi<strong>de</strong>red in our case. The strategy <strong>de</strong>veloped in [Torres and Infante, 2001] consists in<br />

eliminating all horizontal wavelet coefficients of the m highest resolution levels, m being<br />

<strong>de</strong>termined heuristically. This procedure is equivalent to a hard thresholding where all<br />

horizontal coefficients are set to zero. Its thresholding function is :<br />

{ 0 if k = 1 and 1 ≤ j ≤ m<br />

S λ (d k j )=<br />

otherwise<br />

d k j<br />

(3.51)<br />

The <strong>de</strong>striping quality <strong>de</strong>pends on the parameter m. When m is small, only small scale<br />

<strong>de</strong>tails of the striping effect are removed. If m takes high values, the thresholding function<br />

3.51 removes the horizontal low-frequency component of the image and introduces strong<br />

blurring. The visual analysis of successive approximations shows that the impact of striping<br />

tends to diminish through lower resolutions. The hard thresholding should then be limited<br />

to resolutions where the stripe noise signature is still persistent (see figure 3.17). Wavelet<br />

<strong>de</strong>striping through hard thresholding can provi<strong>de</strong> good visual results <strong>de</strong>pending on the<br />

manipulation of horizontal coefficients. However it inevitably eliminates <strong>de</strong>tails related to<br />

the image sharp structures.<br />

3.8 Assessing <strong>de</strong>striping quality<br />

In the previous sections, several approaches have been <strong>de</strong>scribed and applied to MODIS<br />

data severely affected with stripe noise. Visual examination of <strong>de</strong>striped results indicates<br />

that equalization methods, either based on statistical or radiometric consi<strong>de</strong>rations, fail<br />

to completely remove the noise in that a consi<strong>de</strong>rable amount of residual stripes are still


64 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

Figure 3.16 – (Left) Noisy image from Terra MODIS band 30 (Right) Destriped result<br />

setting to zero vertical <strong>de</strong>tails d 1 1 , d1 2 and d1 3 (m=3) prior to wavelet reconstruction<br />

Image distortion ID<br />

0.93<br />

0.925<br />

0.92<br />

0.915<br />

0.91<br />

0.905<br />

0.9<br />

0.895<br />

0.89<br />

0.885<br />

0.88<br />

ID<br />

NR<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1 2 3 4 5 6 7 8 0 1000<br />

m (db2)<br />

Noise reduction NR<br />

Image distortion ID<br />

0.935<br />

0.93<br />

0.925<br />

0.92<br />

0.915<br />

0.91<br />

0.905<br />

ID<br />

NR<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

0.9<br />

1000<br />

0.895<br />

1 2 3 4 5 6 7 8 0<br />

m (db4)<br />

Noise reduction NR<br />

Figure 3.17 – Image distortion (ID)* and Noise reduction (NR)* as a function of the<br />

thresholding parameter m for Terra MODIS band 27. The selected wavelets are (Left)<br />

db2 and (Right) db4. The breaking points visible in the NR curves at m = 5 for the<br />

wavelet db 2 and m = 3 for db 4 , translate the reduction of striping from higher to lower<br />

resolution scales. Depending on the number of vanishing moments of the selected wavelet,<br />

the striping will be concentrated in the wavelet coefficients of higher resolutions and its<br />

impact on vertical wavelet coefficients eventually fa<strong>de</strong>s below at given scale, hence the<br />

constant NR in lower resolutions.<br />

visible. This is due to non-linear effects and random stripes, mostly present on Terra MO-<br />

DIS. Despite uncomplete <strong>de</strong>striping, equalization techniques are often prefered for their<br />

ability to preserve the signal radiometry.<br />

On the other hand, filtering methods (frequency filtering, facet filtering and wavelet thre-


65<br />

sholding) can be tuned to regulate the amount of distortion introduced in the restored<br />

image. A rigourous comparison of these methods requires additional criteria other than<br />

visual interpretation. In this section, we recall the <strong>de</strong>finition of several qualitative in<strong>de</strong>xes<br />

used to evaluate the <strong>de</strong>striping quality.<br />

3.8.1 Noise Reduction Ratio and Image Distortion<br />

The spectral analysis conducted in section 3.5.1, takes advantage of the unidirectional<br />

aspect of striping to highlight its spectral signature. Periodic stripes related to <strong>de</strong>tectorto-<strong>de</strong>tector<br />

and mirror banding introduces peaks, clearly visible in the ensemble averaged<br />

power spectrum down the columns. A reliable <strong>de</strong>striping should reduce the stripe peaks<br />

without interfering with the global shape of the original signal’s power spectrum. Let us<br />

<strong>de</strong>note P 0 -resp. P 1 - the power spectrum obtained as an average of the periodograms of<br />

the noisy image -resp. the <strong>de</strong>striped image- columns. We <strong>de</strong>fine :<br />

N 0 =<br />

∑<br />

P 0 (f)<br />

f∈BW N<br />

N 1 =<br />

∑<br />

P 1 (f)<br />

f∈BW N<br />

(3.52)<br />

where BW N is the noisy part of the frequency spectrum (f ∈ BW N corresponds to f ∈<br />

{0.1, 0.2, 0.3, 0.4, 0.5} in the case of <strong>de</strong>tector-to-<strong>de</strong>tector stripes). N 0 and N 1 contain the<br />

spectral components of stripes in both the noisy image and the restored one. The Noise<br />

Reduction ratio (NR) is then <strong>de</strong>fined as :<br />

NR = N 0<br />

N 1<br />

(3.53)<br />

The NR in<strong>de</strong>x measures the attenuation of the frequency peaks in the power spectrum and<br />

hence can only be used for images affected with periodic stripes. Furthermore, evaluation<br />

of a <strong>de</strong>striping techniques with the NR in<strong>de</strong>x has to be imperatively completed with a<br />

distortion in<strong>de</strong>x that quantifies the blur introduced in the <strong>de</strong>striped result. Denoting :<br />

S 0 =<br />

∑<br />

P 0 (f)<br />

f∈BW S<br />

S 1 =<br />

∑<br />

P 1 (f)<br />

f∈BW S<br />

(3.54)<br />

where BW S is the noise-free portion of the spectrum, the image distortion in<strong>de</strong>x used in<br />

[Pan and Chang, 1992] is <strong>de</strong>fined as :<br />

ID = N 0<br />

N 1<br />

(3.55)


66 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

The previous <strong>de</strong>finition contrains the computation of the ID in<strong>de</strong>x to the only portion of<br />

the frequency spectrum not affected with noise and, therefore tends to over estimate its<br />

value. In addition, the sums in the terms N 0 and N 1 are dominated by the amplitu<strong>de</strong> of<br />

low-frequency power spectrum, which are approximatively the same for both noisy and<br />

<strong>de</strong>striped images. This results in ID values very close to unity even in the presence of blur.<br />

An alternative option is to focus on the periodograms of lines. Denoting Q 0 -resp. Q 1 -,<br />

the ensemble averaged power spectrum down the lines of the noisy -resp. <strong>de</strong>striped imagewe<br />

<strong>de</strong>fine :<br />

S 0 =<br />

∑<br />

Q 0 (f)<br />

S 1 =<br />

f∈BW<br />

∑<br />

f∈BW<br />

Q 1 (f)<br />

(3.56)<br />

where BW represents the entire frequency spectrum (BW = BW N + BW S ). The Image<br />

Distortion (ID) in<strong>de</strong>x is then <strong>de</strong>fined as :<br />

ID =1−<br />

1<br />

card(BW)<br />

∑<br />

f∈BW<br />

3.8.2 Radiometric Improvement Factors<br />

|Q 0 (f) − Q 1 (f)|<br />

Q 0 (f)<br />

(3.57)<br />

Results achieved with standard <strong>de</strong>striping techniques display consi<strong>de</strong>rable residual<br />

stripes <strong>de</strong>spite the reduction of the spectral component related to periodic noise. In fact,<br />

both in<strong>de</strong>xes NR and ID are <strong>de</strong>fined in the fourier domain and cannot measure the spatial<br />

regularity of the restored image. In the spatial domain, periodic and random stripes<br />

translate as strong fluctuations of the image values along the vertical axis. Let us <strong>de</strong>note<br />

m Is [j] and mÎ[j] the unidimensional signals associated with the mean values of lines j in<br />

the noisy image I s and the <strong>de</strong>striped image Î. The <strong>de</strong>striping quality in the spatial domain<br />

can then be <strong>de</strong>termined by comparing the values of m Is [j] and mÎ[j]. To this purpose, let<br />

us consi<strong>de</strong>r a reference signal I ref with a smooth cross-track profile. I ref is obtained from<br />

the noisy image I s with a low-pass filter given by :<br />

H(u, v) =exp<br />

(− u2 + v 2 )<br />

σ 2 (3.58)<br />

Since the filter H is only used to provi<strong>de</strong> a reference regularized cross-track profile, it can<br />

also be replaced by a simple averaging filter in the spatial domain. Denoting :<br />

d s [j] =m Is [j] − m Iref [j]<br />

d e [j] =mÎ[j] − m Iref [j]<br />

the first radiometric improvement factor is <strong>de</strong>fined as :<br />

(∑ )<br />

j<br />

IF 1 = 10log d2 s[j]<br />

10 ∑<br />

j d2 e[j]<br />

(3.59)<br />

(3.60)


67<br />

A secondary in<strong>de</strong>x, in<strong>de</strong>pen<strong>de</strong>nt of I ref , consi<strong>de</strong>rs the radiometric errors :<br />

∆ s [j] =m Is [j] − m Isj−1[j]<br />

∆ e [j] =mÎ[j] − mÎ[j − 1]<br />

The second or<strong>de</strong>r radiometric improvement factor is then given by :<br />

(∑ )<br />

j<br />

IF 2 = 10log ∆2 s[j]<br />

10 ∑<br />

j ∆2 e[j]<br />

(3.61)<br />

(3.62)<br />

3.8.3 Conclusion<br />

Destriping techniques presented in this chapter can be classified in two major categories.<br />

The first class is composed of statistical or radiometric equalization techniques such<br />

as moment matching, histogram matching and the OFOV method. All these approaches<br />

rely on the acquisition principle of pushbroom imaging systems to equalize the response of<br />

individual <strong>de</strong>tectors. Their application to MODIS data illustrates a significant amount of<br />

residual stripes. Figure 3.20 shows that even after <strong>de</strong>striping, rapid fluctuations can still<br />

be seen in the cross-track profiles of the <strong>de</strong>noised data. The analysis of column ensemble<br />

averaged power spectrums (figure ) indicates that the amplitu<strong>de</strong> of peaks related to periodic<br />

stripes is reduced. To improve visual analysis of noise reduction, power spectrums<br />

are plotted with a logarithmic scale, as a function of normalized frequency.<br />

Equalization techniques do not account for non linear effects which are predominant in<br />

MODIS data. Residual stripes in the images affect its power spectrum in two ways. Strong<br />

non linear effects result in poor <strong>de</strong>tector-to-<strong>de</strong>tecotr stripes reduction and appear as persisting<br />

peaks with a lower magnitu<strong>de</strong> located at the same frequencies of 0.1, 0.2, 0.3, 0.4 and<br />

0.5 pixels per cycles. Random stripes translate as strong variations of the power spectrum<br />

in the high frequency range. Despite residual stripes, equalization techniques maintain the<br />

image distortion in<strong>de</strong>x close to 1.<br />

Alternative approaches to equalization are based on filtering techniques. The periodicity<br />

of stripes can be exploited by a band-pass filter that cuts-off stripe related frequencies.<br />

This can be achieved by placing narrow wells in the fourier response of the filter, centered<br />

on the coordinates associated with periodic stripes. As seen in figure e, periodic peaks<br />

are properly removed from the power spectrums. However, the overall shape of the column<br />

power spectrum is strongly distorted. This is even more visible on the line ensemble<br />

averaged power spectrum, which should be i<strong>de</strong>ntical for both noisy and <strong>de</strong>striped images.<br />

Wavelet analysis also offers an interesting perspective on the striping issue. The multdirectional<br />

representation provi<strong>de</strong>d by wavelet transform is well suited to process striping<br />

since only horizontal wavelet coefficients are contaminated with stripes. Setting to zero the<br />

wavelet coefficients of the m highest resolution scales prior to wavelet reconstruction can<br />

reduce the visual impact of stripe noise. The choice of the parameter m is fixed according<br />

to the NR and ID in<strong>de</strong>xes values and highly <strong>de</strong>pends on the number of vanishing moments


68 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

of the selected wavelet. If the analysing wavelet has enough vanishing moments, regular<br />

areas will have small wavelet coefficients and the stripe noise will be isolated in the highest<br />

resolution scales (figure 3.17b). On the contrary, if the number of vanishing moments is<br />

weak, stripe noise will also affect wavelet coefficients of lower resolutions (figure 3.17a)<br />

and therefore, wavelet <strong>de</strong>striping provi<strong>de</strong>s images with a weak ID in<strong>de</strong>x.<br />

The facet filtering mo<strong>de</strong>l proposed in [Rakwatin et al., 2007] is used over the histogram<br />

matching techniques to process random stripes. Due to its hybrid aspect, it is not compared<br />

to other techniques.<br />

Filtering techniques provi<strong>de</strong> results that are visually better than equalization methods because<br />

the removal of stripes can be tuned with σ for band-pass frequency filtering or m for<br />

wavelet thresholding. However, blurring and ringing artifacts introduced in the corrected<br />

signal discards any further quantitative analysis based on radiometric values.<br />

The limitations of standard techniques exposed in this chapter serve as a basis for<br />

the <strong>de</strong>velopment of a robust <strong>de</strong>striping technique. An optimal <strong>de</strong>striping algorithm should<br />

satisfy the following requirements :<br />

- Complete removal of stripe noise, whether periodic or random<br />

- Minimization of the distortion introduced in the restored image<br />

From the <strong>de</strong>finition of the NR and ID in<strong>de</strong>xes, an optimal <strong>de</strong>striping algorithm increases<br />

NR while leaving ID close to 1. To achieve such results, we explore in the next chapter<br />

the striping issue from a variational perspective.


Figure 3.18 – Destriped results on Terra MODIS band 30 (TL) Original image (TC)<br />

Moment matching (CL) Histogram matching (CR) IFOV method (BL) Frequency filtering<br />

(BR) Wavelet thresholding. Reflectances over oceanic regions are highlighted to<br />

emphasize the presence of residual stripes.<br />

69


70 3. Standard <strong>de</strong>striping techniques and application to MODIS<br />

13500<br />

13500<br />

Mean value<br />

13000<br />

12500<br />

Mean value<br />

13000<br />

12500<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

13500<br />

13500<br />

Mean value<br />

13000<br />

12500<br />

Mean value<br />

13000<br />

12500<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

13500<br />

13500<br />

Mean value<br />

13000<br />

12500<br />

Mean value<br />

13000<br />

12500<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

12000<br />

0 100 200 300 400 500<br />

Line Number<br />

Figure 3.19 – Cross track profiles computed for (TL) Original image (TR) Moment<br />

matching (CL) Histogram matching (CR) IFOV method (BL) Frequency filtering (BL)<br />

Wavelet thresholding.


71<br />

10<br />

10<br />

9<br />

9<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

10<br />

10<br />

9<br />

9<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

10<br />

10<br />

9<br />

9<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

Power spectrum<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

Figure 3.20 – Ensemble averaged power spectrums downn the columns for (TL) Original<br />

image (TR) Moment matching (CL) Histogram matching (CR) IFOV method (BL)<br />

Frequency filtering (BR) Wavelet thresholding.


Table 3.4 – Noise Reduction (NR), Image distortion (ID), radiometric Improvement Factors IF 1 and<br />

IF 2 for the Terra MODIS band 27, 30 and 33<br />

– Terra band 27 Terra band 30 Terra band 33<br />

in<strong>de</strong>x NR ID IF 1 IF 2 NR ID IF 1 IF 2 NR ID IF 1 IF 2<br />

moment matching 45.8895 0.6305 7.3085 11.5732 9.4280 0.9963 6.2656 6.6818 – 0.9995 0.1637 0.1724<br />

histogram (IMAPP) 101.0629 0.9459 8.5463 20.1003 15.3316 0.8816 5.7455 12.8977 – 0.9773 4.8320 11.1566<br />

IFOV 100.5751 0.9692 5.8563 13.7338 5.4195 0.8991 4.6190 5.7684 – 0.9388 -0.1046 0.5466<br />

Frequency filtering 1784 0.9868 9.8866 21.5316 25.5615 0.9884 7.3092 14.8566 – – – –<br />

Wavelet thresholding 5749 0.9326 9.8649 38.3728 48.9917 0.9050 6.2421 18.6632 – 0.8892 4.7349 17.0134<br />

Table 3.5 – Noise Reduction (NR), Image distortion (ID), radiometric Improvement Factors IF 1 and<br />

IF 2 for the Aqua MODIS band 27, 30 and 36<br />

– Aqua band 27 Aqua band 30 Aqua band 36<br />

in<strong>de</strong>x NR ID IF 1 IF 2 NR ID IF 1 IF 2 NR ID IF 1 IF 2<br />

moment matching 12.643 0.984 10.4393 12.0686 4.2363 0.9991 6.4180 12.6573 1.90 0.9997 3.1844 2.3252<br />

histogram (IMAPP) 33.8517 0.9828 11.5966 18.1798 4.4847 0.9839 5.5991 12.6135 1.9593 0.9957 3.2537 2.4242<br />

IFOV 1 12.1310 0.9890 9.3335 11.9212 1.8756 0.9646 -3.6762 8.9016 1.1396 0.8627 -1.8263 2.3097<br />

frequency filtering 510.5828 0.9854 10.3039 24.8529 7.8478 0.9874 6.35 17.8977 3.2048 0.9868 0.7912 10.6393<br />

wavelet thresholding 1131 0.9280 9.6601 25.4008 13.4574 0.9288 6.1015 17.9192 5.2528 0.8918 -0.1918 11.3379<br />

72 3. Standard <strong>de</strong>striping techniques and application to MODIS


73<br />

Chapitre 4<br />

A Variational approach for the<br />

<strong>de</strong>striping issue<br />

4.1 PDEs and variational methods in image processing<br />

Many disciplines from physical sciences such as thermodynamics and fluids mechanics<br />

have inspired the application of Partial Differential Equations (PDEs) in the field<br />

of image processing. Early <strong>de</strong>noising methods are mostly based on smoothing operations,<br />

either computed directly in the spatial domain via a convolution with a filter, or in the<br />

frequency domain. Assuming the noise to be contained in the high frequencies of the observed<br />

signal, a common restoration technique consists in convolving the noisy image with<br />

a linear operator. Denoting u 0 the noisy signal <strong>de</strong>fined in a boun<strong>de</strong>d domain Ω of R 2 , an<br />

estimate of the true image u is obtained by consi<strong>de</strong>ring the scale-space generated by u 0<br />

as :<br />

∫<br />

u(x, y, t) = G(x − ξ, y − η, t)u 0 (ξ, η)dxdy, (x, y) ∈ Ω (4.1)<br />

Ω<br />

Typically, the operator G is a bidimensional gaussian kernel :<br />

G(x, y, t) = 1 ( −(x 2<br />

4πt exp + y 2 )<br />

)<br />

4t<br />

(4.2)<br />

where the variance σ 2 =2t of the gaussian operator controls the <strong>de</strong>gree of smoothing in<br />

the restored image u. The work of [Koen<strong>de</strong>rink, 1984] reformulates the convolution with<br />

a gaussian kernel as a diffusion process where the value of a pixel can be expressed with<br />

respect to a neighboorhood which size is <strong>de</strong>termined by the variance σ 2 of the operator<br />

G. The noise free image u can be seen as the solution of a simple parabolic PDE :<br />

∂u<br />

∂t (x, y, t) u<br />

=∂2 ∂ 2 x (x, y, u<br />

t)+∂2 ∂ 2 (x, y, t)<br />

y<br />

u(x, y, 0) = u 0 (x, y)<br />

(4.3)


74 4. A Variational approach for the <strong>de</strong>striping issue<br />

The previous PDE, also known as the heat equation, translates an isotropic diffusion process<br />

where all directions are smoothed i<strong>de</strong>ntically. The isotropic property of the heat<br />

equation is a major drawback for <strong>de</strong>noising applications. In fact, noise in homogeneous<br />

regions is effectively removed, however in heterogeneous areas, sharp discontinuities related<br />

to edges are processed similarly to noise. The corresponding high gradient magnitu<strong>de</strong><br />

is therefore reduced which results in a significant loss of contrast in the restored image.<br />

This limitation was first explored by [Perona and Malik, 1990] by means of anisotropic<br />

diffusion mo<strong>de</strong>l, where the intensity of the diffusion process is <strong>de</strong>pen<strong>de</strong>nt on the local<br />

gradient value. The <strong>de</strong>gree of smoothing is inversely proportional to the gradient value<br />

so that homogeneous areas are strongly diffused while edge-discontinuities are preserved.<br />

The anisotropic diffusion PDE proposed by Perona and Malik can be formulated as :<br />

∂u<br />

∂t (x, y, t) =div (φ(|∇u(x, y, t)|)∇u(x, y, t)) (4.4)<br />

where div is the divergence operator, ∇ <strong>de</strong>signates the spatial gradient operator and ψ is<br />

a <strong>de</strong>creasing function. The <strong>de</strong>creasing property of φ ensures that the smoothing process is<br />

weaker in regions with strong gradient values, and stronger in flat areas. Common choices<br />

for φ inclu<strong>de</strong> exponentially <strong>de</strong>creasing functions as :<br />

φ(|∇u|) =exp(− |∇u|2<br />

k 2 ) (4.5)<br />

If φ is a constant function, the PDE equation (4.4) of Perona and Malik is reduced to :<br />

∂u<br />

∂t (x, y, t) =div (∇u(x, y, t)) (4.6)<br />

which is another formulation of the isotropic heat equation. The anisotropic diffusion (4.4)<br />

also faces limitations. The presence of noise introduces strong oscillations which can be<br />

consi<strong>de</strong>red as edges and be subsenquently preserved. To circumvent this issue, alternative<br />

techniques were proposed in<strong>de</strong>pen<strong>de</strong>ntly in [Catté et al., 1992] and [Nitzberg and Shiota,<br />

1992]. The gradient of the image is replaced by a smoothed version and the PDE (4.4)<br />

becomes :<br />

∂u<br />

∂t (x, y, t) =div (φ(|∇(G σ ∗ u(x, y, t))|)∇u(x, y, t)) (4.7)<br />

where G σ is gaussian operator with variance σ. Further research in the field of anisotropic<br />

diffusion introduced in [Alvarez et al., 1992] resulted in a non linear PDE :<br />

∂u<br />

∂t (x, y, t) =g (|∇(G σ ∗ u(x, y, t))|) |∇u(x, y, t)| div<br />

( ∇u(x, y, t)<br />

|∇u(x, y, t)|<br />

)<br />

(4.8)<br />

where g is a <strong>de</strong>creasing function that tends to 0 when ∇u tends to infinity. The( second )<br />

<strong>de</strong>rivative of u in the direction orthogonal to the gra<strong>de</strong>nt ∇u is the term |∇u|div ∇u<br />

|∇u|<br />

,<br />

and coinci<strong>de</strong>s with the image level lines. The intensity of the diffusion process is regulated


75<br />

Figure 4.1 – (Left) Noisy image from Terra MODIS band 30 (Center) Application of the<br />

heat equation (ID=0.41) (Right) Application of the Perona-Malik algorithm (ID=0.84).<br />

by the term g (|∇G σ ∗ u|). In homogeneous areas, strong values of g (|∇G σ ∗ u|) induce an<br />

anisotropic diffusion in the direction orthogonal to the gradient. Along the edges of the<br />

images (strong gradient values), the weak weighting due to the <strong>de</strong>screasing property of g<br />

disables the diffusion process.<br />

In [Nordstrom, 1990], the original PDE of Perona and Malik is formulated in a way that<br />

forces the estimated solution to be close to the observed image u 0 . The resulting PDE<br />

takes the following form :<br />

∂u<br />

∂t (x, y, t) − div(φ(|G σ ∗∇u(x, y, t)|)∇u(x, y, t)) = β (u(x, y, t) − u 0 (x, y)) (4.9)<br />

where the coefficient β regulates the fi<strong>de</strong>lity of the estimated solution to the orignal noisy<br />

sinal u 0 .<br />

In the unifying framework proposed in [Deriche and Faugeras, 1995], the authors show<br />

that image restoration based on the resolution of PDEs can be reformulated as the minimization<br />

of energy functionals. Such formulation is a<strong>de</strong>quate to many ill-posed inverse<br />

problems encountered in computer vision applications such as <strong>de</strong>noising, <strong>de</strong>convolution or<br />

inpainting. According to Hadamard’s <strong>de</strong>finition, a mathematical problem can be consi<strong>de</strong>red<br />

as ill-posed if at least one of the following conditions is not satisfied :<br />

- A solution exists<br />

- If a solution exists, it is unique<br />

- The solution is stable ; A small perturbation in the observed data induces a small perturbation<br />

in the estimated solution.<br />

Ill-posed inverse problems are systematically regularized to impose well-posedness and<br />

are tackled un<strong>de</strong>r the assumption of a specific image formation mo<strong>de</strong>l. Let us <strong>de</strong>note by<br />

K a linear transformation that accounts for the sensor multiple <strong>de</strong>gradations (diffraction,<br />

<strong>de</strong>focalisation, mouvement blur). The observed image f resulting from the acquisition


76 4. A Variational approach for the <strong>de</strong>striping issue<br />

process is given by :<br />

f = Ku + n (4.10)<br />

where n is generally assumed to be a zero mean gaussian noise. Equation (4.10) is refered to<br />

as the direct problem and the inverse problem then consists in <strong>de</strong>termining the real image<br />

u from the noisy observation f. This can not be achieved through the inversion of the<br />

operator K, since existence and stability of the inverse of K are not garanteed. Instead,<br />

a stable approximation of K −1 can be consi<strong>de</strong>red by introducing a priori regularizing<br />

information to enlarge the space of admissible solutions. The estimation of the true image<br />

u from (4.10) reduces to an optimization problem based on the minimization of an energy<br />

or cost function as <strong>de</strong>scribed in (Deriche and Faugeras, 1995) :<br />

E(u) =E 1 (u)+λE 2 (u) (4.11)<br />

The term E 1 (u) translates the fi<strong>de</strong>lity of the solution to the observed image while E 2 (u)<br />

is a regularizing term that smoothes the estimated solution and is often related to the<br />

gradient of u. The parameter λ balances the tra<strong>de</strong>off between a good fit to the observed<br />

image and a regularized solution. Typically,<br />

E 1 (u) =‖f − Ku‖ 2<br />

∫<br />

E 2 (u) = φ(|∇u|)<br />

Ω<br />

(4.12)<br />

In the particular case (4.12), an estimate of the true image û is obtained as the solution<br />

to the following optimization problem :<br />

∫<br />

û = argmin u ‖u − f‖ 2 + λ φ(|∇u|)dΩ (4.13)<br />

The selection of φ(|∇u|) =|∇u| 2 corresponds to the well-known Tikhonov regularization.<br />

4.2 Rudin, Osher and Fatemi Mo<strong>de</strong>l<br />

Among many variational based image restoration procedures, the total variation regularization<br />

introduced in the seminal work of [Rudin et al., 1992] provi<strong>de</strong>s efficient results<br />

for image <strong>de</strong>noising because it is well suited for piecewise constant signals and therefore<br />

allows the retrieval of sharp discontinuities. Total variation regularization was first proposed<br />

in the context of image <strong>de</strong>noising, and have been successively applied to a variety<br />

of fields including image recovery [Blomgren et al., 1997], image interpolation [Guichard<br />

and Malgouyres, 1998], image <strong>de</strong>composition (see section 4.3), scale estimation [Luo et al.,<br />

2007], image inpainting [Chan et al., 2005] and super resolution [D.Babacan et al., 2008].<br />

Let us recall the image formation mo<strong>de</strong>l where we assume the linear operator K to be<br />

unity :<br />

f = u + n (4.14)<br />

Ω


77<br />

The goal is to find the true image u from the noisy observation f assuming n to be white<br />

gaussian noise with a known variance σ 2 . This ill-posed problem can be regularized by<br />

assuming that the true image u belongs to the space of functions of boun<strong>de</strong>d variations,<br />

<strong>de</strong>noted BV and <strong>de</strong>fined as :<br />

BV (Ω) =<br />

{<br />

u ∈ L 1 (Ω)|sup ϕ∈C 1 c ,|ϕ|≤1<br />

(∫<br />

) }<br />

u divϕ < ∞<br />

(4.15)<br />

The total variation of a signal u ∈ BV (Ω) is equivalent to the L 1 norm of its gradient<br />

norm and is given by :<br />

{∫<br />

}<br />

TV(u) =sup u divϕ, ϕ ∈Cc 1 , |ϕ| ≤ 1<br />

(4.16)<br />

Hereafter, we consi<strong>de</strong>r u to be in BV (Ω) ∩C 1 , and the total variation simplifies into :<br />

∫<br />

TV(u) = |∇u|dΩ (4.17)<br />

To retrieve u from f, Rudin, Osher and Fatemi propose to solve a constrained optimisation<br />

problem :<br />

minimize<br />

TV(u)<br />

subject to ∫ Ω f = ∫ Ω u, and ‖u − f‖2 = σ 2 (4.18)<br />

In the previous formulation the equality constraint is not convex and can be replaced by<br />

an inequality constraint as :<br />

Ω<br />

minimize<br />

TV(u)<br />

subject to ∫ Ω f = ∫ Ω u, and ‖u − f‖2 ≤ σ 2 (4.19)<br />

Chambolle and Lions have shown in [Chambolle and Lions, 1997] that if ‖f− ∫ Ω fdΩ‖2 >σ 2<br />

then problems (4.18) and (4.19) are equivalent. (4.19) can then be reformulated as the<br />

minimization of :<br />

E(u) =λ‖u − f‖ 2 + TV(u) (4.20)<br />

where the lagrangian multiplier λ> 0 regulates the compromise between the fi<strong>de</strong>lity term<br />

and the regularizing term. For a given value of λ, Karush-Kuhn-Tucker conditions ensure<br />

the equivalence between (4.19) and (4.20). The energy functional (4.20) will be refered to<br />

hereafter as the ROF mo<strong>de</strong>l. The minimization of ROF mo<strong>de</strong>l is obtained via its Euler-<br />

Lagrange equation :<br />

⎛<br />

⎛<br />

−2λ(u − f)+ ∂<br />

∂x<br />

⎜<br />

⎝<br />

√ ( ∂u<br />

∂x<br />

∂u<br />

∂x<br />

) 2<br />

+<br />

(<br />

∂u<br />

∂y<br />

⎞<br />

⎟<br />

) 2 ⎠ + ∂ ⎜<br />

∂y ⎝<br />

√ ( ∂u<br />

∂x<br />

∂u<br />

∂x<br />

) 2<br />

+<br />

(<br />

∂u<br />

∂y<br />

⎞<br />

⎟<br />

) 2 ⎠ =0 (4.21)


78 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.2 – (Left) Noisy image from Terra MODIS band 30 (Right) Denoising with<br />

TV regularization (ID=0.51)<br />

which can take the simple form :<br />

−2λ(u − f) + div<br />

( ) ∇u<br />

= 0 (4.22)<br />

|∇u|<br />

From a computational point of view, the non-differentiability of the term |∇u| for ∇u =0<br />

is problematic. In [Acar and Vogel, 1994], the authors suggest a smoothed version of the<br />

energy functional where the total variation norm is relaxed with a small positive parameter<br />

ɛ. The original ROF energy functional (4.20) becomes :<br />

∫<br />

√<br />

E ɛ (u) =λ‖u − f‖ 2 + |∇u| 2 + ɛdΩ (4.23)<br />

which leads to the following Euler-Lagrange Equation :<br />

(<br />

)<br />

∇u<br />

−2λ(u − f) + div √ = 0 (4.24)<br />

|∇u| 2 + ɛ 2<br />

From the following inequality :<br />

∫<br />

∫<br />

∀u ∈ L 1 (Ω), |∇u|dΩ ≤<br />

Ω<br />

Ω<br />

Ω<br />

√<br />

∫<br />

|∇u| 2 + ɛdΩ ≤<br />

Ω<br />

|∇u|dΩ+ √ ɛ|Ω| (4.25)<br />

it follows that :<br />

√<br />

∫<br />

lim |∇u|<br />

ɛ→0<br />

∫Ω<br />

2 + ɛdΩ = |∇u|dΩ (4.26)<br />

Ω<br />

When the value of ɛ tends to zero, the solution of the optimisation problem (4.20) converges<br />

to that of (4.25).


79<br />

Fast and exact minimization algorithms for TV-based energy functionals have been subject<br />

to intense research, and many methods can be used to solve ROF mo<strong>de</strong>l. In their original<br />

work, Rudin, Osher and Fatemi relied on a fixed step gradient <strong>de</strong>scent. Quasi-Newton<br />

schemes have been investigated in [Chambolle and Lions, 1997], [Dobson and Vogel, 1997],<br />

and [Nikolova and Chan, 2007]. The dual formulation of TV minimization was explored<br />

in [Cha] and later lead to Antonin Chambolle’s projection algorithm [Chambolle, 2004].<br />

The approach <strong>de</strong>veloped by Chambolle was the first to provi<strong>de</strong> a solution to the exact<br />

TV minimization problem (4.20) instead of a smoothed version. More recently, techniques<br />

based on graph cuts have proved successfull in [Boykov et al., 2001] and [Dar]. In this<br />

section, we propose a fixed point technique based on a Gauss-Sei<strong>de</strong>l iterative method. The<br />

image domain Ω is discretized into a grid where (x i = ih) and (y i = jh), h being the cell<br />

size. Let us <strong>de</strong>note D + , D − and D 0 the forward, backward and central finite differences.<br />

In the x-direction for exemple, we can write :<br />

(D ±x u) i,j = ± u i±1,j − u i,j<br />

h<br />

(D 0x u) i,j = u i+1,j − u i−1,j<br />

2h<br />

(4.27)<br />

The divergence operator is implemented using the non-negative discretization which offers<br />

more stability than a discretization based solely on central differences. The discret form<br />

of Euler-Lagrange equation (4.24) is :<br />

u i,j = f i,j + 1<br />

2λ D −x<br />

+ 1<br />

2λ D −y<br />

(<br />

(<br />

)<br />

D +x u i,j<br />

√<br />

(D+x u i,j ) 2 +(D 0y u i,j ) 2 + ɛ 2<br />

)<br />

D +y u i,j<br />

√<br />

(D0x u i,j ) 2 +(D +y u i,j ) 2 + ɛ 2<br />

(<br />

)<br />

= f i,j + 1<br />

u i+1,j − u i,j<br />

2λh 2 √<br />

(D+x u i,j ) 2 +(D 0y u i,j ) 2 + ɛ − u i,j − u i−1,j<br />

√ 2 (D−x u i,j ) 2 +(D 0y u i−1,j ) 2 + ɛ 2<br />

(<br />

)<br />

+ 1<br />

u i,j+1 − u i,j<br />

2λh 2 √<br />

(D0x u i,j ) 2 +(D +y u i,j ) 2 + ɛ − u i,j − u i,j−1<br />

√ 2 (D0x u i,j−1 ) 2 +(D −y u i,j ) 2 + ɛ 2<br />

(4.28)<br />

The linearization of the previous equation leads to :<br />

⎛<br />

⎞<br />

u n+1<br />

i,j<br />

= f i,j + 1<br />

u<br />

⎝<br />

n i+1,j − un+1 i,j<br />

u n+1<br />

i,j<br />

− u n i−1,j<br />

2λh 2 − √<br />

⎠<br />

√(D +x u n i,j )2 +(D 0y u n i,j )2 + ɛ 2 (D −x u n i,j )2 +(D 0y u n i−1,j )2 + ɛ 2<br />

⎛<br />

⎞<br />

+ 1<br />

u<br />

⎝<br />

n i,j+1 − un+1 i,j<br />

u n+1<br />

i,j<br />

− u n i,j−1<br />

2λh 2 − √<br />

⎠<br />

√(D 0x u n i,j )2 +(D +y u n i,j )2 + ɛ 2 (D 0x u n i,j−1 )2 +(D −y u n i,j )2 + ɛ 2<br />

(4.29)


80 4. A Variational approach for the <strong>de</strong>striping issue<br />

To simplifie notations, we introduce :<br />

1<br />

C 1 = √<br />

(D +x u n i,j )2 +(D 0y u n i,j )2 + ɛ 2<br />

1<br />

C 2 = √<br />

(D −x u n i,j )2 +(D 0y u n i−1,j )2 + ɛ 2<br />

(4.30)<br />

1<br />

C 3 = √<br />

(D 0x u n i,j )2 +(D +y u n i,j )2 + ɛ 2<br />

1<br />

C 4 = √<br />

(D 0x u n i,j−1 )2 +(D −y u n i,j )2 + ɛ 2<br />

The solution of (4.24) is obtained with the following iterative scheme :<br />

u n+1<br />

i,j<br />

= 2λh2 f i,j + C 1 u n i+1,j + C 2u i−1,j + C 3 u n i,j + C 4u n i,j−1<br />

2λh 2 + C 1 + C 2 + C 3 + C 4<br />

(4.31)<br />

To satisfy Neumann boundary condition ∂u<br />

∂n = 0, u i,j is exten<strong>de</strong>d by reflection outsi<strong>de</strong> the<br />

domain Ω.<br />

Since it’s introduction in 1992, total variation regularization has grown very popular in the<br />

field of image processing. Nevertheless, for <strong>de</strong>noising purposes, its application is limited<br />

to the removal of isotropic noises such as gaussian or speckle noise [Sheng et al., 2005].<br />

Given the unidirectionality of stripe noise, tackling the striping issue with TV regularization<br />

might not be appropriate. The wavelet analysis conducted in chapter 2 un<strong>de</strong>rscored<br />

another geometrical feature of striping ; on most of MODIS emisssive bands, the amplity<strong>de</strong><br />

of stripe noise is of the same or<strong>de</strong>r as the edges of the image. Discontinuities due to<br />

striping are perceived as image sharp structures and are therefore preserved by the TV<br />

mo<strong>de</strong>l. Distinction between stripe noise and image edges can not be achieved directly with<br />

the TV mo<strong>de</strong>l and regardless the choice of the lagrange multiplier λ, reduction of striping<br />

effect is inevitably followed by a loss of contrast (figure 4.2).<br />

Nonetheless, an alternative use of ROF mo<strong>de</strong>l is conceivable. Recently, a variational approach<br />

was proposed in the context of image <strong>de</strong>striping and inpainting. It is based on a<br />

maximum a posteriori (MAP) algorithm applied to a modified image formation mo<strong>de</strong>l<br />

where the noisy observation f is given by :<br />

f = Au + B + n (4.32)<br />

where Au is a point to point multiplication. The <strong>de</strong>gradation process is assumed to be<br />

linear and inclu<strong>de</strong>s the parameters A and B associated with gain and offset values for<br />

every pixel. A and B are matrices of the same size as the image. n is assumed to be zero<br />

mean gaussian noise. The true image u is <strong>de</strong>duced from a MAP estimate :<br />

û = argmax<br />

u<br />

p(u|f) (4.33)


81<br />

Using Bayes’rule the previous equations becomes :<br />

û = argmax<br />

u<br />

p(f|u)p(u)<br />

p(f)<br />

(4.34)<br />

The a posteriori term p(u|f) being in<strong>de</strong>pen<strong>de</strong>nt on p(f), problem (4.33) reduces to :<br />

û = argmax<br />

u<br />

Application of the logarithm function on (4.35) then gives :<br />

p(f|u)p(u) (4.35)<br />

û = argmax<br />

u<br />

{log (p(f|u)) + log (p(u))} (4.36)<br />

Un<strong>de</strong>r the assumption of white additive gaussian noise, the conditional <strong>de</strong>nsity of f given<br />

u is given by :<br />

p(f|u) = 1 (<br />

Z exp − 1 )<br />

2 ‖K− 1 2 (f − Au − B)‖<br />

2<br />

(4.37)<br />

where Z is a normalizing constant and K is a diagonal matrix containing the variance<br />

values of the noise n. The prior <strong>de</strong>nsity probability function p(u) is <strong>de</strong>pen<strong>de</strong>nt on the prior<br />

contraint imposed on the image. The Markov prior for example is given by :<br />

⎛<br />

⎞<br />

p(u) = 1 Z exp ⎝− 1 ∑ ∑<br />

ρ(d c (u i,j )) ⎠ (4.38)<br />

2λ<br />

where C is the set of image cliques and d c (u i,j ) is a spatial measure of u at pixel (i, j),<br />

expressed as a function of first or second or<strong>de</strong>r differences. In [Shen et al., 2008], the<br />

authors selected an edge-preserving prior based on the Huber potential function <strong>de</strong>fined<br />

as :<br />

{<br />

x<br />

ρ(x) =<br />

2 if |x| ≤ µ<br />

2µ|x|− µ 2 (4.39)<br />

if |x| >µ<br />

The maximization of the posterior probablility distribution (4.36) is equivalent to the<br />

minimization of the following energy :<br />

i,j<br />

c∈C<br />

E(u) =λ‖K − 1 2 (f − Au − B)‖ 2 + ∑ i,j<br />

∑<br />

ρ(d c (u i,j ))<br />

c∈C<br />

(4.40)<br />

The work of [Shen et al., 2008] can be directly transposed to the TV framework. Since<br />

the Huber-Markov mo<strong>de</strong>l is used mainly for its edge-preserving ability, the prior <strong>de</strong>nsity<br />

probability function p(u) can also relie on the TV norm as :<br />

p(u) = 1 exp (−λT V (u)) (4.41)<br />

Z


82 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.3 – (Left) Image from Terra MODIS band 30 <strong>de</strong>noised with TV regularization<br />

(ID=0.5750) (Right) Denoised with the variational mo<strong>de</strong>l () (ID=0.7768)<br />

If we discard the presence of gaussian noise n or assume it is implicitly accounted for in<br />

the gain and offset parameters A and B, the covariance matrix K reduces to the i<strong>de</strong>ntity<br />

matrix and can be discar<strong>de</strong>d. Then, a <strong>de</strong>striping approach similar to [Shen et al., 2008]<br />

can be achieved by minimizing :<br />

E(u) =λ‖(f − Au − B)‖ 2 + TV(u) (4.42)<br />

Destriping via the minimization of (4.42) relies on the observational mo<strong>de</strong>l (4.32)) where<br />

A and B are assumed to be known. This is however not the case in pratice and a preprocessing<br />

stage is required to estimate the values of A and B. In their work, [Shen et al.,<br />

2008] used the moment matching technique to evaluate the values of A and B from the<br />

image mean value and variance. The limitations of such approach are directly attached to<br />

its hybrid aspect. In fact, the TV or Huber-Markov regularization acts as a post-processing<br />

stage, smoothing residual stripes that moment matching or any other technique based on<br />

linear adjustment (OFOV method for exemple) fails to remove. Nonetheless, the <strong>de</strong>striped<br />

results are much less distorted than those obtained with a direct application of the ROF<br />

mo<strong>de</strong>l (figure 4.3).<br />

4.3 Striping as a texture <br />

The main characteristic of stripping effect lies on it’s unidirectional aspect. To this<br />

extent, stripe noise can be consi<strong>de</strong>red as a structured texture with sharp fluctuations<br />

along a single axis of the image. This is particularly the case on MODIS spectral bands<br />

where striping is periodic. Texture discriminating variational mo<strong>de</strong>ls inspired from Yves


83<br />

Meyer’s work can then be used in hope of isolating the stripe noise from other structures<br />

present in the true scene.<br />

4.3.1 Yves Meyer’s mo<strong>de</strong>l for oscillatory functions<br />

The total variation regularization used in the previous section as a <strong>de</strong>noising technique<br />

can be seen from a different perspective as an image <strong>de</strong>composition mo<strong>de</strong>l, where<br />

the observed image f is approximated by a sketchy version u that lies in the BV space<br />

and the component v = f − u contains small sace <strong>de</strong>tails such as noise and/or texture.<br />

Many variational methods have the explicit goal of extracting an image u composed of<br />

homogeneous areas separated by sharp discontinuities but do not retain the component<br />

v as it is consi<strong>de</strong>red to be noise. This can be problematic for images containing texture<br />

because noise and texture are both oscillatory functions, processed equally with the TV<br />

regularization. From an image <strong>de</strong>composition perspective, ROF mo<strong>de</strong>l can be formulated<br />

as the following minimization :<br />

inf<br />

(u,v)∈BV (Ω)×L 2 (Ω)/u+v=f<br />

(<br />

)<br />

TV(u)+λ‖v‖ 2 L 2 (Ω)<br />

(4.43)<br />

In his investigation of the standard ROF mo<strong>de</strong>l [Meyer, 2002], Yves Meyer pointed out<br />

that small values of λ can remove fine <strong>de</strong>tails related to texture. To overcome this issue,<br />

he proposes a different <strong>de</strong>composition, where the classical L 2 norm associated with the<br />

residual v = f − u is replaced by a weaker-norm, more sensitive to oscillatory functions.<br />

Yves Meyer suggests finding a component v in the space G <strong>de</strong>fined as the Banach space<br />

composed of all distributions v which can be written as :<br />

v(x, y) =∂ x g 1 (x, y)+∂ y g 2 (x, y) (4.44)<br />

where g 1 and g 2 both belong to the space L ∞ (R 2 ). The space G is endowed with the norm<br />

‖v‖ G <strong>de</strong>fined as the lower bound of all L ∞ norms of the functions |⃗g| where ⃗g =(g 1 ,g 2 )<br />

and |⃗g(x, y)| = √ g 1 (x, y) 2 + g 2 (x, y) 2 . Additionnaly to G, which can be viewed as the<br />

dual space of BV , Meyer introduces the spaces E and F also suited to mo<strong>de</strong>l texture.<br />

The space E (dual of Ḃ 1,1<br />

1 ) is <strong>de</strong>fined similarly to G except that g 1,g 2 are in the space<br />

of boun<strong>de</strong>d mean oscillations functions <strong>de</strong>noted by BMO(R 2 ). For the space F (dual of<br />

H 1 ), g 1 ,g 2 belong to Besov space B∞<br />

−1,∞ (R 2 ). When the texture component v is assumed<br />

to lie in the space G, Yves Meyer proposes the following <strong>de</strong>composition mo<strong>de</strong>l :<br />

(<br />

TV(u)+λ‖v‖G(Ω ))<br />

2 (4.45)<br />

inf<br />

(u,v)∈BV (Ω 2 )×G(Ω 2 )/u+v=f<br />

The ‖.‖ G -norm can efficiently capture oscillating patterns because it is weaker than the<br />

‖.‖ 2 -norm (L 2 (Ω) ⊂ G(Ω) ). However, due to its mathematical form, the Euler-Lagrange<br />

equation of (4.45) cannot be expressed explicitly and several u + v <strong>de</strong>composition mo<strong>de</strong>ls<br />

have been later introduced as an approximation to Yves Meyer mo<strong>de</strong>l.


84 4. A Variational approach for the <strong>de</strong>striping issue<br />

4.3.2 Vese-Osher’s Mo<strong>de</strong>l<br />

Vese and Osher were the first to overcome the difficulty of Meyer’s functional minimization.<br />

They proposed a pratical resolution in [Vese and Osher, 2002] where they<br />

approximate the L ∞ norm of |⃗g| with :<br />

√<br />

∥ ‖|⃗g|‖ L ∞ =<br />

∥ g1 2 + ∥∥∥ g2 2∥ = lim<br />

√g 2<br />

L ∞ p→∞<br />

1 + g2 ∥<br />

2<br />

∥<br />

L p<br />

(4.46)<br />

As pointed out in [Meyer, 2002], the residual v = f − u in the original ROF mo<strong>de</strong>l, can<br />

be expressed as the divergence of a vector field ⃗g ∈ L ∞ (Ω) since :<br />

v = − 1 ( ) ∇u<br />

2λ div |∇u|<br />

Vese and Osher then consi<strong>de</strong>r the space of generalized functions :<br />

(4.47)<br />

G p (Ω) = {v = div(⃗g), ⃗g =(g 1 ,g 2 ), g 1 ,g 2 ∈ L p (Ω)} (4.48)<br />

induced by the norm :<br />

‖v‖ Gp(Ω) =<br />

inf<br />

v=div(⃗g), g 1 ,g 2 ∈L p (Ω)<br />

√<br />

∥ g1 2 + g2 2∥ (4.49)<br />

As an approximation to Yves Meyer mo<strong>de</strong>l, Vese and Osher propose the following <strong>de</strong>composition<br />

:<br />

inf<br />

(u,⃗g)∈BV (Ω)×L p (Ω) 2 |u| BV (Ω) + λ‖f − (u + div(⃗g))‖ 2 L 2 (Ω) + µ ‖|⃗g|‖ L p (Ω) (4.50)<br />

which approximates the original <strong>de</strong>composition mo<strong>de</strong>l of Yves Meyer when λ →∞and<br />

p →∞. The minimization problem () also writes as :<br />

∫<br />

inf<br />

|∇u|dxdy + λ |f − u − ∂ x g 1 − ∂ y g 2 |<br />

(u,(g 1 ,g 2 ))∈BV (Ω)×L p (Ω)<br />

∫Ω<br />

2 dxdy<br />

2 Ω<br />

(∫ √<br />

+ µ<br />

(<br />

Ω<br />

g 2 1 + g2 2 )p dxdy<br />

) 1<br />

p<br />

(4.51)<br />

where λ and µ are tuning parameters. The first term in () forces u to lie in the space<br />

BV (Ω). The second term insures that the noisy image can be approximated as f ≈<br />

u + div(⃗g). The last term is a penalty on the G p norm of v which regulates the amount of<br />

texture extracted in the component v. The minimization of Vese-Osher functional requires<br />

the computation of its partial <strong>de</strong>rivative with respect to u, g 1 and g 2 which leads to three


85<br />

coupled Euler-Lagrange equations :<br />

u = f − ∂g 1<br />

∂x − ∂g 2<br />

∂y + 1 ( ) ∇u<br />

2λ div |∇u|<br />

) ( p−2 ∥∥∥∥ √<br />

) 1−p (<br />

µg 1<br />

(√g1 2 + g2 2 g1 2 + g2 ∂(u − f)<br />

2∥ =2λ<br />

p<br />

∂x<br />

) ( p−2 ∥∥∥∥ √<br />

) 1−p (<br />

µg 2<br />

(√g1 2 + g2 2 g1 2 + g2 ∂(u − f)<br />

2∥ =2λ<br />

p<br />

∂y<br />

)<br />

+ ∂2 g 1<br />

∂ 2 x + ∂2 g 2<br />

∂x∂y<br />

)<br />

+ ∂2 g 1<br />

∂x∂y + ∂2 g 2<br />

∂ 2 y<br />

(4.52)<br />

In [Vese and Osher, 2002], Vese and Osher observed similar results for 1 ≤ p ≤ 10. When<br />

p ≫ 10, only texture associated with small scale <strong>de</strong>tails can be extracted. As suggested<br />

by Vese and Osher, we consi<strong>de</strong>r the case p = 1 for computational speed and the previous<br />

equations reduce to :<br />

u = f − ∂g 1<br />

∂x − ∂g 2<br />

∂y + 1 ( ) ∇u<br />

2λ div |∇u|<br />

( )<br />

g 1<br />

∂(u − f)<br />

µ √ =2λ<br />

+ ∂2 g 1<br />

g<br />

2<br />

1 + g2<br />

2 ∂x ∂ 2 x + ∂2 g 2<br />

∂x∂y<br />

( )<br />

g 2<br />

∂(u − f)<br />

µ √ =2λ<br />

+ ∂2 g 1<br />

g<br />

2<br />

1 + g2<br />

2 ∂y ∂x∂y + ∂2 g 2<br />

∂ 2 y<br />

(4.53)<br />

Neumann conditions are used at the boundary of the image domain Ω. Let us <strong>de</strong>note<br />

by (n x ,n y ) the normal to the boundary ∂Ω of the image domain, Neumann conditions<br />

translate to :<br />

∇u<br />

|∇u| (n x,n y ) = 0<br />

(<br />

)<br />

f − u − ∂g 1<br />

∂x − ∂g 2<br />

∂y<br />

n x =0<br />

(<br />

)<br />

f − u − ∂g 1<br />

∂x − ∂g 2<br />

∂y<br />

n y =0<br />

(4.54)<br />

The three coupled equations <strong>de</strong>rived from Vese-Osher mo<strong>de</strong>l are discretized and linearized<br />

as in section 4.2. Mixed <strong>de</strong>rivatives of g 1 (and g 2 ) are approximated with :<br />

∂ 2 g 1,i,j<br />

∂x∂y = 1<br />

4h 2 (g 1,i+1,j+1 + g 1,i−1,j−1 − g 1,i+1,j−1 − g 1,i−1,j+1 ) (4.55)


86 4. A Variational approach for the <strong>de</strong>striping issue<br />

Using the same notations for C 1 , C 2 , C 3 , C 4 as in section 4.2, estimates for u, g 1 and g 2<br />

are obtained via the fixed point iterative scheme :<br />

(<br />

) (<br />

u n+1<br />

1<br />

i,j<br />

=<br />

1+ 1<br />

f i,j − gn 1,i+1,j − gn 1,i−1,j<br />

(C<br />

2λh 2 1 + C 2 + C 3 + C 4 )<br />

2h<br />

− gn 2,i,j+1 − gn 2,i,j−1<br />

+ 1<br />

)<br />

2h 2λh 2 (C 1u n i+1,j + C 2 u n i−1,j + C 3 u n i,j+1 + C 4 u n i,j−1)<br />

⎛<br />

g1,i,j n+1 = ⎝<br />

4λ<br />

h 2 µ<br />

⎞<br />

(<br />

2λ<br />

u<br />

n<br />

√<br />

⎠ i+1,j − u n i−1,j<br />

g1,i,j n 2 + g2,i,j n 2 2h<br />

− f i+1,j − f i−1,j<br />

2h<br />

+ gn 1,i+1,j − gn 1,i−1,j<br />

h 2 + 1<br />

4h 2 (g 2,i+1,j+1 + g 2,i−1,j−1 − g 2,i+1,j−1 − g 2,i−1,j+1 )<br />

⎞<br />

(<br />

⎝<br />

2λ<br />

u<br />

n<br />

√<br />

⎠ i,j+1 − u n i,j−1<br />

− f i,j+1 − f i,j−1<br />

g1,i,j n 2 + g2,i,j n 2 2h<br />

2h<br />

⎛<br />

g2,i,j n+1 =<br />

4λ<br />

h 2 µ<br />

+ gn 2,i,j+1 − gn 2,i,j−1<br />

h 2 + 1<br />

4h 2 (g 1,i+1,j+1 + g 1,i−1,j−1 − g 1,i+1,j−1 − g 1,i−1,j+1 )<br />

4.3.3 Osher-Solé-Vese’s Mo<strong>de</strong>l<br />

)<br />

)<br />

(4.56)<br />

The mo<strong>de</strong>l proposed by Osher, Solé and Vese in [Osher et al., 2002], provi<strong>de</strong>s another<br />

practical approximation of (4.45). If the texture component can be written as v = f − u =<br />

div(⃗g) with ⃗g ∈ L ∞ (Ω) 2 then the Hodge <strong>de</strong>composition of ⃗g leads to :<br />

⃗g = ∇P + ⃗ Q (4.57)<br />

where ⃗ Q is a divergence free vector. Consi<strong>de</strong>ring the divergence of the previous equation,<br />

the texture component f − u can be written as :<br />

f − u = div(⃗g) =div(∇P + ⃗ Q)=div(∇P )=△P (4.58)<br />

From the previous equation, P can be expressed as P = △ −1 (f − u). Osher et al. then<br />

suggest to replace the L ∞ -norm by the L 2 -norm of |⃗g|. Neglecting Q ⃗ in the Hodge <strong>de</strong>composition<br />

of ⃗g, Osher et al. consi<strong>de</strong>r the simple minimization problem :<br />

∫ ∫<br />

inf<br />

u∈BV (Ω)<br />

Ω<br />

|∇u| + λ<br />

inf<br />

u∈BV (Ω)<br />

Ω<br />

Ω<br />

|∇(△ −1 )(f − u)| 2 dx dy (4.59)<br />

Using the norm in the space H −1 <strong>de</strong>fined by ‖v‖ H −1 = ∫ Ω |∇(△−1 )(v)| 2 dx dy, the previous<br />

minimization problem can be simplified to :<br />

∫<br />

|∇u| + λ‖f − u‖ 2 H−1 (4.60)


87<br />

The minimization of (4.60) is obtained with the Euler-Lagrange equation :<br />

( ) ∇u<br />

2λ △ −1 (f − u) =div<br />

|∇u|<br />

which takes a pratical form after application of the Laplacian operator :<br />

( ( )) ∇u<br />

2λ(f − u) =△ div<br />

|∇u|<br />

(4.61)<br />

(4.62)<br />

Equation (4.62) can be solved using a fixed step gradient <strong>de</strong>scent. The estimation of u at<br />

iteration n + 1 is given by :<br />

( ( ))) ∇u<br />

u n+1 = u n − ∆ t<br />

(2λ(f − u n n<br />

) −△ div<br />

|∇u n |<br />

(4.63)<br />

u 0 = f<br />

It is interesting to point out that Osher-Solé-Vese mo<strong>de</strong>l is very similar to the particular<br />

case of Vese-Osher mo<strong>de</strong>l where p = 2 and λ = ∞. In fact, the space G p (Ω) coinci<strong>de</strong>s<br />

with the Sobolev space W −1,p (Ω) which is the dual of W 1,p′<br />

0 (Ω) where p and p ′ satisfie the<br />

relationship 1 p + 1 p<br />

= 1. If p = 2, then the texture component v lies in G ′ 2 (Ω) = W −1,2 (Ω) =<br />

H −1 (Ω). The <strong>de</strong>composition mo<strong>de</strong>l of Vese and Osher can then be written as :<br />

inf |u| BV (Ω) + λ‖f − (u + v)‖ 2 L 2 (Ω) + µ ‖v‖ H −1 (Ω) (4.64)<br />

(u,v)∈BV (Ω)×H −1 (Ω)<br />

4.3.4 Other u + v mo<strong>de</strong>ls<br />

More image <strong>de</strong>composition mo<strong>de</strong>ls were introduced following [Vese and Osher, 2002]<br />

and [Osher et al., 2002]. In [Aujol et al., 2005], the following energy is consi<strong>de</strong>red :<br />

(<br />

)<br />

TV(u)+λ‖f − (u + v)‖ 2 L 2 (Ω)<br />

(4.65)<br />

inf<br />

(u,v)∈BV (Ω)×G µ(Ω)<br />

where the space G µ is <strong>de</strong>fined as :<br />

G µ = {v ∈ G(Ω)/‖v‖ G ≤ µ)} (4.66)<br />

The minimization of (4.66) is obtained with the projection algorithm of Chambolle.<br />

[Daubechies and Teschke, 2005] introduced an image <strong>de</strong>composition mo<strong>de</strong>l based on a<br />

wavelet framework as :<br />

inf<br />

(u,v)∈B 1 1 (L1 (Ω))×H −1 (Ω)<br />

(<br />

)<br />

2α|u| B 1<br />

1 (L 1 (Ω)) + ‖f − (u + v)‖2 L 2 (Ω) + ‖‖2 H −1 (Ω)<br />

(4.67)<br />

The space BV (Ω) used in most image <strong>de</strong>composition mo<strong>de</strong>ls is replaced by a space suited<br />

for wavelet coefficients, namely the Besov space B 1 1 (Ω).


88 4. A Variational approach for the <strong>de</strong>striping issue<br />

We also refer to the variational mo<strong>de</strong>l initially proposed for 1D signals in [] and later<br />

explored by [Nikolova, 2004]. The proposed functional replaces the L 2 norm of the original<br />

ROF mo<strong>de</strong>l by the L 1 norm as :<br />

inf<br />

(u,v)∈BV (Ω) 1 (Ω)<br />

(<br />

)<br />

TV(u)+λ‖f − (u + v)‖ 1 L 1 (Ω)<br />

(4.68)<br />

4.3.5 Experimental results and discussion<br />

The original ROF mo<strong>de</strong>l, Vese-Osher (VO) and Osher-Solé-Vese (OSV) <strong>de</strong>composition<br />

mo<strong>de</strong>ls have been applied to Terra MODIS images from band 30 and 33, in an attempt to<br />

extract stripe noise in the texture component v. The u + v <strong>de</strong>compositions are illustrated<br />

in figures 4.4 and 4.5. Traditionally, the lagrange multiplier λ is selected so that the L 2<br />

norm of v is the same for every <strong>de</strong>composition mo<strong>de</strong>l. This strategy aims at visually<br />

evaluating the ability of <strong>de</strong>composition mo<strong>de</strong>ls to discriminate texture and noise from the<br />

image main content and is well suited when no requirements are imposed on the cartoon<br />

component u. In our case however, the primary goal is to isolate the striping on v while<br />

preserving an acceptable distortion in u as it constitute an estimate of the <strong>de</strong>striped image.<br />

Consequently, we proceed by selecting a value of λ that ensures approximatively the same<br />

ID in<strong>de</strong>x for u for every <strong>de</strong>composition mo<strong>de</strong>l.<br />

Figures 4.4 and 4.5 un<strong>de</strong>rscore a better ability of (VO) and (OSV) mo<strong>de</strong>ls compared to the<br />

ROF mo<strong>de</strong>l in terms of texture discrimation. In fact, it can be seen that the v component<br />

of the ROF mo<strong>de</strong>l also contains smooth structures related to ocean and clouds, and not<br />

visible with VO and OSV <strong>de</strong>compositions. Nevertheless, <strong>de</strong>spite a strong regularization<br />

(ID=0.7), striping is still visible in the u-component of all three <strong>de</strong>composition mo<strong>de</strong>ls.<br />

Remarquably, we point out that the u-component <strong>de</strong>rived from ROF mo<strong>de</strong>l contains less<br />

stripes than that obtained with (VO) and (OSV) mo<strong>de</strong>ls. This observation is in agreement<br />

with the conlusion drawn from the wavelet analysis of section 3.7. In<strong>de</strong>ed, as a result of its<br />

high intensity (in terms of gradient values), striping tends to be consi<strong>de</strong>red as an image<br />

discontinuity and is therefore better preserved in the cartoon component u of texture<br />

discriminating variational mo<strong>de</strong>ls.<br />

For all three mo<strong>de</strong>ls, the presence of residual stripes <strong>de</strong>spite oversmoothing, is a limitation<br />

related to the contradictive compromise between the terms of the energy functionals. In<br />

fact, in addition to the true image u being attached simply to the noisy image f (regardless<br />

the norm used), the commonly used TV-norm reinforces the anisotropic preservation of<br />

structures and does not distinguish strong gradient values related to striping from those<br />

corresponding to edges. We recall that the initial motivation behind the use of variational<br />

<strong>de</strong>composition mo<strong>de</strong>ls is the textured aspect of striping due to its unidirectionality. This<br />

feature however, is not accounted for neither in the fi<strong>de</strong>lity term nor in the regularizing<br />

term.<br />

We shall see in the following sections how the introduction of directional information in<br />

variational mo<strong>de</strong>ls offers a new perpective for the removal of stripe noise.


Figure 4.4 – Decomposition of the image from Terra MODIS band 30 as u (Left) +<br />

v (Right). λ is selected so that ID(u) is the same for the three mo<strong>de</strong>ls. From top to<br />

bottom, TV mo<strong>de</strong>l (NR=4.56, ID=0.739), VO mo<strong>de</strong>l (NR=4.23, ID=0.735) and OSV<br />

mo<strong>de</strong>l (NR=2.47, ID=0.735)<br />

89


90 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.5 – Decomposition of the image from Terra MODIS band 33 as a cartoon u<br />

(Left) + v (Right). λ is selected so that ID(u) is the same for the three mo<strong>de</strong>ls. From<br />

top to bottom, TV mo<strong>de</strong>l (ID=0.447), VO mo<strong>de</strong>l (ID=0.472) and OSV mo<strong>de</strong>l (ID=0.461)


91<br />

4.4 Destriping via gradient field integration<br />

This far, we introduced variational methods based on PDE’s and energy functional minimization,<br />

with image <strong>de</strong>noising applications in mind. These techniques are actually used<br />

extensively in computer vision/graphics, to solve over-constrained geometric problems.<br />

It is typically the case of Photometric Stereo Methods (PSM) and Shape From Shading<br />

(SFS) applications where the goal is to recover a <strong>de</strong>pth map (or an image) by integrating<br />

a gradient field with discrete values.<br />

Let us <strong>de</strong>note by G =(G x ,G y ) a bidimensional gradient vector <strong>de</strong>fined in a subspace<br />

Ω of R 2 . The problem of gradient field integration consists in <strong>de</strong>termining a function u<br />

whose gradient ∇u is close to G. Two classes of techniques can be used to tackle this<br />

ill-posed problem and estimate the true image u. Local integration techniques [Coleman<br />

et al., 1982], [Healey and Jain, 1984] and [Wu and Li, 1988] rely on a curve integration :<br />

∫<br />

u(x, y) =u(x 0 ,y 0 )+ G x dx + G y dy (4.69)<br />

where γ is the integration path from a starting point (x 0 ,y 0 ) to pixel (x, y) ∈ Ω. Local integration<br />

techniques recover an image u starting with an initial height and then propagating<br />

height values according to the neighborhood gradient values. Local integration techniques<br />

are <strong>de</strong>pen<strong>de</strong>nt on the data accuracy and can propagate error values when <strong>de</strong>aling with<br />

noisy gradient fields. Global integration methods [Horn and Brooks, 1986], [], [Frankot<br />

and Chellappa, 1988] and [Horn, 1990] can be formulated in a variational framework as<br />

the minimization of an energy functional :<br />

∫<br />

E(u) =<br />

Ω<br />

γ<br />

( ) ∂u 2 ( ) ∂u 2<br />

∂x − G x +<br />

∂y − G y dx dy (4.70)<br />

Unlike <strong>de</strong>noising and <strong>de</strong>composition variational mo<strong>de</strong>ls <strong>de</strong>scribed this far, both terms in the<br />

energy functional (4.70) are fi<strong>de</strong>lity terms that measure the L 2 norm difference between the<br />

gradient field components of u and the observed gradient field G. For instance, the energy<br />

functional does not inclu<strong>de</strong> any lagrange multiplier λ. The Euler-Lagrange equations of<br />

the previous functionals leads to :<br />

∂<br />

∂x<br />

( ∂u<br />

∂x − G x<br />

which can be simplified into the following Poisson equation :<br />

)<br />

+ ∂ ( ) ∂u<br />

∂x ∂y − G y = 0 (4.71)<br />

∇ 2 u = ∇.G (4.72)<br />

The previous equation can be solved using a fast marching method [Ho et al., 2006] or<br />

the streaming multigrid method proposed in [Kazhdan and Hoppe, 2008]. Among several<br />

other techniques, we draw a particular attention to the well-known Frankot-Chellapa


92 4. A Variational approach for the <strong>de</strong>striping issue<br />

(FC) algorithm [Frankot and Chellappa, 1988], which provi<strong>de</strong>s a solution by consi<strong>de</strong>ring<br />

the Fourier transform of the gradient field G components. Let us <strong>de</strong>note by (ξ x ,ξ y ) the<br />

frequency components as in [Bracewell, 1986] and recall the differentiation properties of<br />

the Fourier transform :<br />

F<br />

⎧⎪ ( ∂u<br />

( ∂x)<br />

) = j.ξx F(u)<br />

⎨ F ∂u<br />

∂y<br />

= j.ξ y F(u)<br />

( )<br />

F ∂ 2 u<br />

= ξ 2 (4.73)<br />

xF(u)<br />

⎪ ⎩<br />

∂ 2 x<br />

F<br />

(<br />

∂ 2 u<br />

∂ 2 y<br />

)<br />

= ξyF(u)<br />

2<br />

where j is the imaginary unit √ −1. The fourier transform of the Poisson equation (4.72)<br />

can be written as :<br />

(<br />

ξ<br />

2<br />

x + ξy<br />

2 )<br />

F(u) =−jξx F(G x ) − jξ y F(G y ) (4.74)<br />

Denoting u F , G F x and G F y the fourier transform of u, G x and G y , the Frankot-Chellapa<br />

reconstruction algorithm gives :<br />

u F = −jξ xG F x − jξ y G F y<br />

ξ 2 x + ξ 2 y<br />

(4.75)<br />

As we will see, gradient field integration problems and their gradient-based variational<br />

formulation offer an interesting perspective on the <strong>de</strong>striping issue. In fact, compared to<br />

other restoration-oriented variational mo<strong>de</strong>ls, the energy functional (4.70) clearly separates<br />

the information related to vertical gradient and horizontal gradient. Going further<br />

in this gradient-based reasoning, we make the following remark :<br />

If the stripe noise is additive, it mostly affects the vertical gradient of the striped image<br />

Let us the consi<strong>de</strong>r the following image formation mo<strong>de</strong>l :<br />

I s = I + n (4.76)<br />

where I is the stripe free true image and n is the stripe noise. The linear operator K is<br />

consi<strong>de</strong>red to be the i<strong>de</strong>ntity. Let us assume that the unidirectional signature of the stripe<br />

noise n translates on its horizontal gradient as :<br />

∂n(x, y)<br />

∂x<br />

≈ 0 (4.77)<br />

The partial <strong>de</strong>rivatives along the x and y-axis of the image formation mo<strong>de</strong>l (4.76) :<br />

∂I s<br />

∂x = ∂I<br />

∂x + ∂n<br />

∂x<br />

∂I s<br />

∂y = ∂I<br />

∂y + ∂n<br />

∂y<br />

(4.78)


93<br />

Figure 4.6 – (Left) Horizontal and (Right) vertical gradient of the image from Terra<br />

MODIS band 30. The stripe noise is isolated in the vertical gradient<br />

can be simplified un<strong>de</strong>r the assumption (4.77) into :<br />

∂I s<br />

∂x ≈ ∂I<br />

∂x<br />

∂I s<br />

∂y = ∂I<br />

∂y + ∂n<br />

∂y<br />

(4.79)<br />

Since the stripe noise only affects the horizontal gradient of the image (figure 4.6), we<br />

propose a variational mo<strong>de</strong>l where the directional information is distributed separately on<br />

the fi<strong>de</strong>lity term and the regularizing term. In addition, the regularization is limited to the<br />

direction of the stripe noise. Un<strong>de</strong>r these requirements, we minimize the following energy :<br />

∫<br />

E(u) =<br />

Ω<br />

∂(u − I s )<br />

∥ ∂x<br />

∥<br />

2<br />

∫<br />

dx dy + λ<br />

Ω<br />

∂(u − H ⊗ I s )<br />

∥ ∂y<br />

∥<br />

2<br />

dx dy (4.80)<br />

where H is a low-pass filter applied to the noisy image I s only to approximate the true<br />

image cross-track profile. In fact, if the regularizing term in (4.80) is replaced with ‖ ∂u<br />

∂y ‖2 ,<br />

lines of the estimated solution u will have zero mean. Frankot-Chellapa algorithm presented<br />

previously can be used to minimize the previous energy functional which Euler-Lagrange<br />

equation is :<br />

∂ 2 (<br />

u<br />

∂ 2 x − ∂2 I s ∂ 2 )<br />

∂ 2 x + λ u<br />

∂ 2 y − ∂2 (H ⊗ I s )<br />

∂ 2 =0 (4.81)<br />

y<br />

Using the differienciation properties of the fourier transform, the previous equation can<br />

be expressed as :<br />

ξ 2 x(u F − I F s )+λξ 2 y(u F − H F .I F s )=0 (4.82)


94 4. A Variational approach for the <strong>de</strong>striping issue<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

u<br />

u<br />

v<br />

v<br />

Figure 4.7 – Frequency response of the filter ˜H. As the value of λ <strong>de</strong>creases, only frequencies<br />

close to the vertical axis of the fourier domain are reduced.<br />

where u F , Is<br />

F and H F are the fourier transform of u, I s and H. An estimate of the stripe<br />

free true image u is obtained with a simple inverse fourier transform :<br />

)<br />

u = F −1 (<br />

(ξ<br />

2<br />

x + λξ 2 yH F ).I F s<br />

ξ 2 x + λξ 2 y<br />

(4.83)<br />

Remark : If we ignore the term H ⊗I s in (4.80) and <strong>de</strong>note ˜H a filter <strong>de</strong>fined in the fourier<br />

domain as :<br />

ξx<br />

˜H 2 =<br />

ξx 2 + λξy<br />

2 (4.84)<br />

<strong>de</strong>striping via the variational mo<strong>de</strong>l (4.80) is equivalent to filtering with ˜H. The fourier<br />

response of the filter ˜H is illustrated in figure 4.7.<br />

The <strong>de</strong>striping variational mo<strong>de</strong>l (4.80) is applied to images extracted from Terra<br />

MODIS band 27, 30 and 33 and the results are illustrated in figures 4.8 and 4.9. In the case<br />

of homogeneous data, as in band 27, we observe a complete removal of the striping effect<br />

without any visible blur in the <strong>de</strong>striped result. We point out that the overall blur, i.e the<br />

image distortion in<strong>de</strong>x <strong>de</strong>fined in section 3.8, <strong>de</strong>pends on the lagrange multiplier λ, whch<br />

choice is discussed further in section 4.6. Although the results on band 27 are satisfactory,<br />

the <strong>de</strong>striping mo<strong>de</strong>l does not perform as well on images containing strong discontinuities.<br />

Figure 4.9, clearly shows local blurring artifacts along sharp edges at ocean-land and oceanclouds<br />

transitions. This major limitation is, analogically to Tikhonov regularization, due<br />

to the L 2 norm used in (4.80) which is not able to preserve sharp structures.<br />

4.5 A unidirectional variational <strong>de</strong>striping mo<strong>de</strong>l<br />

From a variational perspective, we first tackled the striping issue with ROF total variation<br />

regularization mo<strong>de</strong>l. We also explored image <strong>de</strong>composition mo<strong>de</strong>ls for their ability


95<br />

Figure 4.8 – (Left) Noisy image from Terra MODIS band 27 (Right) Image <strong>de</strong>striped<br />

with the variational mo<strong>de</strong>l (3.76) showing no evi<strong>de</strong>nce of stripes or distortion (ID=0.989)<br />

Figure 4.9 – (Left) Noisy image from Terra MODIS band 30 (b) Image <strong>de</strong>striped with the<br />

variational mo<strong>de</strong>l (3.76). Although stripes have been completely removed, local blurring<br />

artifacts appear along the image discontinuities.<br />

to discriminate texture. The attempt to isolate the stripe noise in the texture component<br />

v fails as the total variation norm in the fi<strong>de</strong>lity term provi<strong>de</strong>s oversmoothed results.<br />

Gradient field integration algorithms based on a variational formulation are suited for the<br />

issue of <strong>de</strong>striping because of the separation between vertical and horizontal information<br />

in the energy functional to be minimized. The <strong>de</strong>striping variational mo<strong>de</strong>l <strong>de</strong>rived in the<br />

previous section performs well on homogeneous areas but introduces local blur artifacts


96 4. A Variational approach for the <strong>de</strong>striping issue<br />

along strong edges. In this section, we overcome this drawback and introduce a new variational<br />

mo<strong>de</strong>l that satisfies the requirements of optimal <strong>de</strong>striping.<br />

In chapter 3, we attributed the limitations of standard <strong>de</strong>striping techniques to the simplistic<br />

assumptions ma<strong>de</strong> on the stripe noise. Equalization based methods (moment matching,<br />

histogram matching and OFOV techniques) assume that the stripe noise can be mo<strong>de</strong>lled<br />

with a gain and offset between the <strong>de</strong>tectors of the sensor. The presence of strong nonlinearities<br />

on MODIS <strong>de</strong>tectors responses, discredits this assumption. A simple and more<br />

realistic assumption lies in the directional aspect of stripe noise. For instance, if we <strong>de</strong>note<br />

by n the stripe noise, it is resonable to assume that :<br />

∣ ∀(x, y) ∈ Ω,<br />

∂n(x, y)<br />

∣∣∣ ∣ ∂x ∣ ≪ ∂n(x, y)<br />

∂y ∣ (4.85)<br />

The integration of the previous inequality over the entire image domain Ω results in :<br />

∫<br />

∫<br />

∂n(x, y)<br />

∣ ∂x ∣ ≪ ∂n(x, y)<br />

∣ ∂y ∣ (4.86)<br />

Ω<br />

which translates a realistic characteristic of stripe noise when written as :<br />

Ω<br />

TV x (n) ≪ TV y (n) (4.87)<br />

where TV x and TV y are respectively horizontal and vertical variations of the noise n.<br />

The L 1 norm being an edge-preserving norm can be introduced in a variational mo<strong>de</strong>l<br />

where the fi<strong>de</strong>lity term takes into account the stripe free horizontal information while<br />

the regularization term only smoothes the result along the vertical axis. We propose a<br />

<strong>de</strong>striping variational mo<strong>de</strong>l based on the minimization of the following energy :<br />

E(u) =TV x (u − I s )+λT V y (u) (4.88)<br />

We anticipate the non-differentiability of this functional at points where ∂u<br />

= 0 by introducing a small parameter ɛ in (4.88) which becomes :<br />

∂u<br />

∂y<br />

∂x = ∂Is<br />

∂x<br />

and<br />

∫<br />

E ɛ (u) =<br />

Ω<br />

√ (∂(u ) − Is ) 2 ∫<br />

+ ɛ 2 + λ<br />

∂x<br />

Ω<br />

√ (∂u ) 2<br />

+ ɛ 2 (4.89)<br />

The mo<strong>de</strong>l (4.89) will be refered to hereafter as the Unidirectional Variational Destriping<br />

Mo<strong>de</strong>l (UVDM). Its Euler-Lagrange equation is given by :<br />

− ∂<br />

∂x<br />

⎛<br />

⎜<br />

⎝<br />

∂(u−I s)<br />

∂x<br />

√ (<br />

∂(u−Is)<br />

∂x<br />

⎞<br />

⎛<br />

⎟<br />

) 2 ⎠ − λ ∂ ⎜<br />

∂y ⎝<br />

+ ɛ 2<br />

√ (<br />

∂u<br />

∂y<br />

∂u<br />

∂y<br />

∂y<br />

⎞<br />

⎟<br />

) 2 ⎠ =0 (4.90)<br />

+ ɛ 2


97<br />

Using the same notations as in sections 4.2, equation (4.90) can be discretized as :<br />

(<br />

) (<br />

)<br />

D +x (u i,j − f i,j )<br />

D +y u i,j<br />

D −x √ + λD −y √ =0<br />

(D+x (u i,j − f i,j )) 2 + ɛ 2 (D+y u i,j ) 2 + ɛ 2<br />

(<br />

)<br />

(u i+1,j − u i,j − f i+1,j + f i,j )<br />

√ − (u i,j − u i−1,j − f i,j + f i−1,j )<br />

√<br />

(D+x (u i,j − f i,j )) 2 + ɛ 2 (D−x (u i,j − f i,j )) 2 + ɛ 2<br />

+λ<br />

(<br />

)<br />

u i,j+1 − u i,j<br />

√<br />

(D+y u i,j ) 2 + ɛ − u i,j − u i,j−1<br />

√ =0<br />

2 (D−y u i,j ) 2 + ɛ 2<br />

(4.91)<br />

We introduce the following linearization :<br />

⎛<br />

⎞<br />

⎝ (un i+1,j − un+1 i,j<br />

− f i+1,j + f i,j )<br />

√<br />

− (un+1 i,j<br />

− u n i−1,j − f i,j + f i−1,j )<br />

√<br />

⎠<br />

(D +x (u n i,j − f i,j)) 2 + ɛ 2 (D −x (u n i,j − f i,j)) 2 + ɛ 2<br />

⎛<br />

⎞<br />

+λ ⎝<br />

un i,j+1 − un+1 i,j<br />

√<br />

− un+1 i,j<br />

− u n i,j−1<br />

√<br />

⎠ =0<br />

(D +y u n i,j )2 + ɛ 2 (D −y u n i,j )2 + ɛ 2<br />

(4.92)<br />

If we <strong>de</strong>note :<br />

C 1 =<br />

C 2 =<br />

C 3 =<br />

C 4 =<br />

1<br />

√<br />

(D +x (u n i,j − f i,j)) 2 + ɛ 2<br />

1<br />

√<br />

(D −x (u n i,j − f i,j)) 2 + ɛ 2<br />

(4.93)<br />

1<br />

√<br />

(D +y u n i,j )2 + ɛ 2<br />

1<br />

√<br />

(D −y u n i,j )2 + ɛ 2<br />

The <strong>de</strong>striped image is obtained with a fixed point iterative scheme :<br />

u n+1<br />

i,j<br />

= C 1(u n i+1,j − f i+1,j + f i,j )+C 2 (u n i−1,j + f i,j − f i−1,j )+λC 3 u n i,j+1 + λC 4u n i,j−1<br />

C 1 + C 2 + λC 3 + λC 4<br />

(4.94)<br />

4.6 Optimal regularization<br />

The limitations of standard <strong>de</strong>striping techniques discussed at the end of chapter 3,<br />

were used as a starting point to establish the requirements of optimal <strong>de</strong>striping. One of


98 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.10 – (Left) Noisy image from Terra MODIS band 30 (Center) Destriped<br />

with histogram matching (IMAPP) showing residual stripes (Right) Destriped with the<br />

UVDM illustrating complete removal of striping without any bluring or ringing artifacts<br />

Figure 4.11 – (Left) Noisy image from Terra MODIS band 33 (Center) Destriped<br />

with histogram matching (IMAPP) showing residual stripes in spite of noisy <strong>de</strong>tectors<br />

being replaced with neighbors (Right) Image <strong>de</strong>striped with the UVDM showing efficient<br />

removal of radom striping without bluring or ringing artifacts<br />

these requirements dictates minimum distortion in the <strong>de</strong>striped image. The ID in<strong>de</strong>x in<br />

clearly <strong>de</strong>pen<strong>de</strong>nt on the choice of the lagrange multiplier λ. The coefficient λ regulates<br />

the smoothness of the solution and its impact on the restored image can be interpreted as<br />

follows : A large value of λ may result in an oversmoothed solution, visually similar to what<br />

could be achieved with a low-pass filter, or a hard thresholding of wavelet coefficients in<br />

the lower scales. On the other hand, if λ is too small, the result might still contain noise. In<br />

fact, if we consi<strong>de</strong>rer the ROF mo<strong>de</strong>l, its minimizer converges to the original noisy image<br />

f when λ → 0. In the case of the UVDM, when λ → 0, the solution converges to that of<br />

the following optimization problem :<br />

√<br />

∫ (∂(u ) − Is ) 2<br />

E ɛ (u) =<br />

+ ɛ<br />

∂x<br />

2 (4.95)<br />

Ω


99<br />

which is not unique. In<strong>de</strong>ed, the solution of (4.95) is the set {u ∈ BV (Ω)/u = f +<br />

C, ∫ ∂C<br />

Ω ∂x<br />

=0} which corresponds to the set of images that differs from f on a constant<br />

per line. Due to the variations of stripe noise along the x-direction, any solution of (4.95)<br />

will still display residual stripes. We are then lead to the question of how to <strong>de</strong>termine a<br />

value of λ that removes all the stripe noise while maintainin the image distortion in<strong>de</strong>x<br />

close to 1.<br />

4.6.1 Tadmor-Nezzar-Vese (TNV) hierarchical <strong>de</strong>composition<br />

In [Tadmor et al., 2004], the authors introduce a multiscale image representation based<br />

on a hierarchical adaptive <strong>de</strong>composition. The ROF mo<strong>de</strong>l is iteratively solved to generate<br />

a sequence of solutions which sum converges to the original image. We propose a<br />

mo<strong>de</strong>st adaptation of this strategy to our variational <strong>de</strong>striping mo<strong>de</strong>l. Let us consi<strong>de</strong>r<br />

the following <strong>de</strong>composition for a striped image I s :<br />

[u 0 ,v 0 ]= argmin<br />

(u,v)/u+v=I s<br />

TV x (v)+λ 0 TV y (u)<br />

(4.96)<br />

If the inital value λ 0 is not too small, then u 0 can be consi<strong>de</strong>red as a cartoon approximation<br />

of the true stripe-free image I, while the component v 0 mainly contains stripe noise.<br />

Following Tadmor et al.’s remark that a texture at a scale λ contains edges at a refined<br />

scale λ 2 , the noisy component v 0 can also be <strong>de</strong>composed using the same variational mo<strong>de</strong>l :<br />

[u 1 ,v 1 ]= argmin<br />

(u,v)/u+v=v 0<br />

TV x (v)+ λ 0<br />

2 TV y(u) (4.97)<br />

The previous <strong>de</strong>composition can be seen as a dyadic refinement to the approximation u 0<br />

and can be iterated as follows :<br />

[u k ,v k ]= argmin<br />

(u,v)/u+v=v k−1<br />

TV x (v)+ λ 0<br />

2 k TV y(u) (4.98)<br />

leading to a simple multilayered representation of the original noisy image I s :<br />

I s = u 0 + v 0<br />

= u 0 + u 1 + v 1<br />

= u 0 + u 1 + u 2 + v 2<br />

(4.99)<br />

= ...<br />

= u 0 + u 1 + u 2 + u 3 + ... + u k + v k<br />

The previous expansion translates the hierarchical <strong>de</strong>composition of I s :<br />

j=k<br />

∑<br />

lim u j = I s (4.100)<br />

k→∞<br />

j=0


100 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.12 – Optimal <strong>de</strong>striping using Tadmor et al. hierarchical <strong>de</strong>composition approach.<br />

From left to right and top to bottom, noisy image from Terra band 30 and successive<br />

<strong>de</strong>striped results for iterations 1 to 7. The value λ 0 = 10 have been selected to ensure<br />

initial oversmoothing<br />

This iterated process provi<strong>de</strong>s a multiscale representation of I s where the successive terms<br />

u k extract <strong>de</strong>tails of the noise-free image I related to the scale λ 0 2 k , while sharper <strong>de</strong>tails<br />

associated with stripe and inclu<strong>de</strong>d in the finer scale λ 0 2 k+1 are isolated in the term v k . The<br />

original TNV hierachical <strong>de</strong>composition was initially proposed as a multiscale framework<br />

in the context of ROF regularization but does not aim at removing noise from the images.<br />

In the later case, a stopping criteria is required so that the term ∑ j=k<br />

j=0 u j does not contain<br />

stripe <strong>de</strong>tails. This is discussed in section 4.6.3. TNV hierarchical <strong>de</strong>composition is applied<br />

to a striped image extracted from Terra MODIS band 30 and illustrated in figure 4.12.<br />

The stripe noise extracted at each iteration is displayed in figure 4.13<br />

4.6.2 Osher et al. iterative regularization method<br />

In the orignal ROF mo<strong>de</strong>l, the term f − u can still contain fine edges and textures if<br />

the lagrange multiplier λ is not choosen carefully. To minimize the removal of structures<br />

from the estimate solution, [Osher et al., 2005] introduced an iterative refinement of the<br />

ROF mo<strong>de</strong>l and proposed a generalization for other variational mo<strong>de</strong>ls based on the use of<br />

Bregman distances. In the case of ROF regularization, the iterative methodology follows<br />

three steps :<br />

Step 1 : Solve the original ROF mo<strong>de</strong>l :<br />

{∫ ∫<br />

u 1 =<br />

inf<br />

u∈BV (Ω)<br />

|∇u| + λ<br />

Ω<br />

Ω<br />

(f − u) 2 }<br />

(4.101)


101<br />

Figure 4.13 – From left to right and top to bottom, extracted stripe noise using Tadmor<br />

et al. hierarchical <strong>de</strong>composition approach for iterations 1 to 8.<br />

This leads to a <strong>de</strong>composition of the image f as f = u 1 + v 1 where v 1 is the estimated<br />

noisy component.<br />

Step 2 : Inject the noise estimated in the first step v 1 in the fi<strong>de</strong>lity term and solve :<br />

{∫ ∫<br />

u 2 =<br />

inf<br />

u∈BV (Ω)<br />

|∇u| + λ<br />

Ω<br />

Ω<br />

(f + v 1 − u) 2 }<br />

(4.102)<br />

This correction step provi<strong>de</strong>s a <strong>de</strong>composition of the form f + v 1 = u 2 + v 2<br />

Step 3 : Solve :<br />

{∫ ∫<br />

}<br />

u k+1 = inf |∇u| + λ (f + v k − u) 2<br />

u∈BV (Ω) Ω<br />

Ω<br />

(4.103)<br />

where v k is obtained after k iterations as v k = v k−1 + f − u k . This iterated methodology<br />

is exten<strong>de</strong>d by Osher et al. to other inverse problems by consi<strong>de</strong>ring a general variational<br />

mo<strong>de</strong>l of the form :<br />

inf {J(u)+H(u, f)} (4.104)<br />

u<br />

where J(u) is a convex non negative regularizing functional and H(u, f) the fi<strong>de</strong>lity term.<br />

The general case consists of iteratively solving :<br />

u k+1 = inf<br />

u<br />

{J(u)+H(u, f)− < u, p k >} (4.105)<br />

where < ., . > is the duality product and p k is a subgradient of J at u k . The adaptation to<br />

the UVDM is straightforward. Replacing J(u) with λT V y (u) and H(u, f) with TV x (u−f),


102 4. A Variational approach for the <strong>de</strong>striping issue<br />

the iterative procedure can be written as :<br />

u k+1 = inf<br />

u<br />

{<br />

TV x (u − f)+λT V y (u) −<br />

∫<br />

Ω<br />

}<br />

u∂J(u k )<br />

(4.106)<br />

In practice, we consi<strong>de</strong>r a smoothed version of TV y , and the subdifferential ∂J can be<br />

replaced by the gradient of J(u) with respect to u. Equation (4.106) then becomes :<br />

⎛ ⎛<br />

⎞⎞<br />

∫<br />

u k+1 = inf TV x (u − f)+λT V y (u) − u ⎝λ ∂ ∂u k<br />

⎝<br />

∂y<br />

√ ⎠⎠ (4.107)<br />

u<br />

Ω ∂y<br />

( ∂u k<br />

∂y )2 + ɛ<br />

4.6.3 Stopping criteria<br />

Whether using TVN <strong>de</strong>composition or Osher et al. methodology, there exists an iteration<br />

k, where the estimate û k is the closest to the true sripe free image. In both cases,<br />

if k →∞, û k converges to the noisy image I s . In [Osher et al., 2005] the discrepancy<br />

principle is used as a stopping rule. Assuming that the noise level δ is known, the iterative<br />

procedure is stopped as soon as the residual term ‖û k − I s ‖ reaches a value of the same<br />

or<strong>de</strong>r as δ. In our case, the stripe noise level is not known and we have to relie on another<br />

approach. The unidirectionality of stripe noise can be exploited, again, to <strong>de</strong>fine a reliable<br />

stopping criteria. The variational mo<strong>de</strong>l (4.84), was <strong>de</strong>signed in or<strong>de</strong>r to constrain the<br />

regularization only to the direction of striping. Nevertheless, if the lagrange multiplier λ is<br />

exessively high, the estimated solution will also be smoothed in the horizontal direction.<br />

If we recall that the unidirectionality of striping translates as the horizontal gradients of<br />

the noisy image I s and the true image I being of the same or<strong>de</strong>r, λ has to be chosen so<br />

that :<br />

∫<br />

∂(û − f)<br />

∣ ∂x ∣ ≤ ɛ (4.108)<br />

Ω<br />

where ɛ is a tolerance parameter that regulates the amount of distortion introduced in<br />

û. In pratice, the difficulties related to the <strong>de</strong>termination of ɛ can be easily overcome. In<br />

fact, an optimal <strong>de</strong>striping is expected to preserve the ensemble averaged power spectrum<br />

down the lines of the noisy image because striping only affects one direction. This means<br />

that the spectral distribution of the information averaged accross the swath should be<br />

approximatively the same for the striped image and the estimated true scene. Conveniently,<br />

such measure is provi<strong>de</strong>d by the Image distortion in<strong>de</strong>x (ID) ; We recall that the ID reflects<br />

a spectral fi<strong>de</strong>lity between the <strong>de</strong>striped and original signals in the direction orthogonal to<br />

striping. As TVN <strong>de</strong>composition and Osher et al. iterative procedures result in a sequence<br />

of solutions {u k } that converges to an image û = I s + C with TV x (C) = 0, we have :<br />

lim<br />

k→∞ ID(u k)=1 (4.109)<br />

A stopping criteria can then be established using a threshold value, ID thres = 0.95.<br />

This threshold was <strong>de</strong>termined heuristically to ensure simultaneously complete removal


103<br />

Image distortion<br />

1<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

1 2 3 4 5 6 7 8<br />

Iteration<br />

Image distortion<br />

1<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

1 2 3 4 5 6 7 8 9 10 11 12<br />

Iteration<br />

Figure 4.14 – Image distortion in<strong>de</strong>x as a function of number of iterations starting with<br />

λ 0 = 10 and using (Left) Tadmor et al. dyadic hierarchical <strong>de</strong>composition (Right) Osher<br />

et al. iterative method<br />

of stripes and minimum distortion. If the threshold value is very close to 1, the estimated<br />

solution might inclu<strong>de</strong> residual striping due to non linearities or random stripes. In fact,<br />

the case ID = 1 corresponds to λ = 0.<br />

4.6.4 Experimental results and discussion<br />

The UVDM was applied to the entire data set used in this study. Experimental comparison<br />

of TVN and Osher et al. procedure shows that TVN provi<strong>de</strong>s a faster convergence<br />

to the solution with an ID close to 1. It also offers more flexibility to reduce the number of<br />

iterations which is <strong>de</strong>pen<strong>de</strong>nt on the initial choice of λ 0 . This can be achieved using tryadic<br />

or higher or<strong>de</strong>r n updates of the lagrange multiplier λ as λ k+1 = λ 0 /n k . The application of<br />

TVN hierarchical <strong>de</strong>composition on the image from Terra MODIS band 30 shows the gradual<br />

extraction of <strong>de</strong>tails through the iterations without retrieval of striping (figure 4.12).<br />

Figure 4.13 illustrates how the extracted striping noise is progressively <strong>de</strong>correlated from<br />

the original signal and isolated from the true scene structures.<br />

Figures 4.10 and 4.11 show the <strong>de</strong>striping results obtained with NASA’s IMAPP software<br />

and the UVDM on images from Terra MODIS band 30 and 33. We point out that the<br />

strategy adopted in the IMAPP software to remove non linear effects and random stripes<br />

is to replace corresponding <strong>de</strong>tectors with neighboors. Despite the distortion introduced by<br />

such procedure, residual stripes are still persistent in the resotred image. Complementary<br />

analysis of cross-track profiles (figure 4.15) and ensemble averaged power spectrum down<br />

the columns (figure 4.16), un<strong>de</strong>rscores the performances of the UVM. Cross-track profiles<br />

are properly smoothed without erasing sharp fluctuations related to transitions between<br />

ocean and land. Futhermore, examination of column power spectrums obtain with the<br />

UVM show that 1) spectral peaks associated with periodic stripes are completely removed<br />

2) rapid fluctuations in the high frequency range due to random stripes are canceled 3)<br />

the overall shape of the column spectrum is preserved. These qualitative observations are


104 4. A Variational approach for the <strong>de</strong>striping issue<br />

supported by quantitative measurements reported in tables 4.1 and 4.2.<br />

Although the UVM was <strong>de</strong>rived intuitively by combining edge-preserving mo<strong>de</strong>ls and gradient<br />

field integration approaches, similar results can be achieved with slightly different<br />

variational mo<strong>de</strong>ls. These are based on the minimization of the following energy functionals<br />

:<br />

E 1 (u) =TV x (u − f)+λT V (u) (4.110)<br />

∫ ( ) ∂ (u − f) 2<br />

E 2 (u) =<br />

+ λT V (u) (4.111)<br />

Ω ∂x<br />

∫ ( ) ∂ (u − f) 2<br />

E 3 (u) =<br />

+ λT V y (u) (4.112)<br />

∂x<br />

Ω<br />

In the mo<strong>de</strong>l (4.112), the L 1 -norm on the x-gradient is replaced by the L 2 -norm. This<br />

modification does not cause bluring artifacts as the edges are still preserved by the regularizing<br />

term which inclu<strong>de</strong>s the L 1 -norm of the y-gradient. Nevertheless, the original<br />

UVDM is more suited to isolate striping due to the L 1 -norm being weaker than the standard<br />

L 2 -norm (L 2 (Ω) ∈ L 1 (Ω)). For u-components holding the same ID in<strong>de</strong>x and <strong>de</strong>rived<br />

from (4.112) and the UVDM, the noisy component v extracted with the UVDM contains<br />

more stripe-related texture. This feature comes in handy in the TVN hierarchical <strong>de</strong>composition<br />

where the <strong>de</strong>striped image is obtained by progressively extracting striping from<br />

oversmoothed u estimates. The L 1 -norm in the fi<strong>de</strong>lity term then ensures a minimal number<br />

of iterations compared to the L 2 -norm.<br />

The mo<strong>de</strong>l (4.110) replaces the regularizing term of the UVM mo<strong>de</strong>l with the commonly<br />

used TV-norm which provi<strong>de</strong>s similar results than the UVM but increases the computational<br />

cost required for the minimization of its energy functional. This drawback, together<br />

with the use of L 2 -norm in the fi<strong>de</strong>lity term are both present in the variational mo<strong>de</strong>l<br />

(4.111).<br />

From a <strong>de</strong>noising perspective, the UVDM and mo<strong>de</strong>ls (4.110), (4.111), (4.112) satisfy the<br />

<strong>de</strong>striping requirements imposed in chapter 3, namely complete stripe removal without<br />

blur/ringing artifacts. In fact, both TVN and/or Osher et al. iterative procedures ensure<br />

that the ID in<strong>de</strong>x of the optimally <strong>de</strong>striped image is close to 1. This emphazises the<br />

importance of a suitable fi<strong>de</strong>lity term which, for all <strong>de</strong>striping mo<strong>de</strong>ls <strong>de</strong>scribed above,<br />

takes into account the unidirectional property of striping.


105<br />

17500<br />

23500<br />

Mean value<br />

16500<br />

15500<br />

Mean value<br />

23000<br />

22500<br />

14500<br />

0 100 200 300 400 500<br />

Line Number<br />

22000<br />

0 100 200 300 400 500<br />

Line Number<br />

17500<br />

23500<br />

Mean value<br />

16500<br />

15500<br />

Mean value<br />

23000<br />

22500<br />

14500<br />

0 100 200 300 400 500<br />

Line Number<br />

22000<br />

0 100 200 300 400 500<br />

Line Number<br />

17500<br />

23500<br />

Mean value<br />

16500<br />

15500<br />

Mean value<br />

23000<br />

22500<br />

14500<br />

0 100 200 300 400 500<br />

Line Number<br />

22000<br />

0 100 200 300 400 500<br />

Line Number<br />

Figure 4.15 – Cross-Track profiles for Terra MODIS band 30 (Left) and band 33 (Right).<br />

From top to bottom : Original image, histogram matching with IMAPP and proposed<br />

UVDM


106 4. A Variational approach for the <strong>de</strong>striping issue<br />

10<br />

8<br />

Power spectrum<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

Power spectrum<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

10<br />

8<br />

Power spectrum<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

Power spectrum<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

10<br />

8<br />

Power spectrum<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

Power spectrum<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

2<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Normalized frequency<br />

Figure 4.16 – Column power spectrum for Terra MODIS band 27 (Left) and band 33<br />

(Right). From top to bottom : Original image, histogram matching with IMAPP and<br />

proposed UVDM


Table 4.1 – Noise Reduction (NR), Image distortion (ID), radiometric Improvement Factors IF 1 and<br />

IF 2 for the Terra MODIS band 27, 30 and 33<br />

– Terra MODIS band 27 Terra MODIS band 30 Terra MODIS band 33<br />

in<strong>de</strong>x NR ID IF 1 IF 2 NR ID IF 1 IF 2 NR ID IF 1 IF 2<br />

moment matching 45.8895 0.6305 7.3085 11.5732 9.4280 0.9963 6.2656 6.6818 – 0.9995 0.1637 0.1724<br />

histogram (IMAPP) 101.0629 0.9459 8.5463 20.1003 15.3316 0.8816 5.7455 12.8977 – 0.9773 4.8320 11.1566<br />

IFOV 100.5751 0.9692 5.8563 13.7338 5.4195 0.8991 4.6190 5.7684 – 0.9388 -0.1046 0.5466<br />

Frequency filtering 1784 0.9868 9.8866 21.5316 25.5615 0.9884 7.3092 14.8566 – – – –<br />

Wavelet thresholding 6194 0.9039 9.9038 38.4647 54.1577 0.8956 6.34871 18.7944 – 0.8765 4.9008 17.1591<br />

Proposed 2602 0.9881 9.7695 42.3641 15.6980 0.9897 11.8861 22.4904 – 0.9877 8.3956 21.3488<br />

Table 4.2 – Noise Reduction (NR), Image distortion (ID), radiometric Improvement Factors IF 1 and<br />

IF 2 for the Aqua MODIS band 27, 30 and 36<br />

– Aqua MODIS band 27 Aqua MODIS band 30 Aqua MODIS band 36<br />

in<strong>de</strong>x NR ID IF 1 IF 2 NR ID IF 1 IF 2 NR ID IF 1 IF 2<br />

moment matching 12.643 0.984 10.4393 12.0686 4.2363 0.9991 6.4180 12.6573 1.90 0.9997 3.1844 2.3252<br />

histogram (IMAPP) 33.8517 0.9828 11.5966 18.1798 4.4847 0.9839 5.5991 12.6135 1.9593 0.9957 3.2537 2.4242<br />

IFOV 1 12.1310 0.9890 9.3335 11.9212 1.8756 0.9646 -3.6762 8.9016 1.1396 0.8627 -1.8263 2.3097<br />

frequency filtering 510.5828 0.9854 10.3039 24.8529 7.8478 0.9874 6.35 17.8977 3.2048 0.9868 0.7912 10.6393<br />

wavelet thresholding 1284 0.9176 9.7060 25.6599 14.8403 0.9198 6.2033 18.1075 5.6922 0.8790 -0.1253 11.6181<br />

Proposed 282.1545 0.9788 12.0818 36.1859 3.8588 0.9985 10.3925 14.4371 2.3648 0.9868 10.4755 15.8115<br />

107


108 4. A Variational approach for the <strong>de</strong>striping issue<br />

Figure 4.17 – Destriping results with the UVDM on the MODIS data set (TL) Terra<br />

band 27 (TR) UVMD on Terra band 27 (CL) Terra band 30 (CR) UVMD on Terra<br />

band 30 (BL) Terra band 33 (BR) UVMD on Terra band 33


Figure 4.18 – Destriping results with the UVDM on the MODIS data set (TL) Aqua<br />

band 27 (TR) UVMD on Aqua band 27 (CL) Aqua band 30 (CR) UVMD on Aqua<br />

band 30 (BL) Aqua band 36 (BR) UVMD on Aqua band 36<br />

109


110 4. A Variational approach for the <strong>de</strong>striping issue


111<br />

Chapitre 5<br />

Application : Restoration of Aqua<br />

MODIS Band 6<br />

5.1 Context<br />

In light of the <strong>de</strong>striping results reported in the chapter 3, it is clear that the quality<br />

of Aqua MODIS data is higher than that collected by Terra MODIS and have been given<br />

preference for many remote sensing applications including ocean colour product generation.<br />

This is mainly due to a better pre-launch calibration and limited non linear effects of<br />

the individual <strong>de</strong>tectors response. Nevertheless, a major issue have been reported on one<br />

of Aqua MODIS bands. 15 of the 20 <strong>de</strong>tectors of Aqua MODIS band 6 (data available at<br />

a resolution of 500 m) are either non functional or noisy. Non functional <strong>de</strong>tectors do not<br />

measure any information and the resulting missing lines, in addition to functional <strong>de</strong>tectors<br />

mis-calibrations, induces a sharp striping pattern accross the entire swath (figure 5.1).<br />

Due to the placement of band 6 (1.6281.652 µm) in the electromagnetic spectrum, many<br />

important applications are affected by this <strong>de</strong>gradation. Two aerosol products, M and<br />

A-aerosols and the corresponding aerosol optical <strong>de</strong>pth are <strong>de</strong>rived over the ocean by the<br />

CERES Science Team using band 6 and band 1 [Tanre et al., 1997], [Ignatov et al., 2005].<br />

Forest biomass estimation and canopy water stress also relie on SWIR range measurements<br />

[Fensholt and Sandholt, 2003]. The relationship between biomass and MODIS band 6 reflectance<br />

was characterized in [Baccini et al., 2004]. More importantly, in the context of<br />

climate change, the missing and noisy scans of Aqua MODIS band 6 is a serious problem<br />

for MODIS snow products. Snow cover information have gained attention of many scientific<br />

studies due to its impact on climate change. The spatial distribution and temporal<br />

evolution of snow provi<strong>de</strong>s a valuable information on the amount of snowmelt, used as<br />

input for water management applications and hydrological cycle studies. In addition, the<br />

Earth’s albedo, amount of solar energy reflected back into the atmosphere, is partially regulated<br />

the snow cover. Depending on their fractional cover, snow and ice high reflectivity<br />

acts as a shield against sun radiation thus reducing the Earth surface temperature. Most


112 5. Application : Restoration of Aqua MODIS Band 6<br />

Figure 5.1 – (Left) Image from Aqua MODIS band 6 showing 4 non functional <strong>de</strong>tectors<br />

(Right) Image from Aqua band 6 after removal of missing lines : functional <strong>de</strong>tectors are<br />

contaminated with stripe noise<br />

approaches related to the analysis of snow cover use the Normalized Difference Snow In<strong>de</strong>x<br />

(NDSI). This in<strong>de</strong>x exploits snow high reflectivity in the visible (0.5-0.7 µm) and its low<br />

reflectance in the SWIR (1-4 µm). Defined as a spectral band ratio, the NDSI allows a<br />

robust separation of snow from clouds and also reduces the influence of atmospheric effects<br />

and viewing geometry [Salomonson and Appel]. On MODIS, the NDSI can be expressed<br />

as the difference of reflectances ρ measured in the visible band 4 (0.555 µm) and a SWIR<br />

band such as band 6 divi<strong>de</strong>d by the sum of the two reflectances :<br />

NDSI 16 = ρ 4 − ρ 6<br />

ρ 4 + ρ 6<br />

(5.1)<br />

The evaluation of the NDSI in<strong>de</strong>x on the Aqua sensor is problematic due to the non<br />

functional <strong>de</strong>tectors. Nevertheless, to ensure complementary observations to Terra MODIS<br />

band 6 and provi<strong>de</strong> continuous monitoring and mapping of snow coverage, scientists have<br />

relied on Aqua MODIS band 7 (2.105-2.155 µm). This is a reasonnable alternative, given<br />

that snow has similar reflectance properties over bands 6 and 7, which happen to display a<br />

high correlation over land surfaces. As a result, Aqua-based snow coverage is <strong>de</strong>termined<br />

using a secondary NDSI <strong>de</strong>fined as :<br />

NDSI 17 = ρ 4 − ρ 7<br />

ρ 4 + ρ 7<br />

(5.2)<br />

It was however pointed out in [Hall et al.], that the reflectance of snow is slightly weaker<br />

in band 7 compared to band 6. Consequently, values of NDSI 17 are higher than NDSI 16<br />

and the estimation of snow coverage is compromised. Given the importance of retrieving


113<br />

Figure 5.2 – Location of the Terra MODIS scenes selected over snow-covered regions<br />

for the study of Aqua MODIS band 6 restoration (Left) Alaska, May 5, 2001 (Center)<br />

Labrador, Canada, November 7, 200 (Right) Siberia, Russia, May 24, 2001<br />

accurate quantitative variables associated with snow and ice components, and used in<br />

climate change numerical mo<strong>de</strong>ls, an alternative option to NDSI 17 is mandatory. To this<br />

purpose, restoration techniques for Aqua MODIS band 6 have been recently proposed.<br />

5.2 Existing restoration techniques<br />

5.2.1 Global interpolation<br />

[L. L. Wang and Nianzeng] were the first to investigate the relationship between bands<br />

6 and 7 and <strong>de</strong>monstrated the feasability of restoring Aqua MODIS band 6 missing data.<br />

The authors first remarqued that the difference of snow reflectance between Terra MODIS<br />

bands 6 and 7 is very close to the same difference on Aqua MODIS. Using observations<br />

consisting of Terra MODIS level 1B calibrated and geolocated radiances at TOA, polynomial<br />

regression was used to quantify the analytical relationship between Terra MODIS<br />

bands 6 and 7. Quantitative analysis conducted by the authors over snow covered areas,<br />

shows that TOA reflectances in Terra MODIS band 6/7, and NDSI 16 /NDSI 17 were highly<br />

correlated with correlation coefficients of 0.9821 and 0.9777 respectively. Linear, quadratic,<br />

cubic, and fourth-<strong>de</strong>gree polynomials were <strong>de</strong>rived from Terra bands 6/7 scatter plots in<br />

or<strong>de</strong>r to be applied to Aqua MODIS. Wang et al. suggest restoring Aqua MODIS band 6<br />

using cubic and quadratic polynomes of the corresponding band 7 reflectances :<br />

ρ 6 =1.6032ρ 3 7 − 1.9458ρ 2 7 +1.7948ρ 7 +0.012396<br />

ρ 6 = −0.70472ρ 2 7 +1.5369ρ 7 +0.025409<br />

(5.3)<br />

These analytical relationships between bands 6 and 7 were <strong>de</strong>rived fom Terra measurements<br />

over snow covered regions. It obviously do not account for land surface cover types,<br />

spectral characteristics, scanning geometry. Application of this approach to restore band 6


114 5. Application : Restoration of Aqua MODIS Band 6<br />

missing data over vegetation, clouds, <strong>de</strong>sert or oceanic surfaces require further refinement<br />

to distinguish surface cover types. In addition, it is assumed that the polynomial relation<br />

established on Terra data is transposable to Aqua. This assumption is all the more sensitive<br />

to unknown calibration differences between Terra and Aqua MODIS, and the striping<br />

noise <strong>de</strong>scribed in the previous chapters. Furthermore, the analytical relation is <strong>de</strong>rived<br />

without prior pre-processing of bands 7 and 6 for stripe noise removal.<br />

5.2.2 Local interpolation<br />

More recently, Rakwatin2009 proposed a restoration procedure for Aqua MODIS band<br />

6 that consists in three step :1) the <strong>de</strong>termination of non functioning <strong>de</strong>tectors ; 2) the<br />

correction of periodic stripes for functioning <strong>de</strong>tectors using histogram matching ; 3) the<br />

estimation of missing pixels via local cubic polynomial regression between band 6 and band<br />

7. Similarly to the approach proposed in [L. L. Wang and Nianzeng], the high correlation<br />

between band 6 and 7 reflectances is quantified using polynomial regression. However, the<br />

fitting is computed locally to account for land cover types. The restoring algorithm can<br />

be summarized with the following :<br />

1) For a <strong>de</strong>ad pixel x in band 6, a initial rectangular window of size 15 × 3 is centered<br />

at x. The minimum and maximum values of the window in band 7, respectively ρ min<br />

7<br />

and ρ max<br />

7 and their location x min , x max are <strong>de</strong>termined.<br />

2) If ρ min<br />

7 ≤ ρ 7 (x) ≤ ρ max<br />

7 , a local cubic polynomial function is calculated from the<br />

values ρ 7 (x min ), ρ 7 (x max ), ρ 6 (x min ) and ρ 6 (x max ) and used to estimated the values of<br />

ρ 6 (x)<br />

3) If ρ 7 (x) < ρ min<br />

7 , ρ 7 (x) > ρ max<br />

7 or if pixels x min or x max in band 6 are also <strong>de</strong>ad,<br />

the size of the analizing window is increased until these criteria are met.<br />

4) Step 1, 2 and 3 are repeated for every <strong>de</strong>ad pixel of band 6<br />

This local cubic interpolation procedure is illustrateed in figure 5.3.<br />

5.3 Proposed approach<br />

We propose here a simple methodology to estimate the value of Aqua MODIS band 6<br />

missing pixels. The approach is based on a concept wi<strong>de</strong>ly used in the field of hyperspectral<br />

image classification, spectral similarity. Data collected from MODIS can be perceived as a<br />

cube, where the third dimension represents the signal’s wavelenght and can also be used to<br />

extract useful information. In hyperspectral remote sensing, the <strong>de</strong>termination of surface<br />

composition requires the analysis of its reflectance spectrum and comparison with known<br />

field spectra via spectral matching techniques [Kruse et al.]. Although conditionned by the<br />

number of available spectral bands, this reasoning also applies to multispectral imagery.


115<br />

ρ 6 (x max )<br />

MODIS Band 6<br />

ρ 6 (x) <br />

ρ 6 (x min )<br />

ρ 7<br />

min<br />

ρ 7 (x)<br />

MODIS Band 7<br />

ρ 7<br />

max<br />

Figure 5.3 – Restoration procedure for Aqua MODIS band 6 proposed in (Rakwatin et<br />

al., 2009) and based on a local cubic interpolation<br />

For the issue of Aqua MODIS band 6 restoration, missing pixels can be estimated using<br />

spectrally similar pixels from functionning <strong>de</strong>tectors. To this purpose, let us first recall<br />

the <strong>de</strong>finition of few spectral similarity measures commonly used in hyperspectral image<br />

classification.<br />

5.3.1 Spectral similarity<br />

We <strong>de</strong>note by ρ(x) the reflectance of a given pixel x, which we consi<strong>de</strong>r as a vector in<br />

a n-dimensional space as :<br />

ρ(x) =(ρ 1 (x),ρ 2 (x), ..., ρ n (x)) T (5.4)<br />

Each component of the vector ρ(x) in (5.4) corresponds to the reflectance of pixel x in a<br />

given spectral band.<br />

5.3.1.1 Spectral Correlation Measure<br />

The Spectral Correlation Measure (SCM) was <strong>de</strong>fined in [<strong>de</strong>r Meero and Bakker] and<br />

is computed for two pixels x and y as :<br />

n ∑ n<br />

1<br />

SCM(x, y) =<br />

ρ(x)ρ(y) − ∑ n<br />

1 ρ(x) ∑ n<br />

√<br />

1<br />

∑ ρ(y)<br />

[n n<br />

1 ρ(x)2 − ( ∑ n<br />

1 ρ(x))2 ][n ∑ n<br />

1 ρ(y)2 − ( ∑ n<br />

(5.5)<br />

1 ρ(y))2 ]<br />

where n is the number of overlapping spectral bands. The SCM measures the correlation<br />

between the two vectors ρ(x) and ρ(y) and takes into account the mean value and variance<br />

of the overall spectral shape. The values of SCM are contained in the interval [-1,1].


116 5. Application : Restoration of Aqua MODIS Band 6<br />

5.3.1.2 Spectral Angle Measure<br />

The Spectral Angle Measure (SAM) is <strong>de</strong>fined in [Kruse et al.] as the following angle :<br />

( ∑n<br />

1<br />

SAM(x, y) = arcos<br />

ρ(x) ∑ )<br />

n<br />

1<br />

√∑ ρ(y)<br />

n<br />

1 ρ2 (x) ∑ n<br />

(5.6)<br />

1 ρ2 (y)<br />

The SAM is not very sensitive to pixel reflectance values and as such tends to reduce<br />

differences due to atmospheric effects or viewing geometry.<br />

5.3.1.3 Euclidian Distance Measure<br />

The Euclidian Distance Measure (EDM) between pixels x and y is computed as :<br />

∑<br />

EDM(x, y) = √ n (ρ(x) − ρ(y)) 2 (5.7)<br />

Unlike the SCM and SAM, the EDM <strong>de</strong>pends on the reflectance differences between pixels<br />

x and y.<br />

5.3.1.4 Spectral Information Divergence Measure<br />

The Spectral Information Divergence Measure (SIDM) is a stochastic in<strong>de</strong>x that measures<br />

the distance between the probability distribution of spectrums associated with pixels<br />

x and y :<br />

SIDM(x, y) =D(x||y)+D(y||x) (5.8)<br />

D(x||y) is the relative entropy of y with respect to x computed as :<br />

1<br />

D(x||y) =<br />

n∑<br />

p i (x)D i (ρ i (x)||ρ i (y)) =<br />

i=1<br />

n∑<br />

p i (x)(I(ρ i (x)) − I(ρ i (y))) (5.9)<br />

i=1<br />

where<br />

p i (x) =<br />

ρ i (x)<br />

∑ n<br />

j=1 ρ j(x) and I(ρ i(x)) = −log p i (x) (5.10)<br />

The measure I(ρ i (x)) is the self-information of x in the spectral band i.<br />

5.3.2 Spectral inpainting<br />

It is clear from the limitations discussed in [L. L. Wang and Nianzeng] that a reliable<br />

restoration of Aqua MODIS band 6 requires accurate distinction of land cover types which<br />

can not be achieved using only the correlation between band 6 and 7. To compute robust<br />

estimate of missing pixels values from band 6, we suggest to rely solely on spectral information<br />

available in other bands where all the <strong>de</strong>tectors are knnown to be functional. Dead


117<br />

pixels can then be restored with a spectral-based non local approach.<br />

Non local or neighborhood filters have been first introduced in the context of image processing<br />

for <strong>de</strong>noising applications [Yaroslavsky and E<strong>de</strong>n, 1996]. In the general case, we<br />

consi<strong>de</strong>r a noisy image f <strong>de</strong>fined in a boun<strong>de</strong>d domain Ω of R 2 . The <strong>de</strong>noised value of a<br />

pixel x is obtained as an average of pixels that have a value close to f(x). Such radiometric<br />

neighborhood is used in the Yarolavsky non local filter and the <strong>de</strong>noised value of a pixel x<br />

is given by :<br />

Y NF h,r (f(x)) = 1 ∫<br />

f(y)e − |f(y)−f(x)|2<br />

h<br />

C(x)<br />

2 (5.11)<br />

B r(x)<br />

where B r (x) is a ball of radius r centered at pixel x, C(x) is a normalization factor<br />

and h is a filtering parameter that controls the <strong>de</strong>cay of the weighting coefficients. More<br />

recently, the high redunduncy observed in natural images was used for texture synthesis<br />

in [Efros and Leung, 1999]. Efros and Leung algorithm exploits non local self similarities<br />

and is particularly suited for inpainting applications. This very principle is used in the<br />

Non Local Means <strong>de</strong>noising algorithm introduced in [Bua<strong>de</strong>s et al., 2005]. With analogy<br />

to the Yaroslavsky filter, the <strong>de</strong>noised value of a pixel x is obtained as a weighted average<br />

of pixels with a similar local configuration as :<br />

NL(f(x)) = 1<br />

C(x)<br />

∫<br />

Ω<br />

f(y)w(x, y)dy (5.12)<br />

where w(x, y) is a weighting between pixels x and y obtained by comparing intensity<br />

patches of windows centered at x and y as :<br />

( ∫<br />

G a (z)|f(x + z) − f(y + z)| 2 )<br />

w(x, y) =exp −<br />

dz<br />

(5.13)<br />

Ω<br />

where G a is a gaussian function with standard <strong>de</strong>viation a.<br />

The concept of neighborhood used in nonlocal filters for image processing applications is<br />

a valuable tool for the restoration of Aqua MODIS band 6. In<strong>de</strong>ed, the value of a <strong>de</strong>ad<br />

pixel can be <strong>de</strong>termined from the set of pixels with similar spectral properties. For a given<br />

Spectral Similarity Measure (SSM), we <strong>de</strong>fine the spectral neighborhood of a pixel x as :<br />

h 2<br />

S(x) ={y ∈ Ω|SSM(x, y)


118 5. Application : Restoration of Aqua MODIS Band 6<br />

Reflectance<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

Land<br />

Ice over land<br />

Sea ice<br />

Coastal waters<br />

Deep ocean<br />

Cloud<br />

Reflectance<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

Ice 1<br />

Ice 2<br />

Ice 3<br />

Ice 4<br />

Ice 5<br />

0.1<br />

0.1<br />

0<br />

3 4 5 6 7<br />

MODIS Spectral band<br />

0<br />

3 4 5 6 7<br />

MODIS Spectral band<br />

Figure 5.4 – Illustration of spectral similarity (Left) Reflectance spectrum of pixels<br />

selected over areas of different geophysical nature (Right) Reflectance spectrum of pixels<br />

all selected over a homogeneous snow-covered area<br />

band 6. For example, the EDM of two pixels x and y is given by :<br />

∑<br />

EDM(x, y) = √<br />

7 (ρ i (x) − ρ i (y)) 2 (5.16)<br />

i=3,i≠6<br />

Additionally, for a missing pixel x from band 6, we <strong>de</strong>fine its spectral neighborhood as :<br />

S(x) ={y ∈ Ω|SSM(x, y)


119<br />

Figure 5.5 – Restoration of Terra Modis Band 6 with synthetic non functional <strong>de</strong>tectors.<br />

From left to right : Original image (Alaska), restored with local interpolation, restored with<br />

gobal interpolation and restored with spectral inpainting. Striping is a clear indication of<br />

poor estimation of missing pixels value<br />

Figure 5.6 – From left to right : Original image (Labrador), restored with local interpolation,<br />

restored with gobal interpolation and restored with spectral inpainting.<br />

Figure 5.7 – From left to right : Original image (Siberia), restored with local interpolation,<br />

restored with gobal interpolation and restored with spectral inpainting.<br />

5.4 Experimental results<br />

5.4.1 Validation of spectral inpainting using NDSI measurements<br />

Three Terra MODIS images have been selected to evaluate the restoration of Aqua<br />

MODIS band 6 obtained with [L. L. Wang and Nianzeng], [Rakwatin et al.] and the pro-


120 5. Application : Restoration of Aqua MODIS Band 6<br />

1<br />

1<br />

0.8<br />

0.8<br />

Simulated NDSI<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

Simulated NDSI<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

−0.4<br />

−0.4<br />

−0.6<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1<br />

Measured NDSI<br />

−0.6<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1<br />

Measured NDSI<br />

Simulated NDSI<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

−0.4<br />

−0.6<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1<br />

Measured NDSI<br />

Figure 5.8 – Scatter plots of simulated and measured NDSI from Terra MODIS band 6<br />

(Labrador) (Left) Local interpolation (Right) Global interpolation (Bottom) Spectral<br />

inpainting<br />

posed approache refered to hereafter as spectral inpainting. The locations of these scenes<br />

coinci<strong>de</strong> with those used in the work of [Salomonson and Appel] where MODIS data were<br />

used to estimate fractional snow cover. We recall that <strong>de</strong>tectors of Terra MODIS band<br />

6 are all functional and therefore, restoring algorithms <strong>de</strong>scribed above are applied on<br />

images where measurements from 4 <strong>de</strong>tectors have been set to zero to simulate non functional<br />

<strong>de</strong>tectors. The true and estimated images for Terra band 6 are then combined with<br />

band 4 data in or<strong>de</strong>r to compute the NDSI in<strong>de</strong>x. The resulting measured and simulated<br />

NDSI in<strong>de</strong>xes are then compared in terms of absolute mean difference, root mean square<br />

error and correlation coefficients to evaluate the efficiency of each approach. Results are<br />

reported in table .<br />

Let us mention few notes on the computational aspect of the spectral inpainting method.<br />

In practice, for a missing pixel x from band 6, a window of limited size (instead of the<br />

entire image domain Ω) is used to <strong>de</strong>termine the spectral similarity between its pixels and<br />

pixel x. Since the proposed approach does not relie on spatial information, standard periodic<br />

or symmetric extensions of boundaries cannot be used. Instead, the search window can


e square or rectangular <strong>de</strong>pending on the position of the pixel with respect to the image<br />

bor<strong>de</strong>rs. Alternatively, the image can be fragmented into blocks of K × K pixels and the<br />

algorithm is applied separately for each block with a search window composed of the entire<br />

domain of the subimage. The choice of K then <strong>de</strong>pends on the available computational<br />

power. Results illustrated in this study were obtained with K = 200.<br />

Another pratical aspect of the spectral inpainting method lies in the selection of the parameter<br />

N in equation (5.20). The choice N = 1 holds a physical meaning and ensures<br />

radiometric coherence as it corresponds to the case where a missing pixel from band 6 is<br />

given the same value as the most spectrally similar available pixel in the same band. Furthermore,<br />

all spectral bands (3, 4, 5, 6 and 7) have been <strong>de</strong>striped prior to the restoration<br />

procedure and, therefore all measurements used in the spectral inpainting can be assumed<br />

to be noise free.<br />

The spectral similarity measures <strong>de</strong>fined in the previous section have all been tested except<br />

the Spectral Correlation Measures which requires overlapping spectral bands.<br />

The two existing techniques for the restoration of Aqua MODIS band 6 are entirely based<br />

on the correlation of measurements between bands 6 and 7. The global polynomial regression<br />

proposed in [L. L. Wang and Nianzeng] only applies to snow covered regions and do<br />

not make any distinction in land cover types. Estimation of band 6 missing lines via equations<br />

(5.3) results in strong striping in the restored image even over snow covered areas.<br />

Improved results are obtained when the global polynomial relationship between band 6<br />

and 7 is estimated using data from the current swath.<br />

Another limitation of the global interpolation method proposed by Wang et al., is attached<br />

to the assumption that the relationship between bands 6 and 7 of Terra MODIS can be<br />

directly transposed to Aqua MODIS. In pratice, this assumption do not hold because of<br />

calibration differences between Terra and Aqua MODIS instruments.<br />

The local interpolation procedure <strong>de</strong>scribed in [Rakwatin et al.] is a refinement of the<br />

global interpolation. To account for land cover types, the estimation of missing pixels is<br />

<strong>de</strong>rived from a polynomial regression between bands 6 and 7, restricted to a local windows.<br />

Results achieved with this technique are often disapointing, as confirmed by the<br />

strong striping visible in the restored band 6. This can be attributed to the incoherent<br />

mixing of spatial and spectral information. In<strong>de</strong>ed, the value of a <strong>de</strong>ad pixel from band<br />

6 is <strong>de</strong>termined from minimum and maximum values of band 7 in a local window, which<br />

actually correspond to spectrally distant pixels. With such a procedure, the spatial locality<br />

does not garantee any distinction in land cover types. On the contrary, the approach<br />

of [Rakwatin et al.] amounts to a spatio-spectral interpolation where pixels used for the<br />

restoration are spectrally distant and spatially close.<br />

The spectral inpainting method <strong>de</strong>scribed above enables a robust estimation of missing<br />

lines as it is only based on spectral information. As illustrated in figures 5.5, 5.6 and 5.7<br />

restoration with spectral inpainting provi<strong>de</strong>s images visually i<strong>de</strong>ntical to the original band<br />

6, and unlike global and local interpolation does not display any stripes. Scatter plots (figure<br />

) and quantitative analysis reported in table indicate reliable results obtained<br />

with the spectral inpainting for all three images of Alaska, Labrador and Siberia.<br />

121


122 5. Application : Restoration of Aqua MODIS Band 6<br />

5.4.2 Assessing the impact of stripe noise<br />

The UVDM <strong>de</strong>scribed in the previous chapter was applied to all bands used by the<br />

spectral inpainting technique. Nevertheless, the striping on Terra MODIS bands 3 to 7<br />

is extremely weak and its impact on the retrieval of NDSI in<strong>de</strong>x is not substantial. We<br />

propose here to quantitatively evaluate the necessity of a robust <strong>de</strong>striping methodology<br />

prior to the generation of high level products. To this purpose, the UVDM was used to<br />

extract a stripe noise from Terra MODIS emissive band 27. This band was specifically<br />

chosen beacause its <strong>de</strong>tectors suffer from strong nonlinear effects that standard <strong>de</strong>striping<br />

techniques fail to correct. The stripe noise from band 27 is then injected in bands 5, 6 and<br />

7 with different intensity levels so that the Peak Signal-to-Noise ratio (PSNR) of images<br />

from each of these bands is equal to 5, 10, 15, 20, 25, 30 and 35 dB. The spectral inpainting<br />

method is then applied to restore band 6 missing lines with and without preliminary<br />

<strong>de</strong>striping. The spectral similarity measure used for this experiment is the EDM as it has<br />

shown to provi<strong>de</strong> better results compared to SAM and SDI. In addition to the UVDM,<br />

histogram matching is also used to illustrate the impact of residual stripes on the retrieval<br />

of NDSI in<strong>de</strong>xes.<br />

Simulated and measured NDSI values have been compared for each level of striping using<br />

three scenarios :<br />

- Images from bands 5, 6 and 7 are not <strong>de</strong>striped prior to spectral inpainting<br />

- Images from bands 5, 6 and 7 are <strong>de</strong>striped with the histogram matching prior to spectral<br />

inpainting<br />

- Images from bands 5, 6 and 7 are <strong>de</strong>striped using the UVDM prior to spectral inpainting<br />

Absolute mean difference (AMD), Root Mean Square Errors (RMSE) and Correlation<br />

(CORR) coefficients for the three scenarios are reported in table . Many interesting<br />

observations can be ma<strong>de</strong> from these results (also plotted in figure 5.10) to un<strong>de</strong>rscore the<br />

benefits of a reliable <strong>de</strong>striping algorithm.<br />

It can be seen that for a reasonable amount of stripe noise (PSNR≥35 dB), the histogram<br />

matching techniques does not provi<strong>de</strong> any improvement in the estimation of the NDSI<br />

in<strong>de</strong>x. In case of very strong striping (PSNR=5dB), the application of the UVDM before<br />

the restoration of band 6 results in NDSI values as accurate as those obtained with the<br />

histogram matching when the stripe noise corresponds to a PSNR=25dB. In addition, the<br />

correlation coefficient obtained with the UVDM, remains above 0.99 for all tested levels of<br />

striping. The extreme case of PSNR=5dB, illustrated in figure 5.9 shows how the UVDM<br />

enables the preservation of the structures contained in the original image.


Figure 5.9 – (TL) Original image from Terra MODIS band 6 (TR) Extreme striping on<br />

band 6 (PSNR=5dB) prior to synthetic non functional <strong>de</strong>tectors. Bands 5 and 7 are contaminated<br />

with a similar striping (BL) Restoration with spectral inpainting after <strong>de</strong>striping<br />

with histogram matching (BR) Restoration with spectral inpainting after <strong>de</strong>striping with<br />

UVDM<br />

123


– Alaska Labrador Siberia<br />

Restoration AMD RMSE CORR AMD RMSE CORR AMD RMSE CORR<br />

Wang et al. 0.0087 9.0000 0.9944 0.0031 12.8145 0.9901 4.3822e-004 5.3724 0.9987<br />

Rakwatin et al. 0.0120 12.4084 0.9893 0.0188 16.2538 0.9859 0.0230 11.9939 0.9950<br />

Proposed SAM 5.5582e-004 6.4964 0.9969 1.7451e-004 6.9496 0.9971 0.0011 7.4547 0.9974<br />

SDIM 6.5480e-004 6.5729 0.9968 3.3377e-004 6.5852 0.9974 0.0016 5.4037 0.9986<br />

EDM 1.7882e-004 3.0273 0.9993 6.5103e-005 4.3314 0.9989 0.0021 2.3593 0.9998<br />

Table 5.1 – Absolute mean difference (AMD), Root Mean Square Error (RMSE) and Correlation<br />

(CORR) between simulated and measured NDSI values using different restoration techniques for<br />

Aqua MODIS band 6<br />

– Without <strong>de</strong>striping Histogram matching UVDM<br />

Stripe noise level AMD RMSE CORR AMD RMSE CORR AMD RMSE CORR<br />

PSNR=5dB 1.1632 7.2247e+004 0.0054 0.0136 36.5744 0.9407 0.0014 10.9336 0.9944<br />

PSNR=10dB 0.3251 4.5093e+003 0.0081 0.0078 24.8579 0.9731 0.0020 7.4031 0.9975<br />

PSNR=15dB 0.0572 81.8125 0.7653 0.0055 17.4133 0.9878 0.0017 5.7920 0.9985<br />

PSNR=20dB 0.0157 39.0451 0.9335 0.0046 12.3446 0.9950 0.0012 4.9865 0.9989<br />

PSNR=25dB 0.0057 22.0486 0.9777 0.0042 9.7847 0.9976 8.1850e-004 3.5721 0.9994<br />

PSNR=30dB 0.0030 13.5149 0.9915 0.0039 8.4876 0.9986 5.0663e-004 3.1400 0.9995<br />

PSNR=35dB 0.0021 7.6345 0.9973 0.0036 7.6683 0.9991 3.8126e-004 2.4931 0.9997<br />

Table 5.2 – Absolute mean difference (AMD), Root Mean Square Error (RMSE) and Correlation<br />

(CORR) between simulated and measured NDSI values using spectral inpainting with EDM and different<br />

levels of stripe noise in bands 5, 6 and 7 prior to restoration<br />

124 5. Application : Restoration of Aqua MODIS Band 6


125<br />

40<br />

1<br />

40<br />

1<br />

35<br />

30<br />

0.99<br />

35<br />

30<br />

0.99<br />

RMSE<br />

25<br />

20<br />

15<br />

RMSE<br />

CORR<br />

0.98<br />

0.97<br />

CORR<br />

RMSE<br />

25<br />

20<br />

15<br />

RMSE<br />

CORR<br />

0.98<br />

0.97<br />

CORR<br />

10<br />

5<br />

0.96<br />

10<br />

5<br />

0.96<br />

0<br />

0.95<br />

5 10 15 20 25 30 35<br />

Level of stripe noise (dB)<br />

0<br />

0.95<br />

5 10 15 20 25 30 35<br />

Level of stripe noise (dB)<br />

Figure 5.10 – RMSE and CORR values between simulated and measured NDSI values<br />

obtained for the Alaska image when different levels of stripe noise are injected in bands<br />

5, 6 and 7. Band 6 is restored using the spectral inpainting method with the EDM. Both<br />

histogram matching (Left) and the UVDM (Right) are used to remove stripe noise prior<br />

to the restoration. In presence of extreme striping (PSNR=5dB) the UVDM allows an<br />

estimation of the NDSI in<strong>de</strong>x as accurate as the one provi<strong>de</strong>d by the histogram matching<br />

technique for a stripe noise level corresponding to a PSNR=25dB


126 5. Application : Restoration of Aqua MODIS Band 6


127<br />

Conclusion<br />

Contribution<br />

Perspectives and furtur work


128 Conclusion


129<br />

Bibliographie<br />

R. Acar and C. R. Vogel. Analysis of boun<strong>de</strong>d variation penalty methods for ill-posed<br />

problems. In Inverse Problems, pages 1217–1229, 1994.<br />

J. Aiken, G. Moore, C. Trees, S. Hooker, and D. Clark. The seawifs czcs-type pigment<br />

algorithm. In SeaWiFS technical report series, page 29, 1995.<br />

L. Alvarez, P. L. Lions, and J. M. Morel. Image selective smoothing and edge <strong>de</strong>tection<br />

by nonlinear diffusion. ii. SIAM J. Numer. Anal., 29 :845–866, 1992.<br />

P. Antonelli, M. di Bisceglie, R. Episcopo, and C. Galdi. Destriping MODIS data using<br />

IFOV overlapping. 7(3) :4568–4571, 2004.<br />

J. F. Aujol, G. Aubert, L. Blanc-Féraud, and A. Chambolle. Image <strong>de</strong>composition into<br />

a boun<strong>de</strong>d variation component and an oscillating component. J. Math. Imaging Vis.,<br />

22 :71–88, 2005.<br />

A. Baccini, M. A. Friedl, C. E. Woodcock, R. Warbington, and L. Remer. Forest biomass<br />

estimation over regional scales using multisource data. J. Geophys. Res. Lett., 31(10) :<br />

L10501.1–L10501.4, 2004.<br />

P. Blomgren, T. F. Chan, P. Mulet, and C. Wong. Total variation image restoration :<br />

Numerical methods and extensions. In in Proc. IEEE ICIP, pages 384–387, 1997.<br />

Y. Boykov, O. Veksler, and R. Zabih. Fast approximate energy minimization via graph<br />

cuts. IEEE Trans. Pattern Anal. Mach. Intell., 23 :1222–1239, 2001.<br />

R. Bracewell. The fourier transform and its applications. In Mcgraw, 1986.<br />

A. Bua<strong>de</strong>s, B. Coll, and J. M. Morel. A review of image <strong>de</strong>noising algorithms, with a new<br />

one. Simul, 4 :490–530, 2005.<br />

F. Catté, P. L. Lions, J. M. Morel, and T. Coll. Image selective smoothing and edge<br />

<strong>de</strong>tection by nonlinear diffusion. SIAM J. Numer. Anal., 29 :182–193, 1992.


130 BIBLIOGRAPHIE<br />

A. Chambolle. An algorithm for total variation minimization and applications. J. Math.<br />

Imaging Vis., 20 :89–97, 2004.<br />

A. Chambolle and P. L. Lions. Image recovery via total variation minimization and related<br />

problems. Numer. Math., 76(2) :167–188, 1997.<br />

T. F. Chan, A. M. Yip, and F. E. Park. Simultaneous total variation image inpainting<br />

and blind <strong>de</strong>convolution. International Journal of Imaging Systems and Technology, 15<br />

(1), 2005.<br />

J. S. Chen, Y. Shao, H. D. Guo, W. Wang, and B. Zhu. Destriping CMODIS data by<br />

power filtering. 41(9) :2119–2124, 2003.<br />

E. N. Coleman, Jr., and R. Jain. Obtaining 3-dimensional shape of textured and specular<br />

surfaces using four-source photometry. Computer Graphics and Image Processing, 18<br />

(4) :309–328, 1982.<br />

C. Cox and W.Munk. Measurement of the roughness of the sea surface from photographs<br />

of the suns glitter. J. Opt. Soc. Am, 44(11) :838–850, 1954.<br />

R. E. Crippen. A simple spatial filtering routine for the cosmetic removal of scan line<br />

noise from Lansat TM P-tape imagery. Photogramm. Eng. Remote Sens., 55(3) :327–<br />

331, 1989.<br />

I. Daubechies and G. Teschke. Variational image restoration by means of wavelets :<br />

simultaneous <strong>de</strong>composition, <strong>de</strong>blurring and <strong>de</strong>noising. Appl. Comp. Harm. Anal., 19 :<br />

1–16, 2005.<br />

S. D.Babacan, R. Molina, and A. K. Katsaggelos. Total variation super resolution using<br />

a variational approach. In IEEE International Conf. on Image Processing 2008, San<br />

Diego, USA, October 2008.<br />

F. V. <strong>de</strong>r Meero and W. Bakker.<br />

R. Deriche and O. Faugeras. Les edp en traitement <strong>de</strong>s images et vision par ordinateur.<br />

In Technical Report 2697, INRIA, 1995.<br />

M. di Bisceglie, R. Episcopo, C. Galdi, and S. L. Ullo. Destriping MODIS data using<br />

overlapping field-of-view method. 47(2) :637–651, 2009.<br />

D. C. Dobson and C. R. Vogel. Convergence of an iterative method for total variation<br />

<strong>de</strong>noising. SIAM J. Numer. Anal., 34 :1779–1791, 1997.<br />

D. L. Donoho and I. M. Johnstone. I<strong>de</strong>al spatial adaptation by wavelet shrinkage. Biometrika,<br />

81(3) :425–455, 1994.


131<br />

A. Efros and T. Leung. Texture synthesis by non-parametric sampling. In In International<br />

Conference on Computer Vision, pages 1033–1038, 1999.<br />

G. S. Fargion and C. R. McClain. Modis validation, data merger and other activities<br />

accomplished by the simbios project : 20022003. In NASA TM-2003-212249.<br />

R. Fensholt and I. Sandholt. Derivation of a shortwave infrared water stress in<strong>de</strong>x from<br />

modis near- and shortwave infrared data in a semiarid environment. Remote Sensing of<br />

Environment, 87(1) :111–121, 2003.<br />

R. T. Frankot and R. Chellappa. A method for enforcing integrability in shape from<br />

shading algorithms. IEEE Transactions on Pattern Analysis and Machine Intelligence,<br />

10 :439–451, 1988.<br />

H. Fukushima, M. Toratani, H. Murakami, P. Deschamps, R. Frouin, and A. Tanaka.<br />

Evaluation of a<strong>de</strong>os-ii gli ocean color atmospheric correction using simbada handheld<br />

radiometer data. J. Oceanogr., 63(3) :533–543, 2007.<br />

W. W. Gregg. Analysis of orbit selection for seawifs : Ascending versus <strong>de</strong>scending no<strong>de</strong>.<br />

In NASA Tech. Memo. 104556, volume 2, 1992.<br />

W. W. Gregg and R. Woodward. Improvements in coverage frequency of ocean color :<br />

Combining data from seawifs and modis. IEEE Trans. Geosci. Remote Sens., 26(4) :<br />

13501353, 2007.<br />

F. Guichard and F. Malgouyres. Total variation based interpolation. In Proceedings of<br />

Eusipcoapos ;98, pages 1741–1744, 1998.<br />

L. Gumley. Scanner Characteristics Image Artifacts Destriping. In Western Australian<br />

Satellite Technology and Applications Consortium MODIS Workshop, Nov. 2002.<br />

S. L. Haines, G. J. Jedlovec, and S. M. Lazarus. A modis sea surface temperature composite<br />

for regional applications. IEEE Trans. Geosci. Remote Sens., 45(9) :2919–2927, 2007.<br />

D. K. Hall, G. A. Riggs, and V. V. Salomonson.<br />

R. M. Haralick and L. T. Watson. A facet mo<strong>de</strong>l for imag data. Comput. Graph. Image<br />

Process., 15(12) :113–129, 1981.<br />

G. Healey and R. Jain. Depth recovery from surface normals. In ICPR’84, 1984.<br />

J. Hedley, A. Harborne, and P. Mumby. Simple and robust removal of sun glint for mapping<br />

shallow-water benthos. Int. J. Remote Sens., 26(10) :2107–2112, 2005.<br />

J. Ho, J. Lim, M. Yang, and D. Kriegman. Integrating surface normal vectors using fast<br />

marching method. In Lecture Notes in Computer Science, page 239250, 2006.


132 BIBLIOGRAPHIE<br />

E. Hochberg, S. Andrefouet, and M. Tyler. Sea surface correction of high spatial resolution<br />

ikonos images to improve bottom mapping in near-shore environments. IEEE Trans.<br />

Geosci. Remote Sens., 41(7) :1724–1729, 2003.<br />

B. K. P. Horn. Height and gradient from shading. International Journal of Computer<br />

Vision, 5(1) :37–75, 1990.<br />

B. K. P. Horn and M. J. Brooks. The variational approach to shape from shading. Comput.<br />

Vision Graph. Image Process., 33 :174–208, 1986.<br />

B. K. P. Horn and R. J. Woodham. Destriping Landsat MSS images by histogram modification.<br />

Comput. Graph. Image Process., 10(1) :69–83, 1979.<br />

A. Huete. Overview of the radiometric and biophysical performance of the modis vegetation<br />

indices. Remote Sensing of Environment, 83(1-2) :195–213, nov 2002.<br />

A. Ignatov, P. Minnis, N. Loeb, B. Wielicki, W. Miller, S. Sun-Mack, D. Tanré, L. Remer,<br />

I. Laszlo, and E. Geier. Two modis aerosol products over ocean on the terra and aqua<br />

ceres ssf datasets. Journal of the Atmospheric Sciences, 62(4) :1008–1031, 2005.<br />

M. Kazhdan and H. Hoppe. Streaming multigrid for gradient-domain operations on large<br />

images. In Proc. of ACM SIGGRAPH, 2008.<br />

J. J. Koen<strong>de</strong>rink. The structure of images. Biological Cybernetics, 50 :363–370, 1984.<br />

F. Kruse, A. Lefkoff, J. Boardman, K. Hei<strong>de</strong>brecht, A. Shapiro, P. Barloon, and A. Goetz.<br />

T. Kutser, E. Vahtmäe, and J. Praks. Sun glint correction method for hyperspectral<br />

imagery containing areas with non-negligible water leaving nir signal. Remote Sens.<br />

Environ., 113(10) :2267–2274, 2009.<br />

E. J. Kwiatkowska and G. S. Fargion. Merger of ocean color data from multiple satellite<br />

missions within the simbios project. In SPIE Symposium : Remote Sensing of the<br />

Atmosphere, Ocean, Environment, and Space. Ocean Remote Sensing and Applications,<br />

page 168182, 2002.<br />

X. S. H. X. X. Y. L. L. Wang, J. J. Qu and C. Nianzeng.<br />

C. H. Li and P. R. S. Tam. A facet mo<strong>de</strong>l for imag data. Pattern Recogn., 33(2) :281–293,<br />

2000.<br />

B. Luo, J. F. Aujol, Y. Gousseau, S. Ladjal, and H. Maitre. Resolution-in<strong>de</strong>pen<strong>de</strong>nt<br />

characteristic scale <strong>de</strong>dicated to satellite images. IEEE transactions on image processing,<br />

16 :2503–2514, 2007.<br />

D. Lyzenga, N. Malinas, and F. Tanis. Multispectral bathymetry using a simple physically<br />

based algorithm. IEEE Trans. Geosci. Remote Sens., 44(8) :2251–2259, 2006.


133<br />

C. M. E. and W. W. Gregg. Comparison of global chlorophyll climatologies : In situ,<br />

czcs, blen<strong>de</strong>d in situ-czcs and seawifs. International Journal of Remote Sensing, 24(5) :<br />

969–991, 2003.<br />

S. Mallat. Une exploration <strong>de</strong>s signaux en on<strong>de</strong>lettes. 2000.<br />

Y. Meyer. Oscillating patterns in image processing and nonlinear evolution equations. In<br />

University Lecture Series, volume 22, 2002.<br />

F. Montagner, V. Billat, and S. Bélanger. Meris atbd 2.13 sun glint flag algorithm. 2003.<br />

URL http://envisat.esa.int/instruments/meris/atbd/atbd_2_13.pdf.<br />

G. P. Nason and B. W. Silverman. The stationary wavelet transform and some statistical<br />

applications. pages 281–300. Springer-Verlag, 1995.<br />

M. Nikolova. A variational approach to remove outliers and impulse noise. J. Math.<br />

Imaging Vis., 20 :99–120, 2004.<br />

M. Nikolova and R. H. Chan. The equivalence of half-quadratic minimization and the<br />

gradient linearization iteration. IEEE Transactions on Image Processing, 16(6) :1623–<br />

1627, 2007.<br />

M. Nitzberg and T. Shiota. Nonlinear image filtering with edge and corner enhancement.<br />

IEEE Trans. Pattern Anal. Mach. Intell., 14 :826–833, 1992.<br />

K. N. Nordstrom. Biased anisotropic diffusion-a unified regularization and diffusion. Image<br />

and Vision Computing, 4 :318–327, 1990.<br />

S. Osher, A. Sole, and L. Vese. Image <strong>de</strong>composition<br />

and restoration using total variation minimization and the<br />

h − 1norm. Multiscale Mo<strong>de</strong>ling and Simulation : A SIAM Interdisciplinary Journal, 1:<br />

349 −−370, 2002.<br />

S. Osher, M. Burger, D. Goldfarb, J. Xu, and W. Yin. An iterative regularization method<br />

for total variation-based image restoration <strong>de</strong>compositions in. Multiscale Mo<strong>de</strong>l. Simul., 4<br />

(2) :460–489, 2005.<br />

M. Ottaviani, H. A. Ei<strong>de</strong>, K. H. Stamnes, W. Li, W. J. Wiscombe, and W. Su. Improving<br />

the <strong>de</strong>scription of sunglint for accurate prediction of top-of-the-atmosphere radiances. J.<br />

Quant. Spectrosc. Radiat. Transf., 109(14) :2364–2375, 2008.<br />

J. J. Pan and C. I. Chang. Destriping of Landsat MSS images by filtering techniques.<br />

Photogramm. Eng. Remote Sens., 58(10) :1417–1423, 1992.<br />

T. W. Parks and J. H. Mcclellan. Chebyshev approximation for nonrecursive digital filters<br />

with linear phase. IEEE Transactions on Circuits and Systems, 19(2) :189–194, 1972.


134 BIBLIOGRAPHIE<br />

P. Perona and J. Malik. Scale-space and edge <strong>de</strong>tection using anisotropic diffusion. IEEE<br />

Trans. Pattern Anal. Mach. Intell., 12 :629–639, 1990.<br />

D. J. Poros and C. J. Peterson. Methods for <strong>de</strong>striping landsat thematic mapper images - a<br />

feasibility study for an online <strong>de</strong>striping process in the thematic mapper image processing<br />

system (tips). Photogrammetric Engineering and Remote Sensing., 51 :1371–1378, 1985.<br />

P. Rakwatin, W. Takeuchi, and Y. Yoshifumi.<br />

P. Rakwatin, W. Takeushi, and Y. Yasuoka. Stripe Noise Reduction in Modis data by<br />

combining histogram matching with facet filter. 45(6) :1844–1856, 2007.<br />

R. W. Reynolds and T. M. Smith. Improved global sea surface temperature analyses using<br />

optimum interpolation. Journal of climate, 7(6) :929–948, 1994.<br />

J. W. Rouse, R. H. Haas, J. A. Schell, and D. W. Deering. Monitoring vegetation systems<br />

in the great plains with erts. In Proceedings of the Third ERTS Symposium, volume 1,<br />

pages 309–317, 1973.<br />

L. I. Rudin, S. Osher, and E. Fatemi. Nonlinear total variation based noise removal algorithms.<br />

Phys. D, 60 :259–268, 1992.<br />

V. V. Salomonson and I. Appel.<br />

H. F. Shen, T. H. Ai, and P. X. Li. Destriping and inpainting of remote sensing images using<br />

maximum a-posteriori method. In ISPRS 2008, volume 37, pages 63–70, 2008.<br />

C. Sheng, Y. Xin, Y. Liping, and S. Kun. Total variation-based speckle reduction using<br />

multi-grid algorithm for ultrasound images. In ICIAP, volume 3617, pages 245–252, 2005.<br />

J. J. Simpson, J. I. Gobat, and R. Frouin. Improved <strong>de</strong>striping of GOES images using finite<br />

impulse response filters. Remote. Sens. Environ., 52(1) :15–35, 1995.<br />

J. J. Simpson, J. R. Stitt, and D. M. Leath. Improved finite impulse response filters for<br />

enhanced <strong>de</strong>striping of geostationary satellite data. Remote. Sens. Environ., 66(3) :235–<br />

249, 1998.<br />

D. S. Sloan and R. Orth. A self-contained landsat data reception and precision cartographic<br />

image poduction system. In Int. Symp. of Comm. III of the Int. Soc. for Photogrammetry,<br />

Graz, Austria, Oct. 1977.<br />

R. Srinivasan, M. Cannon, and J. White. Lansat data <strong>de</strong>striping using power filtering. Opt.<br />

Eng., 27(11) :939–943, 1988.<br />

F. Steinmetz, P. Deschamps, and D. Ramon. Atmospheric correction in presence<br />

of sun glint :application to meris. In 2nd MERIS-(A)ATSR workshop, Frascati,<br />

2008. URL http://earth.esa.int/workshops/meris_aatsr2008/participants/75/<br />

pres_75_Steinmetz.pdf.


135<br />

W. M. Strome and S. S. Vishnubhatla. A sytem for improving the radiometric corrections<br />

for erts-1 mss data. In XXIV Int. Astronautical Congress, Baku, USSR, Oct. 1973.<br />

E. Tadmor, S. Nezzar, and L. Vese. A multiscale image representation using hierarchical<br />

(BV , L 2 ) <strong>de</strong>compositions. Multiscale Mo<strong>de</strong>l. Simul., 2(4) :554–579, 2004.<br />

D. Tanre, Y. J. Kaufman, M. Herman, and S. Matoo. Remote sensing of aerosol properties<br />

over oceans using the modis/eos spectral radiances : Passive remote sensing of tropospheric<br />

aerosol and atmospheric corrections from the new generation of satellite sensors. J.<br />

Geophys. Res., 102(D14) :1697116988, 1997.<br />

J. Torres and S. O. Infante. Wavelet analysis for the elimination of striping noise in satellite<br />

images. Opt. Eng., 40(7) :13091314, 2001.<br />

L. Vese and S. Osher. Mo<strong>de</strong>ling textures with total variation minimization and oscillating<br />

patterns in image processing. Journal of Scientific Computing, 19 :553–572, 2002.<br />

M. Wang and S. W. Bailey. Correction of sun glint contamination on the seawifs ocean and<br />

atmosphere products. Appl. Opt., 40(27) :4790–4798, 2001.<br />

M. Wegener. Destriping multiple <strong>de</strong>tector imagery by improved histogram matching. Int J.<br />

Remote Sens., 11(5) :859–875, 1990.<br />

M. P. Weinreb, R. Xie, J. H. Lienesch, and D. S. Crosby. Destriping GOES images by<br />

matching empirical distribution functions. Remote. Sens. Environ., 29(2) :185–195, 1989.<br />

Z. Wu and L. Li. A line-integration based method for <strong>de</strong>pth recovery from surface normals.<br />

Comput. Vision Graph. Image Process., 43 :53–66, 1988.<br />

Z. Xiaoxiang, F. Tianxi, and H. Qian. Destripe multi-sensor imaging spectroradiometer data.<br />

In CGMS XXXII, 2004. CMA-WP 07, 2007.<br />

Z. Yang, J. Li, W. P. Menzel, and R. A. Frey. Destriping for modis data via wavelet shrinkage.<br />

In Applications with Weather Satellites, volume 4895, pages 187–199, 2003.<br />

L. P. Yaroslavsky and M. E<strong>de</strong>n. Texture synthesis by non-parametric sampling. In Fundamentals<br />

of Digital Optics, Birkhauser Boston, 1996.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!