30.01.2013 Views

Impacts des changements climatiques sur les flux de carbone ...

Impacts des changements climatiques sur les flux de carbone ...

Impacts des changements climatiques sur les flux de carbone ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Thèse <strong>de</strong> l’Université Pierre et Marie Curie – Paris VI<br />

Laboratoire d’Océanographie <strong>de</strong> Villefranche (UMR 7093)<br />

Spécialité<br />

Océanologie – Météorologie – Environnement<br />

Présenté par<br />

Simon Bélanger<br />

Pour l’obtention du gra<strong>de</strong> <strong>de</strong> Docteur <strong>de</strong> l’Université Pierre et Marie Curie<br />

Sujet <strong>de</strong> thèse:<br />

<strong>Impacts</strong> <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong> <strong>sur</strong> <strong>les</strong> <strong>flux</strong> <strong>de</strong> <strong>carbone</strong><br />

stimulés par la lumière dans l'Océan Arctique:<br />

Quantification et suivi <strong>de</strong> la photo-oxydation <strong>de</strong> la matière organique dissoute dans<br />

la Mer <strong>de</strong> Beaufort par télédétection spatiale<br />

Response of light-related carbon <strong>flux</strong>es in the Arctic Ocean to<br />

climate change:<br />

Quantification and monitoring of dissolved organic matter photo-oxidation in the<br />

Beaufort Sea using satellite remote sensing<br />

Devant le jury compose <strong>de</strong> :<br />

Rapporteurs :<br />

Dr. Rainer M. W. Amon<br />

Pr. Dave A. Siegel<br />

Soutenue le 28 novembre 2006<br />

Éxaminateurs :<br />

Dr. Ingrid Obernosterer<br />

Pr. Alain Saliot (Prési<strong>de</strong>nt)<br />

Pr. Warwick F. Vincent (co-directeur <strong>de</strong> thèse)<br />

Dr. Marcel Babin (directeur <strong>de</strong> thèse)


Remerciements<br />

On ne vient pas à bout d’une thèse <strong>de</strong> doctorat sans le soutient financier <strong>de</strong> différent<br />

organismes subventionnaires, <strong>de</strong> celui professionnel <strong>de</strong> nos pairs et personnel <strong>de</strong> nos<br />

proches. Donc pour commencer je tiens à exprimer ma gratitu<strong>de</strong> envers <strong>les</strong> déci<strong>de</strong>urs du<br />

programme <strong>de</strong> bourse <strong>de</strong> thèse du Fond Québécois pour la Recherche <strong>sur</strong> la Nature et <strong>les</strong><br />

Technologies (FQRNT). En plus <strong>de</strong> m’avoir attribué un salaire raisonnable durant mes trois<br />

années <strong>de</strong> thèse, ils m’ont <strong>sur</strong>tout donné carte blanche pour réaliser mon projet <strong>de</strong> recherche<br />

dans un pays étranger. Projet qui fût aussi soutenu à travers le programme <strong>de</strong> recherche<br />

Canadian Arctic Shelf Exchange Study (CASES), financé par le Conseil Canadien <strong><strong>de</strong>s</strong><br />

Sciences Naturel<strong>les</strong> et Génie (CRSNG) et dirigé par le Dr Louis Fortier. Cette plate-forme<br />

m’a permis d’obtenir <strong>les</strong> données nécessaires à mes étu<strong><strong>de</strong>s</strong> et d’acquérir une expérience<br />

« Arctique » si enrichissante. Un autre acteur important dans la réalisation <strong>de</strong> la campagne<br />

CASES, que je remercie pour la place qu’il m’a réservée <strong>sur</strong> le navire, est Mr Pierre Larouche<br />

<strong>de</strong> l’Institut Maurice-Lamontagne à Mont-Joli. Aussi à propos <strong>de</strong> la campagne en mer à bord<br />

du brise-glace canadien CCGS Amundsen, j’en profite aussi pour remercier ses capitaines et<br />

ses membres d’équipages qui ont su mener à bon port nos aventures entre mer, terre et glace,<br />

ainsi que mes collègues scientifiques et amis dont l’enthousiasme m’a permi <strong>de</strong> passer 12<br />

semaines inoubliab<strong>les</strong>.<br />

Sur le plan scientifique, plusieurs personnes ont consacré nombre d’heures <strong>de</strong> leur<br />

temps pour m’ai<strong>de</strong>r dans l’interprétation et la publication <strong>de</strong> mes résultats. Je remercie tous<br />

ces personnes, et plus particulièrement ceux qui, malgré leur éloignement, ont contribué <strong>de</strong><br />

façon très significative à mes travaux : Dr Warwick Vincent, Dr Huixiang Xie, Dr Nick<br />

Krotkov et Jens Ehn. Je dois ajouter au passage que la présence <strong>de</strong> Warwick a été une source<br />

<strong>de</strong> motivation importante lors <strong>de</strong> mes passages à Québec. Merci Warwick d’avoir accepté <strong>de</strong><br />

me co-diriger. Tu seras toujours un modèle <strong>de</strong> motivation, et je compte bien m’en inspirer<br />

durant mon nouveau travail <strong>de</strong> professeur à Rimouski.<br />

Plus près <strong>de</strong> moi, j’ai aussi eu une chance immense <strong>de</strong> bénéficier <strong>de</strong> l’expertise d’un<br />

certain nombre <strong>de</strong> mes collègues du Laboratoire d’Océanographie <strong>de</strong> Villefranche-<strong>sur</strong>-mer.<br />

Plus qu’une chance, ce fût un privilège <strong>de</strong> côtoyer l’un <strong><strong>de</strong>s</strong> plus grand chercheur <strong>de</strong><br />

l’océanographie mo<strong>de</strong>rne, le Prof André Morel. Je n’oublierai jamais <strong>les</strong> nombreux matins<br />

III


passés à discuter <strong>de</strong> tout et <strong>de</strong> rien dans sa « caverne d’Ali Baba », laquelle est pratiquement<br />

toujours éclairée par le chaud soleil <strong>de</strong> Villefranche...<br />

Marcel, quand j’ai pris contact avec toi il y a quatre ans, je ne doutais pas que tes<br />

qualités professionnel<strong>les</strong> m’ai<strong>de</strong>raient à mener à bien mon projet <strong>de</strong> recherche. Mais j’étais<br />

loin d’imaginer à quel point te côtoyer allait m’apporter au niveau personnel. Merci pour le<br />

temps que tu m’as consacré, pour ta patience et ta disponibilité, pour ton sens <strong>de</strong> la<br />

perfection et ta rigueur scientifique, pour m’avoir appris à mieux gérer mon stress (et mes<br />

gestes durant mes présentations!), pour avoir su comment me redonner confiance quand il le<br />

fallait, etc., etc.. Je pourrais continuer ainsi pour <strong><strong>de</strong>s</strong> pages! Je te considère avant tout comme<br />

un ami, et ensuite comme un collègue. Ma porte sera toujours ouverte pour toi à Rimouski…<br />

Il y a plusieurs personnes avec <strong>les</strong>quel<strong>les</strong> je me suis lié d’amitié entre le quai <strong>de</strong> la<br />

darse, la Colle St-Michel, le Mont Boron, le cap Ferrat, Nice, la Corse, le Hockey, Monaco,<br />

Roma, Barcelone, Lavello, Paris, etc. Je <strong>les</strong> quitte avec un gros pincement au cœur. Ces<br />

personnes sauront se reconnaître en lisant ces lignes, nul besoin d’en ajouter.<br />

Il y a une personne en particulier à qui je dois beaucoup plus que tout au mon<strong>de</strong>,<br />

c’est ma blon<strong>de</strong> adorée. Sans toi, je ne m’en serais jamais sorti!! Merci d’avoir toujours été là,<br />

prête à me couper un peu <strong>de</strong> jambon, ou à me cuisiner une p’tite crêpe bretonne.<br />

Pour terminer, je tiens à m’excuser auprès <strong>de</strong> mes proches du Québec qui m’ont très<br />

peu vu durant ces années passées <strong>sur</strong> la côte d’azur. Je <strong>les</strong> retrouve avec empressement et<br />

près à rattraper le temps passé outre Atlantique.<br />

IV


V<br />

À la mémoire <strong>de</strong> père


Résumé <strong>de</strong> la thèse<br />

L’oxydation photochimique <strong>de</strong> la matière organique dissoute colorée (CDOM), et la<br />

production <strong>de</strong> CO 2 qui en résulte, est maintenant reconnue comme étant un processus<br />

majeur dans le cycle du <strong>carbone</strong> dans le système ocean-atmosphère. L’Océan Arctique fait<br />

parti <strong><strong>de</strong>s</strong> environnements où ce processus risque <strong>de</strong> jouer un rôle <strong>de</strong> plus en plus important<br />

dans le contexte <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong> à cause <strong>de</strong> : 1) l’augmentation du CDOM<br />

d’origine terrigène transporté à l’Océan côtier en réponse à la fonte du pergélisol et à<br />

l’augmentation du débit <strong><strong>de</strong>s</strong> rivières; 2) la diminution du couvert <strong>de</strong> glace estivale permettant<br />

au rayonnement solaire <strong>de</strong> pénétrer dans la colonne d’eau; et 3) l’augmentation du<br />

rayonnement ultraviolet (UV) dans cette région.<br />

Un modèle couplé « optique-photochimique » a été utilisé pour examiner le rôle <strong>de</strong> la<br />

photooxydation dans le cycle du <strong>carbone</strong> organique <strong>de</strong> l’Océan Arctique. Pour calculer la<br />

photoproduction <strong>de</strong> CO 2 (P DIC), l’éclairement spectral inci<strong>de</strong>nt, incluant <strong>les</strong> UV, a été<br />

modélisé avec un modèle <strong>de</strong> transfert radiatif nécessitant <strong>les</strong> concentrations <strong>de</strong> glace <strong>de</strong> mer,<br />

d’ozone, d’aérosols atmosphériques et le couvert nuageux observés par satellite entre 1979 et<br />

2004. Les ren<strong>de</strong>ments quantiques apparents pour calculer la P DIC ont été déterminés en<br />

laboratoire <strong>sur</strong> <strong><strong>de</strong>s</strong> échantillons provenant <strong>de</strong> la Mer <strong>de</strong> Beaufort et du Fleuve Mackenzie. La<br />

contribution du CDOM au coefficient d’absorption totale ([a CDOM/a t]), un paramètre clé du<br />

modèle, a été déterminée soit à partir <strong><strong>de</strong>s</strong> me<strong>sur</strong>es in situ, soit dérivée <strong><strong>de</strong>s</strong> données <strong>de</strong> la<br />

couleur <strong>de</strong> l’océan à l’ai<strong>de</strong> d’un nouvel algorithme empirique.<br />

Contrairement aux métho<strong><strong>de</strong>s</strong> semi-analytiques trouvées dans la littérature,<br />

l’algorithme empirique proposé permet <strong>de</strong> séparer, à l’échelle régionale, <strong>les</strong> coefficients<br />

d’absorption du CDOM et <strong><strong>de</strong>s</strong> particu<strong>les</strong> non-alga<strong>les</strong>. Cependant, l’utilisation <strong><strong>de</strong>s</strong> données <strong>de</strong><br />

la couleur <strong>de</strong> l’océan dans <strong>les</strong> hautes latitu<strong><strong>de</strong>s</strong> est souvent compromise par la présence <strong>de</strong><br />

glace <strong>de</strong> mer qui contamine <strong>les</strong> données. Ce problème a été abordé dans le cadre <strong>de</strong> la<br />

présente étu<strong>de</strong>, et une métho<strong>de</strong> a été proposée afin <strong>de</strong> détecter et éliminer <strong>les</strong> données<br />

contaminées par la glace <strong>de</strong> mer.<br />

Enfin, il a été démontré que <strong>les</strong> valeurs <strong>de</strong> P DIC sont similaires aux taux <strong>de</strong> <strong>carbone</strong><br />

organique provenant <strong>de</strong> la production primaire marine qui est séquestré dans <strong>les</strong> sédiments<br />

océaniques; et que l’augmentation <strong><strong>de</strong>s</strong> UV et la diminution <strong>de</strong> la glace <strong>de</strong> mer estivale au<br />

cours <strong><strong>de</strong>s</strong> 26 <strong>de</strong>rnières années a conduit à une augmentation <strong>de</strong> P DIC d’environ 15%. Ces<br />

résultats indiquent que la photooxydation joue un rôle significatif dans le cycle du <strong>carbone</strong><br />

actuel, et que la diminution <strong>de</strong> l'étendue <strong>de</strong> la glace <strong>de</strong> mer, prévue par <strong>les</strong> modè<strong>les</strong><br />

<strong>climatiques</strong>, risque d'amplifier significativement la reminéralisation photochimique du<br />

<strong>carbone</strong> organique dissous d'origine terrigène dans l'Océan Arctique.<br />

VI


Thesis abstract<br />

Photochemical oxidation of colored dissolved organic matter (CDOM), and the<br />

resulting production of CO2, is now known to be a significant process in the cycling of<br />

carbon in the ocean-atmosphere system. One environment where that process may play a<br />

major role in the context of climate change is the Arctic ocean because of: 1) the<br />

increasing amount of terrestrial CDOM released by the melting permafrost and brought to<br />

coastal ocean by rivers, 2) the <strong>de</strong>creasing summer ice cover that allows more solar<br />

radiation to penetrate the water column, and 3) the continuing increase in UV radiation<br />

over that region.<br />

A coupled optical-photochemical mo<strong>de</strong>l was used to assess the role of photooxidation<br />

in the carbon cycle of the Arctic Ocean. To calculate the photoproduction of<br />

CO2 (PDIC), the incoming spectral irradiance, including UV, was mo<strong>de</strong>led with a radiative<br />

transfer mo<strong>de</strong>l that uses satellite observations of sea ice, ozone, aerosols and cloud cover<br />

covering the 1979 to 2004 period. In situ <strong>de</strong>terminations of the apparent quantum yield<br />

for the photoproduction of CO2 ma<strong>de</strong> in the Beaufort Sea were used for the calculations.<br />

A key parameter in the mo<strong>de</strong>l was the contribution of CDOM to the total absorption<br />

coefficient. It was either obtained from in situ mea<strong>sur</strong>ements or <strong>de</strong>rived from Ocean<br />

Color imagery using a new empirical algorithm.<br />

Unlike most semi-analytical approaches found in the literature, the proposed<br />

empirical algorithm provi<strong><strong>de</strong>s</strong> a mean to separate CDOM absorption coefficient from nonalgal<br />

partic<strong>les</strong> absorption coefficient at the regional scale. The use of Ocean Color remote<br />

sensing at high latitu<strong>de</strong> is, however, compromised by the presence of sea ice that<br />

contaminates the data. This problem was addressed in the present study, and a method<br />

was proposed to <strong>de</strong>tect and eliminate contaminated pixels.<br />

Finally, it was shown that the level of PDIC is similar to the level of sequestered<br />

rates of organic carbon in the ocean sediments, which was produced through marine<br />

photosynthesis; and that the increase in UV and <strong>de</strong>crease in summer sea ice over the last<br />

26 years have led to an increased in PDIC by about 15%. These results indicate that the<br />

predicted trend of ongoing contraction of sea ice cover will greatly accelerate the<br />

photomineralization of CDOM in Arctic <strong>sur</strong>face waters.<br />

VII


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

Remerciements ....................................................................................................... III<br />

Résumé <strong>de</strong> la thèse ..................................................................................................VI<br />

Thesis abstract....................................................................................................... VII<br />

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

Liste <strong><strong>de</strong>s</strong> figures..................................................................................................... XII<br />

Liste <strong><strong>de</strong>s</strong> tableaux .............................................................................................. XVIII<br />

Avant propos .........................................................................................................XIX<br />

Chapitre I : Introduction et problématique générale .................................1<br />

I.1. Introduction ......................................................................................................... 2<br />

I.2. Problématique ..................................................................................................... 5<br />

I.2.1. Importance <strong>de</strong> la matière organique dissoute d’origine terrigène dans l’Océan<br />

Arctique et impact du réchauffement <strong><strong>de</strong>s</strong> écosystèmes terrestres nordiques .................... 5<br />

I.2.2. Importance <strong>de</strong> la photooxydation et impact <strong>de</strong> la fonte <strong>de</strong> la glace <strong>de</strong> mer et <strong>de</strong><br />

l’augmentation du rayonnement ultraviolet dans l’Arctique ................................................ 8<br />

I.2.3. Quantification <strong><strong>de</strong>s</strong> processus photochimiques........................................................... 12<br />

I.2.4. La télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan : un outil <strong>de</strong> choix pour quantifier <strong>les</strong><br />

processus photochimiques aux échel<strong>les</strong> océaniques............................................................ 15<br />

I.2.4.1. Propriétés optiques <strong><strong>de</strong>s</strong> eaux arctiques et algorithmes pour l’estimation du coefficient<br />

d’absorption du CDOM........................................................................................................... 17<br />

I.2.4.2. La glace <strong>de</strong> mer : un problème pour la télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan..................... 22<br />

I.3. Objectifs et Organisation <strong>de</strong> la Thèse.............................................................. 25<br />

I.4. Thesis objectives and organization................................................................... 27<br />

Chapitre II : L’Océan Arctique, la Mer <strong>de</strong> Beaufort et le Plateau du<br />

Mackenzie................................................................................................. 29<br />

II.1. Introduction...................................................................................................... 30<br />

II.2. Généralités <strong>sur</strong> l’Océan Arctique..................................................................... 32<br />

II.3. Généralités <strong>sur</strong> la Mer <strong>de</strong> Beaufort et le Plateau du Mackenzie ..................... 35<br />

II.3.1. Cyc<strong>les</strong> Saisonniers <strong>sur</strong> le Plateau du Mackenzie ....................................................... 37<br />

II.3.2. Bilan <strong>de</strong> Carbone Organique <strong>sur</strong> le Plateau du Mackenzie..................................... 41<br />

Chapitre III: Photominéralisation <strong>de</strong> la matière organique dissoute<br />

d’origine terrigène dans <strong>les</strong> eaux côtières arctiques entre 1979 et 2003:<br />

Variabilité interannuelle et implications <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong> 43<br />

III.A Résumé ........................................................................................................... 44<br />

III.B Article publié dans la revue Global Biogeochemical Cyc<strong>les</strong> :<br />

“Photomineralization of terrigenous dissolved organic matter in Arctic coastal<br />

waters from 1979 to 2003: Interannual variability and implications of climate<br />

change” .................................................................................................................... 45<br />

Acknowledgments ................................................................................................... 46<br />

VIII


Abstract .................................................................................................................... 47<br />

III.1. Introduction .................................................................................................... 48<br />

III.2. Materials and methods ................................................................................... 51<br />

III.2.1. Sample collection and storage ................................................................................... 51<br />

III.2.2. Optical mea<strong>sur</strong>ements ................................................................................................ 52<br />

III.2.3. Irradiation experiments .............................................................................................. 53<br />

III.2.4. DIC mea<strong>sur</strong>ement and calculation of φ DIC .............................................................. 54<br />

III.2.5. Mo<strong>de</strong>ling DIC photoproduction............................................................................... 55<br />

III.3. Results and discussion ................................................................................... 59<br />

III.3.1. Apparent Quantum Yield for DIC photoproduction............................................ 59<br />

III.3.2. Role of DIC photoproduction in the organic carbon cycling .............................. 62<br />

III.3.3. Interannual variability in DIC photoproduction .................................................... 65<br />

III.3.4. Implications for tDOC cycling in Arctic coastal waters........................................ 69<br />

III.4. Summary and conclusions ............................................................................. 72<br />

III.5. Auxiliary material: On the <strong>de</strong>termination of the Apparent Quantum Yield for<br />

DIC photoprodcution .............................................................................................. 73<br />

III.6. References ...................................................................................................... 77<br />

Chapitre IV : Quantification améliorée <strong>de</strong> la photooxydation <strong>de</strong> la<br />

matière organique dissoute colorée dans <strong>les</strong> eaux côtières à l’ai<strong>de</strong> <strong><strong>de</strong>s</strong><br />

propriétés optiques inhérentes dérivées <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong><br />

l’océan ....................................................................................................... 83<br />

IV.A Résumé............................................................................................................ 84<br />

IV.B. Article soumis à la revue Journal of Geophysical Research - Oceans (28<br />

décembre 2006): “Improved quantification of Chromophoric Dissolved Organic<br />

Matter photooxidation in coastal waters using satellite-<strong>de</strong>rived inherent optical<br />

properties” ............................................................................................................... 85<br />

Abstract .................................................................................................................... 86<br />

IV.1. Introduction .................................................................................................... 87<br />

IV.2. Materials and Methods ................................................................................... 89<br />

IV.2.1. Data sets <strong><strong>de</strong>s</strong>cription and mea<strong>sur</strong>ements................................................................. 89<br />

IV.2.2. Description of the [aCDOM/at] algorithm............................................................. 96<br />

IV.3. Results and Discussion................................................................................... 96<br />

IV.3.1. Variability of [a CDOM/a t] and [a CDOM/a CDM] in coastal waters.................................. 96<br />

IV.3.2. Retrieval of [a CDOM/a t] from remote sensing reflectance ....................................... 98<br />

IV.3.4. Application to SeaWiFS imagery............................................................................. 113<br />

IV.4. Summary and conclusions.............................................................................116<br />

IV.5. Notations........................................................................................................120<br />

IV.6. References......................................................................................................121<br />

Chapitre V: Impact <strong>de</strong> la glace <strong>de</strong> mer <strong>sur</strong> <strong>les</strong> estimations <strong>de</strong> la<br />

réflectance marine, la concentration <strong>de</strong> chlorophylle a et <strong>les</strong> propriétés<br />

optiques inhérentes dérivés <strong><strong>de</strong>s</strong> données satellita<strong>les</strong> <strong>de</strong> la Couleur <strong>de</strong><br />

l’Océan .....................................................................................................125<br />

IX


V.A. Résumé ...........................................................................................................126<br />

V.B. Article soumis à la revue Remote Sensing of Environment (1 novembre 2006):<br />

“Impact of sea ice on the retrieval of water-leaving reflectance, chlorophyll a<br />

concentration and inherent optical properties from satellite Ocean Color data”..128<br />

Abstract ...................................................................................................................129<br />

V.1. Introduction .....................................................................................................130<br />

V.2. Methodology....................................................................................................132<br />

V.2.1. Definitions ................................................................................................................... 132<br />

V.2.2. Description of the <strong>sur</strong>face reflectance spectra........................................................ 133<br />

V.2.3. Simulations of the adjacency effect .......................................................................... 136<br />

V.2.4. Simulations of sea ice contamination within an ocean pixel................................. 136<br />

V.2.5. Atmospheric correction and bio-optical processing.............................................. 137<br />

V.3. Results and discussion ....................................................................................138<br />

V.3.1. Impact of the adjacency effect on the retrieval of [ρ w] N, CHL and IOPs .......... 138<br />

V.3.2. Impact of sub-pixel sea ice contamination on the retrieval of [ρ w] N, CHL and<br />

IOPs.......................................................................................................................................... 147<br />

V.3.3. Detection of pixels affected by sea ice (adjacency effect and sub-pixel<br />

contamination) ........................................................................................................................ 153<br />

V.4. Summary and conclusions...............................................................................157<br />

Acknowledgment ....................................................................................................158<br />

Appendix V.1. Water reflectance mo<strong>de</strong>l .................................................................158<br />

Appendix V.2. Post-treatment of 6s output ............................................................161<br />

Appendix V.3. Modification of the QAA ................................................................163<br />

V.5. References .......................................................................................................165<br />

Chapitre VI : Utilisation <strong>de</strong> l’imagerie « couleur <strong>de</strong> l’océan » pour la<br />

quantification <strong>de</strong> la photooxydation dans la Mer <strong>de</strong> Beaufort...............169<br />

VI.A Résumé...........................................................................................................170<br />

VI.B. Article en preparation: “The use of Ocean Color imagery for the<br />

quantification of photooxidation in the Beaufort Sea” ..........................................171<br />

Abstract ...................................................................................................................172<br />

VI.1. Introduction ...................................................................................................173<br />

VI.2. Methods .........................................................................................................174<br />

VI.2.1. SeaWiFS Level 2 and 3 processing.......................................................................... 174<br />

VI.2.2. DIC photoproduction calculation........................................................................... 177<br />

VI.3. Results............................................................................................................179<br />

VI.3.1. Validation of the SeaWiFS water-leaving reflectance........................................... 179<br />

VI.3.2. Spatial-temporal variability in [a CDOM/a t](412)....................................................... 183<br />

VI.3.3. Spatial-temporal variability in DIC photoproduction.......................................... 186<br />

VI.4. Discussion......................................................................................................189<br />

VI.4.1. SeaWiFS data quality................................................................................................. 189<br />

VI.4.2. Quantification of DIC photoproduction using satellite Ocean Color .............. 190<br />

VI.5. Conclusions....................................................................................................193<br />

VI.6. References......................................................................................................193<br />

Chapitre VII: Conclusion générale et perspectives ................................196<br />

X


VII.A. Version anglaise: Overall Conclusion and Perspectives ............................ 204<br />

Bibliographie générale ............................................................................ 211<br />

Annexe 1 : Modification <strong>de</strong> l’algorithme Quasi-Analytique pour<br />

l’application à la Mer <strong>de</strong> Beaufort...........................................................231<br />

Annex A1. Modification of the Quasi-Analytical Algorithm for an application to<br />

the Beaufort Sea..................................................................................................... 232<br />

A1.1. Introduction and motivation......................................................................... 232<br />

A1.2. Quasi-Analytical Algorithm <strong><strong>de</strong>s</strong>cription....................................................... 232<br />

A1.3. Results and discussion.................................................................................. 237<br />

A1.3.1. Application of QAA to in situ data set.................................................................. 237<br />

A1.3.2. Estimation of a CDOM.................................................................................................. 240<br />

A1.4. References..................................................................................................... 242<br />

XI


Liste <strong><strong>de</strong>s</strong> figures<br />

Figure I.1. Anomalie débit d’eau douce (trait noir) <strong><strong>de</strong>s</strong> six plus grands fleuves sibériens qui<br />

se déchargent dans l’Océan Arctique <strong>de</strong>puis 1940. L’augmentation (7%) est corrélée à<br />

l’augmentation <strong>de</strong> la température <strong>de</strong> <strong>sur</strong>face à l’échelle globale (trait bleu), ainsi qu’avec<br />

l’indice climatique NAO (trait rouge) (source : Peterson et al., 2002)................................. 7<br />

Figure I.2. Observations <strong>de</strong> l’étendue moyenne du couvert <strong>de</strong> glace saisonnier au cours du<br />

20 ième siècle (source <strong><strong>de</strong>s</strong> données : ACIA, 2005). La diminution <strong>de</strong> la glace est plus<br />

importante au printemps et en été, qu’en automne et en hiver. .......................................... 9<br />

Figure I.3. Étendue <strong>de</strong> la banquise Arctique observée par satellite annuellement au mois <strong>de</strong><br />

septembre entre 1979 et 2005. (Source <strong><strong>de</strong>s</strong> données: National Snow and Ice Data Center,<br />

2006)........................................................................................................................................... 10<br />

Figure I.4. Éclairement spectral dans <strong>les</strong> parties UV-A et UV-B pour trois concentrations<br />

en ozone tel<strong>les</strong> qu’indiquées <strong>sur</strong> la figure (DU = Dobson Units), et un exemple <strong>de</strong><br />

ren<strong>de</strong>ment quantique apparent (voir la section suivante pour plus d’explications)<br />

exprimé en unités relatives, représentant la réactivité photochimique du CDOM en<br />

fonction <strong>de</strong> la longueur d’on<strong>de</strong> (courbe grisée).................................................................... 11<br />

Figure I.5. Trajectoires possib<strong>les</strong> dans le système atmosphère-océan pour un photon émis<br />

par le soleil par temps dégagé (voir aussi encadré 4). De la quantité <strong>de</strong> photon diffusée<br />

dans l’océan qui parvient à sortir <strong>de</strong> l’océan en direction du satellite (a; L w), une partie<br />

sera transmise à travers l’atmosphère jusqu’au capteur (b) alors que l’autre partie sera<br />

perdue par diffusion ou absorption dans l’atmosphère (c). Certains photons en<br />

provenance du soleil sont transmis directement (d) ou indirectement (e) dans le champ<br />

<strong>de</strong> visée du capteur après avoir été réfléchis <strong>de</strong> façon spéculaire par l’interface air-mer<br />

(L r, en anglais sun glint ou glitter). Comme pour L w, une partie sera transmise (g) et l’autre<br />

sera perdue dans l’atmosphère (f). Dans la partie visible du spectre, plus <strong>de</strong> 90% <strong>de</strong><br />

l’énergie reçue au satellite provient <strong>de</strong> l’atmosphère (L p) et comprend : <strong>les</strong> photons<br />

diffusés une (h) ou plusieurs (i) fois par <strong>les</strong> molécu<strong>les</strong> ou <strong>les</strong> aérosols atmosphériques.<br />

Dans <strong>les</strong> régions polaires où la glace <strong>de</strong> mer est présente, il faut ajouter à ce<br />

schéma la contribution <strong><strong>de</strong>s</strong> photons réfléchis par la glace à l’extérieur du champ <strong>de</strong><br />

visée qui sont redirigés vers le capteur (l; effet <strong>de</strong> l’environnement). Enfin, quand la<br />

glace se trouvant dans le champ <strong>de</strong> visée du capteur, une partie <strong><strong>de</strong>s</strong> photons qu’elle<br />

réfléchira sera transmise (m; contamination sub-pixel) et l’autre sera perdue dans<br />

l’atmosphère (n). Les contributions indirectes (l) et directes (m) posent problème<br />

puisqu’ils ne sont pas pris en compte dans le traitement <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong><br />

l’océan......................................................................................................................................... 23<br />

Figure II.1. Patrons <strong>de</strong> circulation générale <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face dans l’Océan Arctique. Les<br />

eaux Atlantique et Pacifique (en rouge) pénètrent dans l’Arctique via la Mer <strong>de</strong> Barents<br />

et le détroit <strong>de</strong> Bering respectivement. Le Tourbillon <strong>de</strong> Beaufort et le courant<br />

Transpolaire (en bleu) sont <strong>les</strong> <strong>de</strong>ux principaux chemins empruntés par <strong>les</strong> eaux<br />

Arctiques, <strong>les</strong>quel<strong>les</strong> sont éventuellement exportées vers l’Atlantique Nord via <strong>de</strong><br />

détroit <strong>de</strong> Fram et l’Archipel Canadien. Ces eaux <strong>de</strong> <strong>sur</strong>face sont fortement influencées<br />

par <strong>les</strong> grands fleuves Sibériens et Nord Américains. Les huit plus importants sont<br />

indiqués avec, entre parenthèses, leur débit annuel (volume d’eau douce en km 3 ans -1 ).<br />

Le cadre noir localise la zone d’étu<strong>de</strong>. (Source <strong><strong>de</strong>s</strong> informations : ACIA, 2005)............ 31<br />

XII


Figure II.2. Carte <strong>de</strong> localisation <strong>de</strong> la Mer <strong>de</strong> Beaufort avec la direction <strong><strong>de</strong>s</strong> principaux<br />

courants : la Gyre <strong>de</strong> Beaufort (en gris); le sous-courant <strong>de</strong> Beaufort (c.f., Beaufort<br />

Un<strong>de</strong>rcurrent; en bleu); la dispersion attendue du panache du Mackenzie (en rouge) qui<br />

varie selon la direction et l’intensité <strong><strong>de</strong>s</strong> vents dominants. L’encadré en gris montre <strong>les</strong><br />

limites <strong>de</strong> l’échantillonnage effectué durant le programme CASES.................................. 36<br />

Figure II.3. Cycle saisonnier <strong>de</strong> l’eau douce <strong>sur</strong> le Plateau du Mackenzie caractérisé par <strong>les</strong><br />

apports continentaux et la glace <strong>de</strong> mer : A) englacement (apports continentaux faib<strong>les</strong>,<br />

la glace commence à se former); B) fin <strong>de</strong> l’hiver (apports continentaux faib<strong>les</strong>, la glace<br />

a atteint son épaisseur maximale et a arrêté <strong>de</strong> croître); C) débâcle printanière (apports<br />

continentaux élevés, la glace <strong>de</strong>meure intacte et reste près <strong>de</strong> la côte); D) été (apports<br />

continentaux élevés, la glace a fondue ou a été exportée vers l’intérieur <strong>de</strong> l’Océan).<br />

(Modifiée <strong>de</strong> Macdonald, 2000).............................................................................................. 37<br />

Figure II.4. Variabilité journalière entre 1976 et 1998 (A) du débit du fleuve Mackenzie, (B)<br />

du <strong>flux</strong> <strong>de</strong> particu<strong>les</strong> en suspension d’origine terrigène, et (C) <strong>de</strong> l’éclairement<br />

photosynthétique inci<strong>de</strong>nt à 71°N et la température <strong>de</strong> l’air me<strong>sur</strong>és à Tuktoyaktuk situé<br />

à l’embouchure du fleuve. (Modifiée <strong>de</strong> O’Brien et al., 2006)............................................ 38<br />

Figure II.5. Image SeaWiFS acquise le 15 Juin 2004 montrant le début <strong>de</strong> la débâcle.<br />

Rapi<strong>de</strong>ment <strong>les</strong> eaux turbi<strong><strong>de</strong>s</strong> du Mackenzie occupent une large partie <strong><strong>de</strong>s</strong> eaux<br />

ouvertes du plateau continental (~15 000-20 000 km²)...................................................... 40<br />

Figure II.6. Coupe transversale effectuée à partir <strong>de</strong> l’embouchure du fleuve Mackenzie<br />

(Kugmalit Bay) jusqu’à la limite du plateau continental en juillet 2004 dans le cadre du<br />

programme CASES : a) <strong>de</strong>nsité; b) salinité........................................................................... 41<br />

Figure III.1. Location of CASES stations sampled for absorption mea<strong>sur</strong>ements in 2004<br />

over the Mackenzie Shelf (open circ<strong>les</strong>), in Amundsen Gulf (open triang<strong>les</strong>), and in<br />

Canada Basin (open squares). ARDEX and CASES samp<strong>les</strong> collected for φDIC <strong>de</strong>termination are shown as closed squares. ......................................................................... 50<br />

Figure III.2. φDIC spectra <strong>de</strong>termined on water samp<strong>les</strong> collected during ARDEX (R5a, R5d,<br />

and R9; solid thin lines) and CASES (108, 406 and 409; dashed thin lines). The grey<br />

curves are the average φDIC spectra published by Vähätalo et al. [2000] for a boreal lake,<br />

and by Johannessen and Miller [2001] for inshore, coastal and offshore waters.................. 58<br />

Figure III.3. (a) Variation of aCDOM(330) with salinity for water samp<strong>les</strong> collected in<br />

Mackenzie Shelf (open circ<strong>les</strong>), Amundsen Gulf (open triang<strong>les</strong>), Canada Basin (open<br />

squares). Samp<strong>les</strong> collected for φDIC are shown as closed squares. (b) Variation of<br />

weighted quantum yield normalized to the integrated irradiance, φ DIC , with salinity. .. 61<br />

Figure III.4. Average DIC photoproduction rates, P DIC, calculated using the φ DIC<br />

representing each sub-region: R5a, R5d and R9 for the Mackenzie Shelf; 108 for the<br />

Amundsen Gulf; 406 and 409 for the Canada Basin........................................................... 62<br />

Figure III.5. Average spectral light fraction absorbed by CDOM in the <strong>sur</strong>face waters at<br />

stations (Fig. III.1) sampled over the Mackenzie Shelf (dashed line), in Amundsen Gulf<br />

(dash-dotted line), and in Canada Basin (thin line).............................................................. 63<br />

Figure III.6. Trends in the annual DIC photoproduction for period 1979-2003 for the<br />

Mackenzie Shelf, the Amundsen Gulf, the Canada Basin, and the sums of the three<br />

sub-regions. ............................................................................................................................... 65<br />

Figure III.7. Correlation between the annual DIC photoproduction and the annual average<br />

area of open water for: (a) Mackenzie Shelf, DIC = 0.91x10 6 * X; (b) Amundsen Gulf,<br />

XIII


DIC = 0.52x10 6 * X ; (c) Canada Basin, DIC = 0.63x10 6 * X. (X = open water area in<br />

km 2 ; slope in Gg C y -1 km -2 ; n=25). ....................................................................................... 66<br />

Figure III.8. Trends in open water area for the three sub-regions as observed by<br />

SMMR/SSMI between 1979 and 2003. The left panels show the day of the year when<br />

>50% of the <strong>sur</strong>face area is ice free. The right panels show the number of days having<br />

open water. ................................................................................................................................ 67<br />

Figure III.9. Trends in monthly O3 concentration over the study area as observed by<br />

TOMS between 1979 and 2001. Note the changes of scale among the panels. ............. 68<br />

Figure III.A.1. Absorption spectra of CDOM for before (black line) and after the<br />

irradiation for the sample un<strong>de</strong>r WG280 cutoff filter (red line). The time of irradiation<br />

and the irradiance level un<strong>de</strong>r the cutoff filter are also shown.......................................... 75<br />

Figure III.A.2. Mo<strong>de</strong>led versus mea<strong>sur</strong>ed DIC photoproduction rates for the Single<br />

exponential (squares) and Quasi-exponential (circ<strong>les</strong>) functional forms.......................... 76<br />

Figure IV.1. Map of the study area showing location of stations where IOPs (dots), SPMR<br />

(triang<strong>les</strong>) and ASD (squares) mea<strong>sur</strong>ements were ma<strong>de</strong>. Contour lines of sea ice<br />

concentration of 50% are also shown for June 1 st (red), July 1 st (blue) and August 1 st<br />

(green)......................................................................................................................................... 91<br />

Figure IV.2. Frequency distributions of the contribution of CDOM to (a) the total light<br />

absorption at 412 nm ([aCDOM/at]), and (b) to the colored <strong>de</strong>trital material ([aCDOM/aCDM]). Note that the graph inclu<strong><strong>de</strong>s</strong> 54 additional COASTlOOC or CASES stations where<br />

only IOP were available........................................................................................................... 98<br />

Figure IV.3. Comparison between retrieved and mea<strong>sur</strong>ed [aCDOM/at] at 412 nm for the<br />

COASTlOOC and CASES datasets. [aCDOM/at](412) was calculated using equation 6<br />

with coefficients obtained using the whole data set (N=255; Table IV.2). ..................... 99<br />

Figure IV.4. Error analysis of [aCDOM/at] at 412, calculated using region-in<strong>de</strong>pen<strong>de</strong>nt<br />

coefficients, as a function of: a) the ratio of ap(412) to Rrs(555); b) the relative<br />

contribution of CDOM to the total CDM absorption coefficient at 412 nm. Only the<br />

statistically significant relationships are shown. ................................................................. 103<br />

Figure IV.5. Same as Fig. IV.3 but for region specific coefficients for Eq. 6 (Table IV.2).104<br />

Figure IV.6. Apparent Quantum Yield spectra for DIC production <strong>de</strong>termined for two<br />

stations located in the southeastern Beaufort Sea. AQY1 was <strong>de</strong>termined on <strong>sur</strong>face<br />

waters influenced the by the Mackenzie River (salinity = 8.2 ‰; Station R5a), while<br />

AQY2 was <strong>de</strong>termined on <strong>sur</strong>face waters collected in the Amundsen Gulf, away from<br />

direct riverine influence (salinity = 30.0 ‰; Station 108) [Bélanger et al., 2006]. ............ 106<br />

Figure IV.7. Spectral production of DIC for different values of [aCDOM/at](412) and for A)<br />

AQY1 and B) AQY2. .............................................................................................................. 107<br />

Figure IV.8. Relative difference in DIC production as a function of the absolute difference<br />

between the retrieved and true [aCDOM/at](412) values for the two AQYs and two initial<br />

values for [aCDOM/at] true (412) of 0.3 and 0.7. The thick line at the bottom illustrate the<br />

95% confi<strong>de</strong>nce interval in the [aCDOM/at] retrieval using Eq 6........................................ 108<br />

Figure IV.9. Variability of SCDOM as a function of the aCDOM(412) value mea<strong>sur</strong>ed in the<br />

<strong>sur</strong>face waters of the southeastern Beaufort Sea during summer 2004 (n=65). A<br />

second-or<strong>de</strong>r polynomial fit between the SCDOM values and the logarithm of aCDOM(412) provi<strong>de</strong>d a strong <strong><strong>de</strong>s</strong>cription of the observations (r 2 =0.79; n=65): .............................. 110<br />

2<br />

S CDOM = 0. 0186 − 0.<br />

0019log[<br />

aCDOM<br />

( 412)<br />

] + 0.<br />

0031log[<br />

aCDOM<br />

( 412)]<br />

................................ 110<br />

Figure IV.10. Spectral absorption by particulate matter normalized to 440 nm value<br />

mea<strong>sur</strong>ed in the <strong>sur</strong>face waters of the southeastern Beaufort Sea during summer 2004<br />

XIV


N<br />

(thin grey lines). The thick black line represent the average ap (λ) values. The<br />

sensitivity the PDIC (eq 1) to the extrapolation to the UV domain is explore within the<br />

limits given by 95% probability of the normal distribution (i.e. ±1.96*Standard<br />

Deviation (σ); dashed lines). The ap(λ) values for λ


Figure V.10. Retrieved Angström exponent parameter (α) using the atmospheric correction<br />

algorithm, as a function of σR ice, for τ a=0.1 and four types of sea ice............................149<br />

Figure V.11. Ratio of CHL estimated from OC4v4 (solid line) and OC4L (dotted line) when<br />

sea ice is present within the water pixel to CHL estimated without sea ice as a function<br />

of σR ice(865). The results are presented for three initial concentrations of chlorophyll<br />

(with f CDOM =50% and SPM NAP =0.0 g m -3 ), as indicated, for τ a=0.1 and four types of sea<br />

ice. The dashed lines represent a 35% error range. ........................................................... 150<br />

Figure V.12. Ratio of a t estimated at 443 nm using QAA when sea ice is present within the<br />

water pixel to a t estimated with no sea ice as a function of σR ice(865). The results are<br />

presented for each water type as shown in figure 2 and for the landfast ice covered by<br />

fresh snow. The dashed lines represent a 35% error range.............................................. 151<br />

Figure V.13. Same as Fig. V.12, but for b bp estimated at 555 nm using QAA....................... 151<br />

Figure V.14. Same as Fig. V.12, but for the ratio [a CDOM/a t] estimated at 412 nm................ 152<br />

Figure V.15. Example of sea ice contamination on the CHL estimation (OC4v4) for a<br />

SeaWiFS scene of the Beaufort Sea acquired on the 8 th of September 2002. Upper panel<br />

is the true color composite showing the sea ice field offshore........................................ 153<br />

Figure V.16. Reflectance ratio between [ρ w] N at 412 and [ρ w] N at 443 nm as a function of<br />

[ρ w] N at 555 nm for: 1) in situ mea<strong>sur</strong>ements of the water-leaving reflectance conducted<br />

in the Southeastern Beaufort Sea [Bélanger et al., submitted manuscript. 2006] (closed<br />

circ<strong>les</strong>); 2) mo<strong>de</strong>led water-leaving reflectance with the bio-optical mo<strong>de</strong>l presented in<br />

Appendix V.1 (closed squares); and 3) SeaWiFS pixels extracted from the image<br />

presented on Fig. V.17 (light grey dots).............................................................................. 155<br />

Figure V.17. Example of application of the adjacency flag to a SeaWiFS scene from the<br />

Southeastern Beaufort Sea acquired on the 16 th of June 1998. The R rs at 555 nm is<br />

shown on colored log scale as indicated on the right; the flagged pixels for adjacency<br />

are i<strong>de</strong>ntified with dark red color. The sea ice and land are shown in dark and light grey<br />

respectively. Because of known problem with the standard SeaWiFS atmospheric<br />

correction over the turbid waters, we apply the algorithm proposed by Ruddick et al.<br />

[2000]........................................................................................................................................ 156<br />

Figure V.A1. Schematic representation of a water pixel located at distance D 0 from an ice<br />

edge........................................................................................................................................... 162<br />

Figure VI.1. A schematic representation of the DIC photoproduction mo<strong>de</strong>ling. The mo<strong>de</strong>l<br />

ingests four inputs: (1) For a given day, when data are available, the spatio-temporal<br />

variability in the [a CDOM/a t] parameter is specified by interpolating the values of the<br />

months before and after. If monthly IOP are not available (due to either sea ice or<br />

cloud cover), the data from the climatology is taken. The spectral interpolation of the<br />

[a CDOM/a t](412) is perform as <strong><strong>de</strong>s</strong>cribed in section IV.3.4. (2) The daily spatial variability<br />

in the area of open water. (3) The spectral daily downward irradiance calculated as<br />

<strong><strong>de</strong>s</strong>cribed in section III.2.5. (4) The spectral apparent quantum yield for the DIC<br />

production (AQY, or φ DIC). Two methods are compared for the choice of the φ DIC for a<br />

given day and pixel. Method 1 is based on the regional <strong>de</strong>finition <strong><strong>de</strong>s</strong>cribed in section<br />

III.2.5. In the Method 2, the φ DIC varies as a function of the magnitu<strong>de</strong> of the CDOM<br />

(see text for <strong>de</strong>tails). ............................................................................................................... 178<br />

Figure VI.2. A) CDOM photoreactivity as a function of a CDOM(412). B) Spectral variability<br />

of the pooled φ DIC spectra for low and high a CDOM(412) value.......................................... 179<br />

XVI


Figure VI.3. Comparison of in situ mea<strong>sur</strong>ements of water-leaving radiance with coinci<strong>de</strong>nt<br />

spectra from the SeaWiFS image processed with the standard and the turbid water AC<br />

algorithms. ............................................................................................................................... 181<br />

Figure VI.4. Scatterplots of the in situ mea<strong>sur</strong>ements versus SeaWiFS-<strong>de</strong>rived variab<strong>les</strong><br />

employed in the empirical algorithm used to estimate [a CDOM/a t](412) (eq. IV.6): a) the<br />

logarithm of water-leaving reflectance ratio of 412 to 555 nm; b) the logarithm of<br />

water-leaving reflectance ratio of 490 to 555 nm............................................................... 182<br />

Figure VI.5. Seasonal variability of the SeaWiFS-<strong>de</strong>rived [a CDOM/a t](412) for 1998 and 2004<br />

(1 x 1 km resolution). Sea ice (or cloud) is shown in grey................................................ 184<br />

Figure VI.6. Variability of the SeaWiFS-<strong>de</strong>rived b bp(555) for June and July 1998 and 2004 (1<br />

x 1 km resolution)................................................................................................................... 185<br />

Figure VI.7. Variability of the SeaWiFS-<strong>de</strong>rived a t(412) for June and July 1998 and 2004 (1 x<br />

1 km resolution)...................................................................................................................... 185<br />

Figure VI.8. Annual maps of DIC photoproduction rates from 1998 to 2004 in<br />

southeastern Beaufort Sea..................................................................................................... 188<br />

Figure A1. Schematic representation of the quasi-analytical algorithm <strong>de</strong>veloped to <strong>de</strong>rive<br />

the partic<strong>les</strong> backscattering coefficient at 555 nm and the spectral total absorption<br />

coefficient. The data (parallelogram) necessary for each steps of the algorithm<br />

(rectangular boxes) are i<strong>de</strong>ntified by the dotted arrows, while the outputs are i<strong>de</strong>ntified<br />

by the thin line with filled arrows. The final outputs, i.e. b bp(555) and a t(λ), are i<strong>de</strong>ntified<br />

by the think line with filled arrows at bottom right of the diagram. The equations<br />

(i<strong>de</strong>ntified with upper script number) are listed in Table A1........................................... 236<br />

Figure A2. Comparison between retrieved and mea<strong>sur</strong>ed a t with the ac9 at A) 412 nm and B)<br />

440 nm. Total absorption coefficients were retrieved using the optimized version of<br />

QAA proposed by Lee et al. [2002] that use λ 0 555 and 640 nm (open circ<strong>les</strong>) and the<br />

tuned QAA with the CASES dataset (inverted triang<strong>les</strong>)................................................. 238<br />

Figure A3. Comparison between QAA-<strong>de</strong>rived b bp(555) and mea<strong>sur</strong>ed b p(555)................... 240<br />

Figure A4. Retrieved versus mea<strong>sur</strong>ed a CDOM(412) in the Beaufort Sea. Statistical parameters<br />

of the Type II regression are also shown............................................................................ 241<br />

XVII


Liste <strong><strong>de</strong>s</strong> tableaux<br />

Table III.1. Chemical and physical properties of samp<strong>les</strong> used for the φDIC........................... 52<br />

Table III.2. CDOM properties and mo<strong>de</strong>l parameters for φDIC. .............................................. 58<br />

Table III.3. Comparison between daily DIC photoproduction, and total and new primary<br />

production rates in open water (in mg C m -2 d -1 )................................................................. 64<br />

Table III.4. Mean annual DIC photoproduction and estimates of tDOC mineralized by<br />

photooxidation.......................................................................................................................... 71<br />

Table IV.1. Relative contribution of CDOM to the colored <strong>de</strong>trital matter<br />

(CDM=CDOM+NAP) and to the total absorption coefficients at 412 nm, respectively,<br />

for the different regions covered by our data set................................................................. 97<br />

Table IV.2. Empirical coefficients of Eq. 6 <strong>de</strong>termined by multiple regression for the<br />

different coastal environments. .............................................................................................. 99<br />

Table IV.3. Performance of Eq. 6 for the retrieval of [aCDOM/at] at 412 nm. For each region,<br />

in<strong>de</strong>pen<strong>de</strong>nt data set is used to obtain the empirical coefficients for Eq. 6. ................. 100<br />

Table VI.4. Coefficient of <strong>de</strong>termination of the linear regression between Δ[aCDOM/at] (shown in Fig. IV.4) and 1) the logarithm of the ratio [ap(412)/Rrs(555)] and 2) the ratio<br />

[ a CDOM / aCDM<br />

] at 412 nm. NS = the slope is not significantly different from 0 at 95%<br />

confi<strong>de</strong>nce level (Type II regression). ................................................................................. 101<br />

Table IV.5. ΔPDIC ( in %), due to error in the spectral extrapolation of [aCDOM/at](412) to the<br />

true<br />

300-600 nm range resulting from variation in the magnitu<strong>de</strong> of at(412). PDIC is<br />

N<br />

calculated with averaged SCDOM , ap (λ), and for different at(412) values. PDIC is<br />

calculated assuming aw(λ) is null........................................................................................... 112<br />

Table IV.6. ΔPDIC ( in %), due to error in the spectral extrapolation of [aCDOM/at](412) to the<br />

N true<br />

300-600 nm range resulting from variation in SCDOM and ap (λ) spectra. PDIC is<br />

N<br />

calculated with averaged SCDOM , ap (λ), and at(412)........................................................... 112<br />

Table V.1. Distance (in km) from ice eg<strong>de</strong> within which the error on CHL > 35% (Rice fresh<br />

snow). Values in parenthesis are for the OCL4 algorithm............................................... 141<br />

Table VI.1. Number of SeaWiFS Level 1A MLAC images processed................................... 175<br />

Table VI.2. Location, date, time interval, and the sun zenith angle at the time of sampling of<br />

the match-ups.......................................................................................................................... 180<br />

Table VI.3. Comparison of in situ with SeaWiFS-<strong>de</strong>rived [aCDOM/at](412) values using<br />

Standard and MUMM AC algorithms. ................................................................................ 183<br />

Table VI.4. Annual DIC photoproduction estimation using various methods. The value in<br />

parenthesis for methods 1 and 2 is the difference relative (in %) to GBC method. .... 186<br />

Table A2. Performance of the QAA to retrieved at(λ) with empirical relationships of Lee et<br />

al. [2002] versus QAA tuned with CASES dataset (n=46)............................................... 239<br />

XVIII


Avant propos<br />

Cette thèse fût d’abord et avant tout motivée par mon désir d’apprendre, mon<br />

émerveillement face à la beauté <strong>de</strong> la Nature, ainsi que pour mon goût d’aventures. En<br />

particulier, vivre l’expérience <strong>de</strong> l’Arctique, où en été <strong>les</strong> jours sont interminab<strong>les</strong>, était pour<br />

moi un rêve qui est <strong>de</strong>venu réalité. Mais au fur et à me<strong>sur</strong>e que je me suis plongé dans mes<br />

étu<strong><strong>de</strong>s</strong>, mon sentiment d’inquiétu<strong>de</strong> face à l’avenir <strong>de</strong> notre planète s’est fait ressentir <strong>de</strong> plus<br />

en plus fort. Si <strong>les</strong> <strong>changements</strong> <strong>climatiques</strong> font partie <strong>de</strong> l’histoire <strong>de</strong> la Terre et sont, par<br />

conséquent, tout à fait naturels, la situation dans laquelle nous nous trouvons aujourd’hui est<br />

sans précé<strong>de</strong>nt dans cette histoire vieille <strong>de</strong> quatre milliards d’années. En exploitant <strong>les</strong><br />

ressources énergétiques fossilisées, nous avons, avec une inconscience presque légitime,<br />

modifié notre environnement au point qu’il déstabilise le fragile équilibre <strong>de</strong> la planète. Mon<br />

sentiment d’inquiétu<strong>de</strong> provient justement du fait que <strong>les</strong> conséquences <strong>de</strong> ce déséquilibre<br />

restent hors <strong>de</strong> portée pour nous, « bipè<strong><strong>de</strong>s</strong> bien intentionnés ». Dans l’état actuel <strong>de</strong> nos<br />

connaissances il est impossible <strong>de</strong> savoir si nous allons ou pas vers une catastrophe naturelle<br />

irrévocable, et ce malgré toutes nos technologies aussi avancées soient-el<strong>les</strong>. Reste<br />

qu’aujourd’hui, lorsque nous faisons une rétrospective environnementale du siècle <strong>de</strong>rnier<br />

(e.g., déforestation, <strong><strong>de</strong>s</strong>truction <strong>de</strong> la couche d’ozone, contamination <strong><strong>de</strong>s</strong> systèmes aquatiques,<br />

augmentation du dioxy<strong>de</strong> <strong>de</strong> <strong>carbone</strong> dans l’atmosphère, etc, etc), nous pouvons nous<br />

attendre au pire 1 . En effet, il a été démontré que la mauvaise gestion <strong>de</strong> l’environnent a<br />

souvent été l’une <strong><strong>de</strong>s</strong> causes premières (pas la seule bien sûre!) expliquant l’effondrement <strong>de</strong><br />

plusieurs sociétés humaines du passé (e.g., l’île <strong>de</strong> Pâques, Viking en Amérique) 2 .<br />

Le sujet que je traite dans ce manuscrit concerne <strong>les</strong> <strong>changements</strong> <strong>climatiques</strong>. Si<br />

cette thèse apporte <strong>de</strong> nouveaux indices qui montrent l’importance <strong><strong>de</strong>s</strong> <strong>changements</strong><br />

<strong>climatiques</strong>, en particulier ceux qui se produisent dans l’Océan Arctique, elle ne propose<br />

aucune solution qui permettrait d’en atténuer ses conséquences. De plus, la quantité<br />

1 Plusieurs scientifiques renommés se sont aujourd’hui engagés dans une campagne <strong>de</strong> conscientisation <strong>de</strong> la<br />

population. À ce sujet, je recomman<strong>de</strong>rais notamment <strong>les</strong> essais d’Hubert Revees (Mal <strong>de</strong> terre, édition Le seuil),<br />

<strong>de</strong> David Suzuki (L’équilibre sacré – redécouvrir sa place dans la nature, édition Fi<strong><strong>de</strong>s</strong>), et <strong>de</strong> Jared Diamond<br />

(Effondrement, édition Gallimard).<br />

2 J. Diamond, Effondrement.<br />

XIX


phénoménale <strong>de</strong> gaz à effet <strong>de</strong> serre qui fut utilisée pour réaliser ce projet <strong>de</strong> thèse (<strong>les</strong><br />

nombreux aller-retour France-Québec, navire océanographique, satellite, Fiat Panda, MBK,<br />

etc., etc.) pèse lourd <strong>sur</strong> ma bonne conscience. Néanmoins, comme le dit si bien David<br />

Suzuki, « le plus important, à l’heure actuelle, est l’échange d’idées; il faut se passer le mot en travaillant<br />

tous à réduire notre effet négatif <strong>sur</strong> la planète, en développant <strong><strong>de</strong>s</strong> infrastructures qui permettront d’adopter<br />

un mo<strong>de</strong> <strong>de</strong> vie durable et <strong>de</strong> susciter dans le public un appui qui changera à la fin <strong>les</strong> priorités politiques ».<br />

J’espère que mon travail <strong>de</strong> thèse puisse me donner au moins la crédibilité nécessaire pour<br />

éduquer et sensibiliser <strong>les</strong> générations futures aux conséquences éventuel<strong>les</strong> que court notre<br />

planète. En tant que « futur » professeur <strong>de</strong> géographie, je compte jouer un rôle en ce sens,<br />

aussi minime soit-il.<br />

Dans mon premier chapitre, qui s’adresse à un large public, j’ai tenté d’expliquer avec<br />

le plus <strong>de</strong> détails possible, dans un nombre limité <strong>de</strong> pages, la problématique générale <strong>de</strong><br />

l’ensemble <strong>de</strong> ma thèse. Ensuite, après un court chapitre qui introduit la zone d’étu<strong>de</strong> où ont<br />

été réalisés mes travaux, on retrouvera trois chapitres qui constituent ce qu’on pourrait<br />

appeler « le cœur <strong>de</strong> la thèse ». La science le voulant ainsi, ces trois chapitres sont assez<br />

pointus et sont écrits en anglais, la langue avec laquelle mes travaux seront le mieux<br />

disséminés au niveau international. Chacun <strong><strong>de</strong>s</strong> chapitres écrits en anglais est précédé d’un<br />

résumé en français qui décrit l’essentiel du contenu. Enfin, une discussion axée <strong>sur</strong> <strong>les</strong><br />

problèmes majeurs qui restent à explorer complètera le travail en guise <strong>de</strong> conclusion<br />

générale. Sur ce, je vous souhaite une lecture agréable et instructive.<br />

XX


Chapitre I : Introduction et problématique<br />

générale<br />

1


I.1. Introduction<br />

La vitesse avec laquelle l’Arctique se réchauffe est sans précé<strong>de</strong>nt dans l’histoire<br />

mo<strong>de</strong>rne <strong>de</strong> la société humaine. Les nombreuses conséquences du réchauffement <strong>de</strong><br />

l’Arctique, et d’autres <strong>changements</strong> environnementaux (e.g., augmentation du rayonnement<br />

ultraviolet), ont récemment fait l’objet d’un important rapport scientifique présenté par<br />

l’Arctic Council (Arctic Climate Impact Assessment, ACIA, 2005). Bien que <strong>les</strong> causes exactes <strong>de</strong><br />

ce réchauffement restent mal comprises, ce d’après le <strong>de</strong>rnier rapport <strong>de</strong> l’Intergovernmental<br />

Panel on Climate Change (IPCC, 2001), il semble que <strong>les</strong> émissions anthropiques <strong>de</strong> gaz à effet<br />

<strong>de</strong> serre soient, en gran<strong>de</strong> partie, responsab<strong>les</strong> <strong>de</strong> la situation qu’on observe aujourd’hui. En<br />

particulier, l’augmentation <strong>de</strong> la concentration atmosphérique en dioxy<strong>de</strong> <strong>de</strong> <strong>carbone</strong> (CO 2)<br />

est significative <strong>de</strong>puis le milieu <strong><strong>de</strong>s</strong> années 50 (Keeling, 1986; Sch<strong>les</strong>inger, 1990). Le CO 2 a<br />

la particularité, comme d’autre gaz à effet <strong>de</strong> serre tel que le méthane (CH 4), d’absorber le<br />

rayonnement infrarouge qui est réémis par la <strong>sur</strong>face <strong>de</strong> la Terre. Une meilleure<br />

compréhension <strong><strong>de</strong>s</strong> processus qui influencent la concentration <strong>de</strong> CO 2 dans l’atmosphère est<br />

donc primordiale dans le contexte <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong>. Le sujet <strong>de</strong> cette thèse<br />

s’intègre dans l’effort international qui vise à mieux comprendre le cycle global du <strong>carbone</strong><br />

afin <strong>de</strong> mieux appréhen<strong>de</strong>r l’évolution du climat.<br />

L’Océan Arctique joue un rôle primordial dans le système climatique global en<br />

permettant à la chaleur accumulée aux basses latitu<strong><strong>de</strong>s</strong> et transportée dans <strong>les</strong> hautes latitu<strong><strong>de</strong>s</strong><br />

<strong>de</strong> s’échapper, pour ainsi équilibrer le bilan énergétique <strong>de</strong> la planète. C’est aussi une région<br />

qui agit comme un « moteur » pour la circulation océanique globale avec la production d’eau<br />

<strong>de</strong>nse. Par exemple, l’Atlantique Nord est le site où se forme la masse d’eau profon<strong>de</strong><br />

connue sous le nom North Atlantic Deep Water (NADW), laquelle tapisse <strong>les</strong> fonds d’une vaste<br />

partie <strong>de</strong> l’Océan Mondial. La circulation générée par la formation <strong>de</strong> cette masse d’eau est<br />

<strong>de</strong> type thermohaline car elle résulte d’un changement <strong>de</strong> <strong>de</strong>nsité causé par <strong><strong>de</strong>s</strong> <strong>changements</strong><br />

<strong>de</strong> température (thermo-) ou <strong>de</strong> salinité (-haline). À l’ai<strong>de</strong> <strong>de</strong> modèle <strong>de</strong> circulation global,<br />

Rahmstorf (2002) a mis en évi<strong>de</strong>nce l’impact <strong>de</strong> la production <strong>de</strong> NADW <strong>sur</strong> le transport <strong>de</strong><br />

chaleur à l’échelle planétaire. Par exemple, <strong>les</strong> températures <strong>de</strong> l’air en <strong>sur</strong>face dans<br />

l’hémisphère Nord seraient jusqu’à 6°C inférieures sans la formation <strong>de</strong> cette masse d’eau<br />

(Rahmstorf, 2002).<br />

2


En plus d’influencer la circulation océanique globale, <strong>les</strong> phénomènes dits <strong>de</strong><br />

«rétroactions» 1 sont très importants dans l’Arctique. Les rétroactions peuvent soit accélérer le<br />

réchauffement global actuel (positive), soit l’inverser (négative). Les trois plus importantes,<br />

i<strong>de</strong>ntifiées dans le rapport <strong>de</strong> l’ACIA (2005) sont l’albédo <strong>de</strong> <strong>sur</strong>face 2 , la circulation<br />

thermohaline, et l’émission <strong>de</strong> gaz à effet <strong>de</strong> serre.<br />

1. Rétroaction due à l’albédo : la glace <strong>de</strong> mer, et son couvert <strong>de</strong> neige, réfléchie<br />

pratiquement 90% du rayonnement solaire vers l’espace. En réduisant l’étendue du<br />

couvert <strong>de</strong> glace, une plus gran<strong>de</strong> quantité d’énergie pénètre dans l’océan, induisant<br />

une augmentation <strong>de</strong> la température <strong>de</strong> la couche d’eau superficielle qui accélère la<br />

fonte <strong>de</strong> la glace. C’est une rétroaction positive.<br />

2. Rétroaction liée à la circulation thermohaline : l’augmentation du volume d’eau<br />

douce dans la couche superficielle <strong>de</strong> l’Arctique (en réponse à l’augmentation du<br />

débit <strong><strong>de</strong>s</strong> rivières Arctiques et <strong>de</strong> la fonte <strong><strong>de</strong>s</strong> glaces) intensifie la stratification <strong><strong>de</strong>s</strong><br />

eaux <strong>de</strong> <strong>sur</strong>face <strong>de</strong> l’Atlantique Nord, ce qui diminue la formation d’eaux <strong>de</strong>nses et<br />

un ralentissement global <strong>de</strong> la circulation thermohaline profon<strong>de</strong>. De tels<br />

<strong>changements</strong> risquent d’entraîner un refroidissement généralisé dans l’Hémisphère<br />

Nord (localisé), et peut être même à l’échelle globale. On parlera donc d’une<br />

rétroaction négative, mais cela reste très incertain à ce jour.<br />

3. Rétroaction due à l’émission <strong>de</strong> gaz à effet <strong>de</strong> serre d’origine naturelle : une gran<strong>de</strong><br />

quantité <strong>de</strong> <strong>carbone</strong> organique se trouve « séquestré» dans le pergélisol (c.f. permafrost<br />

en anglais; le sol qui est gelé en permanence) situé en marge <strong>de</strong> l’Océan Arctique. Le<br />

pergélisol, qui représente 25% <strong>de</strong> la <strong>sur</strong>face continentale <strong>de</strong> l’hémisphère nord, subit<br />

<strong>de</strong>puis une quarantaine d’années une augmentation <strong>de</strong> sa température et, à plusieurs<br />

endroits, une fonte progressive. Cette fonte conduit à une mobilisation du <strong>carbone</strong><br />

organique séquestré, qui est alors accessible à la minéralisation 3 microbienne<br />

aérobique (production <strong>de</strong> CO 2) et anaérobique (production <strong>de</strong> CH 4). Comme on<br />

s’attend à ce que la minéralisation excè<strong>de</strong> la production primaire (fixation <strong>de</strong> CO 2),<br />

on parlera <strong>de</strong> rétroaction positive.<br />

1<br />

Dans le contexte particulier <strong>de</strong> l’étu<strong>de</strong> <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong>, on utilise le terme feedback pour signifier<br />

un phénomène physique, biologique ou chimique qui, soumis à un réchauffement climatique, produit une<br />

action qui augmente (positif) ou diminue (négatif) le réchauffement (Woodwell et al., 1998).<br />

2<br />

Albédo planétaire : proportion <strong>de</strong> l’énergie solaire inci<strong>de</strong>nte qui est simplement réfléchie vers l’espace par le<br />

système terre.<br />

3<br />

Minéralisation : transformation <strong>de</strong> la matière organique en composés inorganiques (e.g. CO2, CH4).<br />

3


Dans cette thèse je m’intéresserai <strong>sur</strong>tout au <strong>de</strong>rnier point, i.e. à la rétroaction<br />

positive due à l’émission <strong>de</strong> gaz à effet <strong>de</strong> serre d’origine naturelle. Dans <strong>les</strong> sections<br />

suivantes, on détaillera la problématique <strong>de</strong> la thèse. Brièvement, la fonte du pergélisol et<br />

l’augmentation du débit <strong><strong>de</strong>s</strong> rivières <strong>de</strong> l’Arctique <strong>de</strong>vraient mobiliser une partie du <strong>carbone</strong><br />

organique sous forme dissoute vers l’Océan Arctique (section I.2.1). Dans l’océan, le <strong>carbone</strong><br />

organique dissous d’origine terrigène peut être minéralisé soit par <strong>les</strong> microorganismes, soit<br />

par la lumière. La minéralisation par la lumière est un phénomène physico-chimique<br />

complexe appelé photooxydation (pour oxydation photochimique) (section I.2.2). Les<br />

rayons ultraviolets (UV) sont <strong>les</strong> principaux responsab<strong>les</strong> <strong>de</strong> la photooxidation. Dans<br />

l’Arctique, <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face sont « protégées » <strong><strong>de</strong>s</strong> UV grâce à la présence <strong>de</strong> la glace <strong>de</strong><br />

mer. En parallèle, la diminution <strong>de</strong> l’ozone stratosphérique permet à une plus gran<strong>de</strong><br />

quantité <strong>de</strong> rayons UV d’atteindre la <strong>sur</strong>face <strong>de</strong> la Terre (section I.2.2). L’hypothèse centrale<br />

<strong>de</strong> cette thèse est que la photooxydation, suite à l’augmentation <strong><strong>de</strong>s</strong> apports<br />

continentaux et du rayonnement UV pénétrant dans la colonne d’eau, est un<br />

mécanisme clé qui contribue à accélérer la minéralisation du <strong>carbone</strong> organique<br />

terrigène, produisant du CO 2 dans la couche superficielle <strong>de</strong> l’océan Arctique, qui<br />

autrement aurait été « séquestré » dans l’océan profond.<br />

L’objectif fixé est <strong>de</strong> quantifier <strong>les</strong> processus <strong>de</strong> photooxydation dans une région<br />

<strong>de</strong> l’Arctique afin d’établir l’importance <strong>de</strong> ce phénomène dans le contexte actuel et<br />

d’appréhen<strong>de</strong>r <strong>les</strong> conséquences potentiel<strong>les</strong> du réchauffement climatique <strong>sur</strong> cette<br />

rétroaction positive.<br />

4


I.2. Problématique<br />

I.2.1. Importance <strong>de</strong> la matière organique dissoute d’origine terrigène dans l’Océan<br />

Arctique et impact du réchauffement <strong><strong>de</strong>s</strong> écosystèmes terrestres nordiques<br />

Bien qu’il soit connu <strong>de</strong>puis le début <strong><strong>de</strong>s</strong> années 70 que <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face <strong>de</strong><br />

l’Arctique sont riches en <strong>carbone</strong> organique dissous (DOC 1 ) (Kinney et al., 1971), ce n’est<br />

que récemment qu’on a démontré l’importance <strong>de</strong> la contribution <strong><strong>de</strong>s</strong> apports terrigènes. En<br />

effet, <strong>de</strong> 5 à 40% du DOC <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face <strong>de</strong> l’Arctique serait d’origine terrigène<br />

(Wheeler et al., 1997; Guay et al., 1999; Opsahl et al., 1999). L’Océan Arctique ne représente<br />

que 1% du volume <strong>de</strong> l’Océan global mais reçoit 11% <strong><strong>de</strong>s</strong> apports fluviaux mondiaux<br />

annuels (Rachold et al., 2004). En plus <strong>de</strong> contribuer au maintient <strong>de</strong> la stratification haline<br />

caractéristique <strong>de</strong> l’Arctique (voir Chap. 2), ces eaux douces, riches en <strong>carbone</strong> organique<br />

dissous, expliquent donc en partie <strong>les</strong> fortes concentrations <strong>de</strong> DOC observées dans<br />

plusieurs secteurs <strong>de</strong> l’Arctique (e.g., Kinney et al., 1971; An<strong>de</strong>rson et al., 1998; Kattner et al.,<br />

1999; Bussmann et Kattner, 2000; An<strong>de</strong>rson 2002; Amon, 2004; Benner et al., 2005). En plus<br />

<strong>de</strong> la gran<strong>de</strong> quantité <strong>de</strong> DOC déchargée dans l’Arctique, le mélange conservatif du DOC<br />

observé dans <strong>les</strong> estuaires – e.g. Lena (Cauwet et Sidorov, 1996), Yenisei et Ob (Köhler et al.,<br />

2003; Amon et Meon, 2004) - suggère que <strong>les</strong> processus <strong>de</strong> minéralisation sont pratiquement<br />

inexistants dans <strong>les</strong> environnements estuariens polaires. En aval <strong>de</strong> la zone <strong>de</strong> floculation, où<br />

<strong>les</strong> <strong>changements</strong> ioniques provoquent la transformation d’une fraction du DOC en particu<strong>les</strong><br />

(i.e. dans <strong>les</strong> eaux <strong>de</strong> salinité ~0-4; Droppo et al., 1998; Köhler et al., 2003), le DOC est<br />

éliminé soit par l’oxydation microbienne, soit par la photooxydation. L’importance<br />

relative <strong>de</strong> ces <strong>de</strong>ux processus <strong>de</strong> minéralisation est inconnue. Pour expliquer que seulement<br />

~40% <strong><strong>de</strong>s</strong> apports terrigènes annuels en DOC est exporté à l’extérieur <strong>de</strong> l’Arctique, Hansell<br />

et al. (2004) ont suggéré que <strong>les</strong> processus microbiens <strong>de</strong>vaient être responsab<strong>les</strong> <strong>de</strong> la<br />

majeure partie <strong>de</strong> la dégradation observée. Cependant, aucune étu<strong>de</strong> n’a permis <strong>de</strong> vérifier<br />

cette hypothèse. Au contraire, d’après l’étu<strong>de</strong> <strong>de</strong> Meon et Amon (2004), le DOC d’origine<br />

terrigène est très réfractaire et ne constitue pas une source significative <strong>de</strong> <strong>carbone</strong> organique<br />

pour <strong>les</strong> bactéries (Mer <strong>de</strong> Kara). Concernant la photooxydation, <strong>les</strong> étu<strong><strong>de</strong>s</strong> d’Amon et al.<br />

(2003), Amon et Meon (2004) et <strong>de</strong> Benner et al. (2004; 2005) suggèrent que ce mécanisme<br />

n’est pas important dans <strong>les</strong> zones polaires, sans toutefois le démontrer quantitativement.<br />

1 DOC : dissolved organic carbon.<br />

5


Les processus responsab<strong>les</strong> <strong>de</strong> la minéralisation <strong>de</strong> la matière organique d’origine terrigène<br />

dans l’Océan Arctique restent énigmatiques.<br />

Une meilleure compréhension <strong><strong>de</strong>s</strong> processus <strong>de</strong> minéralisation <strong>de</strong> la matière<br />

organique d’origine terrigène (et/ou <strong>de</strong> sa préservation) est cruciale dans le contexte<br />

<strong><strong>de</strong>s</strong> <strong>changements</strong> environnementaux récemment observés dans <strong>les</strong> écosystèmes<br />

terrestres nordiques. En effet, la tendance selon laquelle <strong>les</strong> écosystèmes terrestres <strong>de</strong><br />

l’Arctique agissent comme puits net <strong>de</strong> CO 2 et <strong>de</strong> CH 4 atmosphérique <strong>de</strong>puis la fin <strong>de</strong> la<br />

<strong>de</strong>rnière glaciation pourrait être renversée avec le réchauffement. Ces écosystèmes pourraient<br />

alors <strong>de</strong>venir une source significative <strong>de</strong> <strong>carbone</strong> pour l’atmosphère (e.g. Oechel et al., 2000).<br />

Les tourbières arctiques ont accumulées approximativement 455 x 10 15 g <strong>de</strong> <strong>carbone</strong><br />

organique <strong>de</strong>puis la <strong>de</strong>rnière glaciation (Gorham, 1991), soit près d’un tiers <strong>de</strong> la quantité <strong>de</strong><br />

<strong>carbone</strong> organique contenu dans <strong>les</strong> sols à l’échelle globale (Dixon et al., 1994). La majeur<br />

partie <strong>de</strong> ce <strong>carbone</strong> (~400 x 10 15 g C) se trouve actuellement emprisonnée dans le pergélisol<br />

(Davidson et Janssens, 2006; et références citées).<br />

Plusieurs étu<strong><strong>de</strong>s</strong> récentes suggèrent que la mobilisation d’une partie du <strong>carbone</strong><br />

contenu dans <strong>les</strong> sols arctiques se fera sous forme dissoute par le réseau hydrographique<br />

(Freeman et al., 2001, 2004; Mack et al., 2004; Billett et al., 2004). En accord avec ces étu<strong><strong>de</strong>s</strong>,<br />

Frey et Smith (2005) prévoient que la quantité <strong>de</strong> DOC exportée <strong>de</strong> la Sibérie vers l’Océan<br />

Arctique augmentera <strong>de</strong> 29 à 46% au cours du siècle présent. Ce scénario n’est pas<br />

<strong>sur</strong>prenant vu l’augmentation prévue du débit <strong><strong>de</strong>s</strong> rivières (jusqu’à 31%; Arnell, 2005) et la<br />

fonte progressive du pergélisol (Camill, 2005; Hinzman et al., 2005). Bien que la mobilisation<br />

du <strong>carbone</strong> « séquestré » n’a pas été détectée avec <strong>les</strong> analyses d’isotopes stab<strong>les</strong> du DOC<br />

dans <strong>les</strong> fleuves arctiques (Benner et al., 2004; Amon et Meon, 2004; Guo et Macdonald,<br />

2006), l’augmentation du débit <strong><strong>de</strong>s</strong> fleuves observée <strong>de</strong>puis <strong>les</strong> années 60 suggère une<br />

augmentation significative <strong>de</strong> l’export <strong>de</strong> <strong>carbone</strong> <strong><strong>de</strong>s</strong> écosystèmes terrestres vers l’océan<br />

(Figure I.1; Peterson et al., 2002, 2006; McClelland et al., 2006). Dans ce contexte, on peut se<br />

poser <strong>les</strong> questions suivantes :<br />

• Quel est le <strong><strong>de</strong>s</strong>tin <strong>de</strong> ce <strong>carbone</strong> organique dissous d’origine terrigène dans<br />

l’océan Arctique?<br />

• Quels sont <strong>les</strong> processus <strong>de</strong> dégradation dominants?<br />

6


Figure I.1. Anomalie débit d’eau douce (trait noir) <strong><strong>de</strong>s</strong> six plus grands<br />

fleuves sibériens qui se déchargent dans l’Océan Arctique <strong>de</strong>puis 1940.<br />

L’augmentation (7%) est corrélée à l’augmentation <strong>de</strong> la température <strong>de</strong><br />

<strong>sur</strong>face à l’échelle globale (trait bleu), ainsi qu’avec l’indice climatique NAO 1<br />

(trait rouge) (source : Peterson et al., 2002).<br />

1 NAO: north atlantic oscillation. L’indice NAO décrit la variabilité <strong>de</strong> pression atmosphérique entre l’Islan<strong>de</strong><br />

et <strong>les</strong> Açores. Le NAO est positif quand la pression atmosphérique aux Açores est plus élevé que la normale et<br />

plus faible en Islan<strong>de</strong> que la normale.<br />

7


I.2.2. Importance <strong>de</strong> la photooxydation et impact <strong>de</strong> la fonte <strong>de</strong> la glace <strong>de</strong> mer et <strong>de</strong><br />

l’augmentation du rayonnement ultraviolet dans l’Arctique<br />

En raison <strong>de</strong> son caractère réfractaire, l’absence d’un fort signal indiquant la présence<br />

<strong>de</strong> matière organique dissoute d’origine terrigène (tDOM 1 ) dans l’océan (Meyers-Schulte et<br />

Hedges, 1986) intrigue <strong>les</strong> océanographes <strong>de</strong>puis <strong><strong>de</strong>s</strong> décennies (voir Hedges et al., 1997 et<br />

réf. citées). À l’ai<strong>de</strong> d’un traceur moléculaire caractéristique <strong><strong>de</strong>s</strong> végétaux terrestres, la lignine,<br />

Opsahl et Benner (1997) et Hernes et Benner (2002; 2006) ont estimé le temps moyen <strong>de</strong><br />

rési<strong>de</strong>nce <strong>de</strong> la tDOM à 21-93 et ~90 ans dans <strong>les</strong> océans Atlantique et Pacifique,<br />

respectivement, ce qui suggère une minéralisation relativement rapi<strong>de</strong> <strong>de</strong> cette matière<br />

organique dans le milieu océanique.<br />

En 1977, Oliver C. Zafiriou suggérait que <strong>les</strong> processus photochimiques pouvaient<br />

jouer un rôle significatif dans le maintient <strong>de</strong> la transparence <strong>de</strong> l’eau <strong>de</strong> mer, et comme puits<br />

pour le DOC réfractaire dans l’océan. Ce n’est que récemment qu’on a pu mettre en<br />

évi<strong>de</strong>nce le rôle prépondérant que jouaient <strong>les</strong> réactions photochimiques dans la<br />

transformation et la minéralisation du DOC d’origine terrigène dans l’océan côtier (Kieber et<br />

al., 1990; Miller et Zepp, 1995; Amon et Benner, 1996; Opsahl et Benner 1998; Benner et<br />

Opsahl 2001; Hernes et Benner 2003). Puisque <strong>les</strong> processus photochimiques n’affectent que<br />

la couche superficielle <strong>de</strong> l’océan (i.e.


2003), <strong>les</strong>quels sont <strong>les</strong> principaux responsab<strong>les</strong> <strong><strong>de</strong>s</strong> réactions photochimiques en milieu<br />

aquatique (Zafiriou, 1977).<br />

On assiste <strong>de</strong>puis la fin <strong><strong>de</strong>s</strong> années 50 à une réduction <strong>de</strong> l’étendue du couvert estival<br />

<strong>de</strong> glace dans l’Arctique (Fig. I.2). La diminution observée est particulièrement importante en<br />

été, saison pendant laquelle l’étendue moyenne est passée <strong>de</strong> ~11 à ~7.5 10 6 km 2 en un <strong>de</strong>mi<br />

siècle. Cette tendance semble même s’accélérer d’après <strong>les</strong> observations satellita<strong>les</strong> qui ont<br />

enregistré une étendue minimale du couvert <strong>de</strong> glace pluriannuelle 1 durant quatre années<br />

consécutives <strong>de</strong>puis 2002 (Serreze et al., 2003; Stroeve et al., 2005; National Snow and Ice Data<br />

Center, 2006). En septembre 2005, le couvert <strong>de</strong> glace atteignait seulement ~5.5 10 6 km 2 , ce<br />

qui représente une réduction <strong>de</strong> l’ordre <strong>de</strong> 27% <strong>de</strong>puis le début <strong><strong>de</strong>s</strong> observations satellita<strong>les</strong>,<br />

il y a 26 ans (Figure I.3).<br />

Figure I.2. Observations <strong>de</strong> l’étendue moyenne du couvert <strong>de</strong> glace<br />

saisonnier au cours du 20 ième siècle (source <strong><strong>de</strong>s</strong> données : ACIA, 2005). La<br />

diminution <strong>de</strong> la glace est plus importante au printemps et en été, qu’en<br />

automne et en hiver.<br />

1 La glace pluriannuelle est celle qui <strong>sur</strong>vit à au moins un cycle <strong>de</strong> gel-dégel. Elle est observée <strong>de</strong> manière<br />

distincte quand l’étendue du couvert est minimale en septembre.<br />

9


Figure I.3. Étendue <strong>de</strong> la banquise Arctique observée par satellite<br />

annuellement au mois <strong>de</strong> septembre entre 1979 et 2005. (Source <strong><strong>de</strong>s</strong> données:<br />

National Snow and Ice Data Center, 2006).<br />

En parallèle au rétrécissement du couvert <strong>de</strong> glace, on assiste à une diminution <strong>de</strong> la<br />

concentration en ozone stratosphérique en réponse aux émissions anthropiques <strong>de</strong><br />

chlorofluorocarbures (CFC) qui, par une série <strong>de</strong> réactions chimiques avec <strong>les</strong> composés<br />

chlorés et bromés, conduit à la <strong><strong>de</strong>s</strong>truction <strong>de</strong> l’ozone stratosphérique. Entre 1979 et 2000, la<br />

concentration d’ozone a diminué <strong>de</strong> 11% en moyenne au printemps au <strong><strong>de</strong>s</strong>sus <strong>de</strong> l’Arctique<br />

(ACIA, 2005). L’augmentation du rayonnement ultraviolet reçu à la <strong>sur</strong>face en est une<br />

conséquence directe. La Figure I.4 montre un exemple <strong>de</strong> l’impact <strong>de</strong> l’ozone <strong>sur</strong><br />

l’éclairement UV reçu à la <strong>sur</strong>face <strong>de</strong> l’océan au midi solaire, à la latitu<strong>de</strong> <strong>de</strong> 70°N, durant le<br />

solstice d’été. Dans la littérature, le rayonnement ultraviolet comprend <strong>les</strong> longueurs d’on<strong>de</strong><br />

allant <strong>de</strong> 100 nm à 400 nm et se divise en trois sous-groupes : UV-C (100-280 nm), UV-B<br />

(280-315 nm) et UV-A (315-400 nm). Si l’ozone absorbe presque entièrement <strong>les</strong> UV-C, elle<br />

laisse pénétrer une partie <strong><strong>de</strong>s</strong> UV-B et n’affecte pratiquement pas <strong>les</strong> UV-A.<br />

En milieu aquatique, le rayonnement UV-B agit <strong>sur</strong> <strong>les</strong> cyc<strong>les</strong> biogéochimiques. Par<br />

exemple, il diminue la productivité primaire, accélère la décomposition du DOC et du<br />

tDOM, entraîne la production <strong><strong>de</strong>s</strong> gaz volati<strong>les</strong> et favorise le recyclage <strong><strong>de</strong>s</strong> micronutriments<br />

tel que le fer et le cuivre (voir la récente revue par Zepp et al., 2003). Gibson et al. (2000) se<br />

10


sont intéressés aux facteurs contrôlants l’exposition <strong><strong>de</strong>s</strong> microorganismes aquatiques aux<br />

radiations UV dans l’Arctique. Ils ont démontré que <strong>les</strong> variations <strong>de</strong> la concentration en<br />

matière organique dissoutes colorée (CDOM 1 ) jouent un rôle plus important que l’ozone <strong>sur</strong><br />

cette exposition. L’impact <strong><strong>de</strong>s</strong> <strong>changements</strong> environnementaux <strong>sur</strong> <strong>les</strong> réactions<br />

photochimiques impliquant le CDOM reste cependant inexploré.<br />

Récemment, Amon et Meon (2004), en se basant <strong>sur</strong> <strong><strong>de</strong>s</strong> expériences <strong>de</strong><br />

photooxidation réalisées <strong>sur</strong> <strong><strong>de</strong>s</strong> eaux provenant <strong><strong>de</strong>s</strong> fleuves Ob et Yenisei, confirmèrent la<br />

forte réactivité du CDOM, et suggèrent que « […] with reduced ice cover in the Arctic Ocean the fate<br />

of terrestrial DOM could be significantly different compared to present situation ». Ces diverses<br />

considérations m’amènent à formuler <strong>les</strong> questions suivantes :<br />

• Peut-on quantifier l’impact du rétrécissement du couvert <strong>de</strong> glace et <strong>de</strong><br />

l’augmentation du rayonnement UV-B <strong>sur</strong> la photooxidation du CDOM dans<br />

l’Arctique?<br />

• Quels sont <strong>les</strong> principaux facteurs (i.e. couvert <strong>de</strong> glace, UV-b, nébulosité,<br />

photoréactivité et concentration du CDOM) à prendre en compte?<br />

Figure I.4. Éclairement spectral dans <strong>les</strong> parties UV-A et UV-B pour trois<br />

concentrations en ozone tel<strong>les</strong> qu’indiquées <strong>sur</strong> la figure (DU = Dobson<br />

Units), et un exemple <strong>de</strong> ren<strong>de</strong>ment quantique apparent (voir la section<br />

suivante pour plus d’explications) exprimé en unités relatives, représentant la<br />

réactivité photochimique du CDOM en fonction <strong>de</strong> la longueur d’on<strong>de</strong><br />

(courbe grisée).<br />

1 CDOM : chromophoric dissolved organique matter. On quantifie sa concentration en me<strong>sur</strong>ant <strong>les</strong> propriétés<br />

d’absorption <strong>de</strong> l’eau <strong>de</strong> mer filtrée avec <strong><strong>de</strong>s</strong> filtres dont la porosité est d’environ 0.2 – 0.7 μm (on parlera du<br />

coefficient d’absorption du CDOM : a CDOM).<br />

11


I.2.3. Quantification <strong><strong>de</strong>s</strong> processus photochimiques<br />

Dans <strong>les</strong> bilans <strong>de</strong> <strong>carbone</strong> globaux ou régionaux, il est primordial <strong>de</strong> quantifier <strong>les</strong><br />

<strong>flux</strong> <strong>de</strong> <strong>carbone</strong> entre <strong>les</strong> réservoirs organique et inorganique, et particulaire et dissous. Dans<br />

le présent travail, je me suis intéressé à quantifier la transformation du <strong>carbone</strong> organique<br />

dissous (DOC) en <strong>carbone</strong> inorganique. Pour se faire, j’ai quantifié la production<br />

photochimique (i.e. photoproduction) <strong>de</strong> CO 2, le composé inorganique le plus important <strong><strong>de</strong>s</strong><br />

réactions photochimiques en milieux aquatiques (e.g. Miller et Zepp 1995). En réalité, la<br />

photoproduction <strong>de</strong> CO 2 est légèrement inférieure à minéralisation du DOC par<br />

photooxydation en milieu marin (Miller et Zepp 1995; Miller et Moran, 1997; Gao et Zepp,<br />

1998; Mopper et Kieber, 2002). En effet, le CO 2 représente le produit final d’une série <strong>de</strong><br />

réactions physiques et chimiques déclanchés par l’absorption <strong>de</strong> l’énergie radiative par le<br />

CDOM (Zafiriou, 1977). Parmi <strong>les</strong> mécanismes connus conduisant à la production<br />

photochimique <strong>de</strong> CO 2, on retrouve la photo<strong>de</strong>carboxylation (Mi<strong>les</strong> et Brezonik, 1981;<br />

Budac et Wan, 1992), où <strong><strong>de</strong>s</strong> groupes carboxyliques (-COOH) rattachés à <strong><strong>de</strong>s</strong> chaînes <strong>de</strong><br />

<strong>carbone</strong> organique sont clivés. Par ailleurs, il a récemment été démontré que le clivage <strong><strong>de</strong>s</strong><br />

structures aromatiques <strong><strong>de</strong>s</strong> substances humiques 1 pouvait aussi être responsable <strong>de</strong> la<br />

production <strong>de</strong> CO 2 dans <strong>les</strong> eaux riches en DOC d’origine terrigène (Vähätalo et al., 1999;<br />

Xie et al., 2004). Si <strong>les</strong> processus exacts menant à la production <strong>de</strong> CO 2 sont complexes et<br />

toujours mal connus (Mopper et Kieber, 2002), il est néanmoins possible d’estimer la<br />

photooxydation du DOC en milieu marin si <strong>les</strong> paramètres suivant sont connus (Zafiriou,<br />

1977) (voir aussi encadré 1.1):<br />

• L’éclairement solaire inci<strong>de</strong>nt à la <strong>sur</strong>face <strong>de</strong> l’océan, E(λ), qui dépend<br />

principalement <strong><strong>de</strong>s</strong> conditions atmosphériques (concentration en H 2O, O 3, aérosol,<br />

nuage, etc);<br />

• le spectre du coefficient d’absorption <strong><strong>de</strong>s</strong> chromophores organiques dissous, a CDOM;<br />

• l’efficacité avec laquelle le CO 2 est produit suite à l’excitation <strong><strong>de</strong>s</strong> chromophores par<br />

<strong>les</strong> photons absorbés, i.e. le ren<strong>de</strong>ment quantique apparent (AQY 2 ou φ).<br />

1 Substances humiques : mélange <strong>de</strong> matières organiques, <strong>de</strong> couleur foncée, fortement polymérisées, que l'on<br />

peut extraire <strong>de</strong> l'humus du sol par l'action <strong>de</strong> solutions alcalines diluées et qui sont précipitab<strong>les</strong> par <strong>les</strong> aci<strong><strong>de</strong>s</strong>.<br />

2 AQY: Apparent Quantum Yield.<br />

12


Encadré 1. Modèle <strong>de</strong> production photochimique <strong>de</strong> <strong>carbone</strong> inorganique à partir du<br />

<strong>carbone</strong> organique dissous<br />

En milieu aquatique, le CO2 se dissous et forme <strong><strong>de</strong>s</strong> ions bicabonatés (H2CO3 et HCO3 -). L’ensemble<br />

forme le <strong>carbone</strong> inorganique dissous ou DIC1. Le taux <strong>de</strong> photoproduction <strong>de</strong> DIC, PDIC, à une<br />

profon<strong>de</strong>ur z et suite à l’absorption <strong>de</strong> photons <strong>de</strong> longueur d’on<strong>de</strong> λ, peut être exprimé<br />

mathématiquement par :<br />

0<br />

( λ , z)<br />

= E ( λ,<br />

z)<br />

a ( λ,<br />

z)<br />

φ ( λ,<br />

z)<br />

, (E.1)<br />

PDIC CDOM DIC<br />

où E 0 est l’éclairement scalaire en mo<strong>les</strong> <strong>de</strong> photon m -2 s -1 (i.e., l’intégrale <strong>de</strong> l’énergie provenant <strong>de</strong><br />

toute <strong>les</strong> directions dans l’espace tridimensionnel), aCDOM est le coefficient d’absorption du CDOM en<br />

m -1(voir Encadré 2) et φDIC est le ren<strong>de</strong>ment quantique pour la photoproduction <strong>de</strong> DIC en mo<strong>les</strong><br />

DIC (mo<strong>les</strong> photon) -1. φDIC est le rapport entre le nombre <strong>de</strong> mo<strong>les</strong> <strong>de</strong> DIC produit par mo<strong>les</strong> <strong>de</strong><br />

photons absorbés par le CDOM.<br />

Dans le cadre <strong>de</strong> cette étu<strong>de</strong>, je m’intéresserai à la photoproduction intégrée dans toute la colonne<br />

d’eau (<strong>sur</strong> la verticale) par unité <strong>de</strong> <strong>sur</strong>face. Comme on ne connaît pas la variabilité verticale <strong>de</strong> tous<br />

<strong>les</strong> paramètres <strong>de</strong> l’équation E.1, j’adopterai <strong><strong>de</strong>s</strong> hypothèses simplificatrices, sans perdre pour autant<br />

trop <strong>de</strong> précision. Tout d’abord, <strong>les</strong> propriétés optiques <strong>de</strong> l’eau <strong>de</strong> mer sont supposées homogènes<br />

dans la couche où <strong>les</strong> réactions photochimiques prennent place. Cette hypothèse est vali<strong>de</strong> tant que la<br />

couche <strong>de</strong> mélange est plus gran<strong>de</strong> que la couche éclairée par le rayonnement UV et bleu, et que φDIC<br />

est indépendant <strong>de</strong> l’intensité <strong>de</strong> E 0 (contrairement à la production primaire par exemple). Ainsi, il<br />

suffira <strong>de</strong> connaître la quantité d’énergie qui pénètre la <strong>sur</strong>face <strong>de</strong> l’océan et la fraction qui est<br />

absorbée par le CDOM dans la colonne d’eau. En abandonnant la dimension verticale, on peut<br />

réécrire plus simplement l’équation E.1 comme :<br />

− aCDOM<br />

( λ)<br />

PDIC ( λ) = Ed<br />

( λ,<br />

0 ) [ 1−<br />

R(<br />

λ)<br />

] φDIC<br />

( λ)<br />

, (E.2)<br />

at<br />

( λ)<br />

où Ed est l’éclairement <strong><strong>de</strong>s</strong>cendant juste sous l’interface air-mer (indiqué par 0-), R est la réflectance<br />

<strong>de</strong> l’océan (i.e. le pourcentage d’énergie rétrodiffusée vers l’atmosphère, voir Encadré 2) et at est le<br />

coefficient d’absorption totale <strong>de</strong> tout <strong>les</strong> constituants optiquement efficaces (i.e. l’eau <strong>de</strong> mer pure,<br />

<strong>les</strong> particu<strong>les</strong> détritiques et minéra<strong>les</strong>, le phytoplancton et le CDOM). Dans la plupart <strong><strong>de</strong>s</strong> conditions<br />

océaniques, le facteur [1-R] approche 1 et peut donc être abandonné. Enfin, en intégrant l’équation<br />

E.2 dans le domaine spectral photochimiquement efficace, i.e. entre 290 nm et 600 nm, on obtient la<br />

production intégrée <strong>de</strong> DIC dans la colonne d’eau,<br />

600<br />

− aCDOM<br />

( λ)<br />

PDIC = ∫ Ed<br />

( λ,<br />

0 ) φDIC<br />

( λ)<br />

dλ<br />

. (E.3)<br />

λ=<br />

290<br />

a ( λ)<br />

Parmi <strong>les</strong> paramètres <strong>de</strong> l’Eq. E.3, le ren<strong>de</strong>ment quantique apparent pour la<br />

photoproduction <strong>de</strong> CO 2 (ou DIC), φ DIC(λ), est <strong>de</strong> loin le moins bien connu. Le manque <strong>de</strong><br />

données <strong>sur</strong> φ DIC dans <strong>les</strong> eaux océaniques s’explique par <strong><strong>de</strong>s</strong> difficultés expérimenta<strong>les</strong><br />

diffici<strong>les</strong> à <strong>sur</strong>monter. Tout d’abord, étant donné que la concentration <strong>de</strong> DIC dans l’eau <strong>de</strong><br />

mer s’éleve à ~2000 μM, une production photochimique <strong>de</strong> moins <strong>de</strong> 10 μM <strong>de</strong> DIC<br />

1 DIC: dissolved inorganic carbon.<br />

13<br />

t


(pendant une expérience d’une durée <strong>de</strong> ~12 h) s’avère indétectable (Johannessen, 2000). Les<br />

premières me<strong>sur</strong>es <strong>de</strong> φ DIC(λ) proviennent donc <strong><strong>de</strong>s</strong> eaux douces récoltées soit dans <strong>les</strong><br />

rivières (Gao et Zepp, 1998), soit dans <strong><strong>de</strong>s</strong> lacs (Vähätalo et al., 2000) où <strong>les</strong> concentrations<br />

initia<strong>les</strong> en DIC sont faib<strong>les</strong>, ce qui rend possible la me<strong>sur</strong>e directe <strong>de</strong> la production<br />

photochimique <strong>de</strong> DIC. Gao et Zepp (1998) ont publié <strong>les</strong> premières valeurs <strong>de</strong> φ DIC à trois<br />

longueurs d’on<strong>de</strong> dans la partie UV du spectre (290, 350, et 390 nm) pour <strong><strong>de</strong>s</strong> eaux très<br />

riches en DOC provenant <strong>de</strong> la rivière Satilla (Georgia, USA). Vähätalo et al. (2000) ont<br />

proposé une métho<strong>de</strong> pour estimer le spectre <strong>de</strong> φ DIC à partir d’incubation in situ à différentes<br />

profon<strong>de</strong>urs. Grâce à un modèle spectral <strong>de</strong> propagation <strong>de</strong> la lumière couplé à un modèle<br />

<strong>de</strong> photooxidation, la photoproduction <strong>de</strong> DIC est calculée en exprimant φ DIC(λ)=x x 10 yλ .<br />

Les coefficients x et y sont retrouvés en minimisant la différence entre <strong>les</strong> productions<br />

calculées et me<strong>sur</strong>ées (par une procédure d’optimisation mathématique). La métho<strong>de</strong> <strong>de</strong><br />

Vähätalo et al. (2000), qui <strong>de</strong>man<strong>de</strong> <strong><strong>de</strong>s</strong> temps d’incubation in situ <strong>de</strong> l’ordre d’une semaine,<br />

n’est pas adaptée aux applications océanographiques où le temps passé à une station<br />

d’échantillonnage est relativement court (souvent


I.2.4. La télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan : un outil <strong>de</strong> choix pour quantifier <strong>les</strong><br />

processus photochimiques aux échel<strong>les</strong> océaniques<br />

Durant <strong>les</strong> étés <strong>de</strong> 1967 et 1968, <strong>les</strong> chercheurs américains George L. Clarke, Gifford<br />

C. Ewing et Carl J. Lorenzen effectuaient <strong><strong>de</strong>s</strong> <strong>sur</strong>vols <strong>de</strong>puis Cape Cod jusqu’à la mer <strong><strong>de</strong>s</strong><br />

Sargasses à bord d’un avion équipé d’un spectromètre me<strong>sur</strong>ant la lumière rétrodiffusée par<br />

l’océan. Ainsi <strong>les</strong> premiers spectres <strong>de</strong> lumière sortant <strong>de</strong> la mer étaient me<strong>sur</strong>és à distance; la<br />

télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan voyait le jour. Avec <strong><strong>de</strong>s</strong> me<strong>sur</strong>es in situ simultanées <strong>de</strong><br />

concentration en chlorophylle a, Clarke et al. (1970) démontraient la possibilité d’utiliser la<br />

télédétection pour cartographier, <strong>de</strong>puis l’espace, la distribution <strong>de</strong> ce pigment à <strong><strong>de</strong>s</strong> échel<strong>les</strong><br />

spatia<strong>les</strong> encore jamais explorées à cette époque. Pour expliquer <strong>les</strong> anomalies <strong><strong>de</strong>s</strong> spectres<br />

ne pouvant pas être attribuées à la chlorophylle a, Clarke et al. (1970) suggéraient entre autre<br />

la présence <strong>de</strong> particu<strong>les</strong> en suspension et <strong>de</strong> « biochromes ». Ces travaux pionniers<br />

ouvraient la voie, avec le lancement en octobre 1978 du satellite Nimbus-7 équipé du capteur<br />

Coastal Zone Color Scanner (CZCS) (Hovis et al., 1980; Gordon et al., 1980), à la télédétection<br />

spatiale <strong>de</strong> la couleur <strong>de</strong> l’océan. Depuis, <strong>les</strong> applications utilisant <strong>les</strong> données « couleur <strong>de</strong><br />

l’océan » acquises <strong>de</strong>puis l’espace se sont multipliées. Ne sont citées ici que <strong>les</strong> plus<br />

courantes : étu<strong>de</strong> <strong>de</strong> la distribution spatiale et <strong>de</strong> l’évolution temporelle <strong>de</strong> la biomasse<br />

phytoplanctonique aux échel<strong>les</strong> régionale et globale, production primaire à l’échelle globale,<br />

étu<strong>de</strong> <strong><strong>de</strong>s</strong> structures physiques océaniques tels <strong>les</strong> tourbillons et <strong>les</strong> méandres du Gulf Stream.<br />

À l’heure actuelle, on compte neuf capteurs en orbite autour <strong>de</strong> la terre qui fournissent <strong><strong>de</strong>s</strong><br />

données <strong>de</strong> façon quasi-continue (voir site internet <strong>de</strong> l’IOCCG 1 ).<br />

En 1997, Cullen et al. proposaient l’idée d’utiliser <strong><strong>de</strong>s</strong> données couleur <strong>de</strong> l’océan<br />

pour estimer l’atténuation <strong><strong>de</strong>s</strong> UV en milieu marin afin <strong>de</strong> quantifier <strong>les</strong> <strong>flux</strong> photochimiques<br />

et <strong>les</strong> dommages causés par <strong>les</strong> UV aux microorganismes planctoniques. Quelques années<br />

plus tard, Johannessen (2000) développait une métho<strong>de</strong> empirique pour estimer l’atténuation<br />

diffuse (voir Encadré 2) <strong>de</strong> la lumière dans l’UV à partir <strong><strong>de</strong>s</strong> me<strong>sur</strong>es du capteur SeaWiFS<br />

(Sea Wi<strong>de</strong> Field-of-View Sensor), un capteur en orbite <strong>de</strong>puis septembre 1997. Les valeurs du<br />

coefficient d’atténuation diffuse dans l’UV permettaient ainsi d’i<strong>de</strong>ntifier trois types d’eaux le<br />

long <strong>de</strong> la côte Est américaine, pour <strong>les</strong>quel<strong>les</strong> <strong>les</strong> spectres <strong>de</strong> φ DIC(λ) avaient été déterminés<br />

(Johannessen et Miller, 2001). Il s’agissait là <strong>de</strong> la première tentative d’utiliser la télédétection<br />

1 IOCCG : International Ocean Colour Coordinating Group, www.ioccg.org/sensors/current.html .<br />

15


<strong>de</strong> la couleur <strong>de</strong> l’océan comme un outil pour la quantification <strong><strong>de</strong>s</strong> processus<br />

photochimiques. Plus récemment, Cédric G. Fichot proposait une métho<strong>de</strong> statistique<br />

robuste permettant d’estimer, à partir <strong><strong>de</strong>s</strong> données SeaWiFS, l’atténuation <strong><strong>de</strong>s</strong> UV,<br />

l’absorption du CDOM et le taux <strong>de</strong> photoproduction <strong>de</strong> monoxy<strong>de</strong> <strong>de</strong> <strong>carbone</strong> dans la<br />

colonne d’eau à l’échelle globale. Les travaux <strong>de</strong> Fichot (2004), bien que peu connus,<br />

représentent probablement le meilleur exemple <strong>de</strong> l’utilité <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan<br />

pour l’estimation <strong><strong>de</strong>s</strong> <strong>flux</strong> photochimiques à partir <strong>de</strong> l’espace. Il s’agit, néanmoins, d’une<br />

application encore peu exploitée dont le plein potentiel reste à découvrir.<br />

Les données couleur <strong>de</strong> l’océan peuvent servir à améliorer la quantification <strong>de</strong><br />

la photoproduction <strong>de</strong> CO 2 (ou DIC) dans l’Arctique. Cependant, l’exploitation <strong>de</strong> ces<br />

données dans <strong>les</strong> régions polaires pose plusieurs problèmes qui <strong>de</strong>man<strong>de</strong>nt <strong><strong>de</strong>s</strong> étu<strong><strong>de</strong>s</strong> plus<br />

approfondies. Dans le cas <strong>de</strong> l’Arctique, on peut mentionner :<br />

1) le manque <strong>de</strong> connaissance <strong><strong>de</strong>s</strong> propriétés optiques <strong><strong>de</strong>s</strong> eaux Arctiques, plus<br />

particulièrement <strong><strong>de</strong>s</strong> eaux <strong><strong>de</strong>s</strong> plateaux continentaux, et l’absence d’algorithmes pour<br />

estimer le CDOM;<br />

2) la présence <strong>de</strong> glace <strong>de</strong> mer qui empêche la télédétection <strong><strong>de</strong>s</strong> propriétés optiques <strong>de</strong><br />

l’eau sous-jacente et complique celle <strong><strong>de</strong>s</strong> eaux environnantes;<br />

3) l’éclairement solaire faible, voire nul, pour la majeur partie <strong>de</strong> l’année limite la<br />

pério<strong>de</strong> <strong>de</strong> détection entre mai et août;<br />

4) l’angle zénithal solaire 1 élevé qui représente un problème particulier pour l’estimation<br />

<strong>de</strong> la contribution <strong>de</strong> l’atmosphère au signal me<strong>sur</strong>é au niveau du satellite, laquelle est<br />

essentielle afin <strong>de</strong> retrouver le signal marin.<br />

Dans mon travail <strong>de</strong> thèse, j’ai examiné <strong>les</strong> <strong>de</strong>ux premiers points. Je présenterai d’abord une<br />

revue <strong>de</strong> la littérature <strong>sur</strong> <strong>les</strong> propriétés optiques me<strong>sur</strong>ées dans <strong>les</strong> eaux <strong>de</strong> l’Arctique. Je<br />

montrerai pourquoi il est difficile d’estimer le coefficient d’absorption du CDOM à partir <strong>de</strong><br />

la télédétection. Enfin, j’expliquerai comment la glace <strong>de</strong> mer peut ruiner la qualité <strong><strong>de</strong>s</strong><br />

données couleur <strong>de</strong> l’océan.<br />

1 L’angle zénithal solaire est l’angle qui se me<strong>sur</strong>e entre la normale sortant <strong>de</strong> la <strong>sur</strong>face <strong>de</strong> la mer et la direction<br />

du soleil. Un angle <strong>de</strong> 0° correspond à un soleil au zénith.<br />

16


I.2.4.1. Propriétés optiques <strong><strong>de</strong>s</strong> eaux arctiques et algorithmes pour l’estimation du<br />

coefficient d’absorption du CDOM<br />

Les premières me<strong>sur</strong>es optiques réalisées dans l’océan Arctique ont été rapportées<br />

par Raymond C. Smith en 1973, qui a déterminé <strong>les</strong> valeurs spectra<strong>les</strong> du coefficient<br />

d’atténuation diffuse, K d(λ) et d’atténuation (c(λ)) (voir encadré 2). Dans la partie verte du<br />

spectre, ces valeurs s’avéraient comparab<strong>les</strong> aux eaux <strong>les</strong> plus claires connues à l’époque.<br />

Avec ces me<strong>sur</strong>es, toutes réalisées sous le couvert <strong>de</strong> glace en mai, on a longtemps cru que<br />

<strong>les</strong> eaux arctiques faisaient partie <strong><strong>de</strong>s</strong> eaux <strong>les</strong> plus transparentes au mon<strong>de</strong>.<br />

Il a fallu attendre plus <strong>de</strong> quinze ans avant que d’autres me<strong>sur</strong>es spectra<strong>les</strong> soient<br />

réalisées dans l’Arctique. Topliss et al. (1989) ont analysé <strong>les</strong> me<strong>sur</strong>es <strong>de</strong> K d(λ) acquises dans<br />

le nord <strong>de</strong> la Baie <strong>de</strong> Baffin (Ouest du Groënland) suggérant que ces eaux diffèrent <strong><strong>de</strong>s</strong> eaux<br />

du Cas 1 tel<strong>les</strong> que définies par Morel et Prieur (1977). Topliss et al. (1989) ont noté que le<br />

coefficient <strong>de</strong> diffusion <strong><strong>de</strong>s</strong> particu<strong>les</strong> était plus élevé que pour <strong>les</strong> eaux du Cas 1, et que la<br />

contribution <strong><strong>de</strong>s</strong> produits <strong>de</strong> dégradation était relativement importante. Peu après, <strong>les</strong><br />

données <strong>de</strong> K d(λ) acquises dans la Mer <strong>de</strong> Barents et dans le détroit <strong>de</strong> Fram (Fram Strait),<br />

publiées par Mitchell (1992), montraient clairement la présence d’un fort signal provenant <strong>de</strong><br />

matière non-corrélée avec le phytoplancton (voir Figure 3 <strong>de</strong> Mitchell, 1992). En comparant<br />

ces observations avec d’autres faites en Antarctique, Mitchell (1992) concluait que <strong>les</strong> eaux<br />

<strong>de</strong> l’Arctique étaient probablement plus influencées par le CDOM d’origine terrigène. Cette<br />

hypothèse a largement été vérifiée récemment grâce aux me<strong>sur</strong>es in situ <strong>de</strong> fluorescence 1 du<br />

CDOM (F CDOM) et d’absorption du CDOM (a CDOM) réalisées dans différents secteurs <strong>de</strong><br />

l’Arctique :<br />

• Arctique central (F CDOM, Guay et al., 1999; a CDOM Pegau 2002);<br />

• Détroit <strong>de</strong> Fram (F CDOM, Amon et al., 2003);<br />

• Mers <strong>de</strong> Chukchi et Beaufort (a CDOM, Wang et al., 2005a; F CDOM et a CDOM,Guéguen et al.,<br />

2005);<br />

• Polynie <strong><strong>de</strong>s</strong> eaux du Nord (a CDOM, Scully et Miller, 2000).<br />

1 Une fraction <strong>de</strong> l’énergie absorbée par le CDOM fait passer <strong>les</strong> électrons d’un chromophore à un niveau<br />

d’énergie supérieur. En retournant très rapi<strong>de</strong>ment à leur niveau d’énergie initial, ces molécu<strong>les</strong> émettent <strong><strong>de</strong>s</strong><br />

photons. C’est la fluorescence. La fluorescence du CDOM est souvent utilisée pour estimer la concentration en<br />

DOC (e.g. Vodacek et al.,1997; Ferrari, 2000; Guay et al., 1999; Amon et al., 2003)<br />

17


Dans la partie Ouest <strong>de</strong> l’Arctique, i.e. la région qui nous intéressera dans cette thèse<br />

(Chap. 2), <strong>les</strong> étu<strong><strong>de</strong>s</strong> <strong>de</strong> Pegau (2002), Wang et al., (2005a) et Guéguen et al. (2005) ont<br />

confirmé que l’absorption dans la partie bleue du spectre était largement dominée par le<br />

CDOM. Pegau (2002) attribue <strong>les</strong> fortes valeurs <strong>de</strong> a CDOM dans <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face à<br />

l’abondance <strong><strong>de</strong>s</strong> apports terrigènes et <strong>les</strong> faib<strong>les</strong> taux <strong>de</strong> photooxydation dans ces eaux<br />

couvertes <strong>de</strong> glace.<br />

Une forte absorption <strong>de</strong> la lumière par le CDOM caractérise <strong>les</strong> propriétés<br />

optiques <strong><strong>de</strong>s</strong> eaux Arctiques; par rapport aux autres bassins océaniques (e.g., Barnard et al.,<br />

1998). Il est reconnu que la présence <strong>de</strong> CDOM d’origine terrigène influence la réflectance<br />

<strong>de</strong> la mer (e.g., Car<strong>de</strong>r et al., 1989, 1991), et que l’estimation <strong>de</strong> a CDOM à partir <strong><strong>de</strong>s</strong> images<br />

satellita<strong>les</strong> est possible (e.g., Sathyendranath et al., 1989). Cependant aucune métho<strong>de</strong> n’est<br />

actuellement disponible pour estimer a CDOM à partir du spectre <strong>de</strong> réflectance. Or il s’agit d’un<br />

paramètre nécessaire, avec le coefficient d’absorption totale (en fait, le rapport [a CDOM/a t]),<br />

pour quantifier <strong>les</strong> <strong>flux</strong> photochimiques (Eq. E.2).<br />

18


Encadré 2. Définition <strong><strong>de</strong>s</strong> propriétés optiques inhérentes et apparentes<br />

En optique marine il est utile <strong>de</strong> distinguer <strong>les</strong> propriétés optiques dites inhérentes et apparentes.<br />

Cette distinction fut proposée par Preisendorfer (1961) afin <strong>de</strong> faciliter l’interprétation <strong><strong>de</strong>s</strong> me<strong>sur</strong>es<br />

optiques et <strong>de</strong> la théorie du transfert radiatif en milieu marin. Les propriétés inhérentes (IOP1) sont<br />

indépendantes <strong><strong>de</strong>s</strong> <strong>changements</strong> dans la distribution du champ radiatif et ne dépen<strong>de</strong>nt que <strong><strong>de</strong>s</strong><br />

propriétés intrinsèques <strong>de</strong> la matière. Il s’agit <strong><strong>de</strong>s</strong> coefficients d’absorption (a) et <strong>de</strong> l’indicatrice<br />

angulaire <strong>de</strong> diffusion (β(ψ, Φ); avec ψ, l’angle <strong>de</strong> diffusion, compris entre 0 et π, et Φ, l’angle<br />

azimutal compris entre 0 et 2π). Le coefficient <strong>de</strong> diffusion (b) se calcule en intégrant l’indicatrice <strong>de</strong><br />

diffusion dans toutes <strong>les</strong> directions :<br />

∫<br />

π<br />

b ( λ)<br />

= 2π<br />

β ( λ,<br />

ψ ) sin( ψ ) dψ<br />

(E.4)<br />

0<br />

Le coefficient <strong>de</strong> rétrodiffusion (bb) se calcule <strong>de</strong> la même manière, mais en n’intégrant que la partie<br />

arrière <strong>de</strong> β, i.e. avec π/2 < ψ < π. La somme <strong>de</strong> a et b donne le coefficient d’atténuation (c). a, b et c<br />

varient en fonction <strong>de</strong> la longueur d’on<strong>de</strong> (λ) et l’unité <strong>de</strong> me<strong>sur</strong>e est le m -1. Dans <strong>les</strong> eaux naturel<strong>les</strong>,<br />

<strong>les</strong> IOP tota<strong>les</strong> peuvent être obtenues en additionnant <strong>les</strong> coefficients <strong>de</strong> l’eau pure, du<br />

phytoplancton, <strong><strong>de</strong>s</strong> détritus, <strong><strong>de</strong>s</strong> particu<strong>les</strong> minéra<strong>les</strong>, et du CDOM.<br />

Les propriétés optiques apparentes (AOP1) dépen<strong>de</strong>nt à la fois <strong><strong>de</strong>s</strong> IOP et <strong><strong>de</strong>s</strong> <strong>changements</strong> dans la<br />

distribution du champ radiatif (e.g., élévation du soleil, présence <strong>de</strong> nuages, état <strong>de</strong> la <strong>sur</strong>face <strong>de</strong> la<br />

mer). Les AOP s’obtiennent par la combinaison <strong>de</strong> quantité d’énergie me<strong>sur</strong>able tels l’éclairement (E)<br />

et la luminance (L). La luminance est le rayonnement électromagnétique enregistré dans <strong>de</strong> petits<br />

ang<strong>les</strong> soli<strong><strong>de</strong>s</strong> pointés <strong>sur</strong> <strong><strong>de</strong>s</strong> directions données, alors que l’éclairement est l’intégrale <strong><strong>de</strong>s</strong><br />

luminances dans un espace tridimensionnel donné. Par exemple, l’éclairement <strong><strong>de</strong>s</strong>cendant, Ed(λ),<br />

me<strong>sur</strong>é par un capteur cosinus est<br />

2π<br />

π / 2<br />

∫ = 0 ∫ =<br />

Ed ( λ) = L(<br />

θ,<br />

ϕ,<br />

λ)<br />

cosθ<br />

sinθdθdϕ<br />

(E.5)<br />

ϕ<br />

θ 0<br />

où θ et ϕ représentent <strong>les</strong> ang<strong>les</strong> zénithaux et azimutaux respectivement. L’éclairement ascendant,<br />

Eu(λ), se calcule <strong>de</strong> la même façon mais en intégrant L avec θ allant <strong>de</strong> π/2 à π. Le coefficient<br />

d’atténuation diffus, Kd(λ) est souvent utilisé en optique marine pour désigner le taux <strong>de</strong> décroissance<br />

verticale <strong>de</strong> Ed(λ) dans la colonne d’eau :<br />

d ln( Ed<br />

( λ))<br />

K d ( λ) = − (m<br />

dz<br />

-1). (E.6)<br />

Pour quantifier <strong>les</strong> variations spectra<strong>les</strong> <strong>de</strong> la couleur <strong>de</strong> la mer, on fait référence à la réflectance qui<br />

se calcule soit comme le rapport entre l’éclairement montant et l’éclairement <strong><strong>de</strong>s</strong>cendant (sans unité):<br />

Eu<br />

( λ)<br />

R ( λ)<br />

= , (E.7)<br />

E ( λ)<br />

d<br />

soit comme le rapport entre la luminance sortant <strong>de</strong> l’eau (Lw) dans une direction donnée (θ, ϕ) et<br />

l’éclairement <strong><strong>de</strong>s</strong>cendant qui est en quelque sorte une réflectance « directionnelle » (en anglais remote<br />

sensing reflenctance):<br />

Lw<br />

( λ,<br />

θ , ϕ)<br />

R rs ( λ)<br />

= . (sr<br />

E ( λ)<br />

-1) (E.8)<br />

d<br />

À partir d’un jeu limité <strong>de</strong> données acquises dans <strong>les</strong> mers <strong>de</strong> Chukchi et <strong>de</strong> Beaufort,<br />

Wang et Cota (2003) ont évalué <strong>de</strong>ux algorithmes permettant <strong>de</strong> retrouver <strong>les</strong> IOP <strong><strong>de</strong>s</strong><br />

1 IOP and AOP : Inherent and Apparent Optical Property.<br />

19


principaux constituants <strong>de</strong> l’eau. Ces métho<strong><strong>de</strong>s</strong>, dites semi-analytiques, se basent <strong>sur</strong> <strong><strong>de</strong>s</strong><br />

formes simplifiées <strong><strong>de</strong>s</strong> équations du transfert radiatif qui relient la réflectance <strong>de</strong> la mer aux<br />

IOP (voir encadré 3 et aussi Mobley, 1994). Les algorithmes semi-analytiques <strong>de</strong> Lee et al.<br />

(1999, 2001) et <strong>de</strong> Maritorena et al. (2002) testés par Wang et Cota (2003) se sont avérés<br />

fiab<strong>les</strong> pour l’estimation <strong><strong>de</strong>s</strong> coefficients d’absorption du phytoplancton et <strong>de</strong> rétrodiffusion<br />

<strong><strong>de</strong>s</strong> particu<strong>les</strong> en suspension, mais peu performantes quant à l’estimation du coefficient<br />

d’absorption par <strong>les</strong> matières détritiques colorées (CDM 1 ). En effet, la plupart <strong><strong>de</strong>s</strong> métho<strong><strong>de</strong>s</strong><br />

d’inversion regroupent le a CDOM avec l’absorption par <strong>les</strong> particu<strong>les</strong> non-alga<strong>les</strong> (NAPs)<br />

(a CDM=a CDOM+a NAP) en raison <strong>de</strong> la similarité <strong>de</strong> leur spectre d’absorption. Ceux-ci peuvent<br />

être décris avec une fonction exponentielle du type (Bricaud et al., 1981; Roesler et al., 1989),<br />

−S<br />

( λ−λ0<br />

)<br />

a cdm = acdm<br />

( λ0<br />

) e<br />

(I.1)<br />

où λ 0 est la longueur d’on<strong>de</strong> <strong>de</strong> référence et le paramètre S est connu comme étant la pente<br />

spectrale. Seigel et al. (2002) publiaient récemment <strong><strong>de</strong>s</strong> cartes <strong>de</strong> distribution globale <strong>de</strong> l’a CDM<br />

(dont la variabilité était attribuée principalement au CDOM qui représente en moyenne<br />

~82% du a CDM) qui apportèrent <strong>de</strong> nouveaux indices <strong>sur</strong> <strong>les</strong> processus régulant sa<br />

concentration dans <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face (i.e. transport vertical, blanchiment). Dans l’objectif<br />

d’utiliser <strong>les</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan pour l’étu<strong>de</strong> <strong><strong>de</strong>s</strong> processus<br />

photochimiques, il est essentiel <strong>de</strong> distinguer a CDOM <strong>de</strong> a NAP. Parmi <strong>les</strong> 12 algorithmes<br />

semi-analytiques récemment évalués par l’IOCCG (2005) pour estimer <strong>les</strong> IOPs, aucun ne<br />

permettait <strong>de</strong> faire cette distinction. Quelques étu<strong><strong>de</strong>s</strong> ont proposé <strong><strong>de</strong>s</strong> algorithmes<br />

empiriques régionaux pour estimer a CDOM (Kahru et Michtell, 2001; D’Sa et Miller, 2003;<br />

Johannessen et al., 2003; Kowalczuk et al. 2005). Cependant, dans <strong>les</strong> eaux côtières<br />

optiquement complexes où la contribution <strong><strong>de</strong>s</strong> NAPs à l’absorption peut être supérieure à<br />

celle du CDOM (Babin et al., 2003b), <strong>les</strong> algorithmes empiriques sont en général peu fiab<strong>les</strong><br />

(IOCCG, 2006). Enfin pour pouvoir calculer la photoproduction <strong>de</strong> DIC intégrée dans la<br />

colonne d’eau (encadré E.1), il nous faudra connaître le rapport entre l’absorption du<br />

CDOM et l’absorption totale. Par conséquent, une métho<strong>de</strong> pour estimer la contribution<br />

du CDOM à l’absorption totale se doit d’être mise au point pour améliorer <strong>les</strong><br />

estimations du taux <strong>de</strong> photoproduction <strong>de</strong> DIC intégrée dans la colonne d’eau.<br />

1 CDM : colored <strong>de</strong>trital material.<br />

20


Encadré 3. Modè<strong>les</strong> semi-analytiques <strong>de</strong> réflectance <strong>de</strong> la mer.<br />

Des solutions numériques (e.g. Monte Carlo) aux équations <strong>de</strong> transfert radiatif permettent <strong>de</strong><br />

déduire <strong><strong>de</strong>s</strong> relations simplifiées entre <strong>les</strong> IOPs et <strong>les</strong> AOPs (Gordon et al., 1975; Morel et Prieur,<br />

1977). Plusieurs auteurs ont proposé <strong><strong>de</strong>s</strong> expressions simplifiées pour exprimer la réflectance (Rrs;<br />

voir encadré 2), en fonction du rapport bb/a+bb en se basant <strong>sur</strong> <strong><strong>de</strong>s</strong> analyses théoriques et <strong><strong>de</strong>s</strong><br />

simulations numériques. Les plus connues ont été proposées par Morel et ses collègues (Morel et<br />

Gentili, 1991, 1993, 1996; Loisel et Morel, 2001; Morel et al., 2002),<br />

− f ⎛ bb<br />

⎞ f ' ⎛ bb<br />

⎞<br />

R = ⎜ ⎟ = ⎜<br />

⎟<br />

rs ( 0 )<br />

(E.9)<br />

Q ⎝ a ⎠ Q ⎝ a + bb<br />

⎠<br />

et par Gordon et al. (1988a),<br />

− ⎛ bb<br />

⎞ ⎛ bb<br />

⎞<br />

R ( 0 ) = 1 ⎜<br />

⎟ + 2 ⎜<br />

⎟<br />

rs g g<br />

, (E.10)<br />

⎝ a + bb<br />

⎠ ⎝ a + bb<br />

⎠<br />

où <strong>les</strong> facteurs f/Q et f’/Q dépen<strong>de</strong>nt <strong>de</strong> la géométrie <strong>de</strong> visé, <strong>de</strong> l’angle zénithal solaire et <strong>de</strong> la<br />

concentration en chlorophylle (valable pour <strong>les</strong> eaux du Cas 1; Morel et al., 2002), et g1 et g2 sont <strong><strong>de</strong>s</strong><br />

constantes (0.0949, 0.0794; Gordon et al., 1988a). Plusieurs modè<strong>les</strong> d’inversion utilisés en<br />

télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan, i.e. <strong>les</strong> modè<strong>les</strong> permettant <strong>de</strong> retrouver <strong>les</strong> IOPs à partir <strong><strong>de</strong>s</strong><br />

me<strong>sur</strong>es d’AOPs, s’appuient <strong>sur</strong> la relation proposée par Gordon et al., (1988a). On peut citer à titre<br />

d’exemple <strong>les</strong> métho<strong><strong>de</strong>s</strong> : d’optimisation spectrale <strong>de</strong> Garver et Siegel (1997) et <strong>de</strong> Lee et al., (1996;<br />

1999); d’inversion matricielle <strong>de</strong> Hoge et al. (1995) et <strong>de</strong> Hoge et Lyon (1996); et d’algorithme quasianalytique<br />

<strong>de</strong> Lee et al. (2002).<br />

Les IOP <strong>de</strong> l’équation E.10 sont simplement la somme <strong><strong>de</strong>s</strong> IOP <strong>de</strong> chacun <strong><strong>de</strong>s</strong> constituants <strong>de</strong> l’eau.<br />

En générale, l’absorption est<br />

a = a + a + a = a + a + a + a<br />

(E.11)<br />

w<br />

p<br />

CDOM<br />

où <strong>les</strong> indices w, p, NAP, phy et CDOM représentent, l’eau, <strong>les</strong> particu<strong>les</strong>, <strong>les</strong> particu<strong>les</strong> non-alga<strong>les</strong>,<br />

le phytoplancton et le CDOM, respectivement. Quant à la rétrodiffusion, on fait la somme entre <strong>les</strong><br />

coefficients <strong><strong>de</strong>s</strong> particu<strong>les</strong> et <strong>de</strong> l’eau pure :<br />

b = b + b<br />

(E.12)<br />

b<br />

bw<br />

21<br />

w<br />

bp<br />

NAP<br />

2<br />

phy<br />

CDOM


I.2.4.2. La glace <strong>de</strong> mer : un problème pour la télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan<br />

La télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan dans <strong>les</strong> hautes latitu<strong><strong>de</strong>s</strong> fait face à un<br />

problème additionnel, comparé aux basses et moyennes latitu<strong><strong>de</strong>s</strong>, directement lié à la<br />

présence <strong>de</strong> glace <strong>de</strong> mer. En effet, la glace <strong>de</strong> mer est la plupart du temps recouverte <strong>de</strong><br />

neige pouvant réfléchir plus <strong>de</strong> 90% <strong>de</strong> la lumière visible et du proche-infrarouge inci<strong>de</strong>nte<br />

(e.g. Grenfell et Maykut, 1977; Perovich et al., 1998, 2002). En comparaison, l’océan réfléchi<br />

en général moins ~10% <strong>de</strong> la lumière dans le visible, et moins <strong>de</strong> 3% <strong>de</strong> la lumière du proche<br />

infrarouge (e.g. Morel, 1988; Morel et Maritorena, 2001). À l’époque du capteur CZCS, cette<br />

différence <strong>de</strong> réflectance causait <strong><strong>de</strong>s</strong> problèmes pour la télédétection <strong><strong>de</strong>s</strong> eaux le long <strong>de</strong> la<br />

glace en « aveuglant » <strong>les</strong> détecteurs du capteur (Comiso et al., 1990). L’amélioration <strong><strong>de</strong>s</strong><br />

performances radiométriques <strong><strong>de</strong>s</strong> nouveaux capteurs (e.g., le capteur SeaWiFS; Gordon et<br />

Wang, 1994) permet aujourd’hui <strong>de</strong> faire <strong><strong>de</strong>s</strong> observations même à proximité <strong>de</strong> la glace <strong>de</strong><br />

mer (Wang et al., 2005b). Cependant, la présence <strong>de</strong> glace génère <strong><strong>de</strong>s</strong> problèmes d’autres<br />

natures. En comparant <strong>les</strong> estimations <strong>de</strong> chlorophylle a obtenues par SeaWiFS avec <strong><strong>de</strong>s</strong><br />

me<strong>sur</strong>es in situ couvrant la plupart <strong><strong>de</strong>s</strong> bassins océaniques, Gregg et Casey (2004) ont<br />

i<strong>de</strong>ntifié un biais <strong>de</strong> l’ordre <strong>de</strong> 100% dans <strong>les</strong> eaux situées en marge <strong>de</strong> l’océan Arctique (mer<br />

<strong>de</strong> Barents), soit la pire performance <strong>de</strong> SeaWiFS à l’échelle régionale. Pour expliquer cette<br />

piètre performance, ces auteurs suggéraient que « […] contamination by drifting, subpixel-scale sea<br />

ice un<strong>de</strong>tected in the processing algorithms […] » <strong>de</strong>vait en être la cause. Pour illustrer le problème,<br />

j’ai représenté schématiquement <strong>les</strong> trajectoires possib<strong>les</strong> d’un photon émis par le soleil par<br />

temps dégagé, dans un système atmosphère-océan influencé par la présence <strong>de</strong> glace <strong>de</strong> mer<br />

(Fig. I.5). En plus <strong><strong>de</strong>s</strong> trajectoires connues, et prises en compte dans le traitement <strong><strong>de</strong>s</strong><br />

données couleur <strong>de</strong> l’océan (voir encadré 4), la glace <strong>de</strong> mer peut contribuer au signal détecté<br />

au niveau du satellite <strong>de</strong> façon 1) directe, quand, <strong>de</strong> taille bien inférieure à celle d’un pixel,<br />

elle se trouve à l’intérieure dans un pixel (contamination sub-pixel), ou 2) indirecte, quand<br />

elle se trouve à proximité et que la lumière qu’elle réfléchie est redirigée par diffusion vers le<br />

capteur qui vise un pixel occupé que par <strong>de</strong> l’eau (effet dit <strong>de</strong> l’environnement; Tanré et al.,<br />

1979, 1981).<br />

L’impact <strong>de</strong> la contamination <strong><strong>de</strong>s</strong> données couleur <strong>de</strong> l’océan par la glace <strong>de</strong><br />

mer n’a jamais fait l’objet d’étu<strong>de</strong>. Ayant comme objectif d’utiliser <strong>les</strong> données acquises<br />

par télédétection pour estimer le taux <strong>de</strong> photoproduction <strong>de</strong> DIC dans <strong>les</strong> eaux arctiques,<br />

22


une étu<strong>de</strong> dédiée au problème spécifique que pose la glace <strong>de</strong> mer vis-à-vis <strong>de</strong> l’exploitation<br />

<strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan, était nécessaire.<br />

Figure I.5. Trajectoires possib<strong>les</strong> dans le système atmosphère-océan pour un<br />

photon émis par le soleil par temps dégagé (voir aussi encadré 4). De la<br />

quantité <strong>de</strong> photon diffusée dans l’océan qui parvient à sortir <strong>de</strong> l’océan en<br />

direction du satellite (a; L w), une partie sera transmise à travers l’atmosphère<br />

jusqu’au capteur (b) alors que l’autre partie sera perdue par diffusion ou<br />

absorption dans l’atmosphère (c). Certains photons en provenance du soleil<br />

sont transmis directement (d) ou indirectement (e) dans le champ <strong>de</strong> visée du<br />

capteur après avoir été réfléchis <strong>de</strong> façon spéculaire par l’interface air-mer (L r,<br />

en anglais sun glint ou glitter). Comme pour L w, une partie sera transmise (g) et<br />

l’autre sera perdue dans l’atmosphère (f). Dans la partie visible du spectre,<br />

plus <strong>de</strong> 90% <strong>de</strong> l’énergie reçue au satellite provient <strong>de</strong> l’atmosphère (L p) et<br />

comprend : <strong>les</strong> photons diffusés une (h) ou plusieurs (i) fois par <strong>les</strong><br />

molécu<strong>les</strong> ou <strong>les</strong> aérosols atmosphériques. Dans <strong>les</strong> régions polaires où la<br />

glace <strong>de</strong> mer est présente, il faut ajouter à ce schéma la contribution <strong><strong>de</strong>s</strong><br />

photons réfléchis par la glace à l’extérieur du champ <strong>de</strong> visée qui sont<br />

redirigés vers le capteur (l; effet <strong>de</strong> l’environnement). Enfin, quand la glace<br />

se trouvant dans le champ <strong>de</strong> visée du capteur, une partie <strong><strong>de</strong>s</strong> photons qu’elle<br />

réfléchira sera transmise (m; contamination sub-pixel) et l’autre sera<br />

perdue dans l’atmosphère (n). Les contributions indirectes (l) et directes (m)<br />

posent problème puisqu’el<strong>les</strong> ne sont pas prises en compte dans le traitement<br />

<strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan.<br />

23


Encadré 4. Introduction au traitement <strong><strong>de</strong>s</strong> données couleur <strong>de</strong> l’océan<br />

Un instrument dédié à la télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan doit enregistrer une certaine quantité<br />

d’énergie radiative en la convertissant en un courant électrique dont l’intensité sera fonction <strong>de</strong><br />

l’énergie reçue. Le signal ainsi créé est ensuite digitalisé et traité pour être converti en une quantité<br />

physique en W m -2 nm -1 (corrections radiométriques élémentaires: première calibration, application<br />

<strong><strong>de</strong>s</strong> gains, etc.) La quantité d’énergie enregistrée dépend <strong>de</strong> la portion du spectre électromagnétique<br />

concernée et également du champ <strong>de</strong> vue instantané (IFOV en anglais pour Instantaneous field-of-view),<br />

qui correspond à l’élément <strong>de</strong> base observable (pixel). Le rayonnement électromagnétique enregistré,<br />

i.e. la luminance, est qualifiée <strong>de</strong> luminance totale (Lt), dans la me<strong>sur</strong>e où son intensité dans une<br />

direction donnée est déterminée aussi bien par <strong>les</strong> propriétés optiques <strong>de</strong> l’océan que par cel<strong>les</strong> <strong>de</strong><br />

l’interface air-mer et <strong>sur</strong>tout cel<strong>les</strong> <strong>de</strong> l’atmosphère (y compris <strong>les</strong> nuages ou la glace s’ils sont<br />

présents). Comme cette luminance contient <strong><strong>de</strong>s</strong> informations <strong>sur</strong> la totalité du système océanatmosphère,<br />

il faut en premier lieu retrancher <strong>de</strong> ce signal la contribution atmosphérique.<br />

L’ensemble <strong><strong>de</strong>s</strong> opérations nécessaires pour extraire la « luminance marine » (Lw) du signal total est<br />

connu sous le terme <strong>de</strong> « corrections atmosphériques ». El<strong>les</strong> consistent en effet à éliminer la partie<br />

du rayonnement qui a été rétrodiffusée par <strong>les</strong> molécu<strong>les</strong> (LR) et <strong>les</strong> aérosols <strong>de</strong> l’atmosphère (LA et<br />

LRA), éventuellement aussi réfléchie par la <strong>sur</strong>face <strong>de</strong> l’océan (Lr), mais qui n’a jamais pénétré l’océan.<br />

Les corrections atmosphériques sont en général basées <strong>sur</strong> une décomposition du signal en<br />

différentes contributions tel<strong>les</strong> que<br />

Lt = (LR + LA + LRA) + TLr + tLw . (E.13)<br />

Les <strong>de</strong>ux contributions majeures au signal me<strong>sur</strong>é proviennent <strong>de</strong> la diffusion par <strong>les</strong> molécu<strong>les</strong> <strong>de</strong><br />

l’air et <strong>de</strong> la diffusion par <strong>les</strong> aérosols (LR + LA + LRA = Lp). L’estimation <strong>de</strong> la part du signal due à la<br />

diffusion par <strong>les</strong> molécu<strong>les</strong> (LR) ne pose pas <strong>de</strong> problèmes majeurs (e.g., Gordon et al., 1988b ;<br />

Gordon et Wang, 1992a,b). Le terme comprenant la réflexion spéculaire du soleil (TLr) par l’interface<br />

air-mer peut être estimé en connaissant la vitesse et la direction du vent, le coefficient <strong>de</strong> réflexion <strong>de</strong><br />

Fresnel, et l’angle par rapport au soleil (Cox et Munk, 1954; Ebuchi et Kizu, 2002).<br />

L’estimation <strong>de</strong> la part du signal dû à la diffusion par <strong>les</strong> aérosols (LA + LRA) est, en revanche, la<br />

difficulté majeure <strong>de</strong> la correction atmosphérique, puisque ni la concentration, ni l’indicatrice <strong>de</strong><br />

diffusion sont connues a priori, c’est-à-dire avant <strong>de</strong> réaliser la correction du signal. Pour déterminer<br />

ces <strong>de</strong>ux inconnues, au moins <strong>de</strong>ux équations doivent être posées, ce qui signifie dans la pratique que<br />

<strong><strong>de</strong>s</strong> informations doivent être obtenues à au moins <strong>de</strong>ux longueurs d’on<strong>de</strong>. La plupart <strong><strong>de</strong>s</strong> techniques<br />

actuellement utilisées reposent <strong>sur</strong> l’observation du système océan + atmosphère dans au moins <strong>de</strong>ux<br />

canaux du proche infrarouge, pour <strong>les</strong>quels le signal océanique est nul (en tout cas dans <strong>les</strong> eaux du<br />

Cas 1). Une fois corrigé <strong>de</strong> l’effet <strong>de</strong> la diffusion moléculaire et <strong>de</strong> la réflexion spéculaire, le signal<br />

restant est entièrement dû aux aérosols. À partir <strong>de</strong> l’intensité <strong>de</strong> ce signal et <strong>de</strong> sa dépendance<br />

spectrale entre <strong>les</strong> <strong>de</strong>ux longueurs d’on<strong>de</strong> considérées, on obtient suffisamment d’informations <strong>sur</strong><br />

l’aérosol en présence pour pouvoir en extrapoler la contribution vers <strong>les</strong> longueurs d’on<strong>de</strong> du<br />

domaine visible, et ainsi corriger <strong>les</strong> luminances me<strong>sur</strong>ées <strong><strong>de</strong>s</strong> effets atmosphériques (pour plus<br />

détails voir : Gordon et Wang, 1994 ; Gordon, 1997 ; Antoine et Morel 1999).<br />

24


I.3. Objectifs et Organisation <strong>de</strong> la Thèse<br />

La problématique décrite ci-<strong><strong>de</strong>s</strong>sus a été abordée dans le cadre du programme<br />

CASES (Canadian Arctic Shelf Exchange Study). CASES est un vaste programme <strong>de</strong> recherche<br />

canadien mis <strong>sur</strong> pied en 2000 avec la collaboration <strong>de</strong> scientifiques provenant <strong>de</strong> sept pays<br />

différents (Canada, État-Unis, Japon, Danemark, Norvège, Espagne, France). Le but ultime<br />

<strong>de</strong> CASES est <strong>de</strong> mieux comprendre <strong>les</strong> conséquences biogéochimiques et écologiques<br />

résultant <strong>de</strong> la fonte <strong>de</strong> la glace présente <strong>sur</strong> le plateau continental du Mackenzie. Entre<br />

octobre 2003 et août 2004, le navire <strong>de</strong> recherche canadien CCGS Amundsen est resté <strong>sur</strong> le<br />

site d’étu<strong>de</strong> <strong>de</strong> CASES. Les données optiques et photochimiques qui furent acquises dans ce<br />

cadre ont permis <strong>de</strong> poursuivre <strong>les</strong> objectifs spécifiques suivants :<br />

1. Déterminer, pour la première fois, le ren<strong>de</strong>ment quantique pour la production <strong>de</strong><br />

DIC <strong>sur</strong> <strong><strong>de</strong>s</strong> échantillons d’eau provenant <strong>de</strong> l’Arctique.<br />

2. Afin <strong>de</strong> déterminer l’importance relative <strong>de</strong> la photooxydation dans le bilan <strong>de</strong><br />

<strong>carbone</strong> actuel, comparer le taux <strong>de</strong> photoproduction <strong>de</strong> DIC aux taux <strong>de</strong><br />

production primaire totale, nouvelle et séquestrée, ainsi qu’à la <strong>de</strong>man<strong>de</strong> en <strong>carbone</strong><br />

par <strong>les</strong> bactéries.<br />

3. À partir <strong>de</strong> l’archive d’observations satellita<strong>les</strong> <strong>de</strong> la glace <strong>de</strong> mer, <strong>de</strong> l’ozone<br />

stratosphérique et du couvert nuageux obtenue entre 1979 et 2003, évaluer l’impact<br />

<strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong> <strong>sur</strong> la photoproduction annuelle <strong>de</strong> DIC.<br />

4. Développer une métho<strong>de</strong> pour estimer le rapport entre <strong>les</strong> coefficients d’absorption<br />

du CDOM et total ([a CDOM/a t]) dans <strong>les</strong> eaux côtières à partir <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur<br />

<strong>de</strong> l’océan. Évaluer la sensibilité du modèle <strong>de</strong> photoproduction <strong>de</strong> DIC à<br />

l’extrapolation [a CDOM/a t], estimé à 412 nm, entre l’UV et le vert.<br />

5. Décrire et comprendre l’influence <strong>de</strong> la glace <strong>de</strong> mer <strong>sur</strong> l’estimation <strong>de</strong> la luminance<br />

marine à partir <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan, ainsi que <strong>les</strong> erreurs résultantes <strong>sur</strong><br />

<strong>les</strong> produits géophysiques qui sont dérivés <strong>de</strong> ces luminances contaminées (i.e. la<br />

chlorophylle, <strong>les</strong> IOPs). Proposer une métho<strong>de</strong> pour détecter <strong>les</strong> pixels contaminés<br />

par la glace <strong>de</strong> mer.<br />

6. Quantifier la photoproduction <strong>de</strong> DIC en utilisant <strong>les</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan.<br />

Afin <strong>de</strong> familiariser le lecteur à la zone d’étu<strong>de</strong>, le Chapitre II présente une brève<br />

revue <strong><strong>de</strong>s</strong> caractéristiques généra<strong>les</strong> <strong>de</strong> l’Océan Arctique et du sud-est <strong>de</strong> la Mer <strong>de</strong> Beaufort.<br />

25


Ensuite, <strong>les</strong> objectifs énumérés ci-<strong><strong>de</strong>s</strong>sus seront abordés à travers quatre chapitres<br />

relativement indépendants et présentés sous forme d’artic<strong>les</strong> scientifiques. Chacun est<br />

constitué <strong><strong>de</strong>s</strong> sections introduction, métho<strong><strong>de</strong>s</strong>, résultats, discussion et conclusion. Le<br />

Chapitre III répond aux trois premiers objectifs concernant le problème <strong>de</strong> la<br />

photooxidation du CDOM dans <strong>les</strong> eaux arctiques. Dans le chapitre IV, je propose une<br />

métho<strong>de</strong> qui permet d’estimer la fraction <strong>de</strong> l’énergie radiative absorbée par le CDOM dans<br />

la colonne d’eau à partir <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan. Le problème supplémentaire que<br />

pose la présence <strong>de</strong> glace <strong>de</strong> mer <strong>sur</strong> la qualité <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan acquises en<br />

milieu polaire est développé dans le chapitre V. Le Chapitre VI démontre l’utilité <strong><strong>de</strong>s</strong><br />

données couleur <strong>de</strong> l’océan pour l’estimation <strong>de</strong> la production photochimique <strong>de</strong> DIC dans<br />

<strong>les</strong> eaux côtières <strong>de</strong> la mer <strong>de</strong> Beaufort.<br />

En guise <strong>de</strong> conclusion, quelques perspectives d’avenir <strong>sur</strong> <strong>les</strong> travaux nécessaires à<br />

faire dans <strong>les</strong> domaines d’étu<strong><strong>de</strong>s</strong> abordés au cours <strong>de</strong> ma thèse seront discutées.<br />

26


I.4. Thesis objectives and organization<br />

The problematic <strong><strong>de</strong>s</strong>cribed above has been addressed within the frame work of the<br />

CASES program (Canadian Arctic Shelf Exchange Study). CASES is an important<br />

Canadian initiative created in 2000 with the collaboration of scientists coming from seven<br />

different countries (Canada, United States, Japan, Norway, Spain and France). The<br />

ultimate objective of CASES is to better un<strong>de</strong>rstand the ecological and biogeochemical<br />

consequences resulting from the reduction in the sea ice cover over the Mackenzie Shelf<br />

and in the Amundsen Gulf. Between October 2003 and August 2004, the Canadian<br />

research vessel CCGS Amundsen stayed in the CASES study area. The optical and<br />

photochemical mea<strong>sur</strong>ements collected during this field campaign allowed me to address<br />

the following specific objectives:<br />

1. To <strong>de</strong>termine, for the first time, the apparent quantum yield for the<br />

photoproduction of dissolved inorganic carbon (DIC) on samp<strong>les</strong> collected in the<br />

Arctic.<br />

2. To assess the relative importance of CDOM photooxidation in the carbon budget,<br />

by comparing the rate of DIC photoproduction to 1) the rates of the total, new and<br />

sequestered primary production, and 2) the rate of bacterial carbon <strong>de</strong>mand.<br />

3. To assess the impact of environmental changes on the annual DIC<br />

photoproduction using archived satellite observations of sea ice, ozone and cloud<br />

cover from 1979 to 2003.<br />

4. To <strong>de</strong>velop a method to <strong>de</strong>rive the ratio between chromophoric dissolved organic<br />

matter (CDOM) and total absorption coefficients ([aCDOM/at]) from Ocean Color<br />

data in coastal waters, which is a parameter of the DIC photoproduction mo<strong>de</strong>l.<br />

To assess the impact of the spectral extrapolation of [aCDOM/at], mea<strong>sur</strong>ed at 412<br />

nm, from the UV to the green part of the spectrum on the DIC photoproduction<br />

estimation.<br />

5. To <strong><strong>de</strong>s</strong>cribe and to better un<strong>de</strong>rstand the effect of sea ice on the retrieval of the<br />

water leaving radiance from satellite Ocean Color data, and the resulting errors on<br />

the geophysical parameters <strong>de</strong>rived from the contaminated spectrum (i.e.<br />

27


chlorophyll a concentration, IOP, [aCDOM/at]). To propose a method to <strong>de</strong>tect the<br />

contaminated pixels by the sea ice.<br />

6. To quantify the <strong>de</strong>pth-integrated DIC photoproduction using Ocean Color data.<br />

Except the Chapter II, which provi<strong><strong>de</strong>s</strong> a brief overview (in french) of the main<br />

characteristics of the Arctic Ocean, the Beaufort Sea and the Mackenzie Shelf, each<br />

Chapter is presented un<strong>de</strong>r the form of an usual scientific paper with: abstract,<br />

introduction, materials and methods, results, discussion and conclusions sections. Chapter<br />

III addresses the first three objectives enumerated above, which are related to the role of<br />

CDOM photooxidation in the Arctic waters. In Chapter IV, I propose a method to<br />

estimate the contribution of CDOM to the total light absorption coefficient from Ocean<br />

Color data in coastal waters. The specific problems posed by the presence of sea ice in<br />

the exploitation on Ocean Color imagery at high latitu<strong>de</strong> are studied in Chapter V. Using<br />

the results of Chapters III to V, the Chapter VI <strong>de</strong>monstrates the utility of Ocean Color<br />

remote sensing to quantify the DIC photoproduction in the coastal waters of the Beaufort<br />

Sea.<br />

In the overall conclusions and perspectives, I will discuss a number of issues<br />

related to Ocean Color remote sensing and to the quantification of the CDOM<br />

photooxidation in the Arctic waters that required more investigation in the future.<br />

28


Chapitre II : L’Océan Arctique, la Mer <strong>de</strong> Beaufort<br />

et le Plateau du Mackenzie<br />

29


II.1. Introduction<br />

La Mer <strong>de</strong> Beaufort est située dans la partie ouest <strong>de</strong> l’Arctique, le long <strong><strong>de</strong>s</strong> côtes <strong><strong>de</strong>s</strong><br />

Territoires du nord-ouest du Canada et <strong>de</strong> l’Alaska (Fig. II.1). Afin <strong>de</strong> mieux comprendre le<br />

contexte géographique et océanographique <strong>de</strong> la Mer <strong>de</strong> Beaufort, ce chapitre offre d’abord<br />

une <strong><strong>de</strong>s</strong>cription générale <strong>de</strong> l’Océan Arctique : <strong>les</strong> masses d’eau, la circulation générale, <strong>les</strong><br />

différents types <strong>de</strong> glace et la biologie au sens large. Ensuite je décris un cycle annuel typique<br />

<strong><strong>de</strong>s</strong> conditions physiques qui caractérisent le Plateau du Mackenzie, qui se situe dans le sud-<br />

est <strong>de</strong> la Mer <strong>de</strong> Beaufort. L’emphase sera mise <strong>sur</strong> la variabilité <strong>de</strong> la glace <strong>de</strong> mer et <strong>de</strong> la<br />

distribution <strong><strong>de</strong>s</strong> eaux douces qui proviennent du Fleuve Mackenzie et/ou <strong>de</strong> la fonte <strong>de</strong> la<br />

glace. Enfin je résume, en quelques lignes, le bilan <strong>de</strong> <strong>carbone</strong> organique du Plateau du<br />

Mackenzie. Ce bilan, tel que synthétisé par Macdonald et al. (1998 ; 2004b), intègre <strong>les</strong><br />

connaissances accumulées <strong>de</strong>puis <strong>les</strong> années 70 à travers <strong>de</strong> nombreuses étu<strong><strong>de</strong>s</strong> réalisées dans<br />

cette région connue pour ses réserves en hydrocarbures.<br />

30


Figure II.1. Patrons <strong>de</strong> circulation générale <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face dans l’Océan<br />

Arctique. Les eaux Atlantique et Pacifique (en rouge) pénètrent dans<br />

l’Arctique via la Mer <strong>de</strong> Barents et le détroit <strong>de</strong> Bering respectivement. Le<br />

Tourbillon <strong>de</strong> Beaufort et le courant Transpolaire (en bleu) sont <strong>les</strong> <strong>de</strong>ux<br />

principaux chemins empruntés par <strong>les</strong> eaux Arctiques, <strong>les</strong>quel<strong>les</strong> sont<br />

éventuellement exportées vers l’Atlantique Nord via <strong>de</strong> détroit <strong>de</strong> Fram et<br />

l’Archipel Canadien. Ces eaux <strong>de</strong> <strong>sur</strong>face sont fortement influencées par <strong>les</strong><br />

grands fleuves Sibériens et Nord Américains. Les huit plus importants sont<br />

indiqués avec, entre parenthèses, leur débit annuel (volume d’eau douce en<br />

km 3 ans -1 ). Le cadre noir localise la zone d’étu<strong>de</strong>. (Source <strong><strong>de</strong>s</strong> informations :<br />

ACIA, 2005)<br />

31


II.2. Généralités <strong>sur</strong> l’Océan Arctique<br />

L’Océan Arctique représente 2.6% <strong>de</strong> la <strong>sur</strong>face <strong>de</strong> l’Océan mondial, mais seulement<br />

1% en terme <strong>de</strong> volume. Cela s’explique par sa géomorphologie caractérisée par un large<br />

plateau continental occupant 52.7% <strong>de</strong> son étendue (Jakobsson et al., 2004). Les échanges<br />

avec <strong>les</strong> océans Pacifique et Atlantique sont relativement limités et confinées entre <strong>les</strong><br />

continents Eurasien, Américain et le Groenland (Fig. II.1), ce qui fait <strong>de</strong> l’Arctique un Océan<br />

dit « méditerranéen ». La majeure partie <strong>de</strong> ces échanges se fait avec l’Atlantique à travers le<br />

détroit <strong>de</strong> Fram (Fram Strait; voir Ru<strong>de</strong>ls et al., 2005) et la mer <strong>de</strong> Barents (voir Indvaldsen et<br />

al., 2004). Les eaux d’origine Atlantique sont plus salées (S~35; Indvaldsen et al., 2004) que<br />

cel<strong>les</strong> d’origine Pacifique (31.9-33; Woodgate et al., 2005b). Lors <strong>de</strong> son passage dans la mer<br />

<strong>de</strong> Barents et le détroit <strong>de</strong> Fram, une partie <strong><strong>de</strong>s</strong> eaux Atlantiques se refroidit et plonge pour<br />

renouveler <strong>les</strong> masses d’eau profon<strong><strong>de</strong>s</strong> <strong>de</strong> l’Arctique. Le temps <strong>de</strong> rési<strong>de</strong>nce <strong><strong>de</strong>s</strong> eaux<br />

profon<strong><strong>de</strong>s</strong> arctiques est <strong>de</strong> ~25 ans entre 200 et 1000 m, et <strong>de</strong> ~300 ans au <strong>de</strong>là <strong>de</strong> 1000 m<br />

<strong>de</strong> profon<strong>de</strong>ur (Macdonald et al., 2004a). Les eaux profon<strong><strong>de</strong>s</strong> sont isolées <strong>de</strong> la <strong>sur</strong>face par la<br />

une couche connue sous le nom <strong>de</strong> halocline (50 à 200 m). Cette <strong>de</strong>rnière est alimentée par<br />

<strong><strong>de</strong>s</strong> eaux formées <strong>sur</strong> <strong>les</strong> plateaux durant la formation <strong><strong>de</strong>s</strong> glaces (e.g., Aagaard et al., 1981), et<br />

par <strong>les</strong> eaux du Pacifique qui pénètrent dans l’Arctique via le détroit <strong>de</strong> Bering (e.g.,<br />

Woodgate et al., 2005a). La halocline joue un rôle important dans la stabilité <strong>de</strong> l’Arctique car<br />

elle isole l’atmosphère et la glace superficielle d’un important réservoir <strong>de</strong> chaleur sous-jacent<br />

(e.g., Aagaard et al., 1981; Ru<strong>de</strong>ls et al., 1996).<br />

La couche <strong>de</strong> <strong>sur</strong>face, appelée « couche mélangée polaire » (Polar Mixed Layer, PML),<br />

s’étend en 0 et 50 m avec une salinité caractéristique <strong>de</strong> ~31.6 (Aagaard et al., 1981). Elle est<br />

maintenue en gran<strong>de</strong> partie par <strong>les</strong> apports continentaux (e.g., Bauch et al., 1995). Les grands<br />

fleuves <strong>de</strong> Sibérie et <strong>de</strong> l’Amérique du Nord déchargent annuellement un volume d’eau<br />

douce <strong>de</strong> 4000 km 3 dans <strong>les</strong> différentes mers environnantes (Fig. II.1; Barents, Kara, Laptev,<br />

Est Sibérienne, Chukchi, Bering et Beaufort), soit ~11% <strong><strong>de</strong>s</strong> apports continentaux globaux<br />

(Aagaard et Carmack, 1989; Rachold et al., 2004). Ces propriétés varient saisonnièrement<br />

avec la formation et la fonte <strong><strong>de</strong>s</strong> glaces, ainsi qu’avec la crue estivale. En été, la <strong>sur</strong>face est<br />

souvent occupée par une couche <strong>de</strong> 5 à 10 mètres d’eau relativement douce résultant <strong>de</strong> la<br />

fonte <strong>de</strong> la glace <strong>de</strong> mer et <strong><strong>de</strong>s</strong> apports fluviaux (Macdonald et al., 2004a).<br />

32


La circulation <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face dans l’Arctique est relativement bien connue grâce<br />

au suivi <strong><strong>de</strong>s</strong> glaces dérivantes. Les trois composantes principa<strong>les</strong> <strong>de</strong> la circulation sont (Fig.<br />

II.1) :<br />

1) le courant anticyclonique <strong>de</strong> la Gyre <strong>de</strong> Beaufort située dans la partie<br />

ouest <strong>de</strong> l’Arctique;<br />

2) le courant « Transpolaire » (Transpolar Drift) qui s’écoule <strong>de</strong>puis la côte<br />

Sibérienne jusqu’au détroit <strong>de</strong> Fram;<br />

3) <strong>les</strong> courants cycloniques côtiers induits par la force <strong>de</strong> Coriolis.<br />

Le Tourbillon <strong>de</strong> Beaufort et le Courant Transpolaire sont influencés fortement par <strong>les</strong><br />

différentes phases (cyclonique et anticyclonique) <strong>de</strong> la circulation atmosphérique. Lorsque la<br />

circulation anticyclonique prédomine (e.g. dans <strong>les</strong> années 80), le Tourbillon <strong>de</strong> Beaufort<br />

s’élargit et tend à faire recirculer <strong>les</strong> eaux <strong>sur</strong>faces du Courant Transpolaire dans le Bassin<br />

Canadien. Sous ces conditions, <strong>les</strong> glaces s’accumulent dû à la convergence et à la formation<br />

<strong>de</strong> crêtes <strong>de</strong> pression (ice ridges). Lorsque la circulation cyclonique prédomine (e.g. <strong>de</strong> 1989 à<br />

1996), le Tourbillon <strong>de</strong> Beaufort s’affaiblit, <strong>les</strong> glaces sont transportées le long <strong><strong>de</strong>s</strong> côtes <strong>de</strong><br />

l’Alaska et <strong>de</strong> l’Eurasie et sont exportées hors <strong>de</strong> l’Arctique via le Courant Transpolaire (voir<br />

Rigor et al., 2002). Sous le PML, <strong>les</strong> courants sont cycloniques et sont influencés par <strong>les</strong><br />

apports provenant <strong>de</strong> l’Atlantique et du Pacifique (e.g., Pickart, 2004).<br />

Une caractéristique unique <strong>de</strong> l’Arctique par rapport aux autres océans est la<br />

présence d’un couvert <strong>de</strong> glace permanent (multi-year ice) dont l’épaisseur varie <strong>de</strong> 1.8 à 5 m<br />

(e.g., Wadhams et Davis, 2000). Aujourd’hui, la banquise pluriannuelle s’étend <strong>sur</strong> ~5.5 10 6<br />

km 2 (Fig. I.3). La variabilité saisonnière du couvert <strong>de</strong> glace résulte <strong>de</strong> la formation/fonte <strong>de</strong><br />

glace annuelle (first-year ice) dont l’étendue peut atteindre entre 7 et 9 millions <strong>de</strong> km 2 au<br />

début du printemps. Comme la glace pluriannuelle, la majeure partie <strong>de</strong> la glace annuelle est<br />

mobile. Le long <strong><strong>de</strong>s</strong> côtes arctiques, cependant, une banquise continentale (landfast ice),<br />

immobile, se développe progressivement à partir d’octobre ou novembre. En hiver,<br />

néanmoins, une frange libre <strong>de</strong> glace se forme entre la banquise continentale et le couvert<br />

mobile <strong>de</strong> glace présent plus au large. L’étendue <strong>de</strong> ces zones varie selon le mouvement <strong><strong>de</strong>s</strong><br />

glaces provoqué par <strong>les</strong> vents et <strong>les</strong> courants. Ces zones d’échanges directs <strong>de</strong> chaleur entre<br />

l’océan et l’atmosphère se nomment polynies ou leads (Smith et al., 1990).<br />

La glace <strong>de</strong> mer joue un rôle limitant pour: 1) la pénétration <strong>de</strong> la lumière dans la<br />

colonne d’eau; 2) <strong>les</strong> échanges <strong>de</strong> chaleur entre l’atmosphère et l’océan; 3) l’action mécanique<br />

33


<strong><strong>de</strong>s</strong> vents et le mélange vertical. À première vue, la présence <strong>de</strong> glace dans l’Arctique est<br />

défavorable au développement <strong>de</strong> la vie marine. Bien que la présence <strong>de</strong> glace <strong>de</strong> mer limite<br />

la pénétration <strong>de</strong> la lumière dans la majeur partie <strong>de</strong> l’Océan Arctique, la production<br />

biologique y est malgré tout significative (Wheeler et al., 1996). En me<strong>sur</strong>ant <strong>les</strong> taux <strong>de</strong><br />

production primaire et bactérienne, Wheeler et al. (1996) mettaient en évi<strong>de</strong>nce le rôle<br />

prédominant que jouaient <strong>les</strong> microorganismes dans le cycle du <strong>carbone</strong> dans <strong>les</strong> eaux <strong>de</strong><br />

l’Arctique. En particulier, <strong>les</strong> travaux <strong>de</strong> Wheeler et al. (1996) démontrent l’importance <strong><strong>de</strong>s</strong><br />

organismes hétérotrophes dans la reminéralisation relativement rapi<strong>de</strong> <strong><strong>de</strong>s</strong> nutriments en<br />

<strong>sur</strong>face. D’après An<strong>de</strong>rson et al. (2003), >90% <strong>de</strong> la production primaire totale (~15 g C m -2<br />

an -1 ; Gosselin et al., 1997) est minéralisée dans <strong>les</strong> 50 premiers mètres <strong>de</strong> la colonne d’eau. La<br />

production primaire est, cependant, beaucoup plus élevée dans <strong>les</strong> mers situées en marge <strong>de</strong><br />

la banquise centrale (voir Sakshaug, 2004 et réf. citées). Pour <strong><strong>de</strong>s</strong> superficies similaires, la<br />

production primaire <strong>sur</strong> <strong>les</strong> plateaux continentaux Arctiques est environ quatre fois plus<br />

élevée que la production dans <strong>les</strong> bassins profonds (329 vs 50 Tg C an -1 ; Sakshaug, 2004). À<br />

titre d’exemple, dans la polynie <strong><strong>de</strong>s</strong> eaux du Nord (en anglais North Water Polynya), <strong>les</strong> taux <strong>de</strong><br />

production primaire totale peuvent atteindre 377 g C m -2 an -1 (Klein et al., 2002), soit <strong><strong>de</strong>s</strong><br />

valeurs comparab<strong>les</strong> à la plupart <strong><strong>de</strong>s</strong> zones côtières <strong><strong>de</strong>s</strong> basses latitu<strong><strong>de</strong>s</strong> (e.g., Antoine et al.,<br />

1996). Ces observations indiquent que, régionalement, la production biologique varie <strong>de</strong> plus<br />

d’un ordre <strong>de</strong> gran<strong>de</strong>ur dans l’Arctique.<br />

34


II.3. Généralités <strong>sur</strong> la Mer <strong>de</strong> Beaufort et le Plateau du Mackenzie<br />

La Figure II.2 montre <strong>les</strong> principaux courants rencontrés dans la Mer <strong>de</strong> Beaufort.<br />

En <strong>sur</strong>face, au large, <strong>les</strong> eaux se déplacent d’est en ouest dans le mouvement anticyclonique<br />

généré par le Tourbillon <strong>de</strong> Beaufort. L’intensité <strong>de</strong> ce courrant dépend fortement <strong>de</strong> la<br />

direction générale <strong><strong>de</strong>s</strong> vents <strong>de</strong> <strong>sur</strong>face (section précé<strong>de</strong>nte; Rigor et al., 2002). Près <strong><strong>de</strong>s</strong> côtes,<br />

<strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face peuvent circuler soit vers l’Est par temps calme sous l’influence <strong>de</strong> la<br />

force <strong>de</strong> Coriolis, soit vers l’Ouest quand <strong>les</strong> vents dominants proviennent <strong>de</strong> l’Est ou du<br />

Sud-Est (Carmack et Macdonald, 2002). À une centaine <strong>de</strong> mètres <strong>de</strong> profon<strong>de</strong>ur, à la limite<br />

du Plateau continental <strong>de</strong> l’Alaska, un fort courant cyclonique <strong>de</strong> sub-<strong>sur</strong>face transporte <strong><strong>de</strong>s</strong><br />

eaux provenant du Pacifique et <strong>de</strong> la Mer <strong>de</strong> Chukchi (Pickart, 2004).<br />

Le sud-est <strong>de</strong> la Mer <strong>de</strong> Beaufort est caractérisé par la présence d’un large plateau<br />

continental occupant plus <strong>de</strong> 60 000 km 2 , bordé à l’est par le Golfe d’Amundsen, à l’ouest<br />

par le canyon du Mackenzie, au sud par le <strong>de</strong>lta du fleuve Mackenzie, et par la Mer <strong>de</strong><br />

Beaufort au nord (Fig. II.2). D’après Macdonald (2000), le Plateau du Mackenzie est le plus<br />

estuarien <strong><strong>de</strong>s</strong> plateaux Arctiques où le temps <strong>de</strong> rési<strong>de</strong>nce <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face est inférieur à<br />

une année. La circulation y est très variable, souvent reliée à l’importante saisonnalité du<br />

débit du fleuve Mackenzie. Ce <strong>de</strong>rnier décharge ~330 km 3 d’eau douce chaque année<br />

(Macdonald, 2000), soit le quatrième plus important à l’échelle <strong>de</strong> l’Arctique et le quinzième à<br />

l’échelle global (Mckee et al., 2004).<br />

35


Figure II.2. Carte <strong>de</strong> localisation <strong>de</strong> la Mer <strong>de</strong> Beaufort avec la direction <strong><strong>de</strong>s</strong><br />

principaux courants : la Gyre <strong>de</strong> Beaufort (en gris); le sous-courant <strong>de</strong><br />

Beaufort (c.f., Beaufort Un<strong>de</strong>rcurrent; en bleu); la dispersion attendue du panache<br />

du Mackenzie (en rouge) qui varie selon la direction et l’intensité <strong><strong>de</strong>s</strong> vents<br />

dominants. L’encadré en gris montre <strong>les</strong> limites <strong>de</strong> l’échantillonnage effectué<br />

durant le programme CASES.<br />

36


II.3.1. Cyc<strong>les</strong> Saisonniers <strong>sur</strong> le Plateau du Mackenzie<br />

Inspirée du modèle conceptuel présenté par Macdonald et al. (1995), Macdonald<br />

(2000) et Carmack et Macdonald (2002), cette section présente un cycle typique <strong>de</strong> la<br />

dispersion <strong>de</strong> l’eau douce et la formation/fonte <strong>de</strong> la glace <strong>de</strong> mer <strong>sur</strong> le Plateau du<br />

Mackenzie (Fig. II.3).<br />

Figure II.3. Cycle saisonnier <strong>de</strong> l’eau douce <strong>sur</strong> le Plateau du Mackenzie<br />

caractérisé par <strong>les</strong> apports continentaux et la glace <strong>de</strong> mer : A) englacement<br />

(apports continentaux faib<strong>les</strong>, la glace commence à se former); B) fin <strong>de</strong><br />

l’hiver (apports continentaux faib<strong>les</strong>, la glace a atteint son épaisseur maximale<br />

et a arrêté <strong>de</strong> croître); C) débâcle printanière (apports continentaux élevés, la<br />

glace <strong>de</strong>meure intacte et reste près <strong>de</strong> la côte); D) été (apports continentaux<br />

élevés, la glace a fondue ou a été exportée vers l’intérieur <strong>de</strong> l’Océan).<br />

(Modifiée <strong>de</strong> Macdonald, 2000).<br />

Avant la formation <strong><strong>de</strong>s</strong> premières glace à l’automne, i.e. en septembre, l’eau douce<br />

provenant <strong>de</strong> la fonte <strong><strong>de</strong>s</strong> glace et du fleuve est mélangée dans <strong>les</strong> 15-20 premiers mètres <strong>de</strong><br />

la colonne d’eau sous l’action <strong><strong>de</strong>s</strong> vents (Carmack et al., 1989). L’intensité du mélange<br />

influence la salinité <strong><strong>de</strong>s</strong> eaux <strong>de</strong> <strong>sur</strong>face, et par conséquent, la formation d’eau profon<strong>de</strong><br />

(Macdonald et al., 1995). Par exemple, si le mélange est intense, <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face seront<br />

plus salées et, par conséquent, la <strong>de</strong>nsité <strong><strong>de</strong>s</strong> eaux saumâtres produites lors <strong>de</strong> la formation<br />

<strong>de</strong> la glace sera plus élevée (formation d’eau profon<strong>de</strong> plus efficace).<br />

37


Figure II.4. Variabilité journalière entre 1976 et 1998 (A) du débit du fleuve<br />

Mackenzie, (B) du <strong>flux</strong> <strong>de</strong> particu<strong>les</strong> en suspension d’origine terrigène, et (C)<br />

<strong>de</strong> l’éclairement photosynthétique inci<strong>de</strong>nt à 71°N et la température <strong>de</strong> l’air<br />

me<strong>sur</strong>és à Tuktoyaktuk situé à l’embouchure du fleuve. (Modifiée <strong>de</strong> O’Brien<br />

et al., 2006).<br />

38


À partir du mois d’octobre, la glace se forme rapi<strong>de</strong>ment à me<strong>sur</strong>e que <strong>les</strong><br />

températures diminuent (Fig. II.4C). Pendant ce temps, le débit du fleuve diminue et ses<br />

eaux se dispersent sous la glace. Au large, la divergence <strong><strong>de</strong>s</strong> glaces crée <strong><strong>de</strong>s</strong> zones d’eau<br />

ouverte. À la fin <strong>de</strong> l’hiver, la banquise continentale fait environ 2 m d’épaisseur et s’étend<br />

jusqu’à l’isobathe d’environ 20 m (Fig. II.3B). La marge externe <strong>de</strong> la banquise continentale<br />

est soumise aux mouvements <strong>de</strong> la banquise Arctique qui se déplace sous l’effet <strong><strong>de</strong>s</strong> vents.<br />

Ces mouvements favorisent l’accumulation <strong>de</strong> glace <strong>de</strong> mer pouvant atteindre plusieurs<br />

mètres d’épaisseur (>20m). La zone d’accumulation, communément appelé stamukhi, forme<br />

un barrage naturel <strong>de</strong> glace retenant près <strong>de</strong> la côte <strong>les</strong> eaux du Fleuve Mackenzie. Au début<br />

du printemps, <strong>les</strong> eaux accumulées forment un « lac » saisonnier contenant jusqu’à 70 km 3<br />

d’eau douce (le lac Herlinveaux). Au-<strong>de</strong>là du stamukhi, on retrouve un corridor d’eau ouverte<br />

entre <strong>les</strong> banquises continentale et Arctique qui varie selon <strong>les</strong> conditions <strong>de</strong> vents et qui fait<br />

quelques dizaines <strong>de</strong> km <strong>de</strong> large (flaw lead polynya). Cette zone, libre <strong>de</strong> glace et en contact<br />

direct avec l’atmosphère, est le lieu d’importants échanges <strong>de</strong> chaleur favorisant la<br />

convection profon<strong>de</strong> <strong><strong>de</strong>s</strong> eaux situées à la marge continentale. La production primaire y est<br />

faible au cours <strong>de</strong> cette pério<strong>de</strong> car limitée par la lumière (cycle solaire, glace, mélange<br />

vertical intensif).<br />

La débâcle printanière commence au sud par la crue du fleuve Mackenzie et<br />

progresse graduellement vers le nord (Fig. II.3C). En juin, le débit du fleuve est maximal (Fig.<br />

II.4A) et inon<strong>de</strong> la zone côtière jusqu’à ce que le stamukhi cè<strong>de</strong> sous la pression générée par<br />

la masse d’eau douce se trouvant en amont. Cette poussée d’eau relativement chau<strong>de</strong> et très<br />

turbi<strong>de</strong> intervient souvent avant même que la glace n’ait commencé à fondre sous l’action du<br />

rayonnement solaire. Rapi<strong>de</strong>ment, le panache s’étend <strong>sur</strong> plusieurs milliers <strong>de</strong> kilomètres<br />

carrés au-<strong><strong>de</strong>s</strong>sus du plateau continental. Une image acquise par le capteur SeaWiFS au<br />

moment <strong>de</strong> la débâcle illustre bien la dispersion <strong><strong>de</strong>s</strong> eaux du Fleuve Mackenzie au moment<br />

<strong>de</strong> la débâcle (Fig. II.5). La couleur brunâtre <strong><strong>de</strong>s</strong> eaux du Mackenzie indique la présence<br />

d’une forte concentration <strong>de</strong> particu<strong>les</strong> en suspension à cette époque <strong>de</strong> l’année. L’étendue<br />

du panache arrive même au-<strong>de</strong>là du plateau continental, sous la banquise Arctique.<br />

39


Figure II.5. Image SeaWiFS acquise le 15 Juin 2004 montrant le début <strong>de</strong> la<br />

débâcle. Rapi<strong>de</strong>ment <strong>les</strong> eaux turbi<strong><strong>de</strong>s</strong> du Mackenzie occupent une large<br />

partie <strong><strong>de</strong>s</strong> eaux ouvertes du plateau continental (~15 000-20 000 km²).<br />

En été, la glace <strong>de</strong> mer font graduellement avec la chaleur apportée par <strong>les</strong> eaux du<br />

Fleuve Mackenzie (~15°C) et l’énergie solaire absorbée en <strong>sur</strong>face. En règle générale, <strong>les</strong><br />

eaux sont fortement stratifiées, dû à l’importance <strong>de</strong> l’eau douce présente (Macdonald et al.,<br />

1989; Carmack et al., 1989). La Figure II.6 montre l’étendue <strong>de</strong> la stratification <strong><strong>de</strong>s</strong> 40<br />

premiers mètres <strong>de</strong> la colonne d’eau au-<strong><strong>de</strong>s</strong>sus du plateau continental. Le déplacement du<br />

panache turbi<strong>de</strong> du Fleuve Mackenzie dépend principalement <strong><strong>de</strong>s</strong> conditions <strong>de</strong> vent (Fig.<br />

II.2). Quand <strong>les</strong> vents sont calmes, la force <strong>de</strong> Coriolis tend à maintenir <strong>les</strong> flots d’eau douce<br />

le long <strong>de</strong> la côte en direction est, avant d’éventuellement d’entrer dans le Golfe<br />

d’Amundsen (Macdonald et al., 1987; Carmack et Macdonald, 2002). Par vent d’est,<br />

cependant, <strong>les</strong> eaux du Mackenzie sont transportées au large dans le Bassin Canadien (e.g.<br />

Macdonald et al. 1999, 2002). La direction que prennent <strong>les</strong> eaux du Fleuve Mackenzie est<br />

importante car elle influence le temps <strong>de</strong> rési<strong>de</strong>nce <strong><strong>de</strong>s</strong> eaux douces dans l’Arctique. Il<br />

semble que <strong>les</strong> <strong>changements</strong> dans la circulation atmosphérique (dominante cyclonique),<br />

observés <strong>de</strong>puis la fin <strong><strong>de</strong>s</strong> années 80, aient favorisé l’accumulation <strong><strong>de</strong>s</strong> eaux douces dans le<br />

Bassin Canadien (voir <strong>les</strong> discussions <strong>de</strong> Macdonald et al., 2002 et <strong>de</strong> Proshutinsky et al., 2002<br />

<strong>sur</strong> <strong>les</strong> conséquences potentiel<strong>les</strong> <strong>de</strong> ce changement).<br />

40


Figure II.6. Coupe transversale effectuée à partir <strong>de</strong> l’embouchure du fleuve<br />

Mackenzie (Kugmalit Bay) jusqu’à la limite du plateau continental en juillet<br />

2004 dans le cadre du programme CASES : a) <strong>de</strong>nsité; b) salinité.<br />

II.3.2. Bilan <strong>de</strong> Carbone Organique <strong>sur</strong> le Plateau du Mackenzie<br />

Un bilan <strong>de</strong> <strong>carbone</strong> organique pour le Plateau du Mackenzie a été proposé par<br />

Macdonald et al. (1998, 2004b). Ce budget ne concerne que la fraction particulaire du<br />

<strong>carbone</strong>. Selon ce budget, le Plateau du Mackenzie est une zone majeure d'enfouissement <strong>de</strong><br />

<strong>carbone</strong> organique à l'échelle <strong>de</strong> l'Arctique. Bien que la production primaire génère plus <strong>de</strong><br />

<strong>carbone</strong> organique particulaire que l'apport du Mackenzie, le <strong>carbone</strong> enfoui est presque<br />

entièrement d'origine terrigène. En effet, le Mackenzie exporte annuellement 2.1 x 10 12 g <strong>de</strong><br />

POC d’origine terrigène, soit l’équivalant d’environ <strong>de</strong>ux tiers <strong>de</strong> la production primaire<br />

marine (3.3 x 10 12 g <strong>de</strong> C an -1 ; Macdonald et al., 1998). À l’échelle régionale, si 65% du<br />

<strong>carbone</strong> terrigène est préservé dans <strong>les</strong> sédiments du <strong>de</strong>lta (40%) et du plateau continental<br />

(25%), 97% <strong>de</strong> la production primaire est minéralisé (Macdonald et al. 1998, 2004b).<br />

Cependant, on connaît encore peu <strong>de</strong> chose <strong>sur</strong> <strong>les</strong> processus exacts <strong>de</strong> minéralisation <strong>de</strong> la<br />

41


matière organique (activité hétérotrophique dans la colonne d’eau et <strong>les</strong> sédiments), et/ou le<br />

transfert vers <strong>les</strong> échelons supérieurs <strong>de</strong> la chaîne alimentaire (e.g. zooplankton, etc). En<br />

faisant le bilan, le Plateau du Mackenzie représente une source <strong>de</strong> <strong>carbone</strong> inorganique (i.e.<br />

CO 2) <strong>de</strong> l’ordre 0.64 x 10 12 g <strong>de</strong> C an -1 (l’écosystème est dit « hétérotrophe net »).<br />

Il est intéressant <strong>de</strong> noter que le <strong>carbone</strong> particulaire terrigène transporté par le fleuve<br />

Mackenzie est en majeure partie « ancien » (>7000 ans) (Goñi et al. 2005). Il semble donc<br />

qu’il proviendrait essentiellement du pergélisol (Goñi et al. 2005). Guo et Macdonald (2006)<br />

suggéraient récemment que la dégradation du pergélisol pourrait entraîner la mobilisation <strong>de</strong><br />

<strong>carbone</strong> « ancien », principalement sous forme particulaire. De plus, la modification <strong>de</strong> la<br />

toundra arctique vers un écosystème <strong>de</strong> type « forêt boréale » sous un climat plus chaud<br />

(Hinzman et al., 2005), risque d’augmenter <strong>les</strong> apports en <strong>carbone</strong> « mo<strong>de</strong>rne ». Se <strong>de</strong>rnier est<br />

facilement dégradable en milieu côtier (Goñi et al. 2005). Par conséquent, le réchauffement<br />

global risque <strong>de</strong> modifier sensiblement le cycle du <strong>carbone</strong> particulaire <strong>sur</strong> le Plateau du<br />

Mackenzie en modifiant la quantité et la qualité <strong><strong>de</strong>s</strong> apports allochtones.<br />

Contrairement au <strong>carbone</strong> particulaire, le cycle du <strong>carbone</strong> dissous est pratiquement<br />

inconnu <strong>sur</strong> le Plateau du Mackenzie. D’après Telang et al. (1991), le fleuve Mackenzie<br />

transporte 1.3 x 10 12 g <strong>de</strong> C organique dissous, soit le quatrième plus important à l’échelle <strong>de</strong><br />

l’Arctique. Le <strong><strong>de</strong>s</strong>tin <strong>de</strong> ce <strong>carbone</strong> terrigène reste pratiquement inconnu, mais si l’on se fit<br />

au mélange conservatif du DOC observé ailleurs en Arctique (Cauwet et Sidorov 1996 ;<br />

Köhler et al., 2003), la majeur partie serait exportée vers l’océan Arctique ou l’Archipel<br />

Canadien. D’après Gosselin et al. (1997), la production <strong>de</strong> DOC par <strong>les</strong> algues <strong>de</strong> glace<br />

Arctique représente entre 20 et 65 % <strong>de</strong> la production primaire totale. La production<br />

primaire marine est donc certainement une source significative <strong>de</strong> DOC <strong>sur</strong> le Plateau du<br />

Mackenzie. Cependant, il semble que ce <strong>carbone</strong> soit rapi<strong>de</strong>ment utilisé par <strong>les</strong> organismes<br />

hétérotrophes (Davis et Benner, 2005).<br />

42


Chapitre III: Photominéralisation <strong>de</strong> la matière<br />

organique dissoute d’origine terrigène dans <strong>les</strong><br />

eaux côtières arctiques entre 1979 et 2003:<br />

Variabilité interannuelle et implications <strong><strong>de</strong>s</strong><br />

<strong>changements</strong> <strong>climatiques</strong><br />

43


III.A Résumé<br />

La minéralisation photochimique (photominéralisation) <strong>de</strong> la matière organique dissoute<br />

d’origine terrigène (tDOM) dans l’Océan Arctique est limitée par la présence d’un couvert <strong>de</strong><br />

glace permanant qui réduit la quantité d’énergie ultraviolette (UV) atteignant la colonne d’eau<br />

sous-jacente. En réponse à la diminution <strong>de</strong> l’étendue et <strong>de</strong> l’épaisseur du couvert <strong>de</strong> glace<br />

causée par le réchauffement global, <strong>les</strong> processus biogéochimiques qui dépen<strong>de</strong>nt <strong>de</strong><br />

l’énergie UV <strong>de</strong>vraient s’amplifier dans cette région. Dans cette étu<strong>de</strong>, nous avons réalisé<br />

pour la première fois <strong><strong>de</strong>s</strong> estimations quantitatives <strong>de</strong> la photominéralisation <strong>de</strong> la tDOM<br />

dans un écosystème côtier Arctique sous <strong><strong>de</strong>s</strong> conditions <strong>de</strong> glace actuel<strong>les</strong> et futures. Nous<br />

avons employé un modèle couplé optique-photochimique utilisant en entrée <strong>les</strong> propriétés<br />

optiques inhérentes <strong>de</strong> la colonne d’eau et <strong><strong>de</strong>s</strong> me<strong>sur</strong>es <strong>de</strong> production photochimique<br />

(photoproduction) <strong>de</strong> <strong>carbone</strong> inorganique dissous (DIC), le principal produit <strong>de</strong> la<br />

photooxidation <strong>de</strong> la DOM. Les ren<strong>de</strong>ments quantiques apparents <strong>de</strong> photoproduction <strong>de</strong><br />

DIC ont été déterminés <strong>sur</strong> <strong><strong>de</strong>s</strong> échantillons provenant <strong>de</strong> l’estuaire du Fleuve Mackenzie, du<br />

Plateau du Mackenzie, et du Golfe d’Amundsen. Grâce à un modèle <strong>de</strong> transfert radiatif<br />

atmosphérique, l’éclairement UV journalier inci<strong>de</strong>nt juste sous l’interface air-mer a été estimé<br />

en combinant <strong><strong>de</strong>s</strong> me<strong>sur</strong>es 1) <strong>de</strong> luminance rétrodiffusée par l’atmosphère et la <strong>sur</strong>face et 2)<br />

<strong>de</strong> micro-on<strong><strong>de</strong>s</strong> passives émises par la <strong>sur</strong>face. La photoproduction annuelle <strong>de</strong> DIC dans <strong>les</strong><br />

eaux <strong>de</strong> <strong>sur</strong>face du sud-est <strong>de</strong> la Mer <strong>de</strong> Beaufort où l’absorption <strong><strong>de</strong>s</strong> UV est dominée par la<br />

tDOM colorée déchargée par le Fleuve Mackenzie, est, en moyenne entre 1979 et 2003,<br />

estimé à 66.5 ± 18.5 Gg <strong>de</strong> <strong>carbone</strong>. Cette valeur est équivalente à environ 10% du taux <strong>de</strong> le<br />

respiration bactérienne dans <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face, 8% du taux <strong>de</strong> production primaire, et 2.8 ±<br />

0.6% <strong><strong>de</strong>s</strong> 1.3 Tg <strong>de</strong> <strong>carbone</strong> organique dissous (DOC) déchargé annuellement par le Fleuve<br />

Mackenzie dans cette région. Durant <strong>les</strong> pério<strong><strong>de</strong>s</strong> où le couvert <strong>de</strong> glace est fortement réduit<br />

(comme en 1998), cette <strong>de</strong>rnière estimation peut atteindre 5.1% <strong>de</strong> la masse du DOC<br />

d’origine terrigène déchargé annuellement par le Fleuve. En simulant un été où la région<br />

serait pratiquement libre <strong>de</strong> glace durant la pério<strong>de</strong> estivale, le modèle prédit une<br />

photoproduction annuelle <strong>de</strong> 150.5 Gg <strong>de</strong> DIC, minéralisant 6.2% <strong><strong>de</strong>s</strong> apports annuels en<br />

DOC par le Fleuve Mackenzie. Ces résultats indiquent que si <strong>les</strong> prédictions <strong><strong>de</strong>s</strong> modè<strong>les</strong><br />

<strong>climatiques</strong> concernant la réduction du couvert <strong>de</strong> la Arctique sont correctes, la<br />

photominéralisation <strong>de</strong> la tDOM dans <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face Arctiques s’accélérera gran<strong>de</strong>ment<br />

dans le futur.<br />

44


III.B Article publié dans la revue Global Biogeochemical Cyc<strong>les</strong> :<br />

“Photomineralization of terrigenous dissolved organic matter in Arctic coastal<br />

waters from 1979 to 2003: Interannual variability and implications of climate<br />

change”<br />

Simon Bélanger 1 , Huixiang Xie 2 , Nickolay Krotkov 3 , Pierre Larouche 4 , Warwick F.<br />

Vincent 5 and Marcel Babin 1<br />

1 Université Pierre et Marie Curie-Paris 6, Laboratoire d'océanographie <strong>de</strong> Villefranche,<br />

06230 Villefranche-<strong>sur</strong>-Mer, France ; CNRS, Laboratoire d'océanographie <strong>de</strong><br />

Villefranche, 06230 Villefranche-<strong>sur</strong>-Mer, France<br />

2 Institut <strong><strong>de</strong>s</strong> Sciences <strong>de</strong> la Mer <strong>de</strong> Rimouski, Université du Québec à Rimouski,<br />

Rimouski, Canada<br />

3 Goddard Earth Sciences and Technology Center, University of Maryland Baltimore<br />

County, USA<br />

4 Institut Maurice-Lamontagne, Pêches et Océans Canada, Mont-Joli, QC, Canada G5H<br />

3Z4<br />

5 Département <strong>de</strong> Biologie & Centre d’Étu<strong><strong>de</strong>s</strong> Nordiques, Université Laval, Québec City,<br />

Canada<br />

45


Acknowledgments<br />

This study was ma<strong>de</strong> possible with financial support from the Natural Sciences and Engineering<br />

Research Council of Canada (NSERC), the Fonds Québécois pour la Recherche <strong>sur</strong> la Nature et<br />

<strong>les</strong> Technologies (FQRNT) (to HX), and the Fonds France Canada pour la Recherche (FFCR to<br />

MB and WFV) and Indian and Northern Affairs Canada (to SB). SB receives a doctoral<br />

fellowship from the FQRNT. We thank Dr S.H. Lee for providing the primary production rates<br />

mea<strong>sur</strong>ed in the Gulf of Amundsen. We are grateful to Dr. C. Osburn and Dr. M. Tzortziou for<br />

their constructive comments on the manuscript. We appreciate the efforts of S. Cizmeli in CASES<br />

field mea<strong>sur</strong>ements, T. Lou in DIC mea<strong>sur</strong>ement and J. Caveen in Matlab programming. We<br />

thank the crews of the CCGS Amundsen cruises for their enthusiastic help onboard the ship. This<br />

is a contribution to the Canadian Arctic Shelf Exchange Study (CASES) un<strong>de</strong>r the overall<br />

direction of L. Fortier. S. C. Johannessen and one anonymous referee are acknowledged for<br />

constructive comments on the manuscript.<br />

46


Abstract<br />

Photomineralization of terrigenous dissolved organic matter (tDOM) in the Arctic Ocean<br />

is limited by persistent sea ice cover that reduces the amount of ultraviolet (UV) radiation<br />

reaching the un<strong>de</strong>rlying water column. UV-<strong>de</strong>pen<strong>de</strong>nt processes are likely to accelerate<br />

as a result of shrinking sea ice extent and <strong>de</strong>creasing ice thickness caused by climatic<br />

warming over this region. In this study, we ma<strong>de</strong> the first quantitative estimates of<br />

photomineralization of tDOM in a coastal Arctic ecosystem un<strong>de</strong>r current and future sea<br />

ice regimes. We used an optical-photochemical coupled mo<strong>de</strong>l incorporating water<br />

column optics and experimental mea<strong>sur</strong>ements of photoproduction of dissolved inorganic<br />

carbon (DIC), the main carbon product of DOM photochemistry. Apparent quantum<br />

yields of DIC photoproduction were <strong>de</strong>termined on water samp<strong>les</strong> from the Mackenzie<br />

River estuary, the Mackenzie Shelf, and Amundsen Gulf. UV irradiances just below the<br />

sea <strong>sur</strong>face were estimated by combining satellite backscattered and passive microwave<br />

radiance mea<strong>sur</strong>ements with a radiative transfer mo<strong>de</strong>l. The mean annual DIC<br />

photoproduction between 1979 and 2003 was estimated as 66.5 ± 18.5 Gg carbon in the<br />

<strong>sur</strong>face waters of the southeastern Beaufort Sea, where UV absorption is dominated by<br />

chromophoric dissolved organic matter discharged by the Mackenzie River. This value is<br />

equivalent to 10% of bacterial respiration rates, 8% of primary production rates and 2.8 ±<br />

0.6% of the 1.3 Tg of dissolved organic carbon (DOC) discharged annually by the<br />

Mackenzie River into the area. During periods of reduced ice cover such as 1998, the<br />

latter value could rise to 5.1% of the annual riverine DOC discharge. Un<strong>de</strong>r an ice-free<br />

scenario, the mo<strong>de</strong>l predicted that 150.5 Gg of DIC would be photochemically produced,<br />

mineralizing 6.2% of the DOC input from the Mackenzie River. These results show that<br />

the predicted trend of ongoing contraction of sea ice cover will greatly accelerate the<br />

photomineralization of tDOM in Arctic <strong>sur</strong>face waters.<br />

47


III.1. Introduction<br />

Terrigenous dissolved organic matter (tDOM) is an abundant component of Arctic<br />

polar <strong>sur</strong>face waters [see Benner et al., 2005 and references therein] as a result of the<br />

large amounts of riverine tDOM that enter the Arctic Ocean relative to its volume<br />

[Aagaard and Carmack, 1989]. The refractory character of tDOM, as observed in several<br />

Arctic marginal seas [e.g., Dittmar and Kattner, 2003; Meon and Amon, 2004], may also<br />

contribute to the high concentrations of Arctic Ocean tDOM. Recent studies suggest that<br />

slightly more than 50% of the annual input of terrigenous dissolved organic carbon<br />

(tDOC) to the Arctic Ocean is remineralized before its transport to the North Atlantic<br />

[Hansell et al., 2004]. However, quantitative information on major tDOM loss processes<br />

in the Arctic is still lacking.<br />

There is growing evi<strong>de</strong>nce indicating a major role of photooxidation in the<br />

remineralization of tDOM in the coastal oceanic environment [e.g., Benner and Opsahl,<br />

2001; Hernes and Benner, 2003; Kieber et al., 1990]. Other studies suggested that a large<br />

fraction of the marine plankton-<strong>de</strong>rived dissolved organic carbon (DOC) pool also can be<br />

photochemically mineralized in the open ocean [Johannessen, 2000; Johannessen and<br />

Miller, 2001]. Although there is molecular evi<strong>de</strong>nce that photochemical transformation of<br />

tDOM occurs in the Arctic Ocean [Benner et al., 2005], this mechanism is not consi<strong>de</strong>red<br />

of importance because of the prevailing low solar angle, cloudy conditions and<br />

permanent sea ice cover [Amon and Meon, 2004; Benner et al., 2004]. Sea ice and its<br />

overlying snow cover is known to be a strong attenuator of ultraviolet (UV) radiation in<br />

both polar regions [Vincent and Belzile, 2003].<br />

Recent observations in the Arctic reveal that sea ice conditions are changing, with<br />

a 20% <strong>de</strong>crease in the summer areal extent of sea ice over the last three <strong>de</strong>ca<strong><strong>de</strong>s</strong> [e.g.,<br />

Cavalieri et al., 2003]. In addition, the thinning of sea ice has also been observed in<br />

several parts of the Arctic Ocean [e.g., Wadhams and Davis, 2000], permitting more UV<br />

radiation to reach the water column. Moreover, a complete loss of summer ice is<br />

predicted for later this century [ACIA, 2005]. In parallel, a <strong>de</strong>crease in stratospheric ozone<br />

concentration (O3), which results in an increase of UVB (280-320 nm) radiation at sea<br />

<strong>sur</strong>face, was recently reported over the high northern latitu<strong><strong>de</strong>s</strong> [Fioletov et al., 2004; Rex<br />

48


et al., 2004]. These observations suggest that the amount of photochemically active UV<br />

radiation (280-400 nm) penetrating into the Arctic waters could have significantly<br />

increased in the last <strong>de</strong>ca<strong><strong>de</strong>s</strong>, and that this effect may accelerate in the coming years.<br />

In addition to alteration in sea ice dynamics, a number of other recent<br />

observations suggest that the Arctic is currently experiencing rapid and extensive<br />

environmental changes in response to global climate warming [Overland et al., 2004;<br />

Serreze et al., 2000]. Arctic rivers account for about 10% of global terrestrial organic<br />

carbon export to the ocean [Rachold et al., 2004]. This percentage is likely to increase<br />

markedly during the next century [Frey and Smith, 2005] as a result of increasing river<br />

discharge [Peterson et al., 2002] and thawing of the permafrost [Camill, 2005]. The latter<br />

stores one third of the global soil carbon and contains 7 to 26% of all terrestrial organic<br />

carbon stored since the last glacial maximum [Smith et al., 2004]. To assess and predict<br />

the impact of climate change on the marine carbon cycle, it is critical to <strong>de</strong>termine the<br />

transformation and fate of the exported terrestrial carbon, including photochemical<br />

production of dissolved inorganic carbon (DIC), which is the main carbon product of<br />

chromophoric DOM (CDOM) photochemistry [Miller and Zepp, 1995; Mopper and<br />

Kieber, 2002].<br />

The objectives of this study are to assess the present significance of<br />

photooxidation of CDOM in the tDOM cycling in ice-free Arctic seawaters, and to assess<br />

the potential impact of the reduction of sea ice extent on the removal of tDOM through<br />

this process. Our field study was conducted in the Canadian Shelf of the Beaufort Sea<br />

where a large amount of tDOM is discharged by the Mackenzie River (Figure III.1). In<br />

terms of dissolved organic carbon export, the Mackenzie River is the fourth largest<br />

among the Arctic rivers, discharging ~1.3 Tg of tDOC per year [Telang et al., 1991]. This<br />

large tDOC <strong>flux</strong> explains the high concentration of terrestrial CDOM found in the entire<br />

Western Arctic Ocean [Guay et al., 1999; Guéguen et al., 2005]. As this area has<br />

experienced a significant <strong>de</strong>crease in sea ice areal extent between 1979 and 2000 [Barber<br />

and Hanesiak, 2004], photooxidation may have become a significant component of the<br />

tDOM cycling in this Arctic coastal ecosystem. In this study, experimentally <strong>de</strong>termined<br />

wavelength-specific apparent quantum yields (AQY) of DIC production, along with<br />

remote sensing-<strong>de</strong>rived spectral irradiances (300 to 600 nm) and <strong>sur</strong>face areas of open<br />

49


waters, were used to mo<strong>de</strong>l the <strong>de</strong>pth-integrated photochemical production rates of DIC<br />

in the ice-free waters of the southeastern Beaufort Sea . To assess recent trends in the<br />

DIC production, we calculated the annual photochemical production of DIC for the<br />

period 1979 to 2003.<br />

Figure III.1. Location of CASES stations sampled for absorption<br />

mea<strong>sur</strong>ements in 2004 over the Mackenzie Shelf (open circ<strong>les</strong>), in<br />

Amundsen Gulf (open triang<strong>les</strong>), and in Canada Basin (open squares).<br />

ARDEX and CASES samp<strong>les</strong> collected for φDIC <strong>de</strong>termination are shown<br />

as closed squares.<br />

50


III.2. Materials and methods<br />

III.2.1. Sample collection and storage<br />

Sampling was conducted in June and July 2004 onboard the CCGS Amundsen in<br />

the Mackenzie Shelf, the Gulf of Amundsen, and the Arctic Canada basin (Fig. III.1) as<br />

part as the Canadian Arctic Shelf Exchange Study (CASES) program, and onboard the<br />

CCGS Nahidik in the Mackenzie River estuary in late July during the Arctic River-Delta<br />

Experiment (ARDEX). Surface water samp<strong>les</strong> were collected at 79 locations during<br />

CASES (Fig. III.1) for mea<strong>sur</strong>ement of the absorption coefficients of CDOM (aCDOM) and<br />

suspen<strong>de</strong>d partic<strong>les</strong> (ap). The samp<strong>les</strong> were filtered through 0.2-μm Anotop® syringe<br />

filters (Whatman) and collected into 100-mL acid-cleaned amber glass bott<strong>les</strong>.<br />

Suspen<strong>de</strong>d partic<strong>les</strong> were retained by 25-mm GF/F glass fiber filters (Whatman) by<br />

filtering 0.1 to 3.5 L of seawater. The glass bott<strong>les</strong> and GF/F filters were stored frozen<br />

(seawater: -20 °C; particle: -80 °C) in the dark until being analyzed ca. four months later<br />

in the land-based laboratory at Rimouski (aCDOM) and Villefranche-<strong>sur</strong>-mer (ap).<br />

Six stations were sampled for <strong>de</strong>termination of the apparent quantum yield (AQY)<br />

spectra of DIC photoproduction: two in the Mackenzie River estuary, one on the<br />

Mackenzie Shelf (ARDEX cruise), and three in the Amundsen Gulf (CASES cruise) (Fig.<br />

III.1; Table III.1). Surface water (∼5 m <strong>de</strong>ep) in the Amundsen Gulf was taken with<br />

Teflon-coated Niskin bott<strong>les</strong> and gravity-filtered through a Pall AcroPak 1000 Capsule<br />

sequentially containing 0.8-μm and 0.2-μm polyethersulfone membrane filters. During<br />

the ARDEX cruise, <strong>sur</strong>face water was collected using an acid-cleaned plastic bucket and<br />

transferred to acid-cleaned carboys. Both the bucket and carboys were profusely rinsed<br />

three times with sample water. The <strong>sur</strong>face water was then filtered through sequential 3-<br />

μm and 0.2-μm pore-size polysulfone filters (PALL corporation). All filtered samp<strong>les</strong><br />

were stored cold (4 °C) in the dark in acid-cleaned 4-L clear glass bott<strong>les</strong> and shipped to<br />

the laboratory at Rimouski.<br />

51


Table III.1. Chemical and physical properties of samp<strong>les</strong> used for the φDIC.<br />

Station- Date of Lat Lon Depth S T<br />

Mission sampling (°N) (°W) (m)<br />

(°C)<br />

R5d-ARDEX 31-07 69.28 133.97 0 1.57 16.1<br />

R5a-ARDEX 31-07 69.42 133.5 0 8.19 12.2<br />

R9-ARDEX 28-07 70.05 133.42 0 25.81 9.23<br />

108-CASES 07-06 70.63 123.22 10 30.09 -1.08<br />

406-CASES 16-06 71.30 127.75 5 30.77 0.51<br />

409-CASES 23-07 71.46 127.3 5 28.79 3.46<br />

III.2.2. Optical mea<strong>sur</strong>ements<br />

The frozen seawater samp<strong>les</strong> were thawed and warmed to room temperature.<br />

Their optical <strong>de</strong>nsities (OD), referenced to pure water, were mea<strong>sur</strong>ed in a 10-cm quartz<br />

cuvette between 250 and 800 nm with 1-nm increments using a Perkin-Elmer Lambda 35<br />

dual beam spectrophotometer. A background correction was applied by subtracting the<br />

absorbance value averaged over an interval of 5 nm around 685 nm from all the spectral<br />

values [Babin et al., 2003]. The spectral absorption coefficient of CDOM, aCDOM(λ) (m -1 ),<br />

was calculated as :<br />

a CDOM<br />

2.<br />

303 OD(<br />

λ)<br />

( λ ) = (1)<br />

l<br />

where l is the optical pathlength (0.1 m). Note that aCDOM(λ) spectra of frozen samp<strong>les</strong><br />

were lower in magnitu<strong>de</strong> (~10%) than the refrigerated samp<strong>les</strong> used for irradiation<br />

experiments.<br />

The spectral absorption coefficients of the partic<strong>les</strong> retained on the GF/F filters,<br />

ap(λ) (m -1 ), were <strong>de</strong>termined at 1-nm resolution between 350 and 750 nm according to<br />

the transmittance-reflectance technique <strong><strong>de</strong>s</strong>cribed by Tassan and Ferrari [1995] using a<br />

Perkin-Elmer Lamda-19 spectrophotometer equipped with a 60-mm integrating sphere.<br />

Before the mea<strong>sur</strong>ements of the OD, the frozen filters were hydrated with 1-2 ml of<br />

filtered seawater and warmed to room temperature to avoid phytoplankton cells lyses.<br />

The OD of partic<strong>les</strong> on filter, obtained by subtracting the mean OD of 10 hydrated blank<br />

GF/F filters at each wavelength from the total mea<strong>sur</strong>ed OD with the reference beam in<br />

air, was converted to ap using the expressions given by Tassan and Ferrari [2002].<br />

Assuming that particle absorption is negligible in the near-IR [Babin and Stramski, 2002],<br />

ap(750) was subtracted from ap at all wavelengths < 750 nm.<br />

52


III.2.3. Irradiation experiments<br />

The irradiation setup and procedure for <strong>de</strong>termining the AQY of DIC production,<br />

φDIC (mol C (mol photons) -1 ), were similar to those adopted by Johannessen and Miller<br />

[2001]. Briefly, stored samp<strong>les</strong> were refiltered through 0.2-μm polycarbonate membrane<br />

filters (Millipore), acidified with HCl to pH ~3 and purged with CO2-free air for ~24 h to<br />

reduce the background DIC to un<strong>de</strong>tectable levels. The purging gas was humidified<br />

(100% relative humidity) so that no net loss of water occurred due to evaporation. The<br />

DIC-free water samp<strong>les</strong> were buffered back to approximately the original pH values with<br />

pow<strong>de</strong>red sodium borate (ACS gra<strong>de</strong>) and transferred into pre-combusted (420 °C), gas-<br />

tight quartz-windowed cylindrical cells (internal diameter: 0.02 m; length: 0.14 m) and<br />

irradiated in a thermostated incubator (0 ± 0.5 °C) using a SUNTEST CPS solar<br />

simulator equipped with a 1-kW Xe lamp. Eight spectral treatments were examined<br />

employing successive Schott long-pass glass filters; the Schott numbers of these filters<br />

are WG280, WG295, WG305, WG320, WG345, GG395, GG435 and GG495. Spectral<br />

irradiance un<strong>de</strong>r each filter was mea<strong>sur</strong>ed with an Optronics OL-754 UV-vis<br />

spectroradiometer fitted with a fibreoptic cable. The irradiance level un<strong>de</strong>r the WG280<br />

filter was set to 190 W m -2 for stations R5a, R5d and R9, and to 585 W m -2 for stations<br />

108, 406 and 409, and the irradiation time varied from 24.5 to 46 h, <strong>de</strong>pending on the<br />

abundance of CDOM in the irradiated samp<strong>les</strong> (Fig 1 in Supplementary Materials). The<br />

rationale for choosing these irradiation times and light intensities was to obtain<br />

mea<strong>sur</strong>able DIC while minimizing the extent of photobleaching. Parallel incubations in<br />

the dark were conducted to correct for potential nonphotochemical production of DIC.<br />

Extreme caution was taken to minimize bacterial contamination during sample handling<br />

and transfer. Small increases of DIC in the dark controls were often observed, but they<br />

were always similar to or <strong>les</strong>s than the amounts of DIC formed in the quartz cell un<strong>de</strong>r<br />

the GG495 filter. The dark control values were subtracted during the calculation of DIC<br />

photoproduction in each irradiated vessel.<br />

53


III.2.4. DIC mea<strong>sur</strong>ement and calculation of φDIC<br />

After irradiation, 1.50 mL of sample (in triplicates) were mixed with 1 mL of 10%<br />

DIC-free H3PO4 in an acid trap and purged with ultra-high-purity nitrogen. DIC in the<br />

sample was converted into CO2 which was introduced by a nitrogen carrier stream to a<br />

LI-COR 6262 CO2/H2O analyzer for quantification. The whole system was calibrated<br />

immediately before and after sample mea<strong>sur</strong>ement with varying volumes (usually 0.20,<br />

0.40, 0.60, and 0.8 mL) of a 12-μM DIC standard. This working standard was daily<br />

prepared by diluting a 600-μM sodium carbonate stock solution; the stock solution was<br />

ma<strong>de</strong> by dissolving dried sodium carbonate (Na2CO3, BDH AnalaR) in DIC-free pure<br />

water. All calibration curves resulted in R 2 > 0.999. The analytical precision was<br />

<strong>de</strong>termined to be ∼2 % at a concentration of 4 μM DIC.<br />

The spectral apparent quantum yield of DIC production, φDIC(λ), was <strong>de</strong>fined as<br />

the mo<strong>les</strong> of DIC photochemically produced per mole of photons absorbed by CDOM at<br />

wavelength λ. To <strong>de</strong>rive the φDIC(λ) values, the iterative curve-fitting method of<br />

Johannessen and Miller [2001] was applied. Briefly, this method assumes an appropriate<br />

mathematical functional form with unknown parameters to express the change of φDIC as<br />

a function of wavelength. Decreasing exponential functional forms are usually chosen for<br />

photochemical quantum yield spectra. The amount of DIC produced in an irradiation cell<br />

over the expo<strong>sur</strong>e time can then be predicted as a product of three terms: the assumed<br />

φDIC(λ) functional form, the mea<strong>sur</strong>ed inci<strong>de</strong>nt irradiance, and the sample absorption<br />

coefficient. The optimum values of the unknown parameters in the assumed φDIC(λ)<br />

functional form are obtained by varying these parameters from initial estimates until<br />

minimum difference between the mea<strong>sur</strong>ed and predicted DIC production is achieved.<br />

The absorption coefficient was mea<strong>sur</strong>ed before and after irradiation and was averaged<br />

according to first-or<strong>de</strong>r kinetic <strong>de</strong>cay, which well <strong><strong>de</strong>s</strong>cribes photobleaching [Del Vecchio<br />

and Blough, 2002; Xie et al., 2004]. We followed the recommendation of Hu et al. [2002]<br />

for the calculation of the number of photons absorbed by CDOM and adopted the<br />

following modified exponential form to fit the data:<br />

where k1, k2 and k3 are fitting parameters.<br />

k2<br />

/ ( λ + k3)<br />

φ DIC ( λ)<br />

= k1<br />

e<br />

(2)<br />

54


III.2.5. Mo<strong>de</strong>ling DIC photoproduction<br />

Assuming that the seawater optical properties are vertically constant in the photic<br />

zone and that the <strong>flux</strong> of photons backscattered to the atmosphere is negligible, the daily<br />

<strong>de</strong>pth-integrated DIC photoproduction rate, PDIC (mol C m -2 d -1 ), can be calculated as:<br />

a ( λ)<br />

DIC λ ( λ)<br />

dλ<br />

(3)<br />

P<br />

600<br />

−<br />

= ∫ Ed<br />

( 0 ,<br />

λ=<br />

300<br />

)<br />

CDOM<br />

at<br />

( λ)<br />

φDIC<br />

where at(λ) (m -1 ) is the total absorption coefficient (i.e., sum of absorption by seawater,<br />

−<br />

CDOM and partic<strong>les</strong>) and ( 0 , λ)<br />

E d (mol photons m -2 d -1 ) is the spectral daily<br />

downward irradiance just beneath sea <strong>sur</strong>face. Eq. 3 was used to calculate the temporal<br />

changes in PDIC in different regions of the Beaufort Sea over the period between 1979 and<br />

2003. Constant regional values of aCDOM(λ)/at(λ) and φDIC(λ) <strong>de</strong>termined from our field<br />

and laboratory work were assumed (see below). Note that in the open ocean where the<br />

fraction of light absorbed by the CDOM (i.e., the ratio aCDOM(λ)/at(λ)) approaches the<br />

value of 1 in the UV domain, PDIC can be approximated by integrating the product<br />

−<br />

of E ( 0 , λ)<br />

and φDIC(λ) over λ [e.g. Johannessen, 2000]. When partic<strong>les</strong> compete for<br />

d<br />

light absorption in the UV domain, as in coastal waters, the parameter aCDOM(λ)/at(λ) of<br />

Eq. 3 is necessary to avoid overestimation of PDIC.<br />

By using constant regional value for φDIC(λ), our approach did not explicitly take<br />

into account any potential <strong>de</strong>pen<strong>de</strong>nce on light dose of this quantity; i.e., the effect of<br />

CDOM light history on φDIC(λ). Johannessen and Miller [2001] observed a 2.3 times<br />

increase in the φDIC(λ) after a coastal water sample was extensively photobleached (loss<br />

of absorbance in the UV wavelengths > 50%) while Vähätalo and Wetzel [2004] found a<br />

13% <strong>de</strong>crease in the DOC loss AQY of a lake water sample with a 9% increase in the<br />

photons absorbed by CDOM. Therefore, both the direction and magnitu<strong>de</strong> of change in<br />

photomineralization AQY in response to light dose remain unclear. In the present study,<br />

significant photobleaching was incurred during the AQY experiments (Fig. III1 in<br />

Supplementary Materials) while no appreciable photobleaching seemed to have taken<br />

place during the transport of CDOM from the estuary of the Mackenzie River to the Shelf<br />

or even further seaward (see Results and Discussion). I<strong>de</strong>ally, experimentally <strong>de</strong>termined<br />

55


AQYs only apply to timesca<strong>les</strong> over which the amount and spectral quality of the light<br />

absorbed by CDOM in the field are equivalent to those of the light absorbed by CDOM in<br />

irradiation experiments. On larger geochemically significant timesca<strong>les</strong> (e.g., the<br />

resi<strong>de</strong>nce times of the Mackenzie River’s runoff in the Canada Basin or the entire Arctic<br />

Ocean), the uncertainties in the calculation of PDIC associated with the dose <strong>de</strong>pen<strong>de</strong>nce<br />

of φDIC(λ) should be smaller, since the extent of photobleaching in the field on these<br />

timesca<strong>les</strong>, as compared to the timesca<strong>les</strong> of land-to-sea CDOM transport, are likely<br />

closer to that occurring in our laboratory irradiations.<br />

The spectral daily downward irradiance received at the sea <strong>sur</strong>face was calculated<br />

using a radiative transfer method that uses as inputs satellite mea<strong>sur</strong>ements of total<br />

atmospheric ozone column content and of cloud cover <strong>de</strong>rived from Total Ozone<br />

Mapping Spectrometer (TOMS) data [Herman et al., 1999; Krotkov et al., 2002; Krotkov<br />

et al., 1998; Krotkov et al., 2001]. The calculations were ma<strong>de</strong> for each day between<br />

February 1979 to April 1993, and between August 1996 to November 2001, which<br />

correspond to the periods when TOMS data were available. Calculations were done for<br />

conditions with and without cloud cover. The method was modified in two respects. First,<br />

instead of using the monthly minimum Lambertian equivalent reflectivity (MLER) for<br />

the <strong>sur</strong>face albedo (As) from climatology, we used a linear combination of albedos for<br />

open water and sea ice:<br />

As water<br />

ice<br />

= A ( 1−<br />

SIC)<br />

+ A SIC<br />

(4)<br />

where SIC is the fractional sea ice <strong>sur</strong>face concentration (dimension<strong>les</strong>s, from 0 to 1), and<br />

Awater and Aice are the water and sea ice albedos, respectively. Note that As must be<br />

accounted for in the calculation of E (λ)<br />

because multiple scattering between sea<br />

d<br />

<strong>sur</strong>face and atmosphere can be significant in the UV domain. Daily SIC data were<br />

obtained from the National Snow and Ice Data Center (NSIDC) and were <strong>de</strong>rived from<br />

observations of the Scanning Multi channel Microwave Radiometer (SMMR; 1979-1987)<br />

or Special Sensor Microwave Imager (SSM/I; 1987-2003) using the NASA Team<br />

algorithm [Cavalieri et al., 1990]. The values for Awater (0.05) and Aice (0.76) were<br />

adopted from the MLER climatology for summer (SIC = 0) and winter (SIC = 1) seasons,<br />

respectively, published by Herman and Celarier [1997] and Herman et al. . Second, the<br />

calculations of irradiance were spectrally exten<strong>de</strong>d to cover the range from 300 to 600 nm<br />

56


(with 0.5 nm resolution). As the above procedure yields downward irradiance just above<br />

sea <strong>sur</strong>face, we applied a factor of 0.95 to account for specular reflection and to<br />

−<br />

obtain E ( 0 , λ)<br />

just beneath <strong>sur</strong>face as nee<strong>de</strong>d in Eq. 3.<br />

d<br />

For the calculation of total DIC photoproduction, the Southeastern Beaufort Sea<br />

was divi<strong>de</strong>d into three sub-regions: the Mackenzie Shelf (MS; 80 000 km 2 ; <strong>de</strong>pth < 200m;<br />

west of Cape Bathurst), the Amundsen Gulf (AG; 87 000 km 2 ; >118°W) and the Canada<br />

Basin (CB; 240 000 km 2 ;


Figure III.2. φDIC spectra <strong>de</strong>termined on water samp<strong>les</strong> collected during<br />

ARDEX (R5a, R5d, and R9; solid thin lines) and CASES (108, 406 and<br />

409; dashed thin lines). The grey curves are the average φDIC spectra<br />

published by Vähätalo et al. [2000] for a boreal lake, and by Johannessen<br />

and Miller [2001] for inshore, coastal and offshore waters.<br />

Table III.2. CDOM properties and mo<strong>de</strong>l parameters for φDIC.<br />

φ ( λ)<br />

= k e<br />

Station aCDOM330 SCDOM a DIC<br />

1<br />

k2<br />

/ ( λ + k3<br />

)<br />

(m -1 ) k1 k2 k3<br />

2 b<br />

r φ DIC (x10 -6 )<br />

R5d 5.44 0.0190 1.2x10 -6 726. -176. .998 20.3<br />

R5a 4.50 0.0201 4.9x10 -6 352. -224. 1.00 25.8<br />

R9 1.50 0.0204 9.6x10 -6 128. -257. .984 18.9<br />

108 0.72 0.0200 2.7x10 -10 5232. 88.7 .976 8.8<br />

406 0.69 0.0191 6.7x10 -8 1538. -111. .996 6.6<br />

409 0.62 0.0217 7.44x10 -6 128. -256. .994 14.6<br />

a<br />

Spectral slope of the the mo<strong>de</strong>l aCDOM(λ)=aCDOM(λ0) Scdom*(λ0-λ) .<br />

b 2<br />

The correlation coefficient (r ) is from the linear regression between mo<strong>de</strong>led and mea<strong>sur</strong>ed<br />

DIC production.<br />

58


III.3. Results and discussion<br />

III.3.1. Apparent Quantum Yield for DIC photoproduction<br />

The six φDIC spectra <strong>de</strong>termined during this study are shown in Fig. III.2 and the<br />

parameters of the quasi-exponential mo<strong>de</strong>l (Eq. 2) are given in Table III.2. The<br />

exponential form adopted here has a steeper slope and fits better the mea<strong>sur</strong>ed production<br />

(r 2 > 0.97) than the two-parameter single exponential form used by Johannessen and<br />

Miller and Vähätalo et al. . A <strong>de</strong>tailed comparison between these two functional forms is<br />

provi<strong>de</strong>d in the supplementary online material. To compare the photoreactivity of the<br />

CDOM, we calculated the mean φDIC value weighted by an inci<strong>de</strong>nt solar irradiance<br />

spectrum:<br />

600<br />

∫<br />

E<br />

( 0<br />

=<br />

280<br />

d<br />

600<br />

∫<br />

280<br />

59<br />

−<br />

, λ)<br />

φ<br />

( λ)<br />

dλ<br />

φ DIC . (5)<br />

E<br />

−<br />

( 0 , λ)<br />

dλ<br />

Here Ed (0 - , λ) was mo<strong>de</strong>led using the Tropospheric Ultraviolet Visible (TUV) mo<strong>de</strong>l<br />

[Madronich and Flocke, 1999] for summer solstice at latitu<strong>de</strong> of 70°N with clear sky,<br />

mo<strong>de</strong>rate aerosols concentration (optical thickness at 550 nm of 0.1), and total ozone<br />

column content of 330 DU. Note that only the spectral shape of Ed (0 - , λ) influenced the<br />

value of φ DIC (e.g., for a total ozone column content of 480 DU, the difference in φDIC<br />

35) waters with 174.0 and 786.5 × 10 -6 mol C<br />

(mol photons) -1 , respectively. However, our φDIC<br />

DIC<br />

values for low-salinity water samp<strong>les</strong><br />

(R5a and R5d, salinity < 9) are comparable to the inshore (S < 31) value of 30.4 × 10 -6<br />

mol C (mol photons) -1 mea<strong>sur</strong>ed by Johannessen and Miller [2001] and to the average<br />

is


φDIC<br />

of 25.9 x 10-6 mol C (mol photons) -1 reported by Vähätalo et al. [2000] for water<br />

samp<strong>les</strong> from Valkea-Kotinen, a humic lake located in coniferous forest of Finland<br />

(61°14′N; 25°04′E). While CDOM was mixed conservatively over the shelf estuary, as<br />

suggested by the negative linear correlation between aCDOM and salinity (Fig. III.3a), we<br />

observed a nonlinear <strong>de</strong>crease of φ DIC with increasing salinity (Fig. III3b). This behavior<br />

possibly resulted from a conservative mixing of high-φDIC inshore waters with low-φDIC<br />

offshore waters and a non-conservative changes in the photoreactivity of CDOM across<br />

the salinity gradient. The <strong>de</strong>crease of φ DIC with salinity observed in this study is opposite<br />

to the trend observed by Johannessen and Miller [2001] who proposed that, as the<br />

terrestrial aromatic structures in terrestrial CDOM are eliminated by photochemical<br />

fading or as marine-<strong>de</strong>rived CDOM increasingly dominates the light absorption, the<br />

proportion of DIC-producing to non-DIC-producing chromophores in CDOM increases.<br />

In the present study, photochemical fading of terrestrial CDOM is not supported by the<br />

conservative behavior of aCDOM(λ) (Fig. III.3a). In contrast, the lower φ DIC values were<br />

observed at stations 108 and 406 (S = 30-31) sampled in June and away from direct input<br />

of the Mackenzie River. Overall, Fig III.3b suggests that the photoreactivity of tDOM<br />

drained into the Mackenzie River changed when it was mixed into coastal and offshore<br />

waters (S


Figure III.3. (a) Variation of aCDOM(330) with salinity for water samp<strong>les</strong><br />

collected in Mackenzie Shelf (open circ<strong>les</strong>), Amundsen Gulf (open<br />

triang<strong>les</strong>), Canada Basin (open squares). Samp<strong>les</strong> collected for φDIC are<br />

shown as closed squares. (b) Variation of weighted quantum yield<br />

normalized to the integrated irradiance, φ DIC , with salinity.<br />

61


III.3.2. Role of DIC photoproduction in the organic carbon cycling<br />

The yearly pattern of DIC photoproduction (PDIC), averaged over the period<br />

corresponding to TOMS data availability (01/1979 to 04/1993; 08/1996 to11/2001), is<br />

shown in Fig. III.4. At summer solstice, PDIC reached 6.1, 2.9 and 4.1 mg C m -2 d -1 for<br />

MS, AG and CB, respectively. Given that the spectral irradiance was similar for each<br />

sub-region, the difference in PDIC was mainly due to regional differences in φDIC (Fig.<br />

III.2) and in the fraction of light absorbed by CDOM (Fig. III.5). While CDOM largely<br />

dominated the total light absorption at wavelengths 90%), the<br />

high concentration of particulate matter in the MS reduced light availability to CDOM by<br />

10 to 20% compared with the other two sub-regions. Neverthe<strong>les</strong>s, PDIC was larger in the<br />

MS due to the higher φDIC in the intermediate salinity range (Fig. III.3b).<br />

Figure III.4. Average DIC photoproduction rates, PDIC, calculated using<br />

the φDIC representing each sub-region: R5a, R5d and R9 for the Mackenzie<br />

Shelf; 108 for the Amundsen Gulf; 406 and 409 for the Canada Basin.<br />

62


Figure III.5. Average spectral light fraction absorbed by CDOM in the<br />

<strong>sur</strong>face waters at stations (Fig. III.1) sampled over the Mackenzie Shelf<br />

(dashed line), in Amundsen Gulf (dash-dotted line), and in Canada Basin<br />

(thin line).<br />

To assess the importance of photomineralization relative to other biogeochemical<br />

processes affecting the marine carbon cycle, our DIC photoproduction rates were<br />

compared with published bacterial respiration and primary production data for the same<br />

areas. Based on leucine uptake rates, Garneau et al. [2006] reported low bacterial<br />

production rates of 12 to 112 μg C m -3 d -1 (mean: 48 μg C m -3 d -1 ) in the fall of 2002 in<br />

the vicinity of the Mackenzie River estuary. Assuming a typical growth efficiency of<br />

27% for coastal Arctic bacteria [Meon and Amon, 2004], the bacterial respiration rate<br />

would be ~3.2 mg C m -2 d -1 in the upper 25 m of the water column. For comparison, PDIC<br />

was 0.6 mg C m -2 d -1 at that time of the year (Fig. III.4), equivalent to ~19% of the<br />

bacterial respiration. In late June and early July 2004, when primary productivity and<br />

temperature are higher, the bacterial respiration rate in the upper 25 m ranged from 20.2<br />

to 77.8 mg C m -2 d -1 (mean: 49.0 mg C m -2 d -1 ) on the shelf [M.-È. Garneau, unpub data].<br />

During the same period, PDIC in the MS was 5.1 mg C m -2 d -1 (Fig. III.4; Table III.3),<br />

equivalent to 10.4% of the bacterial respiration.<br />

Wavelength (nm)<br />

63


Table III.3. Comparison between daily DIC photoproduction, and total and new primary<br />

production rates in open water (in mg C m -2 d -1 ).<br />

Sub-region Period Primary Production PDIC<br />

Mackenzie Shelf<br />

(80 000 km 2 )<br />

Canada Basin<br />

(240 000 km 2 )<br />

Late July<br />

Total New<br />

67-134 (1) 3.5-5.0<br />

100-200 (1)<br />

Late August 79-145 (106) (2)<br />

64<br />

14-26 (19) (2) 1.3-1.8<br />

Amundsen Gulf<br />

(87 000 km 2 )<br />

Late August<br />

13.2 (2) 1.0<br />

(1) From Carmack et al. [2004].<br />

(2) From Lee and Whitledge [2005], personal communication of S. Lee for AG values.<br />

91.9 (2)<br />

Table III.3 compares our PDIC values with recent published values of primary<br />

production (PP) and new production (sensus Dugdale and Goering [1967]) for open<br />

waters in the same areas. PP in the MS is typically 100 to 200 mg C m -2 d -1 in late July<br />

while new production (based on nitrate drawdown) is 67 to 134 mg C m -2 d -1 [Carmack et<br />

al., 2004]. Values reported for late August in the Canada Basin range from 79 to 145 mg<br />

C m -2 d -1 for PP, with a mean of 106 mg C m -2 d -1 , and from 14 to 26 mg C m -2 d -1 for<br />

new production (based on nitrate and ammonium uptake), with a mean of 19 mg C m -2 d -1<br />

[Lee and Whitledge, 2005]. In the AG, primary and new productions in late August 2002<br />

were 91.9 and 13.2 mg C m -2 d -1 , respectively [S. H. Lee, personal communication].<br />

Based on our estimates of PDIC, photomineralization of DOC occurs at a rate equivalent<br />

to 2.6-7.8, 6.8-9.4 and 7.6% of the amount of organic carbon produced by new<br />

production in MS, CB, and AG, respectively. Note that for the small area near the Cape<br />

Bathurst, however, satellite-based PP rates can reach up to 2800 mg C m -2 d -1 during the<br />

spring phytoplankton bloom [Arrigo and van Dijken, 2004]. Therefore the relative<br />

contribution of photomineralization to the organic carbon cycling could be much lower<br />

during short periods of bloom conditions.


III.3.3. Interannual variability in DIC photoproduction<br />

Figure III.6 shows the annual photoproduction of DIC from 1979 to 2003. The<br />

mean annual production (± s. d.) was 24.9 ± 5.2, 13.7 ± 3.3 and 27.8 ± 11.8 Gg C in the<br />

MS, AG and CB, respectively. When all regions are pooled together, the photoproduction<br />

was on average 66.5 ± 18.4 Gg C y -1 . Because of the large interannual variability, the<br />

positive slopes of the linear regressions between the production rate and time are not<br />

significantly different from zero. Neverthe<strong>les</strong>s, the positive trends observed over the 25<br />

years are higher for the AG and CB (+ 1.5 and + 2.1 Gg C <strong>de</strong>ca<strong>de</strong> -1 , respectively) than for<br />

the MS (+ 0.5 Gg C <strong>de</strong>ca<strong>de</strong> -1 ). Based on these trends, the DIC photoproduction increased<br />

by 5.0%, 26.3%, and 18.1% between 1979 and 2003 for the MS, AG, and CB,<br />

respectively, with an overall increase of 14.8%.<br />

Figure III.6. Trends in the annual DIC photoproduction for period 1979-<br />

2003 for the Mackenzie Shelf, the Amundsen Gulf, the Canada Basin, and<br />

the sums of the three sub-regions.<br />

65


Figure III.7. Correlation between the annual DIC photoproduction and<br />

the annual average area of open water for: (a) Mackenzie Shelf, DIC =<br />

0.91x10 6 * X; (b) Amundsen Gulf, DIC = 0.52x10 6 * X ; (c) Canada Basin,<br />

DIC = 0.63x10 6 * X. (X = open water area in km 2 ; slope in Gg C y -1 km -2 ;<br />

n=25).<br />

The large interannual variability of the DIC photoproduction is mainly due to<br />

variability in sea ice cover. Figure III.7 shows the strong positive correlation between the<br />

annual photoproduction of DIC and the average area of open water for each sub-region.<br />

Note that the timing of the opening of sea ice is also important due to the strong seasonal<br />

variability of PDIC (Fig. III.4). For instance, the presence or absence of an ice cover in<br />

June has a stronger impact on the annual DIC photoproduction than it does in September.<br />

66


Figure III.8 shows the changes in the number of days with open water (<strong>de</strong>fined as >50 %<br />

of the area free of ice) and in the date when the opening begins (i.e., first appearance of<br />

open water) for each sub-region. The number of days with open water was on average (±<br />

s. d.) 116.8 ± 25.7, 125.7 ± 26.9 and 52.6 ± 47.6, and increased at a rate of 3.3, 12.0, and<br />

9.3 days per <strong>de</strong>ca<strong>de</strong> for MS, AG and CB, respectively. Concurrently, the opening<br />

occurred on average at day 170.9 ± 20.3, 167.1 ± 18.9 and 181.6 ± 74.3, and <strong>de</strong>creased at<br />

rates of 1.9, 14.1 and 6.8 days per <strong>de</strong>ca<strong>de</strong>, for MS, AG and CB, respectively. As expected,<br />

we observed a larger interannual variability in sea ice concentration in the Canada Basin,<br />

which was due in part to the movement of the Arctic central packed ice associated with<br />

dominant anticyclonic or cyclonic circulation regime [e.g., Comiso et al., 2003]. For<br />

instance, the open-water area in the CB remained 120 days of fully open water were observed in 1987,<br />

1993 and 1998.<br />

Figure III.8. Trends in open water area for the three sub-regions as<br />

observed by SMMR/SSMI between 1979 and 2003. The left panels show<br />

the day of the year when >50% of the <strong>sur</strong>face area is ice free. The right<br />

panels show the number of days having open water.<br />

67


In addition to the variability in sea-ice cover and in the timing of its opening, the<br />

interannual variation of the atmospheric ozone concentration and cloud cover also<br />

contributed to the variability of DIC photoproduction. Figure III.9 gives the trends<br />

between 1979 and 2001 in monthly averaged ozone column content as observed by<br />

TOMS over the southeast Beaufort Sea. In May, the ozone content was high and varied<br />

within a narrow range (379.5 ± 4.1 DU). In June and July, it <strong>de</strong>creased down to 322.4 ±<br />

10.5 and 287.8±7.22 DU, respectively. In August, the year-to-year variability was higher,<br />

as indicated by a higher standard <strong>de</strong>viation, 344.1 ± 37.2 DU, and a larger range, 270 to<br />

400 DU. Linear regression gave a significant negative slope (p < 0.05) for July and<br />

August, -5.4 and -24 DU <strong>de</strong>ca<strong>de</strong> -1 , respectively. Neverthe<strong>les</strong>s, the influence of the<br />

<strong>de</strong>creasing trends in the summer ozone column content on the DIC photoproduction was<br />

largely overshadowed by the variability in sea-ice cover (Fig. III.7).<br />

Figure III.9. Trends in monthly O3 concentration over the study area as<br />

observed by TOMS between 1979 and 2001. Note the changes of scale<br />

among the panels.<br />

68


These results show that variation of sea-ice cover is the most important factor<br />

controlling the photochemical production of DIC in the Arctic <strong>sur</strong>face waters. Our<br />

calculations did not account for seasonal variability in φDIC and optical properties of<br />

<strong>sur</strong>face waters since they are unknown. Fortunately, our field sampling took place at the<br />

time of year (June and July) when both irradiance and river plume extent were maximal,<br />

and therefore when photooxidation processes were most intense. Our sampling time,<br />

however, might not a<strong>de</strong>quately capture the influence of particulate matter dynamics on<br />

the availability of light to CDOM photochemistry. For instance, the continental shelf was<br />

sampled in early July, shortly after the ice breakup when the particle-rich river plume was<br />

spread over most of the shelf area [Carmack and Macdonald, 2002]. Later in the season<br />

the river discharge and suspen<strong>de</strong>d particulate matter abundance <strong>de</strong>crease sharply,<br />

allowing more light available for photochemical processes. The PDIC for the MS, based<br />

on the ratio of aCDOM / at obtained in early July, might therefore be un<strong>de</strong>restimated for late<br />

summer. In contrast, spring bloom events, as observed in the area of Cape Bathurst<br />

polynya [Arrigo and van Dijken, 2004], could temporarily <strong>de</strong>crease aCDOM / at in some<br />

areas.<br />

The trend in sea-ice cover extent observed in the southeastern Beaufort Sea is<br />

coherent with the trend observed over the whole Arctic Ocean [e.g., Cavalieri et al., 2003;<br />

Rothrock and Zhang, 2005]. The rate of change in the northern hemisphere sea-ice cover<br />

extent between 1979 and 2002 was estimated to be -0.36 ± 0.05 x 10 6 km 2 <strong>de</strong>ca<strong>de</strong> -1<br />

[Cavalieri et al., 2003]. If the PDIC calculated in this study is applicable to the rest of the<br />

Arctic Ocean, the increase in DIC production since 1979 can be estimated for the whole<br />

Arctic. Using the area-normalized slope (in g C y -1 km -2 ) for the CB (Fig. III.7), we<br />

estimated that photochemical production of DIC in the northern hemisphere increased by<br />

19.6-26.0 Gg C y -1 between 1979 and 2002.<br />

III.3.4. Implications for tDOC cycling in Arctic coastal waters<br />

Physical transport appears the main mechanism responsible for the removal of<br />

tDOC in the Arctic Ocean. Benner et al. [2005] reported that physical transport of tDOC<br />

to the northeast Atlantic Ocean represents 25 to 33% of the 25 Tg tDOC y -1 input into the<br />

Arctic Ocean. In addition, a similar fraction of tDOC may be exported through the<br />

69


Canadian archipelago into the northwest Atlantic [Amon, 2004] and into the Arctic <strong>de</strong>ep<br />

basin resulting from the sinking of brine-enriched waters formed over the shelves<br />

[Dittmar, 2004]. The remaining fraction (~34 to 50%) should accumulate or be removed<br />

by biological and photochemical processes in the Polar Surface Water (PSW). The<br />

relative importance for the removal of tDOC by microbial and photochemical processes<br />

is unknown [Amon, 2004]. Our study is a first attempt to quantify the photochemical sink<br />

of tDOC in an Arctic coastal environment affected by large inputs of terrestrial DOM.<br />

The main source of CDOM in the study area is the terrestrial runoff from the<br />

Mackenzie River [Guay et al., 1999; Guéguen et al., 2005]. To estimate tDOM<br />

mineralization, one needs to know the fraction of light absorbed by the terrigenous<br />

CDOM (aCDOM ter ). This fraction (in %) was approximated as<br />

a<br />

a<br />

ter<br />

CDOM<br />

CDOM<br />

⎛<br />

= ⎜<br />

⎝<br />

6.<br />

5<br />

6.<br />

5 f ⎞<br />

⎟ * 100 , (6)<br />

f + 0.<br />

5(<br />

1−<br />

f ) ⎠<br />

where f is the fraction of runoff in the <strong>sur</strong>face layer, and 6.5 and 0.5 are the aCDOM(330)<br />

values for the riverine and marine end-members, respectively (Fig. III.3a). Based on<br />

conservative estimates for f of ~25% in the MS [Macdonald et al., 1995] and ~8% in the<br />

CB [Macdonald et al., 2002], we calculated that 80 and 50% of the aCDOM is of<br />

terrigenous origin in the MS and CB, respectively (Table III.4). The contribution of<br />

aCDOM ter in the Amundsen Gulf is <strong>les</strong>s certain and <strong>de</strong>pends on the <strong>sur</strong>face currents and the<br />

autochthonous production of CDOM. Un<strong>de</strong>r calm wind conditions, the Mackenzie River<br />

plume tends to flow eastward and enter the gulf [Carmack and Macdonald, 2002].<br />

However, since PP appears relatively important around Cape Bathurst [Arrigo and van<br />

Dijken, 2004], we assume that only 20% of the CDOM in the AG is of terrigenous origin.<br />

This number represents a typical contribution of tDOC to PSW [Benner et al., 2005].<br />

Based on these assumptions, tDOC photomineralization is estimated to be 19.9 ± 4.2,<br />

13.1 ± 5.9 and 2.8 ± 0.7 Gg C y -1 for the MS, CB and AG, respectively (Table III.4),<br />

representing 1.53 ± 0.32%, 1.0 ± 0.5% and 0.21 ± 0.04% of the tDOC <strong>flux</strong> into the<br />

southeastern Beaufort Sea via the Mackenzie River. Put together, these estimates<br />

represent 2.8 ± 0.55% of the annual input of tDOC (1.3 Tg) [Telang et al., 1991].<br />

Therefore, our results corroborate the speculation by Benner et al. [2005] that<br />

70


photochemical transformation of tDOM occurs in the PSW but that this removal process<br />

is a small fraction of the total organic carbon <strong>flux</strong>.<br />

Table III.4. Mean annual DIC photoproduction and estimates of tDOC mineralized by<br />

photooxidation.<br />

Sub-region DIC<br />

photoproduction (1)<br />

Gg C y -1<br />

a<br />

a<br />

ter<br />

CDOM<br />

71<br />

CDOM<br />

(%)<br />

tDOC<br />

mineralized (1)<br />

Gg C y -1<br />

Mackenzie Shelf 24.9 ± 5.2 80 19.9 ± 4.2<br />

Canada Basin 27.8 ± 11.8 50 13.9 ± 5.9<br />

Amundsen Gulf 13.7 ± 3.3 20 2.8 ± 0.7<br />

Total 66.5 ± 18.5 36.6 ± 7.1<br />

(1) Average ± s. d. calculated from 1979 to 2003.<br />

Amon and Meon [2004] suggested that the fate of tDOM un<strong>de</strong>r reduced ice cover<br />

could become significantly different from the present situation, and our results allow<br />

further analysis of this prediction for photomineralization losses. Un<strong>de</strong>r the minimal sea<br />

ice conditions observed during spring-summer 1998, DIC photoproduction nearly<br />

doubled relative to the averaged value for the period from 1979 to 2003 (123.0 vs. 66.5<br />

Gg C y -1 , P


30.0, 32.5 and 4.2 Gg y -1 of tDOC to be photomineralized in 1998 in the MS, CB, and<br />

AG respectively, the sum of these numbers representing 5.1% of the annual tDOC release<br />

from the Mackenzie River. The ice-free scenario gives an equivalent estimate of 6.2%<br />

(assuming no change in tDOC input). For these conditions, we estimated that 43.5 Gg y -1<br />

of allochthonous DOC would be photomineralized in the Canada Basin, which is higher<br />

than the estimates for autochthonous organic carbon sequestered in the <strong>de</strong>ep ocean basin<br />

(22.8 to 42 Gg C y -1 ). This range is based on the assumptions that 1% of the annual PP is<br />

buried in sediments [Stein and Macdonald, 2004] and that PP rate ranges from 9.5 to 17.4<br />

g C m -2 in the open water of CB [Lee and Whitledge, 2005].<br />

In this study, photooxidation of CDOM within or un<strong>de</strong>r sea ice was not<br />

consi<strong>de</strong>red because most of the radiative energy is either backscattered to the atmosphere<br />

or absorbed by the freshly produced biogenic material accumulated at the bottom of sea<br />

ice . Although photochemical reactions within the organic-rich bottom ice layer is likely<br />

to play a role in the marine DOM cycling [Xie and Gosselin, 2005], it does not affect the<br />

un<strong>de</strong>rlying terrigenous DOM. Due to the lack of tDOM data in sea ice [Amon, 2004], it is<br />

impossible at present to estimate the photomineralization of tDOC in sea ice.<br />

III.4. Summary and conclusions<br />

This study represents the first attempt to quantify photomineralization of DOM in<br />

an Arctic coastal ecosystem that receives large riverine inputs of terrigenous materials.<br />

The quantum yield spectra for DIC photoproduction (φDIC) are the first to be obtained for<br />

the Arctic, and among the first for marine waters in general. The non-conservative<br />

behavior of φDIC in the freshwater-saltwater transitional zone suggests that the<br />

photoreactivity of tDOM transported by Mackenzie River changed when it was mixed<br />

into coastal and offshore waters, and/or that marine algae-<strong>de</strong>rived DOM was <strong>les</strong>s<br />

photoreactive than tDOM in terms of photoproduction of DIC.<br />

Photomineralization of DOC in the study area, as estimated from DIC<br />

photoproduction, occurred at a rate equivalent to ~10% of the bacterial respiration rate<br />

and


esponse to the <strong>de</strong>creasing sea ice extent. Our estimates of tDOC photomineralization<br />

confirm that photooxidation presently is not a major removal pathway for tDOM in the<br />

Arctic. The photochemical sink for tDOM is, however, anticipated to grow in the future<br />

due to <strong>de</strong>crease in sea-ice extent as predicted by the Global Circulation Mo<strong>de</strong>l [ACIA,<br />

2005]. In addition to remineralizing CDOM to CO2, photolysis also produces biologically<br />

labile DOM compounds [e.g., Kieber et al., 1989; Mopper et al., 1991]. The latter may<br />

provi<strong>de</strong> another pathway by which Arctic tDOM is removed, thereby amplifying the<br />

effect of <strong>de</strong>crease in sea ice cover on tDOM cycling in the Arctic Ocean.<br />

III.5. Auxiliary material: On the <strong>de</strong>termination of the Apparent Quantum Yield for<br />

DIC photoprodcution<br />

In sections III.2.3 and III.2.4, we <strong><strong>de</strong>s</strong>cribed the method adopted to <strong>de</strong>termine of<br />

the Apparent Quantum Yield for DIC photoproduction. Briefly, we exposed filtered<br />

seawater to artificial sunlight. To obtain spectral value for φDIC, the inci<strong>de</strong>nt light was<br />

filtered using eight Schott long-pass glass filters. For each of the six experiments, spectral<br />

absorption by colored dissolved organic matter (aCDOM) was mea<strong>sur</strong>ed before and after<br />

the irradiation. Figure III.A.1 presents the mea<strong>sur</strong>ed aCDOM for the six stations before<br />

(black) and after (red) the irradiation experiments. The loss of absorption observed is due<br />

to photobleaching. This change was accounted for in the calculation of the AQY by using<br />

an appropriately averaged absorption coefficient. To calculated the amount of energy<br />

absorbed by the CDOM during the course of the experiment, Qa, the following equation<br />

was used:<br />

⎛ a ⎞ CDOM<br />

= E(<br />

0)<br />

× ⎜ × S × [ 1−<br />

exp( −a<br />

× L)<br />

]<br />

a ⎟<br />

, (mol photons s<br />

⎝ t ⎠<br />

-1 nm -1 ),<br />

Qa t<br />

S: cross-section of the irradiation cells;<br />

L: pathlength of the irradiation cells;<br />

E(0): irradiance just below the upper window of the irradiation cells<br />

at: total absorption coefficient, which is only acdom+awater since the sample was filtered<br />

before irradiation.<br />

73


Note that in all samp<strong>les</strong>, the fraction aCDOM/atotal is >90% in the UV-blue domain.<br />

aCDOM was averaged according to first-or<strong>de</strong>r kinetic <strong>de</strong>cay (i.e., mean aCDOM =<br />

exp((ln(ainitial)+ln(aend))/2)).<br />

To <strong>de</strong>rive the φDIC(λ) values, the iterative curve-fitting method of Johannessen<br />

and Miller [2001] was applied. Briefly, this method assumes an appropriate mathematical<br />

functional form with unknown parameters to express the change of φDIC as a function of<br />

wavelength. Decreasing exponential functional forms are usually chosen for<br />

photochemical quantum yield spectra. The amount of DIC produced in an irradiation cell<br />

over the expo<strong>sur</strong>e time can then be predicted as a product of three terms: the assumed<br />

φDIC(λ) functional form, the mea<strong>sur</strong>ed inci<strong>de</strong>nt irradiance, and the sample absorption<br />

coefficient. The optimum values of the unknown parameters in the assumed φDIC(λ)<br />

functional form are obtained by varying these parameters from initial estimates until<br />

minimum difference between the mea<strong>sur</strong>ed and predicted DIC production is achieved.<br />

We tested two exponential form to fit the data<br />

1.<br />

2.<br />

φ<br />

−(<br />

m1<br />

+ m2<br />

( λ + 290)<br />

DIC ( λ)<br />

= e (Single exponential, as in Johannessen and<br />

Miller [2001]);<br />

k2<br />

/ ( λ + k3)<br />

φ DIC ( λ)<br />

= k1<br />

e (Quasi-exponential).<br />

where m1, m2, k1, k2 and k3 are fitting parameters. Figure III.A.2 shows the mea<strong>sur</strong>ed<br />

versus the mo<strong>de</strong>led DIC within each irradiated cells. Table III.A.1 gives the regression<br />

coefficients for each mo<strong>de</strong>l and station. These show that the performance of Quasi-<br />

exponential form mo<strong>de</strong>l was better than the single exponential form as previously used<br />

by Johannessen and Miller [2001].<br />

74


Figure III.A.1. Absorption spectra of CDOM for before (black line) and<br />

after the irradiation for the sample un<strong>de</strong>r WG280 cutoff filter (red line).<br />

The time of irradiation and the irradiance level un<strong>de</strong>r the cutoff filter are<br />

also shown.<br />

75


Table III.A.1. Comparison of statistical results between the single exponential and quasiexponential<br />

functional forms<br />

Station R5d R5a R9 108 406 409<br />

r Single exponential 0.990 0.994 0.964 0.956 0.994 0.988<br />

2<br />

Quasi-exponential 0.998 1.00 0.984 0.976 0.996 0.994<br />

∑(mo<strong>de</strong>led- Single exponential 0.173 0.065 0.028 0.075 0.010 0.011<br />

mea<strong>sur</strong>ed) 2 Quasi-exponential 0.025 0.002 0.009 0.020 0.006 0.003<br />

Figure III.A.2. Mo<strong>de</strong>led versus mea<strong>sur</strong>ed DIC photoproduction rates for<br />

the Single exponential (squares) and Quasi-exponential (circ<strong>les</strong>) functional<br />

forms.<br />

76


III.6. References<br />

Aagaard, K., and E. C. Carmack (1989), The role of sea ice and other fresh water in the<br />

arctic circulation, J. Geophys. Res., 94, 14485-14498.<br />

ACIA (2005), Arctic Climate Impact Assessment, 1042 pp., Cambridge Press University.<br />

Amon, R. M. W. (2004), The role of dissolved organic matter for the organic carbon<br />

cycle in the Arctic Ocean, in The organic carbon cycle in the Arctic Ocean, edited<br />

by R. Stein, and R. W. MacDonald, pp. 83-99, Springer, New York.<br />

Amon, R. M. W., and B. Meon (2004), The biogeochemistry of dissolved organic matter<br />

and nutrients in two large Arctic estuaries and potential implications for our<br />

un<strong>de</strong>rstanding of the Arctic Ocean system, Mar. Chem., 92, 311-330.<br />

Arrigo, K. R., and G. L. van Dijken (2004), Annual cyc<strong>les</strong> of sea ice and phytoplankton<br />

in Cape Bathurst polynya, southeastern Beaufort Sea, Canadian Arctic, Geophys.<br />

Res. Lett., 31, L08304, doi:10.1029/2003GL013978.<br />

Babin, M., and D. Stramski (2002), Light absorption by aquatic partic<strong>les</strong> in the nearinfrared<br />

spectral region, Limnol. Oceanogr., 47, 911-915.<br />

Babin, M., D. Stramski, G. M. Ferrari, H. Claustre, A. Bricaud, G. Obolensky, and N.<br />

Hoepffner (2003), Variations in the light absorption coefficients of phytoplankton,<br />

nonalgal partic<strong>les</strong>, and dissolved organic matter in coastal waters around Europe,<br />

J. Geophys. Res., 108(C7), 3211, doi: 10.1029/2001JC000882.<br />

Barber, D. G., and J. M. Hanesiak (2004), Meteorological forcing of sea ice<br />

concentrations in the southern Beaufort Sea over the period 1979 to 2000, J.<br />

Geophys. Res., 109(C6), C06014, doi:10.1029/2003JC002027.<br />

Benner, R., B. Benitez-Nelson, K. Kaiser, and R. M. W. Amon (2004), Export of young<br />

terrigenous dissolved organic carbon from rivers to the Arctic Ocean, Geophys.<br />

Res. Lett., 31, L05305.<br />

Benner, R., P. Louchouarn, and R. M. W. Amon (2005), Terrigenous dissolved organic<br />

matter in the Arctic Ocean and its transport to <strong>sur</strong>face and <strong>de</strong>ep waters of the<br />

North Atlantic, Global Biogeochem. Cyc<strong>les</strong>, 19, GB2025,<br />

doi:10.1029/2004GB002398.<br />

Benner, R., and S. Opsahl (2001), Molecular indicators of the sources and<br />

transformations of dissolved organic matter in the Mississippi river plume, Org.<br />

Geochem., 32, 597-611.<br />

77


Camill, P. (2005), Permafrost thaw accelerates in boreal peatlands during late-20th<br />

century climate warming, Climatic Change, 68, 135-152.<br />

Carmack, E. C., and R. W. Macdonald (2002), Oceanography of the Canadian Shelf of<br />

the Beaufort Sea: A setting for Marine Life, Arctic, 55, 29-45.<br />

Carmack, E. C., R. W. MacDonald, and S. Jasper (2004), Phytoplankton productivity on<br />

the Canadian Shelf of the Beaufort Sea, Mar. Ecol. Prog. Ser., 277, 37-50.<br />

Cavalieri, D. J., P. Gloersen, and J. Zwally (1990), DMSP SSMI/I daily polar grid<strong>de</strong>d sea<br />

ice concentrations, edited by J. A. Maslanik, and J. Stroeve, pp. Digital media,<br />

National Snow and Ice Data Center, Boul<strong>de</strong>r, CO.<br />

Cavalieri, D. J., C. L. Parkinson, and K. Y. Vinnikov (2003), 30-Year satellite record<br />

reveals contrasting Arctic and Antarctic <strong>de</strong>cadal sea ice variability, Geophys. Res.<br />

Lett., 30, 1970, doi:10.1029/2003GL01803.<br />

Comiso, J. C., J. Yang, H. Susumo, and R. A. Krishfield (2003), Detection change in the<br />

Arctic using satellite and in situ data, J. Geophys. Res., 108(C12), 3384,<br />

doi:10.1029/2002JC001347.<br />

Del Vecchio, R., and N. V. Blough (2002), Photobleaching of chromophoric dissolved<br />

organic matter in natural waters: kinetics and mo<strong>de</strong>lling, Mar. Chem., 78, 231-253.<br />

Dittmar, T. (2004), Evi<strong>de</strong>nce for terrigenous dissolved organic nitrogen in the Arctic<br />

<strong>de</strong>ep sea, Limnol. Oceanogr., 49, 148-156.<br />

Dittmar, T., and G. Kattner (2003), The biogeochemistry of the river and shelf ecosystem<br />

of the Arctic Ocean: a review, Mar. Chem., 83, 105-120.<br />

Dugdale, R. C., and J. J. Goering (1967), Uptake of new and regenerated forms of<br />

nitrogen in primary production, Limnol. Oceanogr., 12, 196-206.<br />

Fioletov, V. E., M. G. Kimlin, N. Krotkov, L. J. B. McArthur, J. B. Kerr, D. I. Wardle, J.<br />

R. Herman, R. Meltzer, T. W. Mathews, and J. Kaurola (2004), UV in<strong>de</strong>x<br />

climatology over the United States and Canada from ground-based and satellite<br />

estimates, J. Geophys. Res., 109(D22), D22308, doi:10.1029/2004JD004820.<br />

Frey, K. E., and L. C. Smith (2005), Amplified carbon release from vast West Siberian<br />

peatlands by 2100, Geophys. Res. Lett., 32, L09401, doi:10.1029/2004GL022025.<br />

Garneau, M.-È., W. F. Vincent, L. Alonso-Saez, Y. Gratton, and C. Lovejoy (2006),<br />

Prokaryotic community structure and hetereotrophic production in a riverinfluence<br />

arctic ecosystem, Aquat. Microb. Ecol., 42, 27-40.<br />

78


Guay, C. K., G. P. Klinkhammer, K. K. Falkner, R. Benner, P. G. Coble, T. E. Whitledge,<br />

B. Black, F. J. Bussell, and T. A. Wagner (1999), High-resolution mea<strong>sur</strong>ements<br />

of dissolved organic carbon in the Arctic Ocean by in situ fiber-optic<br />

spectrometry, Geophys. Res. Lett., 26(8), 1007-1010.<br />

Guéguen, C., L. Guo, and N. Tanaka (2005), Distributions and characteristics of colored<br />

dissolved organic matter in the Western Arctic Ocean, Cont. Shelf Res., 25, 1195-<br />

1207.<br />

Hansell, D. A., D. Kadko, and N. R. Bates (2004), Degradation of Terrigenous Dissolved<br />

Organic Carbon in the Western Arctic Ocean, Science, 304, 858-861.<br />

Herman, J. R., and E. A. Celarier (1997), Earth <strong>sur</strong>face reflectivity climatology at 340-<br />

380 nm from TOMS data, J. Geophys. Res., 102(D23), 28003-28011.<br />

Herman, J. R., N. Krotkov, E. Celarier, D. Larko, and G. Labow (1999), Distribution of<br />

UV radiation at the Earth's <strong>sur</strong>face from TOMS-mea<strong>sur</strong>ed UV-backscattered<br />

radiances, J. Geophys. Res., 104(D10), 12059-12076.<br />

Hernes, P. J., and R. Benner (2003), Photochemical and microbial <strong>de</strong>gradation of<br />

dissolved lignin phenols: Implications for the fate of terrigenous dissolved organic<br />

matter in marine environments, J. Geophys. Res., 108(C9), 3291, doi:<br />

1029/2002JC001421.<br />

Hu, C., F. E. Muller-Karger, and R. G. Zepp (2002), Absorbance, absorption coefficient,<br />

and apparent quantum yield: A comment on common ambiguity in the use of<br />

these optical concepts, Limnol. Oceanogr., 47, 1261-1267.<br />

Johannessen, S. C. (2000), A photochemical sink for dissolved organic carbon in the<br />

ocean, Ph. D. thesis, Dalhousie University, Halifax.<br />

Johannessen, S. C., and W. L. Miller (2001), Quantum yield for the photochemical<br />

production of dissolved inorganic carbon in seawater, Mar. Chem., 76, 271-283.<br />

Kieber, D. J., J. McDaniel, and K. Mopper (1989), Photochemical source of biological<br />

substrates in sea water: implications for carbon cycling, Nature, 341, 637-639.<br />

Kieber, R. J., X. Zhou, and K. Mopper (1990), Formation of carbonyl compounds from<br />

UV-induced photo<strong>de</strong>gradation of humic substances in natural waters: Fate of<br />

riverine carbon in the sea, Limnol. Oceanogr., 35(7), 1503-1515.<br />

Krotkov, N., J. Herman, P. K. Bhartia, C. Seftor, A. Arola, S. Kalliskota, P. Taalas, and I.<br />

V. Geogdzhaev (2002), Version 2 total ozone mapping spectrometer ultraviolet<br />

algorithm: problems and enhancements, Opt. Eng., 41, 3028-3039.<br />

79


Krotkov, N. A., P. K. Bhartia, J. R. Herman, V. Fioletov, and J. Kerr (1998), Satellite<br />

estimation of spectral <strong>sur</strong>face UV irradiance in the presence of tropospheric<br />

aerosols - 1. Cloud-free case, J. Geophys. Res., 103(D8), 8779-8793.<br />

Krotkov, N. A., J. R. Herman, P. K. Bhartia, V. Fioletov, and Z. Ahmad (2001), Satellite<br />

estimation of spectral <strong>sur</strong>face UV irradiance 2. Effects of homogeneous clouds<br />

and snow, J. Geophys. Res., 106(D11), 11743-11759.<br />

Lee, S. H., and T. R. Whitledge (2005), Primary and new production in the <strong>de</strong>ep Canada<br />

Basin during summer 2002, Polar Biology, 28, 190-197.<br />

Macdonald, R. W., F. A. McLaughlin, and E. C. Carmack (2002), Fresh water and its<br />

sources during the SHEBA drift in the Canada Basin of the Arctic Ocean, Deep<br />

Sea Res., Part I, 49, 1769-1785.<br />

Macdonald, R. W., D. W. Paton, E. C. Carmack, and A. Omstedt (1995), The freshwater<br />

budget and un<strong>de</strong>r-ice spreading of Mackenzie River Water in the Canadian<br />

Beaufort Sea based on salinity and 18 O/ 16 O mea<strong>sur</strong>ements in water and ice., J.<br />

Geophys. Res., 100(C1), 895-919.<br />

Madronich, S., and S. Flocke (1999), The role of solar radiation in atmospheric chemistry,<br />

in Handbook of Environmental Chemistry, edited by P. Boule, pp. 1-26, Springer-<br />

Verlag, Hei<strong>de</strong>lberg, Germany.<br />

Meon, B., and R. M. W. Amon (2004), Heterotrophic bacterial activity and <strong>flux</strong>es of<br />

dissolved free amino acids and glucose in the Arctic rivers Ob, Yenisei and the<br />

adjacent Kara Sea, Aquat. Microb. Ecol., 37, 121-135.<br />

Miller, W. L., and R. G. Zepp (1995), Photochemical production of dissolved inorganic<br />

carbon from terrestrial organic matter : Significance to the oceanic carbon cycle,<br />

Geophys. Res. Lett., 22, 417-420.<br />

Mopper, K., and D. J. Kieber (2002), Photochemistry and the Cycling of Carbon, Sulfur,<br />

Nitrogen and Phosphorus, in Biogeochemistry of Marine Dissolved Organic<br />

Matter, edited by D. A. Hansell, and C. A. Carlson, pp. 455-507, Aca<strong>de</strong>mic Press,<br />

San Diego.<br />

Mopper, K., X. Zhou, R. J. Kieber, D. J. Kieber, R. J. Sikorski, and R. D. Jones (1991),<br />

Photochemical <strong>de</strong>gradation of dissolved organic carbon and its impact on the<br />

oceanic carbon cycle, Nature, 353, 60-62.<br />

Overland, J. E., M. C. Spillane, and N. N. Sorei<strong>de</strong> (2004), Integrated analysis of physical<br />

and biological Pan-Arctic change, Climatic Change, 63, 291-322.<br />

80


Peterson, B. J., R. M. Holmes, J. W. McClelland, C. J. Vorosmarty, R. B. Lammers, A. I.<br />

Shiklomanov, and S. Rahmstorf (2002), Increasing river discharge to the Arctic<br />

Ocean, Science, 298, 2171-2173.<br />

Rachold, V., H. Eicken, V. V. Gor<strong>de</strong>ev, M. N. Grigoriev, H.-W. Hubberten, A. P. Lisitzin,<br />

V. P. Shevchenko, and L. Schirrmeister (2004), Mo<strong>de</strong>rn terrigenous organic<br />

carbon input to the Arctic Ocean, in The organic carbon cycle in the Arctic Ocean,<br />

edited by R. Stein, and R. W. MacDonald, pp. 33-41, Springer, New York.<br />

Rex, M., R. J. Salawitch, P. von <strong>de</strong>r Gathen, N. R. P. Harris, M. P. Chipperfield, and B.<br />

Naujokat (2004), Arctic ozone loss and climate change, Geophys. Res. Lett., 31,<br />

L04116, doi:10.1029/2003GL018844.<br />

Rothrock, D. A., and J. Zhang (2005), Arctic Ocean sea ice volume: What explains its<br />

recent <strong>de</strong>pletion?, J. Geophys. Res., 110(C1).<br />

Serreze, M. C., J. E. Walsh, F. S. Chapin, T. Osterkamp, M. Dyurgerov, V. Romanovsky,<br />

W. C. Oechel, J. Morison, T. Zhang, and R. G. Barry (2000), Observational<br />

evi<strong>de</strong>nce of recent change in the northern high-latitu<strong>de</strong> environment, Climatic<br />

Change, 46, 159-207.<br />

Smith, L. C., G. M. MacDonald, A. A. Velichko, D. W. Beilman, O. K. Borisova, K. E.<br />

Frey, K. V. Kremenetski, and Y. Sheng (2004), Siberian peatlands a net carbon<br />

sink and global methane source since the early Holocene, Science, 303(5656),<br />

353-356.<br />

Stein, R., and R. W. Macdonald (2004), Organic carbon budget: Arctic Ocean vs. Global<br />

Ocean, in The Organic Carbon Cycle in the Arctic Ocean, edited by R. Stein, and<br />

R. W. MacDonald, pp. 315-322, Springer, New York.<br />

Tassan, S., and G. M. Ferrari (1995), An alternative approach to absorption<br />

mea<strong>sur</strong>ements of aquatic partic<strong>les</strong> retained on filters, Limnol. Oceanogr., 40,<br />

1358-1368.<br />

Tassan, S., and G. M. Ferrari (2002), A sensitivity analysis of the 'transmittancereflectance'<br />

method for mea<strong>sur</strong>ing light absorption by aquatic partic<strong>les</strong>, J.<br />

Phytoplankton. Res., 24, 757-776.<br />

Telang, S. A., R. Pocklington, A. S. Naidu, E. A. Romankevich, I. I. Gitelson, and<br />

Glacyshev (1991), Carbon and mineral transport in Major North American,<br />

Russian Arctic, and Siberian Rivers: the St Lawrence, the Mackenzie, the Yukon,<br />

the Arctic Alaskan Rivers, the Arctic Basin rivers in the Soviet Union, and the<br />

Yenisei, in Biogeochemistry of Major World Rivers. SCOPE 42, edited by E. T.<br />

Degens, S. Kempe, and J. E. Richey, pp. 74-104, Wiley, New York.<br />

81


Vähätalo, A. V., M. Salkinoja-Salonen, P. Taalas, and K. Salonen (2000), Spectrum of<br />

the Quantum yield for the photochemical mineralization of dissolved organic<br />

carbon in a humic lake, Limnol. Oceanogr., 45, 664-676.<br />

Vähätalo, A. V., and R. G. Wetzel (2004), Photochemical and microbial <strong>de</strong>composition<br />

of chromophoric dissolved organic matter during long (months-years) expo<strong>sur</strong>es,<br />

Mar. Chem., 89, 313-326.<br />

Vincent, W. F., and C. Belzile (2003), Biological UV expo<strong>sur</strong>e in the polar oceans:<br />

Arctic-Antarctic comparisons, in Antarctic Biology in a Global Context, edited by<br />

W. J. Wolff, pp. 176-181., Backhuys Publishers, Lei<strong>de</strong>n, Netherlands.<br />

Wadhams, P., and N. R. Davis (2000), Further evi<strong>de</strong>nce of ice thinning in the Arctic<br />

Ocean, Geophys. Res. Lett., 27, 3973-3975.<br />

Xie, H., and M. Gosselin (2005), Photoproduction of carbon monoxi<strong>de</strong> in first-year sea<br />

ice in Franklin Bay, southeastern Beaufort Sea, Geophys. Res. Lett., 32, L12606,<br />

doi:10.1029/2005GL022803.<br />

Xie, H., O. C. Zafiriou, W.-J. Cai, R. G. Zepp, and Y. Wang (2004), Photooxidation and<br />

its effects on the carboxyl content of dissolved organic matter in two coastal<br />

rivers in the southeastern United States, Environ. Sci. Tech., 38(15), 4113-4119.<br />

82


Chapitre IV : Quantification améliorée <strong>de</strong> la<br />

photooxydation <strong>de</strong> la matière organique dissoute<br />

colorée dans <strong>les</strong> eaux côtières à l’ai<strong>de</strong> <strong><strong>de</strong>s</strong><br />

propriétés optiques inhérentes dérivées <strong><strong>de</strong>s</strong><br />

données <strong>de</strong> couleur <strong>de</strong> l’océan<br />

83


IV.A Résumé<br />

Pour estimer le taux <strong>de</strong> photooxydation <strong>de</strong> la matière organique dans <strong>les</strong> eaux côtières, le<br />

rapport entre <strong>les</strong> coefficients d’absorption <strong>de</strong> la matière organique dissoute colorée (CDOM)<br />

et total ([a CDOM/a t]) se doit d’être connu <strong>de</strong>puis l’ultraviolet (UV) jusqu’à la partie verte du<br />

spectre électromagnétique. La variabilité <strong>de</strong> ce rapport à 412 nm, observée à 307 stations<br />

visitées dans plusieurs environnements côtiers, est très importante, soit entre 20 et 94%. Une<br />

relation empirique est proposée pour estimer [a CDOM/a t](412) à partir <strong><strong>de</strong>s</strong> me<strong>sur</strong>es spatia<strong>les</strong> <strong>de</strong><br />

la réflectance <strong>de</strong> l’océan. L’incertitu<strong>de</strong> absolue <strong>de</strong> l’algorithme <strong>de</strong> [a CDOM/a t](412) est <strong>de</strong> ±<br />

14%. Dans <strong><strong>de</strong>s</strong> environnements tels que la Mer Baltique et la Mer du Nord, l’algorithme<br />

permet <strong>de</strong> séparer le CDOM <strong><strong>de</strong>s</strong> particu<strong>les</strong> non-alga<strong>les</strong> à l’échelle régionale. À l’ai<strong>de</strong> d’un<br />

modèle spectral <strong>de</strong> photoproduction <strong>de</strong> <strong>carbone</strong> inorganique dissous (DIC) intégrée dans la<br />

colonne d’eau, on a évalué l’impact <strong>de</strong> 1) l’erreur <strong>de</strong> [a CDOM/a t](412) retrouvé, et <strong>de</strong> 2)<br />

l’extrapolation <strong>de</strong> ce rapport entre l’UV et le vert, <strong>sur</strong> la photooxydation. L’incertitu<strong>de</strong> <strong>sur</strong> la<br />

valeur absolue <strong>de</strong> [a CDOM/a t](412) retrouvée résulte en une incertitu<strong>de</strong> <strong>sur</strong> la photoproduction<br />

<strong>de</strong> DIC inférieure à 50%. Si la valeur du coefficient d’absorption total à 412 nm est connue,<br />

l’erreur due à l’extrapolation spectrale <strong>de</strong> [a CDOM/a t](412) est inférieure à 20%. L’algorithme a<br />

été appliqué à une image SeaWiFS du sud-est <strong>de</strong> la Mer <strong>de</strong> Beaufort où <strong><strong>de</strong>s</strong> ren<strong>de</strong>ments<br />

quantiques apparents spectraux (AQY) pour la photoproduction <strong>de</strong> DIC furent déterminés<br />

dans <strong>les</strong> principa<strong>les</strong> masses d’eau. En se basant <strong>sur</strong> ces spectres <strong>de</strong> AQY, la variabilité<br />

spatiale du taux <strong>de</strong> photoproduction <strong>de</strong> DIC résultant <strong><strong>de</strong>s</strong> variations du spectre <strong>de</strong> [a CDOM/a t]<br />

était <strong>de</strong> l’ordre d’un facteur trois au-<strong><strong>de</strong>s</strong>sus du plateau continental et au-<strong>de</strong>là. Ces résultats<br />

démontrent clairement qu’il est nécessaire <strong>de</strong> tenir compte <strong><strong>de</strong>s</strong> variations spatia<strong>les</strong> <strong>de</strong><br />

[a CDOM/a t] pour quantifier <strong>les</strong> processus <strong>de</strong> photooxydation dans <strong>les</strong> eaux côtières.<br />

84


IV.B. Article soumis à la revue Journal of Geophysical Research - Oceans (28<br />

décembre 2006): “Improved quantification of Chromophoric Dissolved Organic<br />

Matter photooxidation in coastal waters using satellite-<strong>de</strong>rived inherent optical<br />

properties”<br />

Simon Bélanger 1 , Marcel Babin 1 and Pierre Larouche 2<br />

1 Laboratoire d’Océanographie <strong>de</strong> Villefranche, Centre National <strong>de</strong> la Recherche<br />

Scientifique, Université Pierre et Marie Curie - Paris 6, 06230 Villefranche-<strong>sur</strong>-mer,<br />

France<br />

2 Institut Maurice-Lamontagne, Pêches et Océans Canada, Mont-Joli, QC, Canada G5H<br />

3Z4<br />

85


Abstract<br />

To estimate the rate of organic matter photooxidation in coastal waters, the ratio between<br />

chromophoric dissolved organic matter (CDOM) and total absorption coefficients<br />

([aCDOM/at]) must be known from the UV to the green spectral range. At 307 sites<br />

sampled in various coastal marine environments, the [aCDOM/at] at 412 nm was found to<br />

vary within wi<strong>de</strong> range, between 20 and 94%. An empirical algorithm was <strong>de</strong>veloped to<br />

retrieve [aCDOM/at](412) from satellite remote sensing reflectance. The absolute<br />

uncertainty on the [aCDOM/at] retrieval was ±14%. As exampled with the data from the<br />

Baltic and North Seas, the algorithm provi<strong><strong>de</strong>s</strong> a mean to separate the CDOM from the<br />

colored <strong>de</strong>trital material (i.e. CDM=CDOM+non-algal partic<strong>les</strong>) absorption coefficient at<br />

the regional scale. Using a <strong>de</strong>pth-integrated spectral dissolved inorganic carbon (DIC)<br />

photoproduction mo<strong>de</strong>l, we assessed the impact of 1) the error on the <strong>de</strong>rived [aCDOM/at]<br />

ratio, and 2) the extrapolation of this ratio over the green and UV spectral domains, on<br />

photooxidation. The uncertainty associated with retrieved [aCDOM/at](412) results in an<br />

uncertainty on DIC photoproduction estimates of


IV.1. Introduction<br />

Dissolved organic carbon (DOC) in the ocean is a major pool of carbon in the<br />

earth system of 685 Gt C [Hansell and Carlson, 1998], a value comparable to the mass of<br />

atmospheric carbon dioxi<strong>de</strong> (CO2). In the context of global warming, it is important to<br />

better un<strong>de</strong>rstand and quantify the sources and sinks of DOC in the ocean [Carlson,<br />

2002]. The role of photochemical oxidation (c.f. photooxidation) of the chromophoric<br />

dissolved organic matter (CDOM) has received increasing attention because it may<br />

represent a major sink for DOC from terrestrial [Kieber et al., 1990; Miller and Zepp,<br />

1995; Benner and Opsahl, 2001] and marine [Johannessen, 2000] origins, in the ocean.<br />

Neverthe<strong>les</strong>s, quantitative assessment for this carbon sink at global- or regional-scale are<br />

scarce [e.g., Johannessen, 2000; Bélanger et al., 2006].<br />

Quantitative assessment of CDOM photooxidation requires the knowledge of the<br />

inci<strong>de</strong>nt irradiance in the ultraviolet (UV) and blue part of the spectrum, the spectral<br />

absorption properties of seawater and its constituents, and the spectral apparent quantum<br />

yield of the reaction (AQY). The latter is <strong>de</strong>fined as the ratio of the number of<br />

photoproduct molecu<strong>les</strong> formed per photon absorbed by the CDOM. Carbon monoxi<strong>de</strong><br />

(CO) and carbon dioxi<strong>de</strong> (CO2, mea<strong>sur</strong>ed as dissolved inorganic carbon (DIC)) are the<br />

two major photoproducts of CDOM photooxidation [see review by Mopper and Kieber,<br />

2002]. Assuming that all inci<strong>de</strong>nt radiation penetrating an optically homogenous water<br />

column is absorbed un<strong>de</strong>rneath, the <strong>de</strong>pth-integrated photoproduction rate of DIC (or CO)<br />

can be calculated as:<br />

P<br />

=<br />

λ max<br />

∫<br />

λ min<br />

E<br />

( 0<br />

a ( λ)<br />

, λ) AQY ( λ)<br />

dλ<br />

− CDOM<br />

d , (1)<br />

at<br />

( λ)<br />

where λmin - λmax is the spectral range within which photooxidation is efficient, aCDOM(λ)<br />

and at(λ) are the spectral CDOM and total absorption coefficients, respectively, and<br />

−<br />

E ( 0 , λ)<br />

is the spectral downward irradiance just beneath sea <strong>sur</strong>face (see the list of<br />

d<br />

Notation). In Eq. 1, the ratio of aCDOM(λ) to at(λ) is the fraction of the inci<strong>de</strong>nt light<br />

absorbed by the CDOM in the whole water column. While this ratio approaches 0.9-1.0<br />

in the UV domain in the open ocean [Johannessen, 2000], absorbing partic<strong>les</strong> may<br />

compete for light absorption in coastal environments. Because this is where<br />

87


photooxidation may be particularly important as rivers discharge form buoyant and<br />

shallow plumes that promote the expo<strong>sur</strong>e of CDOM to solar radiation [see Dagg et al.,<br />

2004 and references therein], information on the ratio of aCDOM(λ) to at(λ) is highly<br />

<strong><strong>de</strong>s</strong>irable to improve quantification of photochemical processes.<br />

On a volume basis, the Arctic Ocean receives the highest amount of terrigenous<br />

DOC relative to the world ocean [Rachold et al., 2004]. During the past <strong>de</strong>ca<strong>de</strong>, dramatic<br />

changes have been observed in the Arctic ice climate, ocean circulation, storage of<br />

freshwater, ozone concentration, permafrost thawing and riverine discharge [ACIA, 2005].<br />

Recently, Bélanger et al. [2006] have shown that CDOM photooxidation in the<br />

southeastern Beaufort Sea (western Arctic) could increase significantly in the future in<br />

response to the predicted trend of ongoing reduction of sea ice cover. Because this region<br />

is influenced by the Mackenzie River, which annually supplies enormous amount of<br />

runoff (330 km 3 ), suspen<strong>de</strong>d partic<strong>les</strong> (124 Tg) and DOC (1.3 Tg) [Telang et al., 1991;<br />

Macdonald et al., 1998], optical properties of the <strong>sur</strong>face waters are largely controlled by<br />

terrigenous inputs. The spatial and temporal variability of the optical properties of the<br />

<strong>sur</strong>face waters are, therefore, expected to be high during the spring-summer season when<br />

the photooxidation occurs and the river plume spreads over the continental shelf.<br />

Bélanger et al. [2006] found the ratio [aCDOM/at] in the UV-blue part of the spectrum to<br />

vary between 30 and 95%. In their calculation of photoxidation at the scale of the<br />

Beaufort Sea, they could only use regionally averaged values for [aCDOM/at], and<br />

therefore only partly account for that variability in the optical properties of the <strong>sur</strong>face<br />

waters. This is a limitation of their approach that perhaps impacts on large-scale<br />

photooxidation estimates, and certainly does small ones.<br />

Satellite remote sensing is particularly suited for monitoring CDOM<br />

photooxidation in the Arctic Ocean. Remote observations of sea ice concentration, cloud<br />

cover and total atmospheric ozone content allow the calculation of UV and visible<br />

radiation <strong>flux</strong>es penetrating the water column. For example, satellite-<strong>de</strong>rived UV<br />

radiation has been used to study the impact of the <strong>de</strong>cline in ozone on the oceanic<br />

primary production in Antarctica [Arrigo et al., 2003], and the impact of the <strong>de</strong>cline in<br />

sea ice concentration on the CDOM photooxidation in the Arctic [Bélanger et al., 2006].<br />

Ocean Color remote sensing can provi<strong>de</strong> relevant bio-optical indicators to refine the<br />

88


quantification of CDOM photooxidation. During the last two <strong>de</strong>ca<strong><strong>de</strong>s</strong>, several algorithms<br />

have been proposed to <strong>de</strong>rive inherent optical properties (IOP) from the remote sensing<br />

reflectance spectrum [see review by IOCCG, 2006]. None of them, however, are<br />

specifically <strong><strong>de</strong>s</strong>igned to <strong>de</strong>rive separately the total and CDOM absorption coefficients.<br />

In<strong>de</strong>ed, because CDOM and non-algal partic<strong>les</strong> (NAP) are characterized by similar<br />

exponentially <strong>de</strong>creasing absorption spectrum with increasing wavelength [Bricaud et al.,<br />

1981; Roesler et al., 1989], current algorithms fail to distinguish them and, as a practical<br />

solution, combine them into a unique absorption coefficient referred to as colored <strong>de</strong>trital<br />

material [CDM=CDOM+NAP, see Siegel et al., 2002]. While CDM is dominated by<br />

CDOM at the global scale [~0.817±0.14 at 440 nm; Siegel et al., 2002], the relative<br />

contribution of CDOM to CDM varies within a wi<strong>de</strong> range in coastal waters, i.e. from<br />

0.25 to 0.95 (see below).<br />

In view of mo<strong>de</strong>ling <strong>de</strong>pth-integrated photooxidation in oceanic coastal waters,<br />

our first objective was to document the variability in the ratio [aCDOM/at]. Using an<br />

extensive data set covering various coastal environments, we observed that [aCDOM/at] at<br />

412 nm varies between 0.20 to 0.90, with an average of 0.53 and a standard <strong>de</strong>viation (σ)<br />

of 0.16, which stresses the importance of the variability in [aCDOM/at](412) in coastal<br />

waters. The second objective was to propose an algorithm for the retrieval of [aCDOM/at]<br />

at 412 from the remote sensing reflectance spectrum. In what follows, we first present the<br />

in situ data sets used to <strong>de</strong>velop our empirical algorithm. Next we quantify the sensitivity<br />

of the <strong>de</strong>pth-integrated photoproduction of DIC to 1) the inaccuracy of the retrieved<br />

[aCDOM/at](412), and 2) to the extrapolation of [aCDOM/at] from 412 nm over the whole<br />

spectral range efficient for CDOM photooxidation (300 to 600 nm). Finally, using Sea-<br />

viewing Wi<strong>de</strong> Field-of-view Sensor (SeaWiFS) data, we <strong>de</strong>monstrate the benefit of the<br />

proposed method for the quantification of photooxidation in the Southeastern Beaufort<br />

Sea.<br />

IV.2. Materials and Methods<br />

IV.2.1. Data sets <strong><strong>de</strong>s</strong>cription and mea<strong>sur</strong>ements<br />

Two different data sets collected in various coastal environments are used in this<br />

study: the Coastal Surveillance Through Observation of Ocean Color (COASTlOOC) and<br />

89


the Canadian Arctic Shelf Exchange Study (CASES) data sets. Details on the time period<br />

and sampling locations of COASTlOOC can be found in Babin et al. [2003a; 2003b].<br />

Only the COASTlOOC data collected in coastal (so-called Case 2) waters were retained<br />

for the algorithm <strong>de</strong>velopment and validation: 29 stations in the northern Adriatic Sea; 53<br />

stations in the southwestern Baltic Sea; 68 stations in the North Sea; and 58 stations in<br />

the English Channel along the southern coast of England. Although our prime interest is<br />

photooxidation in the Arctic ocean, we also used the COASTlOOC data to reach more<br />

general conclusions about [aCDOM/at] variability in coastal waters, and to <strong>de</strong>velop a more<br />

robust [aCDOM/at] algorithm.<br />

CASES sampling was conducted in June and July 2004 onboard the CCGS<br />

Amundsen in the southeastern Beaufort Sea (Fig. IV.1). The Amundsen Gulf was first<br />

sampled after the sea ice opening in June and was revisited in late July when most of the<br />

area was free of ice . The <strong>sur</strong>face waters were, therefore, influenced by melt waters at few<br />

stations. The Mackenzie Shelf was sampled during late June and early July when the<br />

Mackenzie River discharge peaks [Macdonald et al., 1998], making those stations typical<br />

Case 2 waters. At 47 stations, our whole set of in situ optical mea<strong>sur</strong>ements was<br />

achieved, while at other stations only IOP mea<strong>sur</strong>ements (52 and 25 spectra of the<br />

particle and CDOM absorption coefficient, respectively) were ma<strong>de</strong> because of either sun<br />

elevation too high for proper apparent optical properties (AOP) mea<strong>sur</strong>ements (sun zenith<br />

angle >70°) or occasional technical problems with instruments (Fig. IV.1).<br />

90


Figure IV.1. Map of the study area showing location of stations where<br />

IOPs (dots), SPMR (triang<strong>les</strong>) and ASD (squares) mea<strong>sur</strong>ements were<br />

ma<strong>de</strong>. Contour lines of sea ice concentration of 50% are also shown for<br />

June 1 st (red), July 1 st (blue) and August 1 st (green).<br />

IV.2.1.1. Inherent optical properties<br />

Details on the IOP mea<strong>sur</strong>ements for COASTlOOC can be found in Babin et al.<br />

[2003a; 2003b]. Similar protocols were adopted during CASES including a few<br />

modifications as <strong><strong>de</strong>s</strong>cribed below. At each station a sample of ~20 L of <strong>sur</strong>face water was<br />

collected with a clean bucket for spectrophotometric analyses. Subsamp<strong>les</strong> for the<br />

<strong>de</strong>termination of aCDOM were filtered through 0.2-μm Anotop® syringe filters (Whatman)<br />

and kept into 100-mL acid-cleaned amber glass bott<strong>les</strong>. For the <strong>de</strong>termination of the<br />

absorption coefficient of partic<strong>les</strong>, ap, suspen<strong>de</strong>d partic<strong>les</strong> were retained onto 25-mm<br />

GF/F glass fiber filters (Whatman; pore size of 0.7-μm) by filtering 0.1 to 3.5 L of<br />

seawater. The glass bott<strong>les</strong> and GF/F filters were stored frozen (seawater: -20 °C; particle:<br />

-80 °C) in the dark until being analyzed two to four months later in the land based<br />

laboratory. Samp<strong>les</strong> treatment and methods applied to <strong>de</strong>termine the ap and aCDOM spectra<br />

are <strong>de</strong>tailed in Bélanger et al. [2006]. Briefly, ap(λ) was <strong>de</strong>termined at 1-nm resolution<br />

between 350 and 750 nm according to the transmittance-reflectance protocol <strong>de</strong>veloped<br />

by Tassan and Ferrari [1995; 2002]. After ap mea<strong>sur</strong>ements, the filters were soaked<br />

during ~30 minutes in 90% methanol to extract phytoplankton pigments [Kishino et al.,<br />

91


1985], and the transmittance-reflectance mea<strong>sur</strong>ements were then repeated for the<br />

<strong>de</strong>termination of non-algal absorption (aNAP). The absorption coefficient by the<br />

phytoplanktonic pigments (aφ) is calculated as ap(λ) – aNAP(λ). The aCDOM(λ) was<br />

mea<strong>sur</strong>ed in 10-cm quartz cuvettes between 250 and 800 nm with 1-nm increments using<br />

a dual beam spectrophotometer (Perkin-Elmer Lambda 35). A baseline correction was<br />

applied by subtracting the absorbance value averaged over an interval of 5 nm around<br />

685 nm from all the spectral values [Babin et al., 2003b]. After applying a baseline<br />

correction to the aCDOM(λ) spectrum, a non-linear least squares curve fitting (Levenberg-<br />

Marquardt) procedure was used to fit an exponential mo<strong>de</strong>l to the data between 300 and<br />

500 nm:<br />

SCDOM<br />

*( λ0<br />

−λ<br />

)<br />

a CDOM ( λ)<br />

= aCDOM<br />

( λ0<br />

) e , (2)<br />

where λ0 is a reference wavelength (e.g. 412 nm) and SCDOM is the spectral slope of the<br />

aCDOM(λ) spectrum.<br />

The spectral absorption and beam attenuation coefficients of seawater constituents<br />

(i.e. excluding pure seawater itself), at-w and ct-w, were <strong>de</strong>termined at nine wavelengths<br />

(412, 440, 488, 510, 532, 555, 650, 676 and 715 nm) using a submersible<br />

spectrophotometer (ac-9, WET Labs Inc.). The ac-9 was either operated onboard the ship<br />

laboratory where ~20 L of <strong>sur</strong>face water were passed through the instrument by gravity,<br />

or <strong>de</strong>ployed from the <strong>de</strong>ck or from a zodiac to obtain vertical profi<strong>les</strong> from <strong>sur</strong>face down<br />

to ~40 m. The manufacturer calibration was checked daily using ~10 L of water purified<br />

onboard (Milli-Q Gradient A10, pre-filtered with a Millipore RiOs 8). Temperature and<br />

salinity corrections were applied using the latest coefficients <strong>de</strong>termined by the<br />

manufacturer [M. Twardowski, pers. comm.]. The ac-9 overestimates absorption<br />

coefficients due to the loss of scattered photons that do not reach the <strong>de</strong>tector [Zaneveld<br />

et al., 1994]. To correct for that error, the following expression was applied:<br />

m<br />

m<br />

a − ( λ)<br />

= a ( λ)<br />

− εb<br />

( λ)<br />

, (3)<br />

t<br />

w<br />

where a m (λ) is the mea<strong>sur</strong>ed absorption coefficient, b m (λ) is the mea<strong>sur</strong>ed scattering<br />

coefficient calculated as the difference ct-w(λ) - a m (λ)), and ε is the fraction of the<br />

scattering coefficient that corresponds to photons not <strong>de</strong>tected by the sensor. We used the<br />

matrix inversion procedure proposed by Gallegos and Neale [2002] to estimate ε. The<br />

92


spectral shapes for aCDOM(λ), aNAP(λ) and aφ(λ), and the relationship between aNAP(440)<br />

and b m (440), which are necessary in the inversion, were <strong>de</strong>termined using mea<strong>sur</strong>ements<br />

on discrete water samp<strong>les</strong> (<strong><strong>de</strong>s</strong>cribed above) following the statistically augmented method<br />

<strong><strong>de</strong>s</strong>cribed by Gallegos and Neale [2002]. When available, the vertical at-w(λ) profi<strong>les</strong><br />

were optically averaged from <strong>sur</strong>face down to the first attenuation length using the<br />

weighting function approach proposed by Gordon [1992].<br />

The ratio [aCDOM/at](412) was computed using the spectrophotometrically<br />

<strong>de</strong>termined IOPs, excepted for seven of the CASES stations where aCDOM(412) was not<br />

available and was computed as at-w(412) – ap(412). A comparison of the two methods<br />

indicates that the [aCDOM/at](412) agree within ±0.11 at 95% confi<strong>de</strong>nce level.<br />

IV.2.1.2. Remote sensing reflectance<br />

During COASTlOOC, irradiance mea<strong>sur</strong>ements were carried out from helicopter<br />

(202) and from ship (6). Vertical profi<strong>les</strong> of the spectral upwelling irradiance, Eu(z, λ),<br />

and the spectral downwelling irradiance just above the sea <strong>sur</strong>face, Es(0 + , λ), were<br />

mea<strong>sur</strong>ed at 13 wavebands (412, 443, 456, 490, 510, 532, 560, 620, 665, 683, 705, 779<br />

and 866 nm) with a Satlantic free-fall SPMR (SeaWiFS Profiling Multichannel<br />

Radiometer) and SMSR (SeaWiFS Multichannel Surface Reference), respectively. Both<br />

radiometers had been calibrated <strong>les</strong>s than 3 months before each cruise by the instrument<br />

manufacturer. Platform (ship or helicopter) shadow was avoi<strong>de</strong>d and, in the specific case<br />

of helicopter, instruments were <strong>de</strong>ployed from an altitu<strong>de</strong> high enough to avoid<br />

perturbation at sea <strong>sur</strong>face created by the rotor air <strong>flux</strong>. After excluding data with an<br />

instrument tilt > 5°, the in-water profi<strong>les</strong> start beyond ca. 2 m when the radiometer was<br />

<strong>de</strong>ployed in free-falling mo<strong>de</strong> from a ship, and beyond ca. 0.2 m when <strong>de</strong>ployed from a<br />

helicopter using a winch. To estimate upward irradiance just below the sea <strong>sur</strong>face (at 0 -<br />

m), Eu(z, λ) profi<strong>les</strong> were extrapolated up to <strong>sur</strong>face by fitting the data to an exponential<br />

function. Eu(0 - , λ) was propagated through water-air interface using a factor of 0.543<br />

[Mueller et al., 2003] to obtained Eu(0 + , λ). The spectral remote sensing reflectance is<br />

calculated as,<br />

93


+<br />

1 Eu<br />

( 0 , λ)<br />

R rs ( λ ) =<br />

, (4a)<br />

+<br />

Q E ( 0 , λ)<br />

where Q factor is the ratio between Eu and Lu. Here a Q value of 3.8 was used, which falls<br />

within calculated range for turbid coastal waters [Loisel and Morel, 2001].<br />

94<br />

s<br />

During CASES, the remote sensing reflectance at 13 wavebands (405, 412, 434,<br />

442, 490, 510, 520, 532, 555, 590, 665, 683 and 700 nm), Rrs(λ), was calculated from the<br />

vertical profile of Lu(z, λ), and Es(0 + , λ) mea<strong>sur</strong>ed with a SPMR and SMSR, respectively.<br />

At each station, three to five SPMR casts were performed at least 50 m away from the<br />

ship. The upwelling radiances were extrapolated to the sub-<strong>sur</strong>face, Lu(0 - , λ), using a<br />

linear fit to all ln[Lu(z, λ)] data points mea<strong>sur</strong>ed within a <strong>de</strong>pth interval of 1-2 m within<br />

the <strong>sur</strong>face layer, and propagated through water-air interface to obtained Lw(0+, λ). Note<br />

that both Eu(0 - , λ) and Lu(0 - , λ) were corrected for instrument self-shading following the<br />

method proposed by Gordon and Ding [1992] as applied by Zibordi and Ferrari [1995]<br />

using at-w(λ) mea<strong>sur</strong>ements ma<strong>de</strong> with the ac-9 and the aw(λ) values published by Pope<br />

and Fry [1997]. The spectral remote sensing reflectance is calculated as,<br />

+<br />

Lw(<br />

0 , λ)<br />

R rs ( λ ) = .<br />

(4b)<br />

+<br />

E ( 0 , λ)<br />

s<br />

For each station, we averaged Rrs(λ) obtained from different casts (2 to 4) after<br />

elimination of the spectra different from the mean value by more than 10% in the blue.<br />

Out of 40 stations with SPMR mea<strong>sur</strong>ements, 34 Rrs(λ) were retained for the algorithm<br />

<strong>de</strong>velopment (section 2.2). Note that four spectra were eliminated because the difference<br />

among all replicates was > 10%, and two because of the <strong>sur</strong>face waters stratification.<br />

In shallow water areas or when <strong>sur</strong>face waters were stratified (i.e. 14 stations), the<br />

Rrs(λ) was mea<strong>sur</strong>ed above the sea <strong>sur</strong>face using a hyperspectral radiometer (336 to 1062<br />

nm with Δλ = 1.42 nm; Analytical Spectral Device, ASD). To make mea<strong>sur</strong>ements away<br />

from the ship shadow and bubble clouds, the instrument was mounted at the end of the<br />

ship bow and <strong>de</strong>ployed either on the port or starboard si<strong>de</strong> <strong>de</strong>pending of the sun position<br />

and the sea state. The mea<strong>sur</strong>ements were ma<strong>de</strong> only when inci<strong>de</strong>nt irradiance conditions<br />

were stable (clear or uniformly cloudy skies). To avoid calibration issues that generally<br />

result from the use of several sensors, the mea<strong>sur</strong>ements of the radiance (acceptance


angle of 10°) coming from sea <strong>sur</strong>face (Lt), downwelling sky (Lsky), and a lambertian<br />

reflector (grey Spectralon) (Lp) were performed successively, within a few minutes, using<br />

an unique sensor. To obtain one Rrs spectrum, five to ten scans of each quantity were<br />

quality checked and averaged. Note that the ASD automatically adjusts the integration<br />

time of the mea<strong>sur</strong>ements to optimize the signal-to-noise ratio (1.09 to 4.35 s for Lt, and <<br />

1 s for Lsky and Lp). The above-water remote sensing reflectance is calculated as [Mobley,<br />

1999]:<br />

R<br />

rs<br />

( Lt<br />

( λ)<br />

− ρ sky ⋅ Lsky<br />

( λ))<br />

⋅ R p<br />

( λ)<br />

= , (5)<br />

πL<br />

( λ)<br />

where ρsky is the air-water interface specular reflection coefficient for radiance, and Rp is<br />

the albedo of the lambertian panel (88.3%). The value of Rp, spectrally constant in the<br />

visible and near infrared parts of the spectrum, was <strong>de</strong>termined using a calibrated<br />

lambertian panel (P. Minnett, U. of Miami; calibration performed on 4 th of May 2004 by<br />

Avian Technologies LLC inc.). As recommen<strong>de</strong>d by Mobley [1999], Lt was mea<strong>sur</strong>ed<br />

with a viewing zenith angle of ~40° and a relative azimuth angle of ~135°. For this<br />

viewing configuration and for a solar zenith angle >45°, i.e. the conditions encountered<br />

during this study, ρsky <strong>de</strong>pends only on wind speed (Figure 9 in Mobley, 1999). ρsky varies<br />

from ~0.0256 to ~0.038 for wind speed of 0 and 15 m s -1 , respectively, un<strong>de</strong>r clear sky<br />

condition [Mobley, 1999]. For cloudy sky condition, a value 0.0256 was adopted for ρsky.<br />

An additional spectrally neutral correction was applied to Rrs(λ) based on the known<br />

spectral shape of Rrs in the near infrared (NIR) part of the spectrum [Ruddick et al., 2006].<br />

This correction aims to correct the Rrs spectra for uncertainty due to the air-sea reflection<br />

correction (e.g. glitter) and the instrument-related errors (e.g. integration time, sky<br />

radiance distribution) [for <strong>de</strong>tails see web appendix 2 in Ruddick et al., 2006]. The<br />

correction assumes that the ratio of water leaving reflectance between two wavelengths in<br />

the NIR is known (here 780 and 870 nm were used), allowing the computation of a<br />

residual error that is then subtracted to Rrs(λ).<br />

95<br />

p


IV.2.2. Description of the [aCDOM/at] algorithm<br />

To improve the quantification of photooxidation (Eq. 1), we need to know the<br />

spectral contribution of aCDOM to the total light absorption coefficient of seawater. As<br />

mentioned in the Introduction, because of the similar spectral absorption spectra of<br />

CDOM and NAP, semi-analytical algorithms currently found in the literature fail to<br />

discriminate them. Here we propose an empirical method to obtain the ratio [aCDOM/at] at<br />

412 nm, directly from the Rrs spectrum. The [aCDOM/at](412) was regressed against<br />

several combinations of Rrs and ratios of Rrs selected according to the following rationa<strong>les</strong>:<br />

in coastal waters, 1) reflectance at 412 nm is the most affected by CDOM absorption<br />

compared with other channels, 2) reflectance at 490 nm is the most affected by<br />

phytoplankton absorption, 3) variations in reflectance at 555 nm is most driven by light<br />

scattering by partic<strong>les</strong>, and 4) a major distinctive property of CDOM is that it does not<br />

contribute to particle scattering. These consi<strong>de</strong>rations led us to propose the following<br />

empirical algorithm for the retrieval of [aCDOM/at], based on multiple regression<br />

conducted using mea<strong>sur</strong>ed [aCDOM/at](412) and Rrs(λ):<br />

⎡a<br />

⎢<br />

⎣ at<br />

⎤<br />

⎛ R ( 412)<br />

⎞ ⎛<br />

rs<br />

R ( 490)<br />

⎞<br />

rs<br />

⎥(<br />

412)<br />

= α + β ⋅log10<br />

⎜ + ⋅log10<br />

+ ⋅log10<br />

( rs ( 555)<br />

)<br />

Rrs<br />

( 555)<br />

⎟ χ ⎜<br />

Rrs<br />

( 555)<br />

⎟ δ<br />

, (6)<br />

⎦<br />

⎝ ⎠ ⎝ ⎠<br />

CDOM R<br />

where α, β, χ and δ are empirical coefficients. When CDOM increases, for instance, the<br />

ratio Rrs(412)/Rrs(555) <strong>de</strong>creases and [aCDOM/at](412) tends to increase (negative slope for<br />

β). In contrast, when phytoplankton pigments increase, the ratio Rrs(490)/Rrs(555)<br />

<strong>de</strong>creases and [aCDOM/at](412) tends to <strong>de</strong>crease (positive slope for χ). Finally, when the<br />

total amount of suspen<strong>de</strong>d partic<strong>les</strong> increases, Rrs(555) increases and [aCDOM/at](412)<br />

tends to <strong>de</strong>crease (negative slope for δ).<br />

IV.3. Results and Discussion<br />

IV.3.1. Variability of [aCDOM/at] and [aCDOM/aCDM] in coastal waters<br />

Figure IV.2a illustrates the observed variability in the ratio [aCDOM/at] in the violet<br />

part of the spectrum in various coastal environments including the Adriatic Sea, Baltic<br />

Sea, English Channel, Gulf of Lyon, North Sea, and Beaufort Sea. Overall, [aCDOM/at] at<br />

412 nm varies between 0.2 and 0.9, with an average of 0.53 and a standard <strong>de</strong>viation (σ)<br />

96


of 0.16, which confirms the large variability of [aCDOM/at](412) in coastal waters. The<br />

frequency distribution is slightly flat compared with a normal distribution (negative<br />

kurtosis).<br />

The contribution of CDOM to the total dissolved and <strong>de</strong>trital material (CDM) also<br />

varies over a wi<strong>de</strong> range in coastal waters, from 0.26 to 0.95 (Fig. IV.2b; Table IV.1),<br />

even at regional scale (e.g. North Sea, 0.26 to 0.91). For comparison, Siegel et al. [2002]<br />

reported a mean value for the ratio [aCDOM/aCDM] of 0.817 at 440 nm (with a σ of 0.13)<br />

from a global collection of absorption mea<strong>sur</strong>ements of aCDOM and aNAP. This is<br />

significantly higher than most values found in coastal waters influenced by river runoff or<br />

sediment resuspension. The GSM01 semi-analytical ocean color mo<strong>de</strong>l [Garver and<br />

Siegel, 1997; Maritorena et al., 2002] is currently implemented in the SeaWiFS data<br />

processing chain to retrieve aCDM(440). The use of this algorithm is of limited utility for<br />

quantifying CDOM photooxidation in coastal waters because it provi<strong><strong>de</strong>s</strong> no information<br />

either on the ratio [aCDOM/aCDM] nor [aCDOM/at]. Since coastal zones may be major sites<br />

for terrigenous CDOM photooxidation [Kieber et al., 1990; Miller and Zepp, 1995;<br />

Benner and Opsahl, 2001], the following method is specifically <strong><strong>de</strong>s</strong>igned to estimate<br />

[aCDOM/at] in turbid coastal waters.<br />

Table IV.1. Relative contribution of CDOM to the colored <strong>de</strong>trital matter<br />

(CDM=CDOM+NAP) and to the total absorption coefficients at 412 nm, respectively, for<br />

the different regions covered by our data set.<br />

[ a CDOM / aCDM<br />

] (412) [ a CDOM / at<br />

] (412)<br />

Area N Min. Max. Average Min. Max. Average<br />

± σ<br />

± σ<br />

Adriatic 29 .38 .87 .63 ± .12 .22 .66 .40 ± .09<br />

Baltic 53 .60 .84 .76 ± .05 .38 .72 .60 ± .06<br />

Eng. Channel 58 .30 .91 .72 ± .13 .24 .69 .50 ± .10<br />

North Sea 68 .26 .91 .61 ± .16 .20 .81 .49 ± .13<br />

Beaufort Sea 47 .41 .95 .82 ± .13 .39 .94 .75 ±.13<br />

All 255 .26 .95 .71 ± .14 .20 .94 .55 ± .15<br />

97


Figure IV.2. Frequency distributions of the contribution of CDOM to (a)<br />

the total light absorption at 412 nm ([aCDOM/at]), and (b) to the colored<br />

<strong>de</strong>trital material ([aCDOM/aCDM]). Note that the graph inclu<strong><strong>de</strong>s</strong> 54 additional<br />

COASTlOOC or CASES stations where only IOP were available.<br />

IV.3.2. Retrieval of [aCDOM/at] from remote sensing reflectance<br />

Table IV.2 provi<strong><strong>de</strong>s</strong> the multiple regression results for the empirical coefficients<br />

of Eq. 6 and the <strong>de</strong>termination coefficients (R 2 ) obtained for the whole dataset (global<br />

algorithm), and separately for the five different regions of the data set (region-specific<br />

98


algorithms). Figure IV.3 compares the [aCDOM/at] at 412 nm retrieved using the global<br />

algorithm with the mea<strong>sur</strong>ed values.<br />

Table IV.2. Empirical coefficients of Eq. 6 <strong>de</strong>termined by multiple regression for the<br />

different coastal environments.<br />

Area N α β χ δ R 2<br />

Adriatic 29 -.015 -.321 .691 -.223 .46<br />

Baltic 53 .078 -.133 .674 -.280 .81<br />

Eng. Channel 58 -.048 -.423 .539 -.204 .33<br />

North Sea 68 -.480 -.255 .526 -.483 .65<br />

Beaufort Sea 47 -.514 -.546 .480 -.454 .42<br />

All 255 -.387 -.387 .577 -.390 .70<br />

Figure IV.3. Comparison between retrieved and mea<strong>sur</strong>ed [aCDOM/at] at<br />

412 nm for the COASTlOOC and CASES datasets. [aCDOM/at](412) was<br />

calculated using equation 6 with coefficients obtained using the whole<br />

data set (N=255; Table IV.2).<br />

To evaluate rigorously the ability of our global algorithm to retrieve<br />

[aCDOM/at](412) in a given region, we conducted the following validation exercise: the<br />

empirical coefficients of Eq. 6 for a given region were also <strong>de</strong>rived using the data from<br />

the four other regions. The algorithm could thereby be tested with a data set different<br />

99


from that used for algorithm tuning. For the Beaufort Sea, for example, the coefficients of<br />

Eq. 6 were <strong>de</strong>rived using the whole COASTlOOC data set (i.e. Adriatic Sea, Baltic Sea,<br />

English Channel and North Sea). Note that these coefficients are not provi<strong>de</strong>d as they<br />

were <strong>de</strong>rived for validation purpose only. Table IV.3 provi<strong><strong>de</strong>s</strong> the results from a<br />

comparison between retrieved and mea<strong>sur</strong>ed [aCDOM/at] for each region, achieved using<br />

Type-II regression. The algorithm captures part of the variability within each region as<br />

the regression slopes are all significantly positive, (p < 0.05), although they are<br />

significantly < 1 (p < 0.05), except for the English Channel. This result indicates that<br />

further tuning is necessary to get the variability of [aCDOM/at](412) right within a given<br />

region. Neverthe<strong>les</strong>s, the overall absolute uncertainty of the global algorithm with the<br />

coefficient <strong>de</strong>rived in<strong>de</strong>pen<strong>de</strong>ntly is ±0.18 at the 95% confi<strong>de</strong>nce interval (Table IV.3).<br />

Despite the relatively high heterogeneity among the five coastal areas consi<strong>de</strong>red here,<br />

the algorithm is able to retrieve the average [aCDOM/at](412) value for each region<br />

(Δ[aCDOM/at]±0.04), except for the Beaufort Sea where the estimation is slightly biased (-<br />

0.078).<br />

Table IV.3. Performance of Eq. 6 for the retrieval of [aCDOM/at] at 412 nm. For each<br />

region, in<strong>de</strong>pen<strong>de</strong>nt data set is used to obtain the empirical coefficients for Eq. 6.<br />

Area N Intercept a<br />

Slope a<br />

Δ[aCDOM/at] b CI 95%c<br />

Adriatic 29 .22 [.11, .31] .44 [.21, .71] 0.001 ±0.13<br />

Baltic 53 .28 [.20, .34] .54 [.43, .67] 0.000 ±0.07<br />

Eng. Channel 58 .00 [-.30,.17] 1.06 [.69, 1.64] 0.039 ±0.20<br />

North Sea 68 .27[.22, .32] .45 [.36, .54] 0.013 ±0.17<br />

Beaufort Sea 47 .30 [.13, .46] .56 [.36, .80] -.078 ±0.20<br />

All d 255 .16 [.12, .20] .70 [.63, .76] -.002 ±0.18<br />

a<br />

Numbers in brackets are for the range of 95 % Confi<strong>de</strong>nce Interval computed for a Type II<br />

regression.<br />

b<br />

[ ] [ ] [ ] mea<strong>sur</strong>ed<br />

1 n<br />

retrieved<br />

.<br />

Δ a CDOM / at<br />

= ∑ aCDOM<br />

/ at<br />

− aCDOM<br />

/ at<br />

n<br />

i<br />

a CDOM / at<br />

retrieval.<br />

d<br />

These results were obtained by comparing the [aCDOM/at] values retrieved within each region<br />

using the global algorithm <strong>de</strong>rived in<strong>de</strong>pen<strong>de</strong>ntly, with the mea<strong>sur</strong>ed values (see text).<br />

c 95% Confi<strong>de</strong>nce Interval of the [ ]<br />

In an attempt to un<strong>de</strong>rstand the above results, we further analyzed the absolute<br />

difference between the retrieved and mea<strong>sur</strong>ed [aCDOM/at](412) values (Δ[aCDOM/at]).<br />

Again for each region, [aCDOM/at](412) was retrieved using Eq. 6 with the coefficients<br />

<strong>de</strong>rived in<strong>de</strong>pen<strong>de</strong>ntly. Figure IV.4a compares Δ[aCDOM/at] with the ratio between the<br />

100


particulate absorption at 412 nm and the remote sensing reflectance at 555 nm<br />

([ap(412)/Rrs(555)]). Positive Δ[aCDOM/at] values indicate an overestimation of<br />

[aCDOM/at](412) by the empirical algorithm. Low values of [ap(412)/Rrs(555)] indicate<br />

that the partic<strong>les</strong> are weakly absorbing blue photons relative to the their scattering<br />

properties in the green, while high values indicate the contrary. The former may indicate<br />

the presence of mineral partic<strong>les</strong>, and the latter the presence of phytoplankton and/or<br />

<strong>de</strong>tritus [Gallegos and Neale, 2002]. Since the absolute value of Rrs(555) is used in the<br />

algorithm as an indicator of the total amount of partic<strong>les</strong>, variability in the<br />

[ap(412)/Rrs(555)] will affect the performance of the algorithm. Statistically significant<br />

positive slopes of the regression (Type II) between Δ[aCDOM/at] and the logarithm of<br />

[ap(412)/Rrs(555)] were observed in the Adriatic Sea, Baltic Sea, English Channel and the<br />

North Sea (Table IV.4). These results suggest that the [aCDOM/at](412) value is, as argued<br />

above, un<strong>de</strong>restimated in presence of weakly absorbing partic<strong>les</strong>, while it is<br />

overestimated in presence of highly absorbing ones. This is because the enhance of<br />

Rrs(555) in presence of weakly absorbing partic<strong>les</strong> results, for instance, in an<br />

overestimation of the contribution of partic<strong>les</strong> to the total light absorption at 412 nm.<br />

Based on these observations, one can ask why in the Baltic Sea, where partic<strong>les</strong> are<br />

predominantly organic (87%; Babin et al. [2003b]) and highly absorbing (Fig. IV.4a), the<br />

[aCDOM/at](412) is not systematically overestimated? The answer can be found in Fig.<br />

IV.4b.<br />

Table VI.4. Coefficient of <strong>de</strong>termination of the linear regression between Δ[aCDOM/at]<br />

(shown in Fig. IV.4) and 1) the logarithm of the ratio [ap(412)/Rrs(555)] and 2) the ratio<br />

[ a CDOM / aCDM<br />

] at 412 nm. NS = the slope is not significantly different from 0 at 95%<br />

confi<strong>de</strong>nce level (Type II regression).<br />

a / (412) a<br />

Area N [ap(412)/Rrs(555)] a [ ]<br />

101<br />

CDOM aCDM<br />

Adriatic 29 .23 (NS) .18 (.17)<br />

Baltic 53 .51 (NS) .34 (NS)<br />

Eng. Channel 58 .46 (.14) NS (.23)<br />

North Sea 68 .46 (.33) .58 (.25)<br />

Beaufort Sea 47 NS (.21) .22 (.27)<br />

All 255 .15 (.08) .21 (.14)<br />

a Values in parenthesis are for the regionally tuned algorithm, while the bold numbers indicate the<br />

parameter that explains most the error for a given region.


Figure IV.4b plots the Δ[aCDOM/at] as a function of the ratio between aCDOM to<br />

aCDM at 412 nm. Statistically significant negative slopes of the linear regression (Type II)<br />

between Δ[aCDOM/at] and [aCDOM/aCDM](412) were observed in the Adriatic Sea, Baltic<br />

Sea, North Sea and Beaufort Sea (Table IV.4). Altogether, the R 2 of the regression<br />

reaches 21%. Generally, when NAP dominates the CDM the empirical algorithm tends to<br />

overestimate the contribution CDOM to the total light absorption. This is because the<br />

ratio of Rrs(412)/Rrs(555) used in the algorithm is an indicator of the total CDM rather<br />

than just the CDOM. The fact that relatively high [aCDOM/aCDM](412) is found in the<br />

Baltic Sea partly explains why we do not observed a systematic overestimation of<br />

[aCDOM/at](412). In the Beaufort Sea, the systematic un<strong>de</strong>restimation of [aCDOM/at](412)<br />

by the empirical algorithm (-0.078; Table IV.3) is probably due to the dominance of<br />

CDOM to CDM (Table IV.1).<br />

102


Figure IV.4. Error analysis of [aCDOM/at] at 412, calculated using regionin<strong>de</strong>pen<strong>de</strong>nt<br />

coefficients, as a function of: a) the ratio of ap(412) to<br />

Rrs(555); b) the relative contribution of CDOM to the total CDM<br />

absorption coefficient at 412 nm. Only the statistically significant<br />

relationships are shown. The color coding is the same as Fig. IV.3.<br />

The trends <strong><strong>de</strong>s</strong>cribed above can be attenuated if the algorithm is tuned with the<br />

region-specific data. Figure IV.5 compares the retrieved and mea<strong>sur</strong>ed [aCDOM/at](412)<br />

values when the coefficients of Eq. 6 are <strong>de</strong>rived for each region individually (Table<br />

IV.2). For the Baltic Sea and North Sea, the regional tuning leads to R 2 > 0.65. In the<br />

103


North Sea, in particular, where the composition of CDM is highly variable (%CDOM =<br />

26 to 91%; Table IV.1), the algorithm is able to discriminate the CDOM from the total<br />

absorption within a CI 95% of 0.13. In the Baltic Sea, where the optical properties of<br />

partic<strong>les</strong> are relatively homogeneous (Fig. IV.4), the algorithm performance is excellent<br />

(R 2 = 0.81, CI 95% = ±0.05). For the English Channel, in contrast, the algorithm poorly<br />

performs as indicated by the low R 2 of 0.33. For the Adriatic Sea and Beaufort Sea,<br />

because most of the mea<strong>sur</strong>ed [aCDOM/at](412) values felt within a narrow range, it is<br />

difficult to draw clear conclusions from this analysis. The absolute uncertainty of the<br />

region-specific algorithm is ±0.14 at the 95% confi<strong>de</strong>nce interval.<br />

Figure IV.5. Same as Fig. IV.3 but for region specific coefficients for Eq.<br />

6 (Table IV.2).<br />

In general, an algorithm based on the ratios Rrs(412)/Rrs(555) and<br />

Rrs(490)/Rrs(555), and on Rrs(555) seems suitable to discriminate aCDOM from at at 412 nm<br />

with good accuracy. We also applied this combination to the synthetic data set <strong>de</strong>veloped<br />

by the International Ocean Color Coordinating Group to test and compare algorithms<br />

(available at: http://www.ioccg.org/groups/OCAG_data.html). The performance of the<br />

algorithm was similar to that found with our in situ data set (R 2 =0.65, CI 95% ±0.15), but<br />

the β and χ coefficients of Eq. 6 were quite different (-.385, -1.105, 1.33 and -.342 for<br />

104


α, β, χ and δ respectively). This difference may be ascribed to the wi<strong>de</strong> range of IOPs<br />

covered by the synthetic data set, and in particular to the highly variable ratio between the<br />

scattering and absorption coefficients of NAP (more than one or<strong>de</strong>r of magnitu<strong>de</strong>). In<br />

addition, the synthetic data set inclu<strong><strong>de</strong>s</strong> clear waters for which the pure water absorption<br />

is no longer negligible at 412 nm (up to 33% of at). This fourth optically active<br />

component complicates the estimation of [aCDOM/at](412) with the proposed algorithm.<br />

Our method was also tested on the NASA bio-Optical Marine Algorithm Data set<br />

(NOMAD) [Wer<strong>de</strong>ll and Bailey, 2005]. The algorithm did not perform as well as on our<br />

in situ data set or on the IOCCG data set (R 2 =0.30, CI 95% ±0.22). Two possible<br />

explanations for this poor result is 1) the relatively low influence of terrestrial inputs in<br />

most of the NOMAD dataset compare to our dataset, and 2) the lack of consistency<br />

among the different sub-datasets that form the NOMAD dataset, in terms of methods to<br />

mea<strong>sur</strong>e the absorption coefficients adopted by different investigator. For some specific<br />

sub-datasets found in NOMAD, however, results similar to the one presented above were<br />

obtained (e.g., R 2 of 0.81 and CI 95% of 0.14 for the ONR-Chesapeake program). These<br />

results suggest that our algorithm works properly only in coastal waters influenced by<br />

terrestrial inputs, and that more data mea<strong>sur</strong>ed following an unique protocols, for the<br />

absorption in particular, are necessary to validate the algorithm.<br />

3.3. Sensitivity of the <strong>de</strong>pth-integrated photooxidation mo<strong>de</strong>l to errors on [aCDOM/at](λ)<br />

In this section, we assess the relative error ma<strong>de</strong> on the <strong>de</strong>pth-integrated<br />

production of dissolved inorganic carbon (PDIC) resulting from the error on the<br />

[aCDOM/at](412) estimate, and from the uncertainty in its extrapolation over the 300-600<br />

nm range. As mentioned in the Introduction, DIC is the main product of CDOM<br />

photooxidation. For our purpose, the sub-<strong>sur</strong>face spectral irradiance, Ed (0 - , λ), was<br />

calculated using the Tropospheric Ultraviolet Visible (TUV) mo<strong>de</strong>l [Madronich and<br />

Flocke, 1999] for the summer solstice, at latitu<strong>de</strong> of 70°N, un<strong>de</strong>r clear sky, mo<strong>de</strong>rate<br />

aerosols concentration (aerosol optical thickness at 550 nm of 0.1), and total ozone<br />

column content of 330 DU. Because the magnitu<strong>de</strong> and spectral shape of apparent<br />

quantum yield for DIC photoproduction vary wi<strong>de</strong>ly between coastal and open ocean<br />

[Johannessen and Miller, 2001; Bélanger et al., 2006], the calculations were ma<strong>de</strong> for<br />

105


two contrasting AQY spectra mea<strong>sur</strong>ed in the Beaufort Sea (Fig. IV.6). The two AQY<br />

spectra show marked differences in shape, with the relative contribution of longer<br />

wavelengths being more important for AQY1.<br />

Wavelength (nm)<br />

Figure IV.6. Apparent Quantum Yield spectra for DIC production<br />

<strong>de</strong>termined for two stations located in the southeastern Beaufort Sea.<br />

AQY1 was <strong>de</strong>termined on <strong>sur</strong>face waters influenced the by the Mackenzie<br />

River (salinity = 8.2; Station R5a), while AQY2 was <strong>de</strong>termined on<br />

<strong>sur</strong>face waters collected in the Amundsen Gulf, away from direct riverine<br />

influence (salinity = 30.0; Station 108) [Bélanger et al., 2006].<br />

Figure IV.7 shows the spectral PDIC(λ) calculated using Eq. 1 with different<br />

values of [aCDOM/at](412). While the maximum PDIC occurs in the spectral range from<br />

325 to 335 nm for both AQYs, the relative contribution of the visible radiation to DIC<br />

photoproduction is different. The <strong>de</strong>crease in [aCDOM/at](412) has a greater impact on<br />

PDIC(λ) when using AQY1 than when using AQY2 (Fig. IV.7). This is because<br />

[aCDOM/at](λ) variability is more pronounced in the visible relative to the UV. So, the<br />

magnitu<strong>de</strong> of the error on PDIC resulting from the uncertainty on [aCDOM/at](412), <strong>de</strong>pends<br />

on the spectral shape of the AQY. To illustrate that, Fig. IV.8 shows the relative error in<br />

spectrally integrated DIC photoproduction (ΔPDIC=100*[PDIC-PDIC true ]/ PDIC true ) as a<br />

106


function of Δ[aCDOM/at](412). If the true [aCDOM/at](412) value of 0.7 is overestimated (or<br />

un<strong>de</strong>restimated) by 0.18, the overall uncertainty of our algorithm, the ΔPDIC reaches ~±24<br />

and ~±18% for AQY1 and AQY2, respectively. When the initial [aCDOM/at](412) value is<br />

0.3, ΔPDIC can reach as much as ~±40%. Overall, PDIC can be predicted with an<br />

uncertainty of 50% (Fig. IV.8). Note that ignoring the impact of partic<strong>les</strong> by using<br />

constant [aCDOM/at](λ) values of one, e.g. as in Johannessen [2000], can lead to an error<br />

in PDIC of up to 3-fold.<br />

Wavelength (nm)<br />

Figure IV.7. Spectral production of DIC for different values of<br />

[aCDOM/at](412) and for A) AQY1 and B) AQY2.<br />

107


Figure IV.8. Relative difference in DIC production as a function of the<br />

absolute difference between the retrieved and true [aCDOM/at](412) values<br />

for the two AQYs and two initial values for [aCDOM/at] true (412) of 0.3 and<br />

0.7. The thick line at the bottom illustrate the 95% confi<strong>de</strong>nce interval in<br />

the [aCDOM/at] retrieval using Eq 6.<br />

The above error analysis assumes that a perfect spectral extrapolation of [aCDOM/at]<br />

from 600 nm to the UV domain is achieved. Here we conduct a sensitivity study of<br />

spectral extrapolation of [aCDOM/at] on the <strong>de</strong>pth-integrated DIC photoproduction. The<br />

extrapolation is achieved as follows:<br />

where<br />

and<br />

a<br />

CDOM<br />

a<br />

t<br />

aCDOM<br />

( λ)<br />

( λ)<br />

= , (7)<br />

a ( λ)<br />

+ a ( λ)<br />

+ a ( λ)<br />

CDOM<br />

p<br />

{ [ a / a ] ( 412)<br />

* a ( 412)<br />

} exp[<br />

S * ( 412 ) ]<br />

a ( λ) − λ<br />

(7a)<br />

CDOM<br />

= CDOM t<br />

t<br />

CDOM<br />

N<br />

ap(λ) = [ at(412) − aCDOM (412) − aw(412) ]* ap (λ). (7b)<br />

In Eq. 7, aw(λ) is the pure water absorption spectrum, SCDOM is the slope of the aCDOM<br />

N<br />

spectrum, and ap (λ) is the particulate absorption spectra normalized at 412 nm<br />

N<br />

( ap (λ) = ap (λ) /a p(412)). As recommen<strong>de</strong>d by Fry [2000], aw(λ) values were taken<br />

108<br />

w


from Quicken<strong>de</strong>n and Irvin [1980] for 300 nm ≤ λ ≤ 320 nm, Pope and Fry [1997] for<br />

380 nm ≤ λ ≤ 600 nm, and were extrapolated for the λ between those ranges. The<br />

N<br />

sensitivity study is conducted for various values for SCDOM, ap (λ) and at(412). These<br />

parameters are required, in addition to [aCDOM/at](412), to operate Eq. 7.<br />

Figure IV.9 shows the variability SCDOM as a function of aCDOM(412) in the<br />

Beaufort Sea. SCDOM ranged from 0.0177 to 0.0303 nm -1 with an average value of 0.022<br />

nm -1 , and a standard <strong>de</strong>viation of 0.002 nm -1 . The SCDOM values observed over the<br />

Mackenzie shelf (~0.018 nm -1 ) were within the range reported by Guéguen et al. [2005]<br />

for this area, and similar to others reported for coastal environments [e.g., Blough and<br />

Del Vecchio, 2002; Babin et al., 2003b]. Offshore, near the Arctic packed ice, SCDOM<br />

reached 0.025 to 0.030 nm -1 , which is higher than the values previously reported for the<br />

Canada Basin [Guéguen et al., 2005], but within the upper range found in the literature<br />

for oligotrophic seawaters [Blough and Del Vecchio, 2002]. Neverthe<strong>les</strong>s, the high SCDOM<br />

may also be due the freezing of our sample during the storage, which is known to affect<br />

to some extent the CDOM absorption spectra (C. Belzile, personal communication, 2005).<br />

Furthermore, when aCDOM is low, the spectral range choosen (e.g. 300 to 500 nm) to<br />

calculate SCDOM using a single exponential greatly affects the results [see Twardowski et<br />

al., 2004]. Interestingly, the relationship we observed between SCDOM and aCDOM(412) is<br />

very similar to the one reported for the Baltic Sea by Kowalczuk et al. [2006] (see their<br />

Figure 4). The sensitivity analysis of the UV extrapolation is conducted using 0.018 and<br />

0.026 nm -1 as the lower and upper values (corresponding to the 95% confi<strong>de</strong>nce interval).<br />

Figure IV.10 shows the 99 ap(λ) spectra normalized to the value at 412 nm<br />

mea<strong>sur</strong>ed in the <strong>sur</strong>face waters during the CASES field campaign. Except for a few<br />

number of spectra where the presence of phytoplankton pigments was dominant (with a<br />

ap(λ) peak at ~440 nm), most spectra show an exponential increase toward shorter<br />

wavelengths, which indicates the dominance of the aNAP in the total particulate matter<br />

N<br />

absorption (62 ± 18% at 443 nm). The average ap (λ) values ± 1.96*σ (i.e. a 95%<br />

confi<strong>de</strong>nce interval) was chosen to represent the limits of the ap spectrum for low and<br />

high spectral <strong>de</strong>pen<strong>de</strong>ncy, respectively.<br />

109


Figure IV.9. Variability of SCDOM as a function of the aCDOM(412) value<br />

mea<strong>sur</strong>ed in the <strong>sur</strong>face waters of the southeastern Beaufort Sea during<br />

summer 2004 (n=65). A second-or<strong>de</strong>r polynomial fit between the SCDOM<br />

values and the logarithm of aCDOM(412) provi<strong>de</strong>d a strong <strong><strong>de</strong>s</strong>cription of<br />

the observations (r 2 =0.79; n=65):<br />

2<br />

S CDOM = 0. 0186 − 0.<br />

0019log[<br />

aCDOM<br />

( 412)<br />

] + 0.<br />

0031log[<br />

aCDOM<br />

( 412)]<br />

.<br />

Figure IV.10. Spectral absorption by particulate matter normalized to 412<br />

nm value mea<strong>sur</strong>ed in the <strong>sur</strong>face waters of the southeastern Beaufort Sea<br />

during summer 2004 (thin grey lines). The thick black line represents the<br />

N<br />

averaged ap (λ) values. The sensitivity the PDIC (eq 1) to the extrapolation<br />

to the UV domain is explored within the limits given by 95% probability<br />

of the normal distribution (i.e. ±1.96*σ; dashed lines). The ap(λ) values<br />

for λ


If we assume that the pure water absorption coefficient is negligible over the 300<br />

to 600 nm range, the spectral extrapolation of [aCDOM/at](412) can be achieve without the<br />

N<br />

absolute value of at(412) by using SCDOM and ap (λ) only. In the Beaufort Sea, the total<br />

absorption coefficient at 412 nm mea<strong>sur</strong>ed in the <strong>sur</strong>face waters varied between 0.048<br />

and 2.7 m -1 with an averaged value of 0.41 m -1 and a standard <strong>de</strong>viation of 0.61 m -1 . So,<br />

these values are used to examine the impact of at(412) on the spectral extrapolation and<br />

PDIC. The null aw(λ) assumption can lead to error > 25% when at(412) is low and the<br />

[aCDOM/at](412) is high (Table IV.5). In most situations, however, the error on PDIC ma<strong>de</strong><br />

using the null aw(λ) assumption is < 10%. In turbid coastal waters where at(412) is larger<br />

than ~0.2 m -1 , aw(λ) could be neglected without losing to much accuracy in PDIC.<br />

N<br />

The variability in the SCDOM and ap (λ) can also lead to some error in PDIC. For<br />

extreme cases, i.e. when [aCDOM/at](412) is low and the spectral <strong>de</strong>pen<strong>de</strong>ncy of both<br />

aCDOM(λ) and ap(λ) are opposite, the ΔPDIC reaches ~10% and ~18% for AQY1 and AQY2<br />

respectively (Table IV.6). Figure IV.11 shows several examp<strong>les</strong> of PDIC(λ) spectra<br />

calculated for relatively high proportions of particulate matter to the total absorption, i.e.<br />

[aCDOM/at](412) of 0.3. Differences in the PDIC(λ) spectra are notable for both AQYs<br />

(note the different y-axis sca<strong>les</strong>). As expected, the sensitivity of PDIC to the spectral<br />

extrapolation <strong>de</strong>creases when the relative contribution of the CDOM to the total<br />

absorption increases. In fact, for a [aCDOM/at](412) value of 0.7, ΔPDIC is


Table IV.5. ΔPDIC ( in %), due to error in the spectral extrapolation of [aCDOM/at](412) to<br />

the 300-600 nm range resulting from variation in the magnitu<strong>de</strong> of at(412). PDIC true is<br />

N<br />

calculated with averaged SCDOM , ap (λ) , and for different at(412) values. PDIC is<br />

calculated assuming aw(λ) is null.<br />

AQY1 AQY2<br />

[aCDOM/at] [aCDOM/at] [aCDOM/at] [aCDOM/at]<br />

= 0.3 = 0.7 = 0.3 = 0.7<br />

at(412) = 2.7 m -1 1.3 3.4 0.4 0.8<br />

at(412) = 0.41 m -1 at(412) = 0.048 m<br />

4.7 10.7 1.5 2.6<br />

-1 14.5 26.4 6.0 8.1<br />

Table IV.6. ΔPDIC ( in %), due to error in the spectral extrapolation of [aCDOM/at](412) to<br />

N<br />

the 300-600 nm range resulting from variation in SCDOM and ap (λ) spectra. PDIC true is<br />

N<br />

calculated with averaged SCDOM , ap (λ), and at(412).<br />

[aCDOM/at]<br />

= 0.3<br />

112<br />

AQY1 AQY2<br />

[aCDOM/at]<br />

= 0.7<br />

[aCDOM/at]<br />

= 0.3<br />

[aCDOM/at]<br />

= 0.7<br />

N<br />

ap (λ) ; SCDOM = 0.018 nm -1 -2.0 4.1 -8.2 -0.9<br />

N<br />

ap (λ) ; SCDOM = 0.026 nm -1 3.2 -3.4 8.1 0.6<br />

N<br />

ap (λ) - 1.96σ ; SCDOM = 0.022 nm -1 7.2 -0.3 11.3 1.9<br />

N<br />

ap (λ) + 1.96σ ; SCDOM = 0.022 nm -1 -3.8 0.9 -8.0 -1.5<br />

N<br />

ap (λ) - 1.96σ ; SCDOM = 0.018 nm -1 4.9 4.1 3.4 1.5<br />

N<br />

ap (λ) + 1.96σ ; SCDOM = 0.018 nm -1 -5.0 4.8 15.9 -2.9<br />

N<br />

ap (λ) - 1.96σ ; SCDOM = 0.026 nm -1 10.2 -3.8 18.7 -2.0<br />

N<br />

ap (λ) + 1.96σ ; SCDOM = 0.026 nm -1 -1.0 -2.3 0.3 -0.4


Figure IV.11. Example of spectral production of DIC for<br />

[aCDOM/at] true (412) value of 0.3 for various spectral shape for aCDOM(λ) and<br />

ap(λ) and for AQY1 (a-c) and AQY2 (d-f). The grey curves on the bottom<br />

N<br />

panels (b-c; e-f) is PDIC(λ) computed using averaged SCDOM and ap (λ).<br />

IV.3.4. Application to SeaWiFS imagery<br />

We applied our algorithm to a full resolution SeaWiFS image of the southeastern<br />

Beaufort Sea acquired on June 21 st 1998 (day 172). On that year, sea ice cover was<br />

exceptionally reduced in early spring over the study area, which allowed good ocean<br />

color observations over the whole Mackenzie Shelf and a large portion of the Canada<br />

Basin. The application of a turbid-water <strong>de</strong>tection algorithm [Morel and Bélanger, 2006]<br />

revealed that almost two thirds of the ice-free waters were turbid. Because the standard<br />

113


atmospheric correction, which is based on invalid assumptions of null water-leaving<br />

radiance in the NIR bands (765 and 865 nm) [Gordon and Wang, 1994], generally gives<br />

rise to negative water leaving radiances in the blue, the scene was processed using the<br />

turbid-water atmospheric correction algorithm <strong><strong>de</strong>s</strong>cribed by Ruddick et al. [2000]. The<br />

[aCDOM/at](412), calculated using Eq. 6 with region-specific coefficients (Table 2), shows<br />

coherent patterns over the continental shelf and beyond (Fig. IV.12). First, [aCDOM/at](412)<br />

values increase from the river mouth to the bor<strong>de</strong>r of the continental shelf (~0.40 to<br />

~0.90). This reflects the <strong>de</strong>creasing contribution of the terrigenous partic<strong>les</strong> to the total<br />

light absorption in the <strong>sur</strong>face waters when the partic<strong>les</strong> gradually settle over the shelf.<br />

Interestingly, the highest values (> 0.85) are found beyond the continental shelf in waters<br />

influence by the Mackenzie runoff. Further offshore in the Canada Basin and in the<br />

Amundsen Gulf, the [aCDOM/at](412) is relatively uniform, with values between 0.7 and<br />

0.8. These lower offshore values probably reflect the absence of direct influence of river<br />

runoff. The values observed there fall within the range observed in those areas during<br />

CASES.<br />

Figure IV.12. Spatial distribution of the [aCDOM/at](412) as <strong>de</strong>rived from<br />

SeaWiFS data acquired the 21 st of June 1998. The grey color co<strong>de</strong><br />

represents sea ice.<br />

Figure IV.13 shows the spatial distribution of the <strong>de</strong>pth-integrated DIC<br />

photoproduction rate (in mg C m -2 d -1 ) for both AQYs, as calculated using the satellite-<br />

<strong>de</strong>rived [aCDOM/at](412) values (Fig. IV.12). The spectral of [aCDOM/at](412) was<br />

114


achieved with the averaged spectral shape of aCDOM and at, and using the total absorption<br />

at 412 <strong>de</strong>rived from the Rrs spectrum using a modified version of the quasi-analytical<br />

algorithm proposed by Lee et al. [2002] (see Annex A1). The daily spectral irradiance<br />

−<br />

just below the sea <strong>sur</strong>face, E ( 0 , λ)<br />

, was obtained using the TUV mo<strong>de</strong>l with the<br />

d<br />

ozone observed by TOMS and aerosols concentrations <strong>de</strong>rived from SeaWiFS data. The<br />

spatial distribution of the PDIC is similar to that of [aCDOM/at](412) with the highest<br />

values of ~10-12 and ~3.6-3.8 mg C m -2 d -1 for AQY1 and AQY2, respectively, observed<br />

at the bor<strong>de</strong>r of the continental shelf. The spatial variability in PDIC is slightly higher for<br />

AQY1 (~2.5-fold) than for AQY2 (~2-fold). These results confirm that the spatial<br />

variability in PDIC resulting from the optical properties of the water column alone is<br />

important.<br />

115


Figure IV.13. Spatial distribution of <strong>de</strong>pth-integrated DIC<br />

photoproduction as calculated using both AQYs and the satellite-<strong>de</strong>rived<br />

[aCDOM/at](412) as shown in Fig. IV.12.<br />

IV.4. Summary and conclusions<br />

This study was motivated by the need to improve the quantification of the<br />

<strong>de</strong>pth-integrated DIC photoproduction in the Arctic coastal waters where a rapid <strong>de</strong>crease<br />

in the summer sea ice cover is observed, and expected to culminate with totally open<br />

waters during September before the end of this century [ACIA, 2005]. This major change<br />

will expose river runoff to light, including UV radiations [Bélanger et al., 2006]. Depth-<br />

integrated DIC photoproduction was assessed here using a simple spectral mo<strong>de</strong>l that<br />

116


needs as input the ratio between the CDOM and total absorption coefficients ([aCDOM/at]).<br />

To account for the important variability in this ratio in coastal waters, an empirical<br />

algorithm was proposed to estimate the [aCDOM/at] ratio at 412 nm from the remote<br />

sensing reflectance spectrum. The algorithm was <strong>de</strong>veloped and validated using an<br />

extensive data set of in situ mea<strong>sur</strong>ements of spectral absorption coefficients and<br />

reflectance ma<strong>de</strong> in contrasting coastal waters. The absolute uncertainty of the algorithm<br />

was found to be ±0.18 (±0.14 for regional tuning). This uncertainty in [aCDOM/at](412)<br />

results in a potential error on the <strong>de</strong>pth-integrated photoproduction of DIC


aspects that need more attention for future semi-analytical algorithms to be successful in<br />

discriminating aCDOM from the total absorption coefficient in coastal waters: the IOP<br />

mo<strong>de</strong>l formulation and the inversion approach.<br />

While the fact that NAP scatters light and CDOM do not is a source of ambiguity<br />

when pooling CDOM and NAP, it provi<strong><strong>de</strong>s</strong> discrimination power when CDOM and NAP<br />

are distinguished in an IOP mo<strong>de</strong>l. So, semi-analytical algorithms need to be based on an<br />

IOP mo<strong>de</strong>l that inclu<strong><strong>de</strong>s</strong> a good link between aNAP and bNAP (or bbNAP). Gallegos and<br />

Neale [2002] successfully applied such an approach to <strong>de</strong>convolve the absorption<br />

contribution by phytoplankton, CDOM and NAP in total absorption spectra mea<strong>sur</strong>ed in<br />

situ. But, obviously, the inversion of remotely sensed reflectance is more challenging<br />

than inverting absorption spectra mea<strong>sur</strong>ed in situ.<br />

Again, the fact that our empirical algorithm succee<strong>de</strong>d in discriminating aCDOM<br />

from aNAP means that the necessary information is contained in the reflectance spectrum.<br />

This information, however, must be extracted properly, and this is probably one<br />

weakness of current semi-analytical mo<strong>de</strong>ls. In coastal waters, variations in reflectance<br />

are primarily due to variations in turbidity [Sathyendranath et al., 1989]. Subtle changes<br />

in the shape of the reflectance spectrum that may result from changes in the proportions<br />

of phytoplankton, CDOM and NAP, are overwhelmed by the change in the magnitu<strong>de</strong> of<br />

reflectance at all wavelengths due to variations in turbidity. So, when some specific<br />

information is nee<strong>de</strong>d from the reflectance spectrum, such as the ratio of aCDOM to at,<br />

some specific spectral regions provi<strong>de</strong> <strong>les</strong>s ambiguous information. The rationa<strong>les</strong><br />

consi<strong>de</strong>red in the <strong>de</strong>velopment of our empirical algorithm are valid for a semi-analytical<br />

algorithm. The latter must go beyond finding the best fit between calculated and observed<br />

reflectance spectra, at all wavelengths [e.g., Roesler and Perry, 1995; Lee et al., 1996;<br />

Garver and Siegel, 1997]. This procedure must not be blurred by the large changes in<br />

reflectance magnitu<strong>de</strong> due to variations in turbidity, and be misled by reflectance at<br />

wavelengths where the information is ambiguous.<br />

The simple method proposed in this study represents, to our knowledge, a first<br />

step toward the integration of Ocean Color information to quantify of CDOM<br />

photooxidation in coastal waters. Further research studies in that field needs to be<br />

conducted co-jointly with specialists of both optics/remote sensing and photochemistry<br />

118


sciences in or<strong>de</strong>r to document in parallel AOPs, IOPs and CDOM photoreactivity (i.e.<br />

AQY) in coastal waters. The latter is critical for the quantification of CDOM<br />

photooxidation and is <strong>de</strong>finitely the least documented at the moment [Johannessen and<br />

Miller, 2001; Mopper and Keiber, 2002; Bélanger et al., 2006].<br />

119


IV.5. Notations<br />

Parameter Description<br />

λ Light wavelength (nm)<br />

a m (λ) Mea<strong>sur</strong>ed absorption coefficient with the ac-9 (m -1 )<br />

ax(λ) Absorption coefficient for constituent x (m -1 ),<br />

where subscript x is either t (total), w (water), CDOM (Chromophoric Dissolved Organic<br />

Matter), p (partic<strong>les</strong>), NAP (Non-algal partic<strong>les</strong>),CDM (Colored Dissolved and Detritic<br />

Material) or φ (phytoplankton).<br />

b(λ) Total scattering coefficient (m -1 )<br />

b m (λ) Mea<strong>sur</strong>ed scattering coefficient with the ac-9 (m -1 ), i.e. c(λ)-a m (λ).<br />

bp(λ) Particulate scattering coefficient (m -1 )<br />

bb(λ) Total backscattering coefficient (m -1 )<br />

bbp(λ) Particulate backscattering coefficient (m -1 )<br />

bbw(λ) Water backscattering coefficient (m -1 )<br />

c(λ) Beam attenuation (m -1 )<br />

Es(λ) Irradiance just above the sea <strong>sur</strong>face (W m -2 nm -1 )<br />

Ed(λ,z) Downwelling irradiance at <strong>de</strong>pth z (W m -2 nm -1 or mol photon m -2 nm -1 )<br />

Lu(λ,z) Upwelling radiance at <strong>de</strong>pth z (W m -2 nm -1 sr -1 )<br />

Lw(λ) Water-leaving radiance (W m -2 nm -1 sr -1 )<br />

Rrs(λ) Remote sensing reflectance above the sea <strong>sur</strong>face (nm -1 sr -1 )<br />

AQY (λ) Apparent quantum yield for dissolved inorganic carbon (mol C nm -1 (mol<br />

PDIC<br />

120<br />

photon) -1 )<br />

Depth-integrated photochemical production of dissolved inorganic carbon<br />

(mol C m -2 s -1 )


IV.6. References<br />

ACIA (2005), Arctic Climate Impact Assessment, 1042 pp., Cambridge Press University.<br />

Arrigo, K. R., D. Lubin, G. L. van Dijken, O. Holm-Hansen, and E. Morrow (2003),<br />

Impact of a <strong>de</strong>ep ozone hole on Southern Ocean primary production, J. Geophys.<br />

Res., 108(C5), doi:1029/2001JC001226.<br />

Babin, M., A. Morel, V. Fournier-Sicre, F. Fell, and D. Stramski (2003a), Light<br />

Scattering properties of marine partic<strong>les</strong> in coastal and open ocean waters as<br />

related to the particle mass concentration, Limnol. Oceanogr., 48(2), 843-859.<br />

Babin, M., D. Stramski, G. M. Ferrari, H. Claustre, A. Bricaud, G. Obolensky, and N.<br />

Hoepffner (2003b), Variations in the light absorption coefficients of<br />

phytoplankton, nonalgal partic<strong>les</strong>, and dissolved organic matter in coastal waters<br />

around Europe, J. Geophys. Res., 108(C7), 3211, doi: 10.1029/2001JC000882.<br />

Bélanger, S., H. Xie, N. Krotkov, P. Larouche, W. F. Vincent, and M. Babin (2006),<br />

Photomineralization of terrigenous dissolved organic matter in Arctic coastal<br />

waters from 1979 to 2003: Interannual variability and implications of climate<br />

change, Global Biogeochem. Cyc<strong>les</strong>, in press, doi:10.1029/2006GB002708.<br />

Benner, R., and S. Opsahl (2001), Molecular indicators of the sources and<br />

transformations of dissolved organic matter in the Mississippi river plume, Org.<br />

Geochem., 32(4), 597-611.<br />

Blough, N. V., and R. Del Vecchio (2002), Chromophoric DOM in the coastal<br />

environment, in Biogeochemistry of Marine Dissolved Organic Matter, edited by<br />

D. A. Hansell, and C. A. Carlson, pp. 509-546, Aca<strong>de</strong>mic Press, San Diego.<br />

Bricaud, A., A. Morel, and L. Prieur (1981), Absorption by dissolved organic matter of<br />

the sea (yellow substance) in the UV and visible domains, Limnol. Oceanogr.,<br />

26(1), 43-53.<br />

Carlson, C. A. (2002), Production and Removal Processes, in Biogeochemistry of Marine<br />

Dissolved Organic Matter, edited by D. A. Hansell, and C. A. Carlson, pp. 91-<br />

151, Aca<strong>de</strong>mic Press, San Diego.<br />

Dagg, M., R. Benner, S. Lohrenz, and D. Lawrence (2004), Transformation of dissolved<br />

and particulate materials on continental shelves influenced by large rivers: plume<br />

processes, Cont. Shelf Res., 24, 833-858.<br />

Fry, E. S. (2000), Visible and near-ultraviolet absorption spectrum of liquid water:<br />

comment, Appl. Opt., 39(16), 2743-2744.<br />

Gallegos, C. L., and P. J. Neale (2002), Partitioning spectral absorption in case 2 waters:<br />

discrimination of dissolved and particulates components, Appl. Opt., 41(21),<br />

4220-4233.<br />

121


Garver, D. L., and D. A. Siegel (1997), Inherent optical property inversion of ocean color<br />

spectra and its biogeochemical interpretation, 1. Time series from the Sargasso<br />

Sea, J. Geophys. Res., 102(C8), 18607-18625.<br />

Guéguen, C., L. Guo, and N. Tanaka (2005), Distributions and characteristics of colored<br />

dissolved organic matter in the Western Arctic Ocean, Cont. Shelf Res., 25(10),<br />

1195-1207.<br />

Gordon, H. R. (1992), Diffuse reflectance of the ocean: influence of nonuniform<br />

phytoplankton pigment profile, Appl. Opt., 31(12), 2116-2129.<br />

Gordon, H. R., and K. Ding (1992), Self-shading of in-water optical instruments, Limnol.<br />

Oceanogr., 37(3), 491-500.<br />

Gordon, H. R., and M. Wang (1994), Retrieval of water-leaving radiance and aerosol<br />

optical thickness over the oceans with SeaWiFS: a preliminary algorithm, Appl.<br />

Opt., 33(3), 443-452.<br />

Hansell, D. A., and C. A. Carlson (1998), Deep-ocean gradients in the concentration of<br />

dissolved organic carbon, Nature, 395(6699), 263-266.<br />

IOCCG (2006), Remote sensing of inherent optical properties: Fundamentals, tests of<br />

algorithms, and applications, edited by Z.-P. Lee, pp. 126, Reports of the<br />

International Ocean-Colour Coordinating Group, No. 5, IOCCG, Dartmouth,<br />

Canada.<br />

Johannessen, S. C. (2000), A photochemical sink for dissolved organic carbon in the<br />

ocean, Ph.D. thesis, 176 pp., Dalhousie University, Halifax, N. S., Canada.<br />

Johannessen, S. C., and W. L. Miller (2001), Quantum yield for the photochemical<br />

production of dissolved inorganic carbon in seawater, Mar. Chem., 76, 271-283.<br />

Kieber, R. J., X. Zhou, and K. Mopper (1990), Formation of carbonyl compounds from<br />

UV-induced photo<strong>de</strong>gradation of humic substances in natural waters: Fate of<br />

riverine carbon in the sea, Limnol. Oceanogr., 35(7), 1503-1515.<br />

Kishino, M., M. Takahashi, N. Okami, and S. Ichimura (1985), Estimation of the spectral<br />

absorption coefficients of phytoplankton in the sea, Bull. Mar. Sc., 37(2), 634-642.<br />

Laurion, I., F. Blouin, and S. Roy (2003), The quantitative filter technique for mea<strong>sur</strong>ing<br />

phytoplankton absorption : Interference by MAAs in the UV waveband, Limnol.<br />

Oceanogr.: Methods, 1(1), 1-9.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and R. A. Arnone (2002), Deriving inherent optical properties<br />

from water color : a multiband quasi-analytical algorithm for optically <strong>de</strong>ep<br />

waters, Appl. Opt., 41(27), 5755-5772.<br />

Lee, Z. P., K. L. Car<strong>de</strong>r, T. G. Peacock, C. O. Davis, and J. L. Mueller (1996), Method to<br />

<strong>de</strong>rive ocean absorption coefficients from remote-sensing reflectance, Appl. Opt.,<br />

35(3), 453-462.<br />

Loisel, H., and A. Morel (2001), Non-isotropy of the upward radiance field in typical<br />

coastal (Case 2) waters, Int. J. Remote Sens., 22(2&3), 275-295.<br />

122


Macdonald, R. W., S. M. Solomon, R. E. Cranston, H. E. Welsh, M. B. Yunker, and C.<br />

Gobeil (1998), A sediment and organic carbon budget for the Canadian Beaufort<br />

Shelf, Mar. Geol., 144, 255-273.<br />

Madronich, S., and S. Flocke (1999), The role of solar radiation in atmospheric chemistry,<br />

in Handbook of Environmental Chemistry, edited by P. Boule, pp. 1-26, Springer-<br />

Verlag, Hei<strong>de</strong>lberg, Germany.<br />

Maritorena, S., D. A. Siegel, and A. R. Peterson (2002), Optimization of a semianalitical<br />

ocean color mo<strong>de</strong>l for global-scale applications, Appl. Opt., 41(15), 2705-2714.<br />

Miller, W. L., and R. G. Zepp (1995), Photochemical production of dissolved inorganic<br />

carbon from terrestrial organic matter : Significance to the oceanic carbon cycle,<br />

Geophys. Res. Let., 22(4), 417-420.<br />

Mobley, C. D. (1999), Estimation of the remote-sensing reflectances from above-water<br />

mea<strong>sur</strong>ements, Appl. Opt., 38(36), 7442-7455.<br />

Mopper, K., and D. J. Kieber (2002), Photochemistry and the Cycling of Carbon, Sulfur,<br />

Nitrogen and Phosphorus, in Biogeochemistry of Marine Dissolved Organic<br />

Matter, edited by D. A. Hansell, and C. A. Carlson, pp. 455-507, Aca<strong>de</strong>mic Press,<br />

San Diego.<br />

Morel, A., and S. Bélanger (2006), Improved <strong>de</strong>tection of turbid waters from ocean color<br />

sensors information, Remote Sens. Environ., 102, 237-249.<br />

Morrison, J. R., and N. B. Nelson (2004), Seasonal cycle of phytoplankton UV<br />

absorption at Bermuda Atlantic Time-series Study (BATS) site, Limnol.<br />

Oceanogr., 49(1), 215-224.<br />

Mueller, J. L., A. Morel, R. Frouin, C. Davis, R. A. Arnone, K. L. Car<strong>de</strong>r, Z.-P. Lee, R. G.<br />

Stewart, S. B. Hooker, B. Holben, C. D. Mobley, S. McLean, M. Miller, C.<br />

Pietras, G. S. Fargion, K. D. Knobelspiesse, and K. Voss (2003), Radiometric<br />

Mea<strong>sur</strong>ements and Data Analysis protocols, in NASA Tech. Memo. 2003-21621,<br />

rev. 4, vol. 3, edited by J. L. Mueller, G. S. Fargion, and C. R. McClain, pp. 78,<br />

NASA Goddard Space Flight Center, Greenbelt, Md.<br />

Pope, R. M., and E. S. Fry (1997), Absorption spectrum (380-700 nm) of pure water. II.<br />

Integrating cavity mea<strong>sur</strong>ements, Appl. Opt., 36, 8710-8723.<br />

Quicken<strong>de</strong>n, T. I., and J. A. Irvin (1980), The ultraviolet absorption spectrum of liquid<br />

water, J. Chem. Phys., 72(8), 4416-4428.<br />

Rachold, V., H. Eicken, V. V. Gor<strong>de</strong>ev, M. N. Grigoriev, H.-W. Hubberten, A. P. Lisitzin,<br />

V. P. Shevchenko, and L. Schirrmeister (2004), Mo<strong>de</strong>rn Terrigenous Organic<br />

Carbon input to the Arctic Ocean, in The organic carbon cycle in the Arctic<br />

Ocean, edited by R. Stein, and R. W. MacDonald, pp. 33-41, Springer, Berlin.<br />

Roesler, C. S., and M. J. Perry (1995), In situ phytoplankton absorption, fluorescence<br />

emission, and particulate backscattering spectra <strong>de</strong>termined from reflectance, J.<br />

Geophys. Res., 100(C7), 13279-13294.<br />

123


Roesler, C. S., M. J. Perry, and K. L. Car<strong>de</strong>r (1989), Mo<strong>de</strong>ling in situ phytoplankton<br />

absorption from total absorption spectra in productive inland marine waters,<br />

Limnol. Oceanogr., 34, 1510-1523.<br />

Ruddick, K. G., V. De Cauwer, Y.-J. Park, and G. Moore (2006), Seaborne<br />

mea<strong>sur</strong>ements of near infrared water-leaving reflectance: The similarity spectrum<br />

for turbid waters, Limnol. Oceanogr., 51(2), 1167-1179.<br />

Ruddick, K. G., F. Ovidio, and M. Rijkeboer (2000), Atmospheric correction of SeaWiFS<br />

imagery for turbid coastal and inland waters, Appl. Opt., 39(6), 897-895.<br />

Sathyendranath, S., L. Prieur, and A. Morel (1989), A three-component mo<strong>de</strong>l of ocean<br />

colour and its application to remote sensing of phytoplankton pigments in coastal<br />

waters, Int. J. Remote Sens., 10(8), 1373-1394.<br />

Siegel, D. A., S. Maritorena, N. B. Nelson, D. A. Hansell, and M. Lorenzi-Kayser (2002),<br />

Global distribution and dynamics of colored dissolved and <strong>de</strong>trital organic<br />

materials, J. Geophys. Res., 107(C12), 3228, doi:10.1029/2001JC000965.<br />

Tassan, S., and G. M. Ferrari (1995), An alternative approach to absorption<br />

mea<strong>sur</strong>ements of aquatic partic<strong>les</strong> retained on filters, Limnol. Oceanogr., 40(8),<br />

1358-1368.<br />

Tassan, S., and G. M. Ferrari (2002), A sensitivity analysis of the 'transmittancereflectance'<br />

method for mea<strong>sur</strong>ing light absorption by aquatic partic<strong>les</strong>, J.<br />

Phytoplank. Res., 24(8), 757-776.<br />

Telang, S. A., R. Pocklington, A. S. Naidu, E. A. Romankevich, I. I. Gitelson, and<br />

Glacyshev (1991), Carbon and Mineral Transport in Major North American,<br />

Russian Arctic, and Siberian Rivers: the St Lawrence, the Mackenzie, the Yukon,<br />

the Arctic Alaskan Rivers, the Arctic Basin Rivers in the Soviet Union, and the<br />

Yenisei, in Biogeochemistry of Major World Rivers. SCOPE 42, edited by E. T.<br />

Degens, S. Kempe, and J. E. Richey, pp. 74-104, Wiley, New York.<br />

Twardowski, M. S., E. Boss, J. M. Sullivan, and P. L. Donaghay (2004), Mo<strong>de</strong>ling the<br />

spectral shape of absorption by chromophoric dissolved organic matter. Marine<br />

Chemistry, 89, 69-88.<br />

Wer<strong>de</strong>ll, P. J., and S. W. Bailey (2005), An improved in-situ bio-optical data set for<br />

ocean color algorithm <strong>de</strong>velopment and satellite product validation, Remote Sens.<br />

Environ., 98, 122-140.<br />

Zaneveld, J. R. V., J. C. Kitchen, and C. Moore (1994), The scattering Error Correction<br />

of Reflecting-Tube Absorption Meters, in Ocean Optics XII, edited by J. S. Jaffe,<br />

pp. 44-55, SPIE, Bergen.<br />

Zibordi, G., and G. M. Ferrari (1995), Instrument Self-Shading in Un<strong>de</strong>rwater Optical<br />

Mea<strong>sur</strong>ements - Experimental-Data, Appl. Opt., 34(15), 2750-2754.<br />

124


Chapitre V: Impact <strong>de</strong> la glace <strong>de</strong> mer <strong>sur</strong> <strong>les</strong><br />

estimations <strong>de</strong> la réflectance marine, la<br />

concentration <strong>de</strong> chlorophylle a et <strong>les</strong> propriétés<br />

optiques inhérentes dérivés <strong><strong>de</strong>s</strong> données<br />

satellita<strong>les</strong> <strong>de</strong> la Couleur <strong>de</strong> l’Océan<br />

125


V.A. Résumé<br />

À l’ai<strong>de</strong> <strong>de</strong> simulations numériques et d’observations, on a décrit <strong>de</strong>ux phénomènes<br />

physiques impliquant la glace <strong>de</strong> mer qui contaminent <strong>les</strong> me<strong>sur</strong>es spatia<strong>les</strong> <strong>de</strong> la couleur <strong>de</strong><br />

l’océan dans <strong>les</strong> hautes latitu<strong><strong>de</strong>s</strong> : 1) l’effet <strong>de</strong> l’environnement qui se produit le long <strong>de</strong> la<br />

bordure <strong>de</strong> la banquise <strong>de</strong> glace et 2) la contamination sub-pixel due à la présence d’une<br />

faible quantité <strong>de</strong> glace à l’intérieur d’un pixel classifié océanique. Le signal reçu au sommet<br />

<strong>de</strong> l’atmosphère (TOA) a été simulé à l’ai<strong>de</strong> un modèle <strong>de</strong> transfert radiatif permettant la<br />

modélisation <strong>de</strong> l’effet <strong>de</strong> l’environnement pour divers types <strong>de</strong> glace <strong>de</strong> mer entourant une<br />

<strong>sur</strong>face d’eau ouverte. Ont été utilisés pour ces simulations <strong><strong>de</strong>s</strong> spectres <strong>de</strong> réflectance <strong>de</strong> la<br />

glace <strong>de</strong> mer obtenus avant et pendant la pério<strong>de</strong> <strong>de</strong> fonte <strong>de</strong> la glace en 2004 dans la Mer <strong>de</strong><br />

Beaufort. Pour la contamination sub-pixel, le signal TOA a été simulé en utilisant <strong>les</strong><br />

propriétés atmosphériques utilisées dans le traitement <strong><strong>de</strong>s</strong> données SeaWiFS, pour différents<br />

spectres <strong>de</strong> réflectance <strong>de</strong> <strong>sur</strong>face obtenus en combinant linéairement <strong>les</strong> spectres <strong>de</strong><br />

réflectance marine (ρ w) et <strong>de</strong> glace <strong>de</strong> mer. Les algorithmes <strong>de</strong> correction atmosphérique<br />

standard <strong>de</strong> SeaWiFS ont été appliqués aux spectres TOA simulés afin <strong>de</strong> retrouver le<br />

spectre ρ w, à partir duquel <strong>les</strong> concentrations en chlorophylle a (CHL) et <strong>les</strong> propriétés<br />

optiques inhérentes (IOP) sont calculées. Lorsque que <strong>les</strong> conditions atmosphériques sont<br />

claires, l’effet <strong>de</strong> l’environnement introduit une <strong>sur</strong>estimation significative (>0.002) <strong>de</strong> ρ w<br />

dans la partie bleue du spectre (i.e. 443 nm), et ce à une distance aussi gran<strong>de</strong> que 24 Km <strong>de</strong><br />

la bordure <strong>de</strong> glace. La CHL, calculée empiriquement à partir d’un algorithme utilisant <strong><strong>de</strong>s</strong><br />

rapports <strong>de</strong> réflectance entre le bleu et le vert, était soit sous-estimée, soit <strong>sur</strong>estimée,<br />

dépendamment <strong>de</strong> la concentration réelle en CHL. Pour <strong><strong>de</strong>s</strong> CHL <strong>de</strong> modérées à élevées (><br />

0.5 mg m -3 ), tout <strong>les</strong> pixels se trouvant à une distance inférieure à 10-20 Km <strong>de</strong> la bordure <strong>de</strong><br />

glace doivent être éliminés. Il <strong>de</strong>vient crucial <strong>de</strong> considérer l’effet <strong>de</strong> l’environnement lorsque<br />

l’on veut estimer le coefficient d’absorption total (a t) <strong>de</strong> l’eau <strong>de</strong> mer à partir d’un algorithme<br />

d’inversion semi-analytique. À 443 nm, a t est sous-estimé par plus <strong>de</strong> 35% à ~20 Km <strong>de</strong> la<br />

bordure <strong>de</strong> glace quand CHL est > 0.5 mg m -3 . L’effet <strong>sur</strong> l’estimation du coefficient <strong>de</strong><br />

rétrodiffusion par <strong>les</strong> particu<strong>les</strong> (b bp) n’est important que lorsque que <strong>les</strong> eaux sont<br />

relativement peu turbi<strong><strong>de</strong>s</strong> (i.e. CHL < 0.05 mg m -3 et matière en suspension < 0.2 g m -3 ).<br />

L’impact <strong>sur</strong> le rapport entre <strong>les</strong> coefficients d’absorption par la matière organique dissoute<br />

colorée (CDOM) et total ([a CDOM/a t]) à 412 nm, calculé empiriquement, est similaire à celui<br />

126


<strong>sur</strong> a t (i.e. sous-estimation). À l’opposé, la contamination sub-pixel produit une sous-<br />

estimation systématique <strong>de</strong> ρ w dans la partie bleue du spectre. En générale, la contamination<br />

sub-pixel résulte en une <strong>sur</strong>estimation <strong>de</strong> la CHL, <strong>de</strong> a t, et <strong>de</strong> [a CDOM/a t], et en une<br />

sousestimation <strong>de</strong> b bp. Une métho<strong>de</strong> simple a été proposée pour i<strong>de</strong>ntifier <strong>les</strong> pixels<br />

contaminés par <strong>les</strong> effets d’environnement.<br />

127


V.B. Article soumis à la revue Remote Sensing of Environment (1 novembre 2006):<br />

“Impact of sea ice on the retrieval of water-leaving reflectance, chlorophyll a<br />

concentration and inherent optical properties from satellite Ocean Color data”<br />

Simon Bélanger a , Jens K. Ehn b and Marcel Babin a<br />

a CNRS, Laboratoire d'océanographie <strong>de</strong> Villefranche, 06230 Villefranche-<strong>sur</strong>-Mer,<br />

France ; Université Pierre et Marie Curie-Paris 6, Laboratoire d'Océanographie <strong>de</strong><br />

Villefranche, 06230 Villefranche-<strong>sur</strong>-Mer, France<br />

b Centre for Earth Observation Science, Department of Environment and Geography,<br />

Clayton H. Rid<strong>de</strong>ll Faculty of Earth, Environment and Resources, 467 Wallace Building,<br />

University of Manitoba, Winnipeg, MB, R3T 2N2, phone: 204-474-6961, fax 272-1532,<br />

umehnjjk@cc.umanitoba.ca<br />

128


Abstract<br />

Through simulations and observations, we <strong><strong>de</strong>s</strong>cribed two physical phenomena by which<br />

satellite remotely sensed ocean color data are contaminated by sea ice at high latitu<strong><strong>de</strong>s</strong>: 1)<br />

the adjacency effect that occurs along sea ice margins and 2) the sub-pixel contamination<br />

by a small amount of sea ice within an ocean pixel. The signal at the top of the<br />

atmosphere (TOA) was simulated using a radiative transfer co<strong>de</strong> that allows mo<strong>de</strong>ling of<br />

the adjacency effect for various types of sea ice <strong>sur</strong>rounding an open water area. In situ<br />

sea ice reflectance spectra obtained prior to and during the melt period are used in the<br />

simulations. For sub-pixel contamination, the TOA signal was simulated using the<br />

SeaWiFS atmospheric look-up tab<strong>les</strong> with various <strong>sur</strong>face reflectances obtained by linear<br />

mixture of both sea ice and water-leaving reflectances (ρw). The standard atmospheric<br />

correction algorithms were applied to the simulated TOA spectra to retrieve the ρw<br />

spectrum from which chlorophyll a concentration (CHL) and inherent optical properties<br />

(IOP) are <strong>de</strong>rived. Un<strong>de</strong>r clear atmospheric conditions, the adjacency effect introduced<br />

large overestimation in ρw (>0.002), as far as 24 km from an ice edge in the blue part of<br />

the spectrum (443 nm). The CHL, estimated using standard blue-to-green reflectance<br />

ratio, is either over- or un<strong>de</strong>restimated <strong>de</strong>pending on the actual concentration. For<br />

mo<strong>de</strong>rate to high CHL (>0.5 mg m -3 ), any pixel located within a distance of ~10-20 km<br />

from the ice edge should be discar<strong>de</strong>d. It is very important to consi<strong>de</strong>r the adjacency<br />

effect when the total absorption coefficient (at) is to be retrieved using a semi-analytical<br />

algorithm. The at(443) was un<strong>de</strong>restimated by more than 35% at a distance of ~20 km<br />

from an ice edge for CHL > 0.5 mg m -3 . The effect on the partic<strong>les</strong> backscattering<br />

coefficient (bbp) retrieval was important only for clear waters (CHL ~0.05 mg m -3 ). The<br />

impact on the ratio between chromophoric dissolved organic matter (CDOM) and total<br />

absorption coefficients ([aCDOM/at]) at 412 nm, estimated empirically, is similar to at(443)<br />

(i.e. un<strong>de</strong>restimation). In contrast, sub-pixel contamination by small amount of sea ice<br />

produces systematic un<strong>de</strong>restimation of ρw in the blue. In general, sub-pixel<br />

contamination results in an overestimation of CHL, at and [aCDOM/at], and an<br />

un<strong>de</strong>restimation of bbp. A simple method was proposed to flag the pixels contaminated by<br />

adjacency effect.<br />

129


V.1. Introduction<br />

Polynyas and flaw leads are areas of open water, or much reduced sea ice cover,<br />

at high latitu<strong><strong>de</strong>s</strong> where full sea ice cover is expected [Smith et al., 1990]. Due to their<br />

high biological productivity, polynyas and leads are sites of reproduction for several<br />

species of birds, fishes and mammals [Stirling, 1997]. At high latitu<strong><strong>de</strong>s</strong>, intense primary<br />

productivity also occur in Marginal Ice Zones (MIZ) where melt water provi<strong><strong>de</strong>s</strong> enough<br />

vertical stability in the water column for phytoplankton to grow un<strong>de</strong>r high-light and<br />

high-nutrients conditions [Smith and Nelson, 1986; Sakshaug and Skjoldal, 1989].<br />

Because polar regions are especially susceptible to global warming (i.e., + ~3°C in 2050)<br />

[ACIA, 2005], the timing and magnitu<strong>de</strong> of the primary productivity in polynyas, leads<br />

and MIZ are likely to be affected in the coming years.<br />

Global warming can also alter the amount and fate of terrigenous material (e.g.,<br />

organic matter, nutrients, and sediments) input into the Arctic Ocean, which is a key<br />

component of the Arctic carbon cycle [Stein and MacDonald, 2004]. For example,<br />

increase in river discharge [McClelland et al., 2006] and accelerate permafrost thawing<br />

[Camill, 2005] could lead to the mobilization of significant amount of terrigenous<br />

dissolved and particulate carbon toward to coastal Arctic ocean during the next century<br />

[Frey and Smith, 2005; Guo and Macdonald, 2006]. Furthermore, amplified<br />

photomineralization of terrigenous dissolved organic carbon in the Arctic coastal waters<br />

is expected as a response to the <strong>de</strong>crease in summer ice cover that allows more solar<br />

radiation to penetrate the water column [Bélanger et al., 2006]. Since global warming can<br />

modify either marine or terrigenous components of the carbon cycle, continuous<br />

monitoring systems for the biological state of north polar seas are urgently nee<strong>de</strong>d.<br />

Satellite remote sensing of the ocean color provi<strong><strong>de</strong>s</strong> a means for studying long-<br />

term and large-scale changes in the biological processes occurring in the polar waters.<br />

The most common bio-optical parameter <strong>de</strong>rived from ocean color is the concentration of<br />

chlorophyll a (CHL), the key molecule of photosynthesis and a common in<strong>de</strong>x of<br />

phytoplankton biomass in the ocean. With a <strong>sur</strong>face estimate of CHL, the gross primary<br />

130


productivity can be calculated using various mo<strong>de</strong>ls [e.g., Antoine and Morel, 1996;<br />

Behrenfeld and Falkowski, 1997; Arrigo et al., 1998]. Furthermore, the present<br />

generation of bio-optical algorithms [Lee et al., 2002; Maritorena et al., 2002] can<br />

provi<strong>de</strong>, in addition to CHL, other optical indicators like some of the inherent optical<br />

properties (IOP).<br />

Good retrieval of CHL and IOP from remotely sensed ocean color data requires (1)<br />

a reliable retrieval of the spectral water leaving radiance, Lw, or reflectance, ρw, which<br />

relies mainly on the quality of the sensor calibration and of the atmospheric correction,<br />

and (2) robust bio-optical algorithms. In polar seas, the former may be strongly<br />

compromised by the presence of sea ice [e.g. Gregg and Casey, 2004]. Despite the fact<br />

that this problem might represent a major obstacle to the use of ocean color in Polar<br />

Regions, that problem has never been addressed quantitatively.<br />

In view of using satellite remote sensing of bio-optical properties to study<br />

biogeochemical processes occurring in polar seas, the purpose of this study is to quantify<br />

the error on the estimate of ρw due to the presence of sea ice, and its consequences on the<br />

retrieval of CHL and IOP. Sea ice can introduce biases in the interpretation of ocean<br />

color data in two ways: 1) the adjacency effect is the process by which a photon, reflected<br />

from a <strong>sur</strong>face adjacent to a targeted pixel, is scattered by the atmosphere between the<br />

sensor and the target, blurring the sharp boundary between sea ice and open water [e.g.,<br />

Tanré et al., 1979; Tanré et al., 1981]; 2) sub-pixel contamination when a small amount<br />

of sea ice within an ocean pixel (a pixel not flagged as cloud or ice), may be wrongly<br />

interpreted as a high aerosol concentration and/or high seawater turbidity. The first<br />

phenomenon is observed throughout the year, but may be more important in spring time<br />

when leads and polynyas start to open and adjacent sea ice is still covered by highly<br />

reflective fresh snow; while the second can occurs during summer when ice floes,<br />

transported by water currents and winds, melt progressively in the ocean.<br />

In what follows, we first <strong><strong>de</strong>s</strong>cribe the methodology which is adopted to simulate<br />

the top-of-atmosphere (TOA) reflectance spectra, ρTOA, with and without the effects of<br />

sea ice. Then we apply the Sea-viewing Wi<strong>de</strong> Field-of-view Sensor (SeaWiFS)<br />

atmospheric correction [Gordon and Wang, 1994] to our simulated TOA spectra to<br />

quantify the errors in water-leaving reflectance resulting from the presence of ice. Next<br />

131


the CHL is calculated using two empirical algorithms: 1) the standard NASA global<br />

algorithm OC4v4 used in both SeaWiFS and MODIS processing [O'Reilly et al., 2000]<br />

and 2) the Arctic’s regionally “tuned” OC4L algorithm [Cota et al., 2004]. For testing the<br />

effect of sea ice on the retrieval of IOPs, we applied an improved version of the algorithm<br />

of Lee at al. [2002] <strong><strong>de</strong>s</strong>igned to estimate the total absorption (at) and particulate<br />

backscattering coefficients (bbp). Finally, we tested an algorithm recently <strong>de</strong>veloped to<br />

retrieve the ratio [aCDOM/at], which is necessary to estimate CDOM photooxidation<br />

[Bélanger et al., Improved quantification of Chromophoric Dissolved Organic Matter<br />

photooxidation in coastal waters using satellite-<strong>de</strong>rived inherent optical properties,<br />

submitted manuscript, 2006; (hereinafter referred to as Bélanger et al., submitted<br />

manuscript, 2006)].<br />

V.2. Methodology<br />

V.2.1. Definitions<br />

The radiance received by a sensor at the TOA in a given direction, LTOA, is<br />

converted into remote sensing reflectance, ρTOA, using the following expression [Gordon,<br />

1997] (note that geometrical <strong>de</strong>pen<strong>de</strong>ncy of the following quantities is omitted for<br />

simplicity):<br />

π LTOA<br />

( λ)<br />

ρ TOA(<br />

λ)<br />

= , (1)<br />

F ( λ)<br />

cos( θ )<br />

132<br />

0 s<br />

where F0 is the extraterrestrial solar irradiance, θs is the sun zenith angle and λ is the<br />

wavelength (in nm). Ocean Color applications commonly use the “normalized” water-<br />

leaving reflectance (or radiance), [ρw]N (or [Lw]N), <strong>de</strong>fined as [Gordon and Clark, 1981]:<br />

π<br />

π L ( λ)<br />

ρ w = ( λ)<br />

=<br />

, (2)<br />

N<br />

) t<br />

[ ] ( λ)<br />

[ L ]<br />

F0 ( λ)<br />

w N<br />

w<br />

F0<br />

( λ)<br />

cos( θs<br />

where Lw is the water-leaving radiance and t0 is the diffuse transmittance from sun to<br />

pixel. Note that [Lw]N/F0 is equivalent to the commonly used remote sensing reflectance<br />

(Rrs) <strong>de</strong>fined as the ratio of water-leaving radiance to the downwelling plane irradiance<br />

just above the sea <strong>sur</strong>face [see Morel and Mueller, 2002].<br />

0


V.2.2. Description of the <strong>sur</strong>face reflectance spectra<br />

V.2.2.1. Sea ice reflectance spectra<br />

The spectral irradiance reflectance spectrum of sea ice (Rice) was <strong>de</strong>termined at<br />

three locations during the spring-summer 2004 onboard the CCGS Amundsen in the<br />

Southeastern Beaufort Sea as part as the Canadian Arctic Shelf Exchange Study (CASES)<br />

program. Radiometric mea<strong>sur</strong>ements were performed at 512 wavelengths ranging from<br />

340 to 1060 nm with a dual-hea<strong>de</strong>d spectroradiometer (FieldSpec, Analytical Spectral<br />

Devices Inc., Boul<strong>de</strong>r, Colorado) equipped with a cosine collector that was placed 0.7 m<br />

above the sea ice <strong>sur</strong>face, yielding mea<strong>sur</strong>ement of upwelling and downwelling<br />

irradiance, Eu and Ed, respectively, and a second cosine collector for simultaneous<br />

mea<strong>sur</strong>ements of<br />

ref<br />

E d to correct for changes in inci<strong>de</strong>nt solar <strong>flux</strong> during the Eu and Ed<br />

mea<strong>sur</strong>ement. The irradiance reflectance is calculated as<br />

ref<br />

Eu<br />

( λ)<br />

Ed<br />

( λ)<br />

R ice ( λ) = . (3)<br />

ref<br />

E ( λ)<br />

E ( λ)<br />

d<br />

d<br />

The Rice spectrum was <strong>de</strong>termined over four different <strong>sur</strong>faces: 1) 2-m-thick<br />

landfast ice overlayed by a 15-cm-thick dry and fresh snow cover mea<strong>sur</strong>ed in early May<br />

at 70.25°N and 126°W (in Franklin Bay, Canadian Arctic); 2) an ice floe covered by<br />

melting snow [Perovich et al., 1998]; 3) a white ice floe with a drained high-scattering<br />

<strong>sur</strong>face layer; 4) a mixed white-grey ice floe with a drained low-scattering <strong>sur</strong>face layers.<br />

The ice floes were mea<strong>sur</strong>ed in the Amundsen Gulf (71.25°N, 127.66°W). The additional<br />

spectrum from Perovich et al. [1998] was inclu<strong>de</strong>d to cover the conditions that may be<br />

observed later during melting season (June-July) on landfast ice when the snow cover<br />

vanishes. The five Rice spectra (Fig. V.1) are representative of Arctic sea ice at different<br />

stages of their seasonal evolution. In early spring, the landfast ice is covered by fresh<br />

snow that reflects >91% at all visible wavelengths. During summer time, the reflectance<br />

can <strong>de</strong>crease below 10% and the spectral variations may become more pronounced (e.g.<br />

the ice floe). During the melting season, however, the <strong>sur</strong>face of an ice floe is a mixture<br />

of different ice <strong>sur</strong>face types. The <strong>sur</strong>face of the ice floes encountered in the Amundsen<br />

Gulf at the end of June was typically composed of about 30% of white ice and the rest of<br />

grey ice (i.e. White and Grey ice spectra shown in Fig. V.1). Therefore, it was more<br />

representative to average the two ice floe spectra mea<strong>sur</strong>ed on June 26 as: 0.3* Rwhite ice +<br />

133


0.7*Rgrey_ice. This spectral composite is called hereinafter “grey ice”. An extensive study<br />

of the sea ice <strong>sur</strong>face reflectance is out of the scope of this study.<br />

Figure V.1. Reflectance spectra for five types of Arctic sea ice at different<br />

stages during the melting season. The spectrum of landfast ice covered by<br />

melting snow was taken from Perovich et al. [1998], while the other<br />

spectra were mea<strong>sur</strong>ed in the Southeastern Beaufort Sea during the<br />

CASES expedition.<br />

V.2.2.2. Water-leaving reflectance spectra<br />

We calculated 20 water-leaving reflectance spectra using the semi-analytical<br />

forward mo<strong>de</strong>l <strong><strong>de</strong>s</strong>cribed in the Appendix V.1 (Fig. V.2). The parameterization of the<br />

mo<strong>de</strong>l was <strong><strong>de</strong>s</strong>igned to simulate spectra typical of both Open Ocean (Case 1) and coastal<br />

(Case 2) waters. The three input parameters used to set the spectral IOP (i.e. absorption<br />

and backscattering coefficients) are: 1) the phytoplankton chlorophyll a concentration<br />

(CHL); 2) the dry weight of partic<strong>les</strong> that do not co-vary with phytoplankton (SPM NAP )<br />

(in g m -3 ); and 3) the contribution of CDOM to the non-water absorption coefficient at<br />

443 nm [fCDOM = aCDOM/(aφ+aNAP+aCDOM)]. To generate Case 1 water reflectance spectra,<br />

fCDOM and SPM NAP are set to 0, and the IOPs are a function of CHL as in Morel and<br />

Maritorena [2001] (for <strong>de</strong>tails, see Appendix V.1). To represent a wi<strong>de</strong> range of trophic<br />

134


and turbidity conditions that could be encountered at high latitu<strong>de</strong>, the following input<br />

parameters were used:<br />

1. 4 values for SPM NAP (0, 0.2, 2.0 and 20 g m -3 );<br />

2. 3 values for CHL (0.05, 0.5, 5.0 mg m -3 ) (excepted when SPM = 20.0 g m -3<br />

because [ρw]N spectra are almost i<strong>de</strong>ntical for any CHL values);<br />

3. 2 levels for fCDOM (50% and 80%).<br />

To be more representative of polar waters where the contribution of the colored dissolved<br />

and <strong>de</strong>trital material (CDM) to the total light absorption at 440 nm is high (>50%) [Siegel<br />

et al., 2005, see their Figure 2a], high values for fCDOM were chosen. In the Arctic, in<br />

particular, the strong CDOM signal is due to the high concentration of DOC that is<br />

maintained by the large amount of river runoff entering into polar <strong>sur</strong>face waters annually<br />

[see Benner et al., 2005 and ref. therein].<br />

Figure V.2. Mo<strong>de</strong>led normalized water-leaving reflectance spectra for<br />

various chlorophyll concentration (mg m -3 ), as indicated. The solid curves<br />

are produced by introducing an fCDOM value of 50% while dotted curves<br />

are produced with fCDOM =80%. Four concentrations of non algal partic<strong>les</strong><br />

(SPM NAP ) are presented: a) 0.0 g m -3 ; b) 0.2 g m -3 ; c) 2.0 g m -3 ; d) 20 g m -3 .<br />

135


V.2.3. Simulations of the adjacency effect<br />

The adjacency effect is significant at high latitu<strong>de</strong> because the sea ice, that<br />

reflects up to 90% of the inci<strong>de</strong>nt light, are adjacent to seawater with a reflectance < 5-<br />

10%. The Second Simulation of the Satellite Signal in the Solar Spectrum (6S) radiative<br />

transfer co<strong>de</strong> [available online at http://www-loa.univ-<br />

lille1.fr/SOFTWARE/Msixs/msixs_gb.html, Vermote et al., 1997] was used to simulate<br />

ρTOA for the eight SeaWiFS spectral bands (i.e. 412, 443, 490, 510, 555, 670, 765, and<br />

865 nm). The ρTOA spectra are simulated for water pixels located at 1 to 50 km from the<br />

coast to offshore along a transect perpendicular to an hypothetical linear ice edge of<br />

infinite length (for <strong>de</strong>tails, see Appendix V.2). The gas concentration and the<br />

atmospheric pres<strong>sur</strong>e, temperature, ozone and water vapor profi<strong>les</strong> for the U.S. standard<br />

atmosphere [Mc Clatchey et al., 1971] were used as input parameters. The 6s maritime<br />

aerosol mo<strong>de</strong>l composed of 95 and 5% of oceanic and water-soluble components,<br />

respectively [Vermote et al., 1997], was used in the simulations. The aerosols optical<br />

thickness at 550 nm (τa(550)) was set to 0.03, 0.1 and 0.2 to cover very clear to turbid<br />

atmospheric conditions (note that global mean is 0.1). Finally, the illumination and<br />

viewing geometry is set for typical SeaWiFS observation in mid-June at 70°N, for a pixel<br />

located in the center of the field of view (i.e., θ0 = 50°; satellite viewing angle from nadir,<br />

θv = 22°; azimuth angle between the sun-pixel and pixel-sensor half planes, Δφ = 8°). The<br />

calculations were performed for four different types of ice and for various water-leaving<br />

spectra as <strong><strong>de</strong>s</strong>cribed above (Section V.2.2).<br />

V.2.4. Simulations of sea ice contamination within an ocean pixel<br />

During the summer season, after the breakup of the ice cover, oceanic pixels may<br />

be contaminated by the presence of sea ice. Physically, the phenomenon can be compared<br />

to the effect of sea foam or whitecaps [Frouin et al., 1996]. In such a case, the TOA<br />

reflectance can be expressed as:<br />

TOA<br />

path<br />

[ ( 1 − σ )[ ρ R ]<br />

ρ = ρ + t ] + σ<br />

0 tv<br />

w N ice , (4)<br />

where ρpath is the atmospheric path reflectance, tv is the diffuse atmospheric transmittance<br />

from sea <strong>sur</strong>face to the sensor and σ is the fractional sea ice <strong>sur</strong>face concentration within<br />

136


the pixel (dimension<strong>les</strong>s, from 0 to 1). Thus in equation 4, the bracketed term represent<br />

<strong>sur</strong>face reflectance, which is simply a linear mixture of both sea ice and water-leaving<br />

reflectances. Here, ρpath, tv and t0 were taken from the SeaWiFS lookup tab<strong>les</strong> for an<br />

atmosphere containing maritime aerosol with relative humidity of 90%. The illumination<br />

and viewing geometry and aerosol concentrations were the same as <strong><strong>de</strong>s</strong>cribed in Section<br />

2.3. Simulations were done for several values of σRice(865) between 0 and 3.2%, a range<br />

at which the pixel may not be masked as cloud. Above this value, the pixel is generally<br />

masked as a cloud. For example, SeaWiFS cloud masking is perform using a single<br />

threshold (here set to 3%) on the TOA reflectance at 865 nm corrected for the molecular<br />

scattering contribution (Rayleigh). So, <strong>de</strong>pending on the magnitu<strong>de</strong> the ice albedo (Fig.<br />

V.1), the σ values ranged from 0 to ~3.5% for the fresh snow, and to ~25% for the grey<br />

ice.<br />

V.2.5. Atmospheric correction and bio-optical processing<br />

In or<strong>de</strong>r to apply the SeaWiFS atmospheric algorithm to the simulated spectra, the<br />

ρTOA were converted into LTOA (Eq. 1) and processed using the SeaWiFS Data Analysis<br />

Software (SeaDAS 4.6). The standard atmospheric correction algorithm based on the<br />

black pixel assumption in the near infrared (NIR) [Gordon and Wang, 1994] was used to<br />

process most simulated spectra. An atmospheric correction algorithm <strong><strong>de</strong>s</strong>igned for turbid<br />

waters [Stumpf et al., 2003] was used to process the simulated spectra un<strong>de</strong>r mo<strong>de</strong>rately<br />

turbid water conditions (Fig. V.2d). For chlorophyll a concentration estimation, we used<br />

the global OC4v4 algorithm, which is a forth-or<strong>de</strong>r polynomial function that uses the<br />

maximum reflectance band ratio between 443, 490 and 510 nm upon 555 [O'Reilly et al.,<br />

2000]:<br />

with,<br />

CHL<br />

a<br />

+ a * R + a * R<br />

2<br />

3<br />

4<br />

= 10 0 1<br />

2<br />

3<br />

4 ,<br />

R =<br />

log 10<br />

⎛ max<br />

⎜<br />

⎝<br />

137<br />

+ a<br />

* R<br />

+ a<br />

{ [ ρ ] ( 443,<br />

490,<br />

510)<br />

}<br />

* R<br />

[ ] ⎟ ⎞<br />

w N<br />

, (5a)<br />

ρw<br />

( 555)<br />

N<br />


and where ai are empirical coefficients (0.366, -3.067, 1.93, 0.649, -1.532, respectively).<br />

We also tested the regionally tuned OC4L algorithm proposed by Cota et al. [2004] for<br />

the Arctic ocean,<br />

b0<br />

b1<br />

* R<br />

CHL 10<br />

+<br />

= , (5b)<br />

where R is the same maximum band ratio that is used in OC4v4 (0.592 and -3.609 for b1<br />

and b2, respectively).<br />

For testing the influence of sea ice on the IOP retrieval, we examine the quasi-<br />

analytical algorithm (QAA) [Lee et al., 2002], as modified here (see Appendix V.3).<br />

Finally, we applied an empirical algorithm <strong><strong>de</strong>s</strong>igned to estimate the ratio between the<br />

CDOM and total absorption coefficients [Bélanger et al., submitted manuscript, 2006]:<br />

⎡ a<br />

[ ρ ] ( 412)<br />

[ ρ ] ( 490)<br />

CDOM ⎤<br />

⎛ w ⎞ ⎛ w ⎞<br />

N<br />

N<br />

( 412)<br />

= c0<br />

+ c1<br />

⋅ log ⎜ ⎟<br />

10<br />

+ c2<br />

⋅ log ⎜ ⎟<br />

10<br />

+ c3<br />

⋅ log10<br />

( [ ρ w ] ( 555)<br />

/ π ) , (6)<br />

⎥<br />

N<br />

a<br />

⎜ [ ρ ] ( 555)<br />

⎟ ⎜ [ ρ ] ( 555)<br />

⎟<br />

⎢<br />

⎣<br />

t<br />

⎦<br />

⎝<br />

w<br />

N<br />

where ci are empirical coefficients (-0.51, -0.54, 0.48 and -0.45, respectively).<br />

V.3. Results and discussion<br />

V.3.1. Impact of the adjacency effect on the retrieval of [ρw]N, CHL and IOPs<br />

⎠<br />

Figure V.3 shows the absolute difference between the retrieved [ρw]N after<br />

application of the atmospheric correction algorithm to the simulated ρTOA in presence and<br />

absence of adjacency effect (Δ[ρw]N), at 443 and 555 nm for clear water (CHL=0.05 mg<br />

m -3 ; fCDOM=50%), five types of sea ice, and for low and high concentrations of aerosols<br />

(τa(550) = 0.03 and 0.2 respectively). The difference is always larger at 443 than at 555<br />

nm because of the properties of the atmospheric light scattering. In<strong>de</strong>ed, the adjacency<br />

effect is a result of increased scattering by atmospheric molecu<strong>les</strong> and aerosols with<br />

<strong>de</strong>creasing wavelength (referred to as Rayleigh and Mie scattering) [Tanré et al., 1979].<br />

Therefore, a larger number of blue photons reflected by the ice environment are scattered<br />

above the targeted water <strong>sur</strong>face compared to the number of green photons. Accordingly,<br />

the adjacency effect is even smaller in the near infrared part of the spectrum, where the<br />

atmospheric correction algorithm estimates the aerosol type and concentration. The<br />

adjacency effect is only partly removed by atmospheric correction. Although the increase<br />

in the NIR signal is interpreted as an increase in aerosol concentration, and thus<br />

138<br />

⎝<br />

w<br />

N<br />


subtracted from ρTOA, the atmospheric correction algorithm does not totally remove the<br />

additional signal at shorter wavelengths that result from the adjacency effect. This is<br />

because the adjacency effect is mainly due to the molecular scattering, which has a<br />

steeper spectral <strong>de</strong>pen<strong>de</strong>ncy than aerosols scattering (λ -4 versus ~λ 0 to λ −2 ). This also<br />

explains why the adjacency effect has <strong>les</strong>s impact when the aerosol concentration<br />

increases and Rayleigh scattering becomes relatively <strong>les</strong>s important (Fig. V.3). Un<strong>de</strong>r<br />

clear skies (i.e., τa(550) = 0.03), the total atmospheric scattering approaches the<br />

molecular spectral <strong>de</strong>pen<strong>de</strong>ncy resulting in a strong spectral <strong>de</strong>pen<strong>de</strong>nce for the<br />

adjacency effect. Un<strong>de</strong>r a turbid sky (i.e., τa(550) = 0.2), in contrast, the spectral<br />

<strong>de</strong>pen<strong>de</strong>ncy of the adjacency effect approaches the aerosol scattering <strong>de</strong>pen<strong>de</strong>ncy (∼λ −1 )<br />

and the contamination is better removed by the atmospheric correction algorithm.<br />

Figure V.3. Error on [ρw]N (<strong>de</strong>noted as Δ[ρw]N) at 443 and 555 nm for four<br />

types of sea ice and two concentrations of aerosol (τa=0.03 and 0.2), as a<br />

function of the distance from the ice edge. The dashed line represents the<br />

acceptable error limit on [ρw]N at 443 nm (i.e. ±2 x 10 -3 ) for precise<br />

chlorophyll estimation [Antoine and Morel, 1999].<br />

139


As expected, the extent of the adjacency contamination also <strong>de</strong>pends on the<br />

magnitu<strong>de</strong> of sea ice reflectance. Δ[ρw]N is larger when the landfast ice is covered by<br />

fresh snow that reflects >90% of the inci<strong>de</strong>nt light, and <strong>de</strong>creases by a factor of ~4 for<br />

<strong>les</strong>s reflective sea ice <strong>sur</strong>faces (Figs. V.3a-d). If we take 0.002 as an acceptable error limit<br />

on [ρw]N(443) for good CHL retrieval [Antoine and Morel, 1999], then the adjacency<br />

effect would be significant within 20-25 km of the landfast ice <strong>sur</strong>face covered by fresh<br />

or melting snow (Figs. V.3a-b), and within 5-15 km for an ice floe <strong>sur</strong>face (Figs. V.3c-d).<br />

The impact of the adjacency effect on the bio-optical parameters retrieval is<br />

examine for τa(550) = 0.1 with an adjacent target covered by dry fresh snow. Therefore,<br />

with this setting, the effect is nearly at its maximum. Figure V.4 shows the impact CHL<br />

estimation. Only the results for water-leaving spectra generated with fCDOM = 50% and<br />

SPM NAP = 0 g m -3 are presented here because they are more representative of the high<br />

latitu<strong>de</strong> open ocean waters (see Section V.2.2.2). The dashed lines represent SeaWiFS<br />

mission requirement of 35% error on CHL [Hooker and McClain, 2000]. Interestingly,<br />

the adjacency effect can lead to an over- and un<strong>de</strong>r-estimation of chlorophyll<br />

concentration <strong>de</strong>pending on the actual chlorophyll a concentration. When the<br />

concentration is low (0.05 mg m -3 ), the CHL is increasingly overestimated when<br />

approaching the ice edge. This is because, at low CHL, the blue-to-green radiance ratio<br />

is higher for seawater than for the photons originating from ice and scattered by the<br />

atmosphere to the sensor. In this case, the adjacency effect tends to <strong>de</strong>crease the blue-to-<br />

green radiance ratio, which leads to CHL overestimates. In contrast, when the initial CHL<br />

is >0.5 mg m -3 , the blue-to-green radiance ratio is lower for seawater than for the photons<br />

reflected by ice and scattered by the atmosphere to the sensor. In this case, the adjacency<br />

effect tends to increase the blue-to-green reflectance ratio, which leads to CHL<br />

un<strong>de</strong>restimates. In general, the error on CHL falls within the 35% range for low initial<br />

concentration (0.05 mg m -3 ) but becomes significant when CHL is >0.5 mg m -3 . In the<br />

worst case simulated (i.e. when CHL = 5.0 mg m -3 , τa = 0.03, and sea ice is covered by<br />

fresh snow), the retrieved CHL are of unacceptable reliability (i.e., error >35%) within 20<br />

km from an ice edge (Table V.1). Compared to OC4v4 algorithm, the Arctic OC4L<br />

algorithm shows somewhat higher sensitivity to the adjacency contamination (Fig. V.4;<br />

Table V.1).<br />

140


Figure V.4. Ratio of CHL estimated from OC4v4 (solid line) and Arctic<br />

OC4L (dotted line) when adjacency effect are present to CHL estimated<br />

without adjacency effect as a function of the distance from the ice edge.<br />

The results are presented for three initial concentrations of chlorophyll<br />

(with fCDOM =50% and SPM NAP =0.0 g m -3 ) as indicated, τa=0.1 and for<br />

four types of sea ice. The dashed lines represent a 35% error range.<br />

Table V.1. Distance (in km) from ice eg<strong>de</strong> within which the error on CHL > 35% (Rice<br />

fresh snow). Values in parenthesis are for the OCL4 algorithm.<br />

Initial CHL τa<br />

(mg m -3 ) 0.03 0.1 0.2<br />

0.05 0 (0) 0 (0) 0 (10)<br />

0.5 12 (14) 10 (17) 8.5 (11)<br />

5.0 19 (20) 17 (19) 17 (18)<br />

Figure V.5 shows the impact of the adjacency effect on at retrieval at 443 nm<br />

when the nearby ice is covered by fresh snow and aerosols are at mo<strong>de</strong>rate concentration<br />

(τa(550) = 0.1). Here the error on at(443) is examined for the 20 water-leaving spectra<br />

presented on Fig. V.2. As for the CHL retrieval, we assume an error acceptable when it is<br />

141


<strong>les</strong>s than 35%. In all cases, at(443) is increasingly un<strong>de</strong>restimated when approaching the<br />

ice edge. At low CHL (0.05 mg m -3 ) and fCDOM (50%), the errors are


Figure V.6 shows results similar to those in Fig. V.5, but for the retrieval of the<br />

particulate backscattering coefficient (bbp) at 555 nm. In general, bbp(555) is increasingly<br />

overestimated when approaching the ice edge when CHL is


Figure V.6. Same as Fig. V.5, but for bbp estimated at 555 nm using the<br />

QAA.<br />

Figure V.7. Same as Fig. V.5, but for the ratio [aCDOM/at] estimated at 412<br />

nm.<br />

144


The impact of the adjacency effect on ocean color data is illustrated by the pixels<br />

extracted along a 37-km transect starting from 70.24°N and 124.6°W to 70.54°N and<br />

124.3°W taken from the SeaWiFS scene acquired on the 31 st of May 2004 over the<br />

Amundsen Gulf (70.25°N; 124.5°W) (Fig. V.8). The SeaWiFS level 1a image with<br />

Local-Area-Coverage (LAC) resolution (1 km) received by the University of Fairbanks in<br />

Alaska was obtained from the Distributed Active Archive Center (DAAC) and processed<br />

to Level 2 using SeaDAS 4.7. Note that this transect is located in clear waters adjacent to<br />

the offshore ice edge. Therefore, it can be reasonably assumed that all bio-optical<br />

variab<strong>les</strong> do not vary much in that area and that any change in them can be ascribed to the<br />

adjacency effect. The aerosol optical thickness (τa) ranges between 0.10 and 0.11 away<br />

from the ice edge (> 40 km), while it increases to ~0.125 toward the ice edge (Fig. V.8a).<br />

This increase in τa is most likely due to the effect of adjacent ice edge, which had not<br />

started to melt significantly at that time, as corroborated by in situ observations in this<br />

area during the CASES expedition (data not shown). At 443 nm, the [ρw]N increased from<br />

~0.01 offshore to 0.038 close to the ice edge, while [ρw]N(555) increased from 0.008 to<br />

0.013 (Fig. V.8b). Consequently, the CHL estimation <strong>de</strong>creases from ~0.55 to ~0.22 mg<br />

m -3 and from ~0.55 to 0.1 mg m -3 using OC4v4 and OC4L, respectively (Fig. V.8c). Note<br />

that if the chlorophyll concentration was really <strong>de</strong>creasing toward the ice edge, [ρw]N at<br />

555 nm should also <strong>de</strong>crease (see Fig. V.2). These results therefore confirm that the<br />

adjacency effect actually occurs and is not corrected for by current SeaWiFS algorithms.<br />

Interestingly, if the correct concentration is ~0.55 mg m -3 (i.e. the offshore value) the<br />

error on CHL is >35% within the first 10 and 15 km from the ice edge for OC4v4 and<br />

OC4L respectively, as predicted by our numerical simulations (see Fig. V.4). The impact<br />

on the IOPs and [aCDOM/at](412) retrievals is also evi<strong>de</strong>nt on this transect (Figs. V.8d-e),<br />

and coherent with the predicted impact of the adjacency effect.<br />

145


Figure V.8. SeaWiFS along-track transect extracted from an image of the<br />

Amundsen Gulf acquired on the 31 st of May 2004: (a) τa; (b) [ρw]N at 443<br />

and 555 nm; (c) CHL; (d) at and bbp at 443 and 555 nm, respectively; (e)<br />

[aCDOM/at] at 412 nm.<br />

146


V.3.2. Impact of sub-pixel sea ice contamination on the retrieval of [ρw]N, CHL and IOPs<br />

The impact of the sub-pixel sea ice contamination is presented here as a function<br />

of the product between the fraction of ice within a given pixel and its reflectance at 865<br />

nm (σRice(865)). This way, the shape of the Rice rather than its magnitu<strong>de</strong>, produce<br />

different results on the retrieval of ρw. Therefore, results presented in this section must be<br />

interpreted with keeping in mind that ice floes can have a fresh snow- or a melting snow-<br />

liked reflectance spectrum (Fig. V.1).<br />

In general, the absolute error in the normalized water-leaving reflectance, Δ[ρw]N,<br />

<strong>de</strong>creases with increasing σRice(865) (Fig. V.9). This is explained by the overestimation<br />

of the aerosol contribution resulting from the enhanced reflectance in the NIR due to sea<br />

ice. As for the adjacency effect, the contamination is more pronounced at 443 nm than at<br />

555 nm, except in presence of grey ice (Fig. V.9d). The larger negative Δ[ρw]N values at<br />

the shorter wavelengths result from the extrapolation of the aerosol reflectance to the<br />

visible from the NIR. Note that un<strong>de</strong>r turbid atmospheric condition (e.g., τa = 0.2), pixels<br />

become flagged as clouds when σRice(865) gets larger than ca. 0.017 if a one-band<br />

threshold test at 865 nm, as for SeaWiFS, is used for cloud <strong>de</strong>tection (resulting in the<br />

truncation of the black full lines in Fig. V.9). Because the spectral shape of Rice in the<br />

NIR influences the selection by the atmospheric correction algorithm of the aerosol<br />

mo<strong>de</strong>ls, the type of sea ice has a significant impact on the retrieval of [ρw]N. For instance,<br />

because the landfast ice covered by melting snow and the grey ice have a steeper<br />

<strong>de</strong>crease towards longer wavelengths (see Fig. V.1), they have the greatest impact on the<br />

retrieval of [ρw]N at all wavelengths (Fig. V.9b,d). Note that the landfast ice cases can<br />

only occur for pixels which are partly overlapping landfast ice, in which case the<br />

adjacency effect would completely mask the sub-pixel contamination effect (e.g., Fig.<br />

V.8) that are computed assuming no contribution from adjacent pixels.<br />

147


Figure V.9. Error on [ρw]N at 443 and 555 nm for four types of sea ice and<br />

two concentrations of aerosol (τa=0.03 and 0.2), as a function of the<br />

product of fractional sea ice <strong>sur</strong>face concentration and its reflectance<br />

values at 865 nm, σRice(865). The dashed line represents the acceptable<br />

error limit on [ρw]N at 443 nm.<br />

To illustrate the impact of sub-pixel contamination on the aerosol mo<strong>de</strong>l selection,<br />

Fig. V.10 shows the angstrom exponent (α) as a function of σRice(865) for the different<br />

types of sea ice. Here, the α refers to the exponent of the power law in λ −α for the spectral<br />

variation in τa. The α increases by a factor >2 for the landfast ice covered by melting<br />

snow and the grey ice floe, which explain the greatest Δ[ρw]N values obtained for these<br />

type of ice (Figs. V.9b,d). For spectrally neutral Rice spectra such as for the landfast ice<br />

covered by fresh snow and the ice floe covered by melting snow (see Fig. V.1), α<br />

remains relatively unchanged in presence of sea ice. Although these sea ice types do not<br />

significantly affect the selected aerosol mo<strong>de</strong>ls, they still introduce an error in [ρw]N. In<br />

fact at TOA, σRice is modulated by the diffuse transmittance (tv and t0 in Eq. 4) which<br />

<strong>de</strong>creases exponentially with <strong>de</strong>creasing wavelengths. For example, the tvt0 values are<br />

148


0.71, 0.86 and 0.96 at 443, 555 and 865 nm, respectively, for a τa(865) value of 0.1 and<br />

the sun-viewing geometry adopted here.<br />

Figure V.10. Retrieved Angström exponent parameter (α) using the<br />

atmospheric correction algorithm, as a function of σRice, for τa=0.1 and<br />

four types of sea ice.<br />

Figure V.11 shows the implication of errors in [ρw]N on the CHL estimation for<br />

the four sea ice types (with τa(550) = 0.1). As for [ρw]N, the error on CHL is dramatic in<br />

presence of the landfast ice (Figs. 11a-b). For CHL > 5.0 mg m -3 , a 10-fold<br />

overestimation could be reached, even for a small amount of sea ice within the pixel<br />

(σRice~0.05-0.01; Fig. V.11b). For low CHL (0.05 mg m -3 ), the un<strong>de</strong>restimation can be<br />

important in the presence of melting snow (Fig. 11b) due to the fact that the band ratio<br />

tends toward infinity. In presence of ice floes, strong overestimation of CHL occur when<br />

the actual concentration is >0.5 mg m -3 (Figs. 11c-d). As for the adjacency effect, the<br />

OC4L algorithm is more sensitive to the contamination than OC4v4 is.<br />

Figures V.12 and V.13 show the impact of the sub-pixel contamination by sea ice<br />

covered by fresh snow on the retrievals of the at(443) and bbp(555), respectively. In<br />

general, the negative Δ[ρw]N result in an overestimation of at(443) (Fig. V.12). The error<br />

increase when both CHL and fCDOM increase. At low CHL (0.05 mg m -3 ), the error is<br />


y


Figure V.12. Ratio of at estimated at 443 nm using QAA when sea ice is<br />

present within the water pixel to at estimated with no sea ice as a function<br />

of σRice(865). The results are presented for each water type as shown in<br />

figure 2 and for the landfast ice covered by fresh snow. The dashed lines<br />

represent a 35% error range.<br />

Figure V.13. Same as Fig. V.12, but for bbp estimated at 555 nm using<br />

QAA.<br />

151


The impact of sub-pixel contamination on the [aCDOM/at](412) retrieval is shown<br />

in Fig. V.14. The un<strong>de</strong>restimation of the blue reflectance produces an overestimation of<br />

[aCDOM/at](412) in all situations. When CHL is >0.5 mg m -3 and SPM NAP is


high aerosol optical thickness (not shown) suggests that it is an artifact resulting from sea<br />

ice contamination. The artifact is, in counterpart, restricted to only a few pixels within the<br />

ice field.<br />

Figure V.15. Example of sea ice contamination on the CHL estimation<br />

(OC4v4) for a SeaWiFS scene of the Beaufort Sea acquired on the 8 th of<br />

September 2002. Upper panel is the true color composite showing the sea<br />

ice field offshore.<br />

V.3.3. Detection of pixels affected by sea ice (adjacency effect and sub-pixel<br />

contamination)<br />

In this section, we propose a simple method to <strong>de</strong>tect pixels contaminated by the<br />

adjacency effect. Due to the highly dynamic nature of summertime sea ice and its quickly<br />

153


changing optical properties, the adjacency effect could not easily be corrected or<br />

predicted without external information. Although sea ice distribution and reflectance<br />

properties may be obtained from other satellite sensors (e.g., SSMI, AVHRR, ERS), the<br />

sensitivity, or time and space sca<strong>les</strong> do not always match those of Ocean Color data.<br />

Therefore, a spectral test on the retrieved [ρw]N spectrum is a straightforward way to<br />

<strong>de</strong>tect and eliminate contaminated pixels. As the adjacency effect increase with<br />

<strong>de</strong>creasing wavelengths (see section V.3.1), the blue part of the spectrum is most<br />

appropriate to <strong>de</strong>tect such contamination. As mentioned above, [ρw]N spectra observed at<br />

high latitu<strong>de</strong> show a typical <strong>de</strong>pression at λ < 450 nm [Reynolds et al., 2001; Wang and<br />

Cota, 2003]. Figure V.16 shows the variation of the ratio [ρw]N(412)/[ρw]N(443) as a<br />

function of [ρw]N(555) (which is an indicator of the water turbidity), for the CASES data<br />

set and for mo<strong>de</strong>led reflectance spectra generated using the reflectance mo<strong>de</strong>l <strong><strong>de</strong>s</strong>cribed<br />

above (with values for CHL and SPM NAP as specified above, and with fCDOM = 20, 50<br />

and 80%). So the ratio of [ρw]N(412)/[ρw]N(443) is generally < 1, and <strong>de</strong>crease as the<br />

turbidity increase. This behavior is opposite to the adjacency contamination for which the<br />

increase in [ρw]N at 555 nm is accompanied by an increase in the reflectance ratio<br />

[ρw]N(412)/[ρw]N(443). The black line on Fig. V.16 represents an upper limit, though<br />

arbitrary, below which the ratio [ρw]N(412)/[ρw]N(443) should be for a given [ρw]N(555)<br />

value. So the contaminated pixels are i<strong>de</strong>ntified as soon as the inequality<br />

[ ρ ]<br />

[ ρ ]<br />

is satisfied. The threshold value is<br />

w N ≥<br />

w<br />

N<br />

( 412)<br />

( 443)<br />

Threshold<br />

Threshold = . 64 − 0.<br />

14log<br />

([ ρ ] ( 555))<br />

. (7)<br />

0 10 w N<br />

An example of application to a SeaWiFS image of the Southeastern Beaufort Sea<br />

is shown in Figs. V.16 and V.17. The image presents a wi<strong>de</strong> variety of water types from<br />

turbid waters inshore to relatively clear waters offshore. Pixels <strong>sur</strong>rounding the central<br />

Arctic packed ice are well i<strong>de</strong>ntified. As expected, the violet-to-blue reflectance ratio of<br />

those pixels increases together with [ρw]N(555) (Fig. V.16). In addition, a number of<br />

turbid pixels, located on the continental shelf east of the Mackenzie River and adjacent to<br />

the remaining landfast ice, are also flagged for adjacency contamination (Fig. V.17).<br />

154


Interestingly, those pixels would not have been <strong>de</strong>tected if a fix threshold (i.e. 1.0) had<br />

been adopted instead of using equation 7. Some pixels within the Amundsen Gulf, away<br />

from the ice edges, were neverthe<strong>les</strong>s flagged, which is likely a result from the adjacency<br />

effect due to the presence ice floe apart from a continuous sea ice cover. The robustness<br />

of the method was examined by applying it to several SeaWiFS scenes from the Beaufort<br />

Sea. In some situations when the actual ratio [ρw]N(412)/[ρw]N(443) was very low (i.e. in<br />

high CDM waters), only pixels where the contamination was extreme were <strong>de</strong>tected.<br />

Clearly, for an application of this flag to images from Antarctica or North Atlantic, an<br />

adjustment of the threshold is likely necessary since the abundance of CDM is lower in<br />

those regions compared to the Arctic coastal waters [Siegel et al., 2005].<br />

Figure V.16. Reflectance ratio between [ρw]N at 412 and [ρw]N at 443 nm<br />

as a function of [ρw]N at 555 nm for: 1) in situ mea<strong>sur</strong>ements of the waterleaving<br />

reflectance conducted in the Southeastern Beaufort Sea [Bélanger<br />

et al., submitted manuscript. 2006] (closed circ<strong>les</strong>); 2) mo<strong>de</strong>led waterleaving<br />

reflectance with the bio-optical mo<strong>de</strong>l presented in Appendix V.1<br />

(closed squares); and 3) SeaWiFS pixels extracted from the image<br />

presented on Fig. V.17 (light grey dots).<br />

155


Figure V.17. Example of application of the adjacency flag to a SeaWiFS<br />

scene from the Southeastern Beaufort Sea acquired on the 16 th of June<br />

1998. The Rrs at 555 nm is shown on colored log scale as indicated on the<br />

right; the flagged pixels for adjacency are i<strong>de</strong>ntified with dark red color.<br />

The sea ice and land are shown in dark and light grey respectively.<br />

Because of known problem with the standard SeaWiFS atmospheric<br />

correction over the turbid waters, we apply the algorithm proposed by<br />

Ruddick et al. [2000].<br />

In this study we do not propose a method to <strong>de</strong>tect the sub-pixel contamination<br />

specifically. It is, however, possible to i<strong>de</strong>ntify contaminated pixel when they are<br />

relatively isolated by applying a spatial analysis of the distribution of the aerosol optical<br />

thickness. In<strong>de</strong>ed, sub-pixel contamination will introduce some noise in the τa which may<br />

be removed using a filtering scheme similar to the one proposed by Franz et al. [2003]<br />

for reducing noise in the SeaWiFS NIR bands. In addition, the overcorrection for aerosol<br />

due to sea ice results in negative water-leaving reflectance at 412 nm which can be used<br />

as a data quality control flag. A visual examination of several individual full resolution<br />

SeaWiFS scenes from the Southeastern Beaufort Sea led us to the conclusion that the<br />

sub-pixel contamination is not an extensive phenomenon in this area. It appears that the<br />

melting sea ice field is generally <strong>de</strong>nse enough to be flagged as cloud, or to produce<br />

adjacency effect. Clearly from this visual analysis, we conclu<strong>de</strong> that adjacency effect is<br />

far more important than sub-pixel contamination and must be accounted for when<br />

processing and interpreting Ocean Color data.<br />

156


V.4. Summary and conclusions<br />

The adjacency effect and the sub-pixel contamination by a small amount of sea<br />

ice within an ocean pixel affect the quality of the water-leaving reflectance retrieval from<br />

satellite Ocean Color data at high latitu<strong>de</strong>. Both contaminations introduce large errors in<br />

[ρw]N in the blue wavelength region reducing the quality of the bio-optical parameters<br />

<strong>de</strong>rived from the water-leaving spectrum. The bio-optical parameters retrieval evaluated<br />

in this study are the chlorophyll a, total absorption and partic<strong>les</strong> backscattering<br />

coefficients respectively, as well as the ratio between CDOM to total absorption<br />

coefficient. While adjacency effect results in an overestimation of the [ρw]N, sub-pixel<br />

contamination result in a systematic un<strong>de</strong>restimation of the water-leaving reflectance in<br />

the blue. These errors result either in un<strong>de</strong>r- and over-estimation of the CHL <strong>de</strong>pending<br />

on the actual chlorophyll a concentration. The impact on IOP retrieval is more<br />

straightforward: 1) adjacency effect results in an un<strong>de</strong>restimation of at and [aCDOM/at],<br />

and an overestimation of bbp; 2) sub-pixel contamination results in an overestimation of at<br />

and [aCDOM/at], and an un<strong>de</strong>restimation of bbp. Because adjacency effect appear to be<br />

more frequent compared to the sub-pixel contamination, a simple spectral test on the<br />

retrieved [ρw]N was proposed to flag contaminated pixels. The method relies on the fact<br />

that adjacency contamination increases both the reflectance ratio between 412 and 443<br />

nm and the [ρw]N value at 555 nm, which is opposite to trend observed when seawater<br />

turbidity increases (i.e. [ρw]N(412)/[ρw]N(443) <strong>de</strong>creases when [ρw]N(555) increases).<br />

Clearly the adjacency effect must be accounted for when the aim is to use Ocean<br />

Color data for quantifying bio-optical properties and biogeochemical processes in ice<br />

covered seas. The method proposed to eliminate contaminated pixel by this effect could<br />

be easily inclu<strong>de</strong>d in the data processing chain, but more data in polar seas are nee<strong>de</strong>d to<br />

adjust the threshold. Neverthe<strong>les</strong>s, because open water nearby sea ice are important<br />

biologically speaking, an extension of the atmospheric correction scheme to inclu<strong>de</strong> a<br />

correction for adjacency effect is <strong><strong>de</strong>s</strong>irable. This might be done by including additional<br />

LUTs generated accounting for the adjacency effect. Those LUTs would be used in the<br />

processing only when contaminated pixels are <strong>de</strong>tected. On the other hand, the sub-pixel<br />

157


contamination, which is not an extensive problem, could probably i<strong>de</strong>ntified using<br />

simultaneous high resolution mea<strong>sur</strong>ements (e.g. MODIS 250-m and MERIS 330-m<br />

channels).<br />

Finally, it must be stressed that similar effect could also be observed world-wi<strong>de</strong><br />

in the <strong>sur</strong>rounding of thick cloud. In<strong>de</strong>ed the adjacency-flagging procedure proposed<br />

above frequently eliminates the pixels located nearby clouds in the study area. Further<br />

study on the adjacency contamination due to thick cloud found at different altitu<strong>de</strong> in the<br />

atmosphere is necessary. The contamination might be particularly important in the<br />

equatorial zones where irradiance and could cover are high.<br />

Acknowledgment<br />

This study was ma<strong>de</strong> possible with financial support from the Fonds Québécois pour la<br />

Recherche <strong>sur</strong> la Nature et <strong>les</strong> Technologies (FQRNT) (to SB), the Fonds France-Canada<br />

pour la Recherche (FFCR to MB) and Indian and Northern Affairs Canada (to SB). We<br />

are grateful to the Laboratoire d’Optique Atmosphérique (LOA) for providing the 6S<br />

radiative transfert co<strong>de</strong>. Pierre Larouche and Dave Barber are kindly acknowledged for<br />

providing instrumentation and field assistance. Yannick Huot, Fabrizio D’Ortenzio and<br />

André Morel are specially thanked for their valuable comments on the manuscript. We<br />

thank the CCGS Amundsen cruise members for their enthusiastic help onboard the ship.<br />

This is a contribution to the Canadian Arctic Shelf Exchange Study (CASES) un<strong>de</strong>r the<br />

overall direction of L. Fortier.<br />

Appendix V.1. Water reflectance mo<strong>de</strong>l<br />

The water-leaving reflectance spectra were generated using the semi-analytical<br />

mo<strong>de</strong>l proposed by Park and Ruddick [2005] valid for both case 1 and case 2 waters. The<br />

remote sensing reflectance (Rrs) is expressed using a forth-or<strong>de</strong>r polynomial parameter as:<br />

158


4 ⎡ b ⎤ b<br />

Rrs ( λ)<br />

= ∑ gi<br />

( λ)<br />

⎢ ⎥ ( λ)<br />

, (A1)<br />

i=<br />

1 ⎣at<br />

+ bb<br />

⎦<br />

where bb and at are the total backscattering and absorption coefficients respectively, and<br />

gi are coefficients tabulated in a look-up table (LUT) for as a function of Sun-Viewing<br />

geometry and of a phase function parameter, γb(λ). γb(λ) is <strong>de</strong>fined as the relative<br />

contribution of partic<strong>les</strong> to the total backscattering (i.e. bbp(λ)/bb(λ)).<br />

The spectral total absorption coefficient is subdivi<strong>de</strong>d into four additive<br />

components as:<br />

159<br />

i<br />

a ( ) = a ( λ)<br />

+ a ( λ)<br />

+ a ( λ)<br />

+ a ( λ)<br />

, (A2)<br />

t λ w ϕ NAP<br />

CDOM<br />

where subscripts w, φ, NAP and CDOM stand for water, phytoplankton, non-algal<br />

partic<strong>les</strong>, and colored dissolved organic matter (CDOM), respectively. Similarly, the<br />

spectral total backscattering coefficient is express with three components as:<br />

b ( ) = b<br />

~<br />

( λ)<br />

+ b<br />

~<br />

b ( λ)<br />

+ b b ( λ)<br />

, (A3)<br />

b λ bw bϕ<br />

φ<br />

bNAP NAP<br />

where bφ and bNAP are the volume scattering coefficients for phytoplankton and NAP<br />

respectively, and b bϕ<br />

~ and b bNAP<br />

~<br />

are the ratio of backscattering to scattering coefficients<br />

for phytoplankton and NAP respectively. Spectral values for aw and bbw are taken from<br />

Pope and Fry [1997] and Morel [1974], respectively.<br />

Spectral absorption coefficient for phytoplankton and its co-varying materials are<br />

computed as a function of CHL using the bio-optical mo<strong>de</strong>l proposed by Morel and<br />

Maritorena [2001] (hereafter referred as MM01). To avoid confusion with current<br />

literature on the phytoplankton absorption [Bricaud et al., 1995], the IOP computed using<br />

MM01 will be <strong>de</strong>noted as acase1 and bbcase1 (thus co-varying CDOM and NAP absorption<br />

are inclu<strong>de</strong>d in acase1(λ)). Non-algal particle absorption varies as a function of SPM NAP<br />

according to:<br />

[ S ( 443 λ)<br />

]<br />

*<br />

NAP<br />

a ( λ) a ( 443)<br />

SPM exp − , (A4)<br />

NAP<br />

= NAP<br />

NAP<br />

*<br />

where a ( 443)<br />

is the mass-specific absorption coefficient <strong>de</strong>fined as ratio of aNAP(443)<br />

NAP<br />

to the SPM NAP (in m 2 g -1 ), and SNAP is the spectral slope of the aNAP(λ) spectrum. Here,<br />

*<br />

a ( 443)<br />

= 0.0534 m 2 g -1 and SNAP = 0.010 nm -1 . Those values were observed in the<br />

NAP<br />

turbid waters of the Mackenzie River [S. Bélanger and P. Larouche, unpublished data].


Spectral CDOM absorption is generated assuming a given contribution of the<br />

CDOM to the non-water constituents absorption coefficients at 443 nm (fCDOM) as:<br />

and<br />

a<br />

CDOM<br />

( 443)<br />

[ a ( 443)<br />

+ a ( 443)<br />

]<br />

f CDOM NAP<br />

case1<br />

= , (A5)<br />

( 1 − f )<br />

CDOM<br />

160<br />

[ S ( 443 ) ]<br />

a ( λ) a ( 443)<br />

exp − λ . (A6)<br />

CDOM<br />

= CDOM<br />

CDOM<br />

Here, SCDOM is fixed at 0.022 nm -1 , the average value mea<strong>sur</strong>ed in the Southeastern<br />

Beaufort Sea [Bélanger et al., submitted manuscript, 2006].<br />

The scattering coefficient at 555 nm for the phytoplankton and its co-varying<br />

partic<strong>les</strong> is obtained using a statistical relationship established for Case 1 Loisel and<br />

Morel [Loisel and Morel, 1998]:<br />

−0.<br />

766<br />

bcase 1(<br />

555)<br />

= 0.<br />

41CHL<br />

. (A7)<br />

To obtain the backscattering coefficient for phytoplankton and its co-varying partic<strong>les</strong>,<br />

the backscattering ratio is computed following MM01 as:<br />

~<br />

= 0.<br />

002 + 0.<br />

01 0.<br />

5 − 0.<br />

25log<br />

( CHL)<br />

. (A8)<br />

bbcase1 [ ]<br />

The spectral backscattering coefficients are obtained using a power law function in<br />

λ ν where the exponent increases from -1 to 0 with respect to CHL as:<br />

calculated as:<br />

= 0. 5*<br />

(log10<br />

( CHL)<br />

− 0.<br />

3)<br />

ν , 0.03 < CHL < 2.0 mg m -3<br />

10<br />

ν = 0, CHL > 2.0 mg m -3 . (A9)<br />

For NAP, which does not covary with phytoplankton, the scattering coefficient is<br />

b ( 555)<br />

*<br />

NAP<br />

*<br />

NAP<br />

= bpNAP<br />

( 555)<br />

SPM , (A10)<br />

*<br />

where b ( 555)<br />

is the mass-specific scattering coefficient <strong>de</strong>fined as ratio of bNAP(443)<br />

pNAP<br />

to the SPM NAP (in m 2 g -1 *<br />

). A b ( 555)<br />

value of 0.5 m 2 g -1 , typical of coastal waters<br />

pNAP<br />

[Babin et al., 2003], was adopted. The backscattering ratio for NAP was assumed to be<br />

1.83%, as mea<strong>sur</strong>ed by Petzold [1972] in the San Diego Harbor. The spectral bbNAP(λ)<br />

values were assumed to follow a λ -1 law.


Appendix V.2. Post-treatment of 6s output<br />

The 6s co<strong>de</strong> computes the ρTOA for a circular target of radius r, <strong>sur</strong>roun<strong>de</strong>d by an<br />

environment of different reflectivity (see Tanré et al. [1981]). Briefly, the reflectance at<br />

the TOA, ρTOA, is express as<br />

TOA<br />

[ ρ + ρ ( r)<br />

ρ ( r)<br />

]<br />

ρ = + , (A11)<br />

tg path `t<br />

arg et env<br />

where tg is the total gas transmission (i.e. O2, H2O and O3), ρpath is the atmospheric path<br />

reflectance, ρtarget is the target reflectance, and ρenv is the environment reflectance. The<br />

reflectances of the target (e.g. water-leaving, ρw) and the environment (e.g. sea ice, ρice)<br />

are expressed as<br />

−τ<br />

T ( θs<br />

) ⎛ cos( θ ) ⎞ v<br />

ρ t arg et ( r) = ⎜ ρwe<br />

⎟<br />

(A12)<br />

1−<br />

〈 ρ〉<br />

( r)<br />

S ⎝ ⎠<br />

and<br />

T ( θs<br />

)<br />

ρ env(<br />

r) = ( 〈 ρ〉<br />

( r)<br />

td<br />

( θv<br />

) ) , (A13)<br />

1−<br />

〈 ρ〉<br />

( r)<br />

S<br />

where T(θs) is the total transmittance (direct + diffuse), S is the spherical albedo of the<br />

atmosphere,<br />

e<br />

−τ<br />

cos( θv<br />

)<br />

is the direct upward transmittance, td is the diffuse transmittance,<br />

and 〈ρ 〉 (r)<br />

is a spatial average of each pixel reflectance over the whole <strong>sur</strong>face. Here<br />

〈ρ 〉 (r)<br />

is a function of the atmospheric conditions and the <strong>sur</strong>face reflectance which is<br />

given by<br />

〈 ρ 〉 ( r ) = ρ F(<br />

r)<br />

+ ( 1−<br />

F(<br />

r))<br />

R , (A14)<br />

w<br />

ice<br />

where F(r) is the environment function, which <strong>de</strong>pends on the atmospheric conditions<br />

(for <strong>de</strong>tails see Tanré et al. [1981] and Vermote et al. [1997]). ρtarget and ρenv were<br />

computed for several r (0.5 to 100 km), and four types of sea ice (Section 2.2.1). Since<br />

more or <strong>les</strong>s linear ice edges rather than circular target is expected, ρtarget(r) and ρenv(r)<br />

were used to estimate the contribution of the environment and the target at TOA of a<br />

pixel located at distance D0 from an ice edge as (see Fig. V.A1)<br />

π<br />

2<br />

edge 1<br />

arg et ( D0) = ∫ ρt<br />

arg et<br />

2π<br />

−π<br />

2<br />

ρt ( D)<br />

dα<br />

(A15)<br />

and<br />

161


π<br />

2<br />

1<br />

D0 ) =<br />

2 ∫ ρenv<br />

π −π<br />

2<br />

edge<br />

ρenv (<br />

( D)<br />

dα<br />

, (A16)<br />

with cos ( )<br />

1 −<br />

D = D α .<br />

0<br />

When 6s is run for an inhomogeneous <strong>sur</strong>face, the specular reflection of the diffuse<br />

skylight toward the sensor is not account for. The sky radiance reflected by the air-water<br />

interface transmitted to the TOA, ρsky, can be estimated as<br />

ρ<br />

sky<br />

π Lsky<br />

⋅ ρ Fresnel<br />

= tv<br />

(A17)<br />

ε F θ )<br />

0<br />

cos( 0<br />

where ρFresnel is the reflectance of a flat air-water interface given by Fresnel’s formula<br />

(typically ~2.3% for satellite viewing angle), Lsky is the radiance at <strong>sur</strong>face coming from<br />

the diffuse sky, and tv is the upward transmittance from pixel to sensor. Assuming a<br />

perfectly isotropic sky, then<br />

diff<br />

Ed<br />

L sky = , (A18)<br />

0.<br />

5π<br />

where<br />

diff<br />

E d is the diffuse downward irradiance at <strong>sur</strong>face and 0.5 is the mean cosine for<br />

an isotropic sky. Since<br />

path reflectance.<br />

diff<br />

Ed and tv are given by 6s, ρsky was estimated and ad<strong>de</strong>d to the<br />

Figure V.A1. Schematic representation of a water pixel located at distance<br />

D0 from an ice edge.<br />

162


Appendix V.3. Modification of the QAA<br />

The quasi-analytical algorithm (QAA) proposed by Lee et al. [2002] was<br />

modified in respect to 1) the formulation of Rrs versus the IOP, 2) the empirical<br />

formulations for the absorption coefficient at reference wavelength (λ0), and 3) the use of<br />

a single value for the exponent of the power law used to mo<strong>de</strong>l the particulate<br />

backscattering spectrum (Y).<br />

First, a semi-analytical remote-sensing reflectance mo<strong>de</strong>l <strong><strong>de</strong>s</strong>igned to account for<br />

the bidirectional effects (Eq. A1) [Park and Ruddick, 2005] is used instead of the Gordon<br />

et al. [1988] formulation. Equation A1 requires a priori information on backscattering<br />

nature of the medium (i.e. molecular versus particulate) to operate. Therefore, the original<br />

QAA is used to provi<strong>de</strong> an initial estimate of γb at a reference wavelength. With the initial<br />

value of γb(λ0) and Rrs(λ0, θs, θv, Δφ, W), [bb/(at+bb)](λ0) is estimated by inverting the<br />

fourth-or<strong>de</strong>r polynomial using a numerical solution [Laguerre's method, or the zroots<br />

function in Chap 9.5 of Press et al., 1992]. Two iterations are sufficient to obtain a<br />

difference between two consecutives estimate of γb(λ0)


available on our ac-9. The value of at(650) is calculated empirically as (r 2 = 0.71, N = 14<br />

at-w(555)>0.5aw(555)):<br />

164<br />

⎡ Rrs<br />

( 650)<br />

⎤<br />

−0.<br />

69+<br />

1.<br />

41log10<br />

⎢ ⎥<br />

⎣ Rrs<br />

( 555)<br />

⎦<br />

a ( 650)<br />

= a ( 650)<br />

+ 10<br />

(A20)<br />

t<br />

w<br />

where aw(650)=0.34 [Pope and Fry, 1997]. For continuity in the at retrieval, when at-<br />

w(555) falls between 0.03 and 0.06 m -1 the a linear combination of the two results<br />

obtained with λ0 = 555 and 650 nm is performed following Eq. 20 in Lee et al. [2002].<br />

Third, the spectral partic<strong>les</strong> backscattering coefficient (bbp) is calculated using λ −Y<br />

law with Y = 1.


V.5. References<br />

ACIA (2005), Arctic Climate Impact Assessment, 1042 pp., Cambridge Press University.<br />

Antoine, D., and A. Morel (1996), Oceanic primary production 1. Adaptation of a<br />

spectral light-photosynthesis mo<strong>de</strong>l in view of application to satellite chlorophyll<br />

observations, Global Biogeochem. Cyc<strong>les</strong>, 10(1), 43-55.<br />

Antoine, D., and A. Morel (1999), A multiple scattering algorithm for atmospheric<br />

correction of remotely sensed ocean colour (MERIS instrument): principle and<br />

implementation for atmospheres carrying various aerosols including absorbing<br />

ones, Int. J. Remote Sens., 20(9), 1875-1916.<br />

Arrigo, K. R., D. Worthen, A. Schnell, and M. P. Lizotte (1998), Primary production in<br />

Southern Ocean waters, J. Geophys. Res., 103(C8), 15587-15600.<br />

Babin, M., A. Morel, V. Fournier-Sicre, F. Fell, and D. Stramski (2003), Light Scattering<br />

properties of marine partic<strong>les</strong> in coastal and open ocean waters as related to the<br />

particle mass concentration, Limnol. Oceanogr., 48(2), 843-859.<br />

Behrenfeld, M. J., and P. G. Falkowski (1997), Photosynthetic rates <strong>de</strong>rived from<br />

satellite-based chlorophyll concentration, Limnol. Oceanogr., 42(1), 1-20.<br />

Bélanger, S., H. Xie, N. Krotkov, P. Larouche, W. F. Vincent, and M. Babin (2006),<br />

Photomineralization of terrigenous dissolved organic matter in Arctic coastal<br />

waters from 1979 to 2003: Interannual variability and implications of climate<br />

change, Global Biogeochem. Cyc<strong>les</strong>, doi:10.1029/2006GB002708.<br />

Benner, R., P. Louchouarn, and R. M. W. Amon (2005), Terrigenous dissolved organic<br />

matter in the Arctic Ocean and its transport to <strong>sur</strong>face and <strong>de</strong>ep waters of the<br />

North Atlantic, Global Biogeochem. Cyc<strong>les</strong>, 19, GB2025,<br />

doi:10.1029/2004GB002398.<br />

Bricaud, A., M. Babin, A. Morel, and H. Claustre (1995), Variability in the chlorophyllspecific<br />

absorption coefficients of natural phytoplankton: Analysis and<br />

parameterization, J. Geophys. Res., 100(C7), 13321-13332.<br />

Camill, P. (2005), Permafrost thaw accelerates in boreal peatlands during late-20th<br />

century climate warming, Climatic Change, 68(1-2), 135-152.<br />

Cota, G. F., J. Wang, and J. C. Comiso (2004), Transformation of global satellite<br />

chlorophyll retrievals with a regionally tuned algorithm, Remote Sens. Environ.,<br />

90, 373-377.<br />

Franz, B. A., R. E. Eplee, S. W. Bailey, and M. Wang (2003), Changes to the<br />

atmospheric correction algorithm and retrieval of oceanic optical properties, Chap.<br />

5, in NASA Tech. Memo. 2003-206892, vol. 22, edited by S. B. Hooker, and E. R.<br />

Firestone, pp. 29-33, NASA Goddard Space Flight Center, Greenbelt, Md.<br />

Frey, K. E., and L. C. Smith (2005), Amplified carbon release from vast West Siberian<br />

peatlands by 2100, Geophys. Res. Let., 32(9).<br />

165


Frouin, R., M. Schwindling, and P. Y. Deschamps (1996), Spectral reflectance of sea<br />

foam in the visible and near-infrared: In situ mea<strong>sur</strong>ements and remote sensing<br />

implications, J. Geophys. Res., 101(C6), 14361-14371.<br />

Gordon, H. R. (1997), Atmospheric correction of ocean color imagery in the Earth<br />

Observation System era, J. Geophys. Res., 102(D14), 17081-17106.<br />

Gordon, H. R., O. B. Brown, R. H. Evans, J. W. Brown, R. C. Smith, K. S. Baker, and D.<br />

K. Clark (1988), A semianalytic radiance mo<strong>de</strong>l of ocean color, J. Geophys. Res.,<br />

93(D9), 10909-10924.<br />

Gordon, H. R., and D. K. Clark (1981), Clear Water Radiances for Atmospheric<br />

Correction of Coastal Zone Color Scanner Imagery, Appl. Opt., 20(24), 4175-<br />

4180.<br />

Gordon, H. R., and M. Wang (1994), Retrieval of water-leaving radiance and aerosol<br />

optical thisckness over the oceans with SeaWiFS: a preliminary algorithm, Appl.<br />

Opt., 33(3), 443-452.<br />

Gregg, W. W., and N. W. Casey (2004), Global and regional evaluation of the SeaWiFS<br />

chlorophyll data set, Remote Sens. Environ., 93, 463-479.<br />

Guo, L., and R. W. Macdonald (2006), Source and transport of terrigenous organic matter<br />

in the upper Yukon River: Evi<strong>de</strong>nce from isotope (δ 13 C, Δ 14 C, and δ 15 N)<br />

composition of dissolved colloidal, and particulate phases, Global Biogeochem.<br />

Cyc<strong>les</strong>, 20, GB2011, doi:10.1029/2005GB002593.<br />

Hooker, S. B., and C. R. McClain (2000), The calibration and validation of SeaWiFS data,<br />

Prog. Oceanog., 45, 427-465.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and R. A. Arnone (2002), Deriving inherent optical properties<br />

from water color : a multiband quasi-analytical algorithm for optically <strong>de</strong>ep<br />

waters, Appl. Opt., 41(27), 5755-5772.<br />

Loisel, H., and A. Morel (1998), Light scattering and chlorophyll concentration in case 1<br />

waters: A reexamination, Limnol. Oceanogr., 43(5), 847-858.<br />

Maritorena, S., D. A. Siegel, and A. R. Peterson (2002), Optimization of a semianalytical<br />

ocean color mo<strong>de</strong>l for global-scale applications, Appl. Opt., 41(15), 2705-2714.<br />

Mc Clatchey, R. A., R. W. Fenn, J. E. A. Selby, F. E. Volz, and J. S. Garing, Optical<br />

properties of the Atmosphere, AFCRL-TR-71-079, L.G. Hancom Fiel, Bedford,<br />

Ms., 1971.<br />

McClelland, J. W., S. J. Dery, B. J. Peterson, R. M. Holmes, and E. F. Wood (2006), A<br />

pan-arctic evaluation of changes in river discharge during the latter half of the<br />

20th century, Geophys. Res. Let., 33(6), L06715, doi:10.1029/2006GL025753.<br />

Morel, A. (1974), Optical properties of pure water and pure sea water, in Optical aspects<br />

of Oceanography, edited by E. S. Nielsen, pp. 1-24, Aca<strong>de</strong>mic, New York.<br />

Morel, A., and S. Maritorena (2001), Bio-optical properties of oceanic waters: A<br />

reappraisal, J. Geophys. Res., 106(C4), 7163-7180.<br />

166


Morel, A., and J. L. Mueller (2002), Normalized Water-Leaving Radiance and Remote<br />

Sensing Reflectance: Bidirectional Reflectance and Other Factors, in Ocean<br />

Optics Protocols For Satellite Ocean Color Sensor Validation, Revision 4,<br />

Volume III, edited by C. R. McClain, pp. 32-59, NASA Goddard Space Flight<br />

Space Center, Greebelt.<br />

O'Reilly, J. E., S. Maritorena, G. Mitchell, D. A. Siegel, K. L. Car<strong>de</strong>r, D. L. Garver, M.<br />

Kahru, and C. R. McClain (1998), Ocean color chlorophyll algorithms for<br />

SeaWiFS, Journal of Geophysical Research, 103(C11), 24937-24950.<br />

O'Reilly, J. E., S. Maritorena, D. Siegel, M. C. O’Brien, D. Toole, B. G. Mitchell, M.<br />

Kahru, F. P. Chavez, P. Strutton, G. Cota, S. B. Hooker, C. R. McClain, K. L.<br />

Car<strong>de</strong>r, F. Muller-Karger, L. Harding, A. Magnuson, D. Phinney, G. F. Moore, J.<br />

Aiken, K. R. Arrigo, R. Letelier, and M. Culver (2000), Ocean color chlorophyll a<br />

algorithms for SeaWiFS, OC2, and OC4: version 4., in NASA Tech. Memo. 2000-<br />

206892, vol. 11, edited by S. B. Hooker, and E. R. Firestone, pp. 9-23, NASA<br />

Goddard Space Flight Center, Greenbelt, Md.<br />

Park, Y.-J., and K. G. Ruddick (2005), Mo<strong>de</strong>l of remote-sensing reflectance including<br />

bidirectional effects for case 1 and case 2 waters, Applied Optics, 44(7), 1236-<br />

1249.<br />

Perovich, D. K., C. S. Roesler, and W. S. Pegau (1998), Variability in Arctic sea ice<br />

optical properties, J. Geophys. Res., 103(C1), 1193-1208.<br />

Petzold, T. J., Volume scattering functions for selected ocean waters, Tech. Rep. SIO, 72-<br />

78 pp., Scripps Institution of Oceanography, San Diego, La Jolla, Ca., 1972.<br />

Pope, R. M., and E. S. Fry (1997), Absorption spectrum (380-700nm) of pure water. II.<br />

Integrating cavity mea<strong>sur</strong>ements, Appl. Opt., 36, 8710-8723.<br />

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery (1992), Numerical<br />

recipes in Fortran 77: The art of scientific computing, Chap 9 pp., Cambridge U.<br />

Press, New York.<br />

Reynolds, R. A., D. Stramski, and B. G. Mitchell (2001), A chlorophyll-<strong>de</strong>pen<strong>de</strong>nt<br />

semianalytical reflectance mo<strong>de</strong>l <strong>de</strong>rived from field mea<strong>sur</strong>ements of absorption<br />

and backscattering coefficients within the Southern Ocean, J. Geophys. Res.,<br />

106(C4), 7125-7138.<br />

Sakshaug, E., and H. R. Skjoldal (1989), Life at the ice edge, Ambio., 18(1), 60-67.<br />

Siegel, D. A., S. Maritorena, and N. B. Nelson (2005), In<strong>de</strong>pen<strong>de</strong>nce and<br />

inter<strong>de</strong>pen<strong>de</strong>ncies among global ocean color properties: Reassessing the biooptical<br />

assumption, J. Geophys. Res., 110(C7), C07011,<br />

doi:10.1029/2004JC002527.<br />

Smith, S. D., R. D. Muench, and C. H. Pease (1990), Polynyas and leads: An overview of<br />

physical processes and environment, J. Geophys. Res. , 95(C6), 9461-9479.<br />

Smith, W. O., and D. M. Nelson (1986), The importance of ice-edge phytoplankton<br />

production in the Southern Ocean, BioScience, 36, 251-257.<br />

167


Stein, R., and R. W. MacDonald (2004), Organic Carbon Budget: Arctic Ocean vs.<br />

Global Ocean, in The Organic Carbon Cycle in the Arctic Ocean, edited by R.<br />

Stein, and R. W. MacDonald, pp. 315-322, Springer, Berlin.<br />

Stirling, I. (1997), The importance of polynyas, ice edges, and leads to marine mammals<br />

and birds, J. Mar. Syst., 10(1-4), 9-21.<br />

Stumpf, R. P., R. A. Arnone, R. W. Gould, R. M. Martinolich, and V. Ransibrahmanakul<br />

(2003), A partially coupled ocean-atmosphere mo<strong>de</strong>l for retrieval of water-leaving<br />

radiance from SeaWiFS in coastal waters, Chap. 9, in NASA/TM-2003-206892,<br />

edited by S. B. Hooker, and E. R. Firestone, pp. 51-59, Greenbelt, MD.<br />

Tanré, D., M. Herman, and P. Y. Deschamps (1981), Influence of the background<br />

contribution upon space mea<strong>sur</strong>ements of ground reflectance, Appl. Opt., 20(20),<br />

3676-3684.<br />

Tanré, D., M. Herman, P. Y. Deschamps, and A. <strong>de</strong> Leffe (1979), Atmospheric mo<strong>de</strong>ling<br />

for space mea<strong>sur</strong>ements of ground reflectances, including bidirectional properties,<br />

Appl. Opt., 18(21), 3587-3594.<br />

Vermote, E., D. Tanré, J. L. Deuzé, M. Herman, and J. J. Morcrette, 6s User Gui<strong>de</strong><br />

Version 2, 218 pp., Laboratoire d'Optique Atmosphérique, Lille, France, 1997.<br />

Wang, J., and G. F. Cota (2003), Remote-sensing reflectance in the Beaufort and Chukchi<br />

seas: observations and mo<strong>de</strong>ls, Appl. Opt., 42(15), 2754-2765.<br />

168


Chapitre VI : Utilisation <strong>de</strong> l’imagerie « couleur <strong>de</strong><br />

l’océan » pour la quantification <strong>de</strong> la<br />

photooxydation dans la Mer <strong>de</strong> Beaufort<br />

169


VI.A Résumé<br />

La production photochimique <strong>de</strong> <strong>carbone</strong> inorganique dissous (DIC) intégrée dans la<br />

colonne d’eau a été estimée l’ai<strong>de</strong> <strong><strong>de</strong>s</strong> propriétés optiques inhérentes dérivées <strong><strong>de</strong>s</strong> données<br />

SeaWiFS <strong>de</strong> la Mer <strong>de</strong> Beaufort obtenues entre 1998 et 2004. Plus <strong>de</strong> 300 images SeaWiFS<br />

ont été traitées à l’ai<strong>de</strong> d’un algorithme <strong>de</strong> correction atmosphérique adapté pour <strong>les</strong> eaux<br />

turbi<strong><strong>de</strong>s</strong> qui a permis <strong>de</strong> retrouver <strong><strong>de</strong>s</strong> spectres <strong>de</strong> réflectance marine <strong>de</strong> bonne qualité. Après<br />

l’élimination <strong><strong>de</strong>s</strong> données contaminées par <strong>les</strong> effets d’environnement dus à la glace <strong>de</strong> mer,<br />

le rapport entre <strong>les</strong> coefficients d’absorption <strong>de</strong> la matière organique dissoute colorée<br />

(CDOM) et totale ([a CDOM/a t]) a été calculé à partir du spectre <strong>de</strong> réflectance marine à l’ai<strong>de</strong><br />

d’un algorithme empirique. La photoproduction annuelle <strong>de</strong> DIC estimée à l’ai<strong>de</strong> <strong>de</strong><br />

[a CDOM/a t] dérivé <strong><strong>de</strong>s</strong> données SeaWiFS est faiblement inférieure (~10%) aux estimations<br />

réalisées avec <strong>les</strong> valeurs régiona<strong>les</strong> constantes <strong>de</strong> [a CDOM/a t] obtenues à partir <strong><strong>de</strong>s</strong> me<strong>sur</strong>es in<br />

situ. En faisant l’hypothèse que le rapport [a CDOM/a t] est <strong>de</strong> 0.9 <strong>sur</strong> tout le spectre, la<br />

photoproduction annuelle <strong>de</strong> DIC était <strong>sur</strong>estimée d’environ 60%. Ces résultats suggèrent<br />

que <strong>les</strong> données <strong>de</strong> Couleur <strong>de</strong> l’Océan peuvent être utilisées pour estimer [a CDOM/a t] <strong><strong>de</strong>s</strong> <strong>les</strong><br />

eaux côtières arctiques, lequel<strong>les</strong> sont souvent inaccessib<strong>les</strong> et <strong>les</strong> me<strong>sur</strong>es in situ n’y sont pas<br />

toujours disponib<strong>les</strong>. En plus <strong>de</strong> fournir [a CDOM/a t], l’estimation <strong>de</strong> la valeure absolue <strong>de</strong><br />

a CDOM, obtenue grâce à l’estimation semi-analytique <strong>de</strong> a t, a aussi permis <strong>de</strong> tenir compte <strong><strong>de</strong>s</strong><br />

variations spatia<strong>les</strong> du ren<strong>de</strong>ment quantique apparent pour la photoproduction <strong>de</strong> DIC (φ DIC).<br />

Dans le futur, d’autres produits dérivés <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan pourront être<br />

utilisés afin d’évaluer la variabilité <strong>de</strong> φ DIC. Le manque <strong>de</strong> données <strong>sur</strong> ce <strong>de</strong>rnier limite,<br />

cependant, le développement d’un tel modèle prédictif <strong>de</strong> φ DIC.<br />

170


VI.B. Article en preparation: “The use of Ocean Color imagery for the<br />

quantification of photooxidation in the Beaufort Sea”<br />

Simon Bélanger and Marcel Babin<br />

Laboratoire d’Océanographie <strong>de</strong> Villefranche, Centre National <strong>de</strong> la Recherche<br />

Scientifique, Université Pierre et Marie Curie - Paris 6, 06230 Villefranche-<strong>sur</strong>-mer,<br />

France<br />

171


Abstract<br />

Depth-integrated photochemical production of dissolved inorganic carbon (DIC) in the<br />

Southeastern Beaufort Sea was assessed for the period from 1998 to 2004 using<br />

SeaWiFS-<strong>de</strong>rived water inherent optical properties. More than 300 SeaWiFS images<br />

were processed using a turbid-water atmospheric correction algorithm that provi<strong><strong>de</strong>s</strong><br />

reliable water-leaving reflectance retrieval. After eliminating data contaminated by the<br />

adjacency effect due to sea ice, the ratio between chromophoric dissolved organic matter<br />

(CDOM) and the total absorption coefficients ([aCDOM/at]) was calculated from the water-<br />

leaving reflectance spectrum using an empirical algorithm. The annual DIC<br />

photoproduction estimated using SeaWiFS-<strong>de</strong>rived [aCDOM/at] was slightly lower (~10%)<br />

than the estimation ma<strong>de</strong> using regionally constant [aCDOM/at] spectra <strong>de</strong>termined from in<br />

situ mea<strong>sur</strong>ements. In contrast, assuming a spectrally constant value for [aCDOM/at] of 0.9,<br />

the annual DIC photoproduction was overestimated by ~60%. These results suggest that<br />

satellite Ocean Color data could be used to estimate [aCDOM/at] in the remote Arctic<br />

coastal waters where in situ mea<strong>sur</strong>ements are still unavailable. In addition, the<br />

estimation of the absolute value in aCDOM, as obtained from at calculated using a semi-<br />

analytical algorithm, was used to account for the spatio-temporal variability in the<br />

apparent quantum yield for the DIC photoproduction (φDIC). In the future, other Ocean<br />

Color products may also be used to account for the variability in φDIC, but such data are<br />

currently lacking to <strong>de</strong>velop a robust parameterization of φDIC.<br />

172


VI.1. Introduction<br />

In Chapter III I quantified, for the first time, the amount of dissolved organic<br />

carbon mineralized through CDOM photooxidation in the Beaufort Sea. A coupled<br />

optical-photochemical was used to calculate the <strong>de</strong>pth-integrated photoproduction of DIC<br />

in ice-free Arctic waters. I showed that CDOM photooxidation in the southeastern<br />

Beaufort Sea (western Arctic) may increase significantly in the future in response to the<br />

predicted trend of ongoing reduction of sea ice cover. Interestingly, the predicted<br />

increased in terrigenous dissolved organic carbon (tDOC) mineralization is nearly<br />

equivalent to the amount of particulate marine organic carbon actually buried into the<br />

<strong>de</strong>ep sediments. This result confirms the relative importance of this <strong>flux</strong> in the balance of<br />

the Arctic Organic Carbon Budget.<br />

In our initial approach, satellite observations of sea ice, ozone, aerosols and cloud<br />

cover were used with a radiative transfer mo<strong>de</strong>l to estimate the spectral irradiance that<br />

penetrates the water column. The regional variability of the <strong>sur</strong>face water optical<br />

properties as <strong>de</strong>rived from in situ mea<strong>sur</strong>ements was used in the calculation, which partly<br />

controlled the rates of CDOM photooxidation in the water column. This variability is<br />

expressed by the [aCDOM/at] parameter, which is the light fraction absorbed by the CDOM<br />

in an optically homogenous water column. In Chapter IV, we <strong>de</strong>veloped an Ocean Color<br />

algorithm to retrieve this parameter from the remote sensing reflectance (Rrs)<br />

mea<strong>sur</strong>ements. The absolute accuracy of the regionally tuned [aCDOM/at] algorithm is<br />

±14%. Interestingly, <strong><strong>de</strong>s</strong>pite UV light are the most efficient radiation in terms of DIC<br />

photoproduction, the variability in [aCDOM/at] in the violet part of the spectrum (412 nm)<br />

is nearly proportional to the spectrally-integrated photoproduction of DIC. We have<br />

shown that using averaged spectral shapes for CDOM and particulate matter, spectral<br />

[aCDOM/at] values can be obtained. The error in terms of DIC photoproduction attributed<br />

to the spectral extrapolation is


to examine the specific problem posed to ocean color remote sensing by the presence of<br />

this almost perfect reflector (Chap. V). The adjacency effect that is due to nearby sea ice<br />

cover is only partly removed by atmospheric correction, resulting in a contamination of<br />

the water-leaving reflectance retrieval and the subsequent estimation of [aCDOM/at](412)<br />

from it. In general, [aCDOM/at](412) is un<strong>de</strong>restimated by more than 0.15 at a distance of<br />

10-15 km from an ice edge. As the adjacency effect increases with <strong>de</strong>creasing<br />

wavelengths, the blue part of the spectrum was used to <strong>de</strong>tect and discard pixels affected<br />

by such contamination.<br />

The main goal of the present chapter is to estimate the DIC photoproduction using<br />

satellite-<strong>de</strong>rived [aCDOM/at] in open waters free of adjacency contamination. The revised<br />

estimations are compare to the ones obtained using in situ and constant [aCDOM/at] values.<br />

Specifically, I want to know if, when in situ mea<strong>sur</strong>ements of [aCDOM/at] in a given<br />

coastal region are not available, ocean color data can be used instead. I first <strong><strong>de</strong>s</strong>cribed the<br />

data processing of a seven years time-series of Sea wi<strong>de</strong> Field-of-View (SeaWiFS)<br />

acquired between 1998 and 2004. Two different methods in which satellite-<strong>de</strong>rived<br />

information is used to calculate the DIC photoproduction are <strong><strong>de</strong>s</strong>cribed. The first one<br />

make used of [aCDOM/at] only, whereas in the second method, the magnitu<strong>de</strong> of the<br />

CDOM absorption coefficient at 412 nm is used to i<strong>de</strong>ntify two water classes with<br />

different photoreactivity. In the Results section, I present a validation of SeaWiFS water-<br />

leaving reflectance by comparing coinci<strong>de</strong>nt satellite and in situ mea<strong>sur</strong>ements obtained<br />

in 2004 in the study area. Then the observed spatio-temporal variability of some IOP is<br />

presented. Finally, I discuss the quality of the SeaWiFS data and the utility of ocean color<br />

imagery to quantify <strong>de</strong>pth-integrated DIC photoproduction in the Arctic coastal zones.<br />

VI.2. Methods<br />

VI.2.1. SeaWiFS Level 2 and 3 processing<br />

SeaWiFS Level 1A MLAC data (merged Local Area Coverage) gathering all<br />

available SeaWiFS HRPT (High Resolution Picture Transmission) and LAC data for a<br />

given orbit were downloa<strong>de</strong>d from the NASA Ocean Color Web site<br />

(www.oceancolor.gsfc.nasa.gov). The Level 1A MLAC data contains raw radiance<br />

values for each of the eight SeaWiFS bands (412, 443, 490, 510, 555, 670, 765, and 865<br />

174


nm) and have a spatial resolution of 1.1 km at nadir. At the latitu<strong>de</strong> of 70°N, a daily<br />

temporal coverage is available with SeaWiFS (up to 3 images per day). However, due to<br />

the presence of sea ice or clouds, only the scenes where open water were present were<br />

processed to Level 2. A total of 335 scenes acquired during the summer months from<br />

1998 to 2004 were processed to Level 2 (Table VI.1).<br />

Table VI.1. Number of SeaWiFS Level 1A MLAC images processed<br />

May June July August Total<br />

1998 14 19 28 17 78<br />

1999 10 11 19 19 59<br />

2000 0 0 9 16 25<br />

2001 6 9 17 15 47<br />

2002 5 10 9 25 39<br />

2003 5 16 11 10 42<br />

2004 1 22 13 9 45<br />

Total 335<br />

The Level 2 processing was performed image-by-image in two steps: 1)<br />

atmospheric corrections (AC) were achieved to subtract the atmospheric scattering<br />

components from the top-of-atmosphere (TOA) reflectance and obtain the water-leaving<br />

reflectance for the visible bands; and 2) the water-leaving reflectance spectrum was<br />

analyzed to <strong>de</strong>rive inherent optical properties (IOP).<br />

The SeaWiFS standard atmospheric correction scheme over clear open ocean<br />

waters is based on the “black pixel assumption” [e.g., Gordon and Wang, 1994; Gordon,<br />

1997; Siegel et al., 2000]. This assumption is based on fact that water-leaving reflectance<br />

for the near-infrared bands at 765 nm and 865 nm is null, and that the mea<strong>sur</strong>ed signal at<br />

those bands in entirely due to the atmosphere. Then the atmospheric aerosols properties<br />

obtained in the NIR can be extrapolated to the visible using recomputed look-up-tab<strong>les</strong>.<br />

Typically over turbid waters, the black pixel assumption results in negative water-leaving<br />

reflectance in the blue part of the spectrum (< 443 nm) due to an overestimation of the<br />

aerosol signal extrapolated from the NIR bands. Because of the violation of the “black<br />

pixel assumption” in turbid coastal waters, the AC was performed using a modified<br />

version of the Multi-Sensor Level-1 to Level-2 (MSL12) software. The modified MSL12<br />

175


was provi<strong>de</strong>d by the Belgian Management Unit of the North Sea Mathematical Mo<strong>de</strong>ls<br />

(MUMM; www.mumm.ac.be/OceanColour/). In the MUMM’s MSL12, the assumption<br />

of null water-leaving reflectance in the NIR bands is replaced by the assumptions of<br />

spatial homogeneity of the 765:865-nm ratios for aerosol reflectance and for water-<br />

leaving reflectance [Ruddick et al., 2000]. These ratios can be fixed as calibration<br />

parameters during the processing. While the ratio of water-leaving reflectance is expected<br />

to be rather in<strong>de</strong>pen<strong>de</strong>nt of region and time [Ruddick et al., 2006], the aerosol reflectance<br />

ratio is expected to vary in space and time in response to variations in aerosol particle<br />

type and concentration. But here, it is assumed that the aerosol type is constante and that<br />

only the amount of aeraosol is variable for a given image. Therefore, the aerosol<br />

reflectance ratio (<strong>de</strong>noted hereafter as "eps78") needs to be assessed on an image-by-<br />

image basis by inspection of the Rayleigh-corrected reflectance scatterplot [Ruddick et al.,<br />

2000]. Instead of inspecting the Rayleigh-corrected reflectance scatterplot, which could<br />

be contaminated by clouds or sea ice, I obtained eps78 by the inspection of the<br />

atmospheric products <strong>de</strong>rived using the standard AC algorithm. In general I chose an<br />

averaged eps78 value obtained in an area away from the turbid waters, but as close as<br />

possible. In general, the aerosol eps78 values were within the range from 1.0 to 1.15,<br />

which is typical of marine aerosols. When large areas were still showing negative water-<br />

leaving reflectance after the application of turbid water AC algorithm, the image was<br />

reprocessed using a slightly lower value for eps78. To avoid negative reflectance in the<br />

blue, therefore, I often had to choose lower values for eps78.<br />

After AC processing, the pixels affected by the adjacency effect (Chap. V), as<br />

well as standard flagged pixels [e.g, Bailey and Wer<strong>de</strong>ll, 2006], were not consi<strong>de</strong>red for<br />

further bio-optical processing. Typically, when the sea ice melting begins in early season,<br />

more than 50% of the open water pixels must be eliminated. This number <strong>de</strong>creases to<br />

~15% during the summer season. Note that the cloud bor<strong>de</strong>rs were also <strong>de</strong>tected by the<br />

adjacency effect <strong>de</strong>tection algorithm (Section V.3.3.), resulting sometime in more than<br />

50% of invalid pixels even in August. For each valid pixels, the ratio [aCDOM/at] at 412<br />

nm was calculated using the empirical algorithm <strong><strong>de</strong>s</strong>cribed in Chapter IV, with the<br />

region-specific coefficients (Table IV.2). The total absorption coefficient at 412 nm (at)<br />

176


and particulate backscattering coefficient at 555 nm (bbp) were computed using a<br />

modified version of the quasi-analytical algorithm (QAA) proposed by Lee et al. [2002]<br />

(see Annex A1). The former coefficient is further used in the DIC photoproduction<br />

processing to extrapolate [aCDOM/at](412) to the full UV-to-green spectrum (300 and 600<br />

nm) (see section IV.3.3) and to estimate the magnitu<strong>de</strong> of aCDOM at 412 nm.<br />

The Level 3 processing consists in binning the Level 2 geophysical products in<br />

time and space. First, the Level 2 data were remapped on an azimuthal projection grid<br />

with pixel resolution of 1 x 1 km. Then, the monthly averages were produced for each<br />

pixel of the grid with all valid pixels available. Finally, for the DIC photoproduction<br />

calculation, the SeaWiFS monthly averages were remapped onto the grid used by the<br />

National Snow and Ice Data Center (NSIDC) for the sea ice concentration data<br />

distribution. The NSIDC grid is polar stereographic projection with grid elements of<br />

approximately 25x25 km resolution.<br />

VI.2.2. DIC photoproduction calculation<br />

The DIC photoproduction calculation is essentially the same as the one <strong><strong>de</strong>s</strong>cribed<br />

in section III.2.5. Here the spatially integrated DIC photoproduction (in mol DIC d -1 ),<br />

however, is done for each element of the NSIDC grid as:<br />

P<br />

×<br />

DIC<br />

600<br />

∫<br />

λ = 300<br />

( x,<br />

y,<br />

d)<br />

= OpenWaterArea(<br />

x,<br />

y,<br />

d)<br />

− ⎡a<br />

⎤ CDOM<br />

Ed<br />

( 0 , λ, d)<br />

⎢ ⎥<br />

⎣ at<br />

⎦<br />

,<br />

( λ,<br />

x,<br />

y,<br />

d)<br />

φDIC<br />

( λ)<br />

dλ<br />

(1)<br />

where the OpenWaterArea (m 2 ) is the area of open water of a given NSIDC pixel,<br />

−<br />

E ( 0 , λ)<br />

(mol photons m -2 d -1 ) is the spectral daily downward irradiance just beneath<br />

d<br />

sea <strong>sur</strong>face, φDIC is the spectral apparent quantum yield for the DIC production, x and y<br />

are the longitu<strong>de</strong> and the latitu<strong>de</strong>, respectively, and d is the day of the year. A schematic<br />

representation of the DIC mo<strong>de</strong>ling is shown and <strong>de</strong>tailed in Fig. VI.1.<br />

In Chapter III, the southeastern Beaufort Sea was sub-divi<strong>de</strong>d in three sub-regions:<br />

Canada Basin, Mackenzie Shelf, and Amundsen Gulf. For each of those, the φDIC spectra<br />

were assumed constant based on the location of the in situ sampling. Here an alternative<br />

method is proposed to select the φDIC as a function of the magnitu<strong>de</strong> of the aCDOM(412). It<br />

appears, at a first approximation, that the CDOM photoreactivity in the study area is<br />

177


higher when the CDOM concentration is higher (Fig. VI.2a). For the low values of<br />

CDOM absorption coefficient at 412 nm (i.e. aCDOM(412) < 0.2 m -1 ) the φDIC for station<br />

108, 406 and 409 were pooled together, while R5a, R5d and R9 were average to represent<br />

the high CDOM concentration (see Fig. III.1 for station locations; Fig. VI.2b). In the<br />

processing, aCDOM(412) is calculated as the product of the empirically-<strong>de</strong>rived<br />

[aCDOM/at](412) and the QAA-<strong>de</strong>rived at(412) (see Annex A1).<br />

Figure VI.1. A schematic representation of the DIC photoproduction<br />

mo<strong>de</strong>ling. The mo<strong>de</strong>l ingests four inputs: (1) For a given day, when data<br />

are available, the spatio-temporal variability in the [aCDOM/at] parameter is<br />

specified by interpolating the values of the months before and after. If<br />

monthly IOP are not available (due to either sea ice or cloud cover), the<br />

data from the climatology is taken. The spectral interpolation of the<br />

[aCDOM/at](412) is perform as <strong><strong>de</strong>s</strong>cribed in section IV.3.4. (2) The daily<br />

spatial variability in the area of open water. (3) The spectral daily<br />

downward irradiance calculated as <strong><strong>de</strong>s</strong>cribed in section III.2.5. (4) The<br />

spectral apparent quantum yield for the DIC production (AQY, or φDIC).<br />

Two methods are compared for the choice of the φDIC for a given day and<br />

pixel. Method 1 is based on the regional <strong>de</strong>finition <strong><strong>de</strong>s</strong>cribed in section<br />

III.2.5. In the Method 2, the φDIC varies as a function of the magnitu<strong>de</strong> of<br />

the CDOM (see text for <strong>de</strong>tails).<br />

178


VI.3. Results<br />

Figure VI.2. A) CDOM photoreactivity as a function of aCDOM(412). B)<br />

Spectral variability of the pooled φDIC spectra for low and high aCDOM(412)<br />

value.<br />

VI.3.1. Validation of the SeaWiFS water-leaving reflectance<br />

To assess the quality of the retrieved water-leaving reflectance (Rrs) with the<br />

MUMM’s algorithm, coinci<strong>de</strong>nt satellite and in situ observations (from hereafter called<br />

“match-ups”) are compared. Out of 48 Rrs spectra mea<strong>sur</strong>ed during CASES (section<br />

IV.2.1.2), six match-ups of good quality were retained for the analysis. The match-ups<br />

used for the assessment were collected within a lag shorter than three hours between in<br />

situ sampling and the satellite overpass, to minimize perturbations due to the temporal<br />

179


variability of seawater optical properties (Table VI.2). Satellite data were obtained from<br />

the average of the 3 x 3 image elements centered on the mea<strong>sur</strong>ement site, and were<br />

consi<strong>de</strong>red as valid for comparison when all of the 9 elements were not flagged for cloud<br />

or adjacency effect. Table VI.2 provi<strong><strong>de</strong>s</strong> the <strong>de</strong>tails on the match-ups location, date, time<br />

interval, and the sun zenith angle at the time of sampling.<br />

Table VI.2. Location, date, time interval, and the sun zenith angle at the time of sampling<br />

of the match-ups<br />

CASES<br />

Station<br />

Date Lat.<br />

(°N)<br />

Long.<br />

(°W)<br />

SeaWiFS image<br />

(SYYYYDDDHHMMSS)<br />

Time interval<br />

(hh:mm)<br />

Sun<br />

zenith<br />

ID<br />

(°)*<br />

CA-20 17-Jul 70.43 126.35 S2004199221315 00:25 51.9 (50.9)<br />

309-1 18-Jul 71.12 125.84 S2004200225405 02:10 54.8 (50.3)<br />

309-2 19-Jul 71.17 126.2 S2004201215611 00:05 52.3 (52.2)<br />

721 26-Jul 69.84 133.29 S2004208232426 01:40 55.5 (62.9)<br />

115 29-Jul 70.84 125.09 S2004211220929 02:40 57.5 (54.9)<br />

215 30-Jul 70.95 123.52 S2004212225019 01:10 54.8 (61.9)<br />

* The value in parenthesis is the sun zenith angle at the time of the in situ mea<strong>sur</strong>ements.<br />

The in situ mea<strong>sur</strong>ements of the water-leaving reflectance spectra are compared to<br />

SeaWiFS data processed with both the standard and the turbid water AC algorithms (Fig.<br />

VI.3). Mo<strong>de</strong>rate turbidity was observed at station 721 located in the Mackenzie Shelf<br />

(Rrs(555)=0.005 sr -1 ), while the turbidity was relatively low at the stations located in the<br />

Amundsen Gulf. The results of the match-ups analysis can be summarized as follows:<br />

1. When the turbidity is low (i.e. Rrs(555) < 0.002 sr -1 ; stations 309 and 215), the<br />

water-leaving reflectance at wavelengths larger than 490 nm retrieved using the<br />

standard algorithm is in good agreement with the in situ mea<strong>sur</strong>ements.<br />

2. The standard AC algorithm produced a systematic un<strong>de</strong>restimation of the<br />

reflectance at 412 and 443 nm, regard<strong>les</strong>s of the water turbidity. For the most<br />

turbid station (721), negative reflectance at 412 nm was return by the standard AC<br />

algorithm, certainly because of the violation of the black pixel assumption.<br />

3. The turbid water AC algorithm (MUMM) produced an overestimation of the Rrs<br />

spectrum all wavelengths (except at stations CA-20 and 115). The overestimation<br />

is in general more pronounced when the turbidity is low (stations 309 and 215).<br />

180


4. In all cases, the spectral shape of the Rrs spectrum retrieved using the turbid water<br />

AC algorithm is good agreement with the in situ mea<strong>sur</strong>ements.<br />

Figure VI.3. Comparison of in situ mea<strong>sur</strong>ements of water-leaving<br />

radiance with coinci<strong>de</strong>nt spectra from the SeaWiFS image processed with<br />

the standard and the turbid water AC algorithms.<br />

Figure VI.4 compares satellite-<strong>de</strong>rived versus in situ reflectance ratios employed<br />

in the [aCDOM/at](412) empirical algorithm, i.e. the log[Rrs(412)/Rrs(555)] and<br />

log[Rrs(490)/Rrs(555)]. The SeaWiFS-<strong>de</strong>rived ratios from both the standard and the turbid<br />

water AC algorithms are shown. Larger discrepancies (> 3 fold) in the<br />

log[Rrs(412)/Rrs(555)] are obtained with the standard relative to the turbid water AC<br />

181


algorithm (Fig. VI.4a). On the other hand, both algorithm provi<strong>de</strong> similar results for the<br />

retrieval of the log[Rrs(490)/Rrs(555)] (Fig. VI.4b).<br />

Figure VI.4. Scatterplots of the in situ mea<strong>sur</strong>ements versus SeaWiFS<strong>de</strong>rived<br />

variab<strong>les</strong> employed in the empirical algorithm used to estimate<br />

[aCDOM/at](412) (eq. IV.6): a) the logarithm of water-leaving reflectance<br />

ratio of 412 to 555 nm; b) the logarithm of water-leaving reflectance ratio<br />

of 490 to 555 nm.<br />

The mean absolute difference between in situ and satellite-<strong>de</strong>rived [aCDOM/at](412)<br />

was, on average, +0.188 and -0.105 for the standard and turbid water AC algorithms,<br />

respectively (Table VI.3). While the use of the standard AC results in systmatic<br />

overestimation in [aCDOM/at](412), the overestimation of the retrieved water-leaving<br />

reflectance at 555 nm using the turbid water AC algorithm results in a systematic<br />

un<strong>de</strong>restimation of [aCDOM/at](412). The [aCDOM/at](412) computed with the reflectance<br />

182


spectrum <strong>de</strong>rived using the standard AC can reach values above 1, which is physically<br />

impossible. Using the turbid water AC algorithm, the [aCDOM/at](412) retrieval falls<br />

within the algorithm accuracy (±0.14 see Chap. IV), except for stations 309-2 and 721.<br />

The latter station was sampled in an area where the spatial variability was very high and<br />

the vertical distribution of IOPs was not homogenous (data not shown). Un<strong>de</strong>r such<br />

conditions large discrepancy between satellite retrievals and in situ Rrs spectra can be<br />

expected (Fig. VI.3). In the next section, only the results using MUMM algorithm are<br />

presented as they appears more reliable than the one obtained using standard AC (see<br />

Discussion VI.4.1 section).<br />

Table VI.3. Comparison of in situ with SeaWiFS-<strong>de</strong>rived [aCDOM/at](412) values using<br />

Standard and MUMM AC algorithms.<br />

CASES Station<br />

ID<br />

[aCDOM/at](412)<br />

in situ Standard MUMM<br />

CA-20 0.88 1.07 (+.19) 0.83 (-.05)<br />

309-1 0.81 1.14 (+.33) 0.69 (-.13)<br />

309-2 0.82 0.96 (+.14) 0.66 (-.16)<br />

721 0.81 NA 0.63 (-.18)<br />

115 0.71 0.83 (+.12) 0.68 (-.03)<br />

215 0.76 0.92 (+.16) 0.68 (-.08)<br />

VI.3.2. Spatial-temporal variability in [aCDOM/at](412)<br />

An example of the seasonal variability in the [aCDOM/at](412) parameter for the<br />

summers of 1998 and 2004 (lowest sea ice concentration of the time serie) is shown on<br />

Fig. VI.5. Most of the study area is characterized by [aCDOM/at](412) values ranging from<br />

0.6 to 0.7. These values may be slightly un<strong>de</strong>restimated for the reasons explained above.<br />

For example, the SeaWiFS averaged value in the Amundsen Gulf in July 2004 was 0.66,<br />

while the in situ mea<strong>sur</strong>ements obtained in this area during the same period range<br />

between 0.60 and 0.94, with an average value of 0.78. In the sector of the Mackenzie<br />

Shelf where the Mackenzie river input are high, however, the SeaWiFS-<strong>de</strong>rived values<br />

ranging from 0.30 to 0.40 are similar to those observed in 2004 near the Mackenzie River<br />

Delta. To better see the turbid river plume, Fig. VI.6 shows the particulate backscattering<br />

coefficient at 555 nm (bbp) observed during June and July. In fact, bbp(555) is a robust<br />

indicator of the total suspen<strong>de</strong>d material. For both years, the plume extent was maximum<br />

183


in June and <strong>de</strong>crease in July. The <strong>de</strong>crease in suspen<strong>de</strong>d material explains the increase in<br />

[aCDOM/at](412) observed over the Mackenzie Shelf during summer, as the river discharge<br />

<strong>de</strong>clines (Fig. VI.5).<br />

Figure VI.5. Seasonal variability of the SeaWiFS-<strong>de</strong>rived [aCDOM/at](412)<br />

for 1998 and 2004 (1 x 1 km resolution). Sea ice (or cloud) is shown in<br />

grey.<br />

184


Figure VI.6. Variability of the SeaWiFS-<strong>de</strong>rived bbp(555) for June and<br />

July 1998 and 2004 (1 x 1 km resolution).<br />

Figure VI.7. Variability of the SeaWiFS-<strong>de</strong>rived at(412) for June and July<br />

1998 and 2004 (1 x 1 km resolution).<br />

185


Further offshore at the Shelf break or along southern coasts of the Amundsen Gulf<br />

and Alaskan Shelf, higher [aCDOM/at](412) values (>0.8) can be observed. These high<br />

values are observed within highly absorbing waters likely coming from the rivers (Fig.<br />

VI.7), but where the amount of suspen<strong>de</strong>d partic<strong>les</strong> is relatively low (Fig. VI.6). The<br />

striking extent of the Mackenzie River plume toward the west in the Canada Basin during<br />

July 1998 is obvious on both [aCDOM/at](412) and at(412) maps. Interestingly, a little<br />

amount of partic<strong>les</strong> is transported beyond the continental shelf at that time (Fig. VI.6). In<br />

2004, the presence of a persistent sea ice cover obscures the visualization of the<br />

Mackenzie River plume extent, but the bbp(555) pattern observed in June and the high<br />

at(412) values around the location 72°N and 140°W (~ 0.2 m -1 ) in July suggest that the<br />

plume was taking approximately the same westward “route” as in 1998.<br />

VI.3.3. Spatial-temporal variability in DIC photoproduction<br />

The DIC photoproduction was reassessed with taking into account the spatio-<br />

temporal variability in [aCDOM/at]. Table VI.4 compares the annual DIC photoproduction<br />

calculated using four different approaches: 1) with regionally constant values for<br />

[aCDOM/at](λ) and φDIC(λ), as <strong>de</strong>fined in our previous study (Chapter III; the “GBC<br />

method”); 2) with SeaWiFS-<strong>de</strong>rived [aCDOM/at](λ) and regional φDIC(λ) values (“Method<br />

1” in Fig. VI.1); 3) with SeaWiFS-<strong>de</strong>rived [aCDOM/at](λ) and variable φDIC(λ) as a<br />

function of aCDOM(412) (“Method 2” in Fig. VI.1); 4) spectrally constant [aCDOM/at] value<br />

of 0.9 and regional φDIC(λ) (“Constant method”).<br />

Table VI.4. Annual DIC photoproduction estimation using various methods. The value in<br />

parenthesis for methods 1 and 2 is the difference relative (in %) to GBC method.<br />

Year GBC Method 1 Method 2 Constant<br />

1998 122.3 108.2 (-11.5) 115.5 (-5.6) 195.7 (+60.0)<br />

1999 71.3 65.6 (-8.0) 72.3 (+1.4) 114.5 (+60.6)<br />

2000 47.9 43.0 (-10.2) 47.2 (-1.5) 77.1 (+61.0)<br />

2001 52.6 48.7 (-7.4) 56.4 (+7.2) 82.3 (+56.5)<br />

2002 60.8 53.5 (-12.0) 54.1 (-11.0) 99.7 (+64.0)<br />

2003 67.7 61.0 (-9.9) 61.5 (-9.1) 109.8 (+62.2)<br />

2004 73.6 69.1 (-6.1) 75.4 (+2.4) 120.6 (+63.9)<br />

186


Comparison between GBC method and Method 1, which differ only in terms<br />

[aCDOM/at], suggest that our former annual DIC photoproduction estimations were<br />

overestimated by 6 to 12%. However, this could be partly attributed to the<br />

un<strong>de</strong>restimation of [aCDOM/at](412) by SeaWiFS (see above). But a regional comparison<br />

reveals that the largest differences were found mainly in the Canada Basin (~25%, data<br />

not shown). This is probably due to the fact that the stations used to represent the Canada<br />

Basin in our former estimation were biased toward higher [aCDOM/at] values. In fact, the<br />

Canada Basin’s stations are located at the fringe of the shelf (see Fig. III.1), where<br />

[aCDOM/at] is generally higher than over the <strong>de</strong>ep basin (e.g., June and July 1998; Fig.<br />

VI.5). In contrast, using a spectrally constant [aCDOM/at] value of 0.90 results in a severe<br />

overestimation of the DIC photoproduction (~+60% compare to GBC method).<br />

Method 2 is an alternative method to account for the variability of φDIC(λ) based<br />

on <strong>de</strong>rived aCDOM(412) magnitu<strong>de</strong> The annual DIC photoproduction is slightly higher<br />

than using Method 1 (Table VI.4). In particular, the production is significantly higher in<br />

the Amundsen Gulf (~30 %) where most of the southern part appears to be affected river<br />

runoff. The spatial variability in the DIC photoproduction rate (in gC m -2 y -1 ) <strong>de</strong>rived<br />

using Method 2 is shown in Fig. VI.8. Although most of the spatial variability is<br />

explained by the sea ice concentration, higher DIC photoproduction is found in the waters<br />

influenced by the river runoff. With the regionally constant φDIC(λ), the runoff was<br />

blurred on the DIC photoproduction maps (data not shown). Finally, the variability in<br />

DIC photoproduction observed between 1998 and 2004 is extremely important both in<br />

time and space, with rates ranging between ~0.05 and 0.65 gC m -2 y -1 . Most of this<br />

variability is, neverthe<strong>les</strong>s, explained by the variability in sea ice concentration.<br />

187


Figure VI.8. Annual maps of DIC photoproduction rates from 1998 to 2004 in southeastern Beaufort Sea.<br />

188


VI.4. Discussion<br />

VI.4.1. SeaWiFS data quality<br />

The visual inspection of several SeaWiFS images confirms that the Rrs(412)<br />

retrieved using the standard AC algorithm was often negative, or unrealistically too low,<br />

even several kilometers away from the turbid waters. It is not clear why the standard AC<br />

produces systematic un<strong>de</strong>restimation of Rrs(412) over clear waters. Because this<br />

systematic un<strong>de</strong>restimation of the water-leaving reflectance at 412 nm results in negative<br />

values or severe un<strong>de</strong>restimation of the log[Rrs(412)/Rrs(555)] ratio (Fig. VI.4a), the<br />

whole SeaWiFS data set was processed using the turbid water AC. I<strong>de</strong>ally, the turbid<br />

water atmospheric correction should be applied only over area where seawater is actually<br />

turbid. For the three match-ups where clear waters were present, the aerosols optical<br />

thickness at 865 nm was reduced by a value of ~0.01 (from ~0.04 to 0.03) using the<br />

turbid water AC, resulting into an un<strong>de</strong>restimation of the aerosol radiances over the<br />

whole visible spectrum and an overestimation of the magnitu<strong>de</strong> of the Rrs (Fig. VI.3).<br />

Several reasons can be invoked to explain this results. First, the radiometric calibration of<br />

the NIR channels may not be perfect, leading to the selection of an aerosol type with<br />

higher spectral <strong>de</strong>pen<strong>de</strong>ncy than it should. This would explained the systematic<br />

un<strong>de</strong>restimation of Rrs < 490 nm, and the relatively good retrieval for channels ≥ 490 nm<br />

obtained with the standard AC algorithm (Fig. VI.3). Second, in clear waters, the<br />

765:865-nm water-leaving reflectance ratio tends to be higher than the value of 1.72 that<br />

is given as a constant in the turbid water AC algorithm (see Figure 4 in Ruddick et al.<br />

[2006]). Consequently, when applying the turbid water AC algorithm over clear water<br />

pixels, the concentration of aerosol is un<strong>de</strong>restimated, obtained by solving the set of two<br />

equations for the near infrared bands (for <strong>de</strong>tails see Ruddick et al. [2000]). While fixing<br />

the type of aerosol significantly reduced the amount of negative reflectance at 412 nm, it<br />

causes an un<strong>de</strong>restimation of the aerosol concentration over the clear waters.<br />

One way to obtain consistent reflectance in the blue part of the spectrum may be<br />

to revise the vicarious calibration coefficients. The vicarious calibration coefficients are<br />

factors applied to the TOA radiances that have been estimated a posteriori to improve the<br />

quality of the water-leaving radiance. The current coefficients were <strong>de</strong>termined using<br />

189


MOBY (Marine Optical Buoy) in situ match-up located west of Hawaiian island of Lanai<br />

characterised by Case 1 water with low pigment concentration. It is known that if the<br />

environmental conditions of a given region, in particular the type and the concentration of<br />

atmospheric aerosols, <strong>de</strong>part from the “typical” conditions encountered at the MOBY site,<br />

then vicarious coefficients may not be valid [e.g. of the Baltic Sea, Oh<strong>de</strong> et al., 2002]. In<br />

addition, the sun zenith angle encountered at the MOBY site is always lower than 50°<br />

(except at the winter). For example, a recent validation of satellite reflectances obtained<br />

over the Adriatic Sea indicates a <strong>de</strong>gradation of the data quality when the sun zenith<br />

angle exceed 40° [Zibordi et al., 2006]. To revise the SeaWiFS vicarious calibration for<br />

the Arctic Ocean, however, a large number of match-ups is required, which are scarce in<br />

this region. Clearly, more work is nee<strong>de</strong>d to better un<strong>de</strong>rstand the reasons of the failure of<br />

the standard AC algorithm over the clear water area.<br />

VI.4.2. Quantification of DIC photoproduction using satellite Ocean Color<br />

To my knowledge, there is only three other studies that proposed the used of<br />

Ocean Color data to calculate photochemical rates in the marine environment. The first<br />

two methods, proposed by Cullen et al. [1997] and Johannessen et al. [2003], were never<br />

tested nor validated on satellite imagery to <strong>de</strong>rive <strong>de</strong>pth-integrated photoproduction of<br />

DIC. More recently, Fichot [2004] proposed a method to <strong>de</strong>rive empirically the diffuse<br />

attenuation coefficients at several UV wavelengths and CDOM absorption coefficient at<br />

320 nm. A data set gathering 335 in situ mea<strong>sur</strong>ements was used to <strong>de</strong>velop his method,<br />

which was validate using a synthetic data set (IOCCG) and 217 SeaWiFS match-ups data<br />

set. Fichot [2004] further apply his method to SeaWiFS imagery to <strong>de</strong>rive <strong>de</strong>pth-resolved<br />

and <strong>de</strong>pth-integrated photoproduction of carbon monoxi<strong>de</strong> (CO) at the global scale.<br />

These methods are briefly <strong><strong>de</strong>s</strong>cribed and discussed below.<br />

Cullen et al. [1997] suggested an approach in several steps which makes use of<br />

the SeaWiFS 412 nm sensor: (1) find empirical relationships between the ratio of<br />

reflectance at two visible wavelengths (412 nm and 555 nm) and diffuse attenuation<br />

coefficient (Kd) at several UV wavelengths; (2) calculate spectrally resolved total UV<br />

absorption from Kd(UV); (3) <strong>de</strong>termine the magnitu<strong>de</strong> of absorption by partic<strong>les</strong>,<br />

particularly phytoplankton; (4) subtract absorption due to partic<strong>les</strong> and water from the<br />

190


total to calculate the CDOM absorption; and (5) apply an action spectrum to calculate<br />

photochemical production. The estimates of absorption and attenuation could be<br />

combined with <strong>sur</strong>face irradiance to calculate absorption by CDOM as a function of<br />

<strong>de</strong>pth.<br />

Johannessen et al. [2003] modified Cullen’s approach by finding direct,<br />

empirical relationships between the Rrs(412)/Rrs(555) and Kd(UV), and then by using a<br />

constant ratio between Kd(UV) and aCDOM(UV). These authors did not, however, used<br />

their method to estimate the <strong>de</strong>pth-integrated DIC photoproduction. But for her DIC<br />

photoproduction calculation, Johannessen [2000] only used the magnitu<strong>de</strong> of the<br />

SeaWiFS-<strong>de</strong>rived Kd to <strong>de</strong>fine three water classes for which φDIC(λ) were <strong>de</strong>termined<br />

[Johannessen and Miller, 2001]. Johannessen [2000] then computed the <strong>de</strong>pth-integrated<br />

DIC photoproduction in the wavelength range going from 300 to 450 nm assuming a<br />

constant [aCDOM/at](λ) value of 1.0.<br />

Fichot [2004] proposed two methods to estimated Kd at 320, 340, 380, 412, 443,<br />

and 490 nm based on the principal components (PC) of the visible Rrs spectrum (at six<br />

SeaWiFS bands 412, 443, 490, 510, 555 and 670 nm). In the first one (SeaUV), the log-<br />

transformed Rrs spectrum is reduced into the first four PCs which are correlated to Kd at<br />

different wavelengths through a multiple regression. Next, aCDOM(λ) is obtained assuming<br />

a constant ratio between Kd(320) and aCDOM(320) and a spectral slope for the CDOM<br />

absorption spectrum (SCDOM) of 0.0194 nm -1 . In the second method (SeaUVc), the water<br />

is first classified using the first 2 PCs into one of the seven pre<strong>de</strong>fined classes. The Kd at<br />

different wavelengths is then calculated empirically using the coefficients <strong>de</strong>termined for<br />

the given class. For each class, a different value for the ratio Kd(323)/aCDOM(323) is used<br />

with SCDOM = 0.0194 nm -1 to obtain aCDOM(λ). With both aCDOM(λ) and Kd(λ), and further<br />

assumptions on the light field distribution (i.e., the conversion of the planar into scalar<br />

irradiance), Fichot [2004] estimated the <strong>de</strong>pth-resolved photoproduction of carbon<br />

monoxi<strong>de</strong> (CO). However, the global maps of <strong>de</strong>pth-integrated CO photoproduction<br />

presented by Fichot [2004] follow the inci<strong>de</strong>nt irradiance distribution and are mostly<br />

in<strong>de</strong>pen<strong>de</strong>nt of the water column optics. This is because the proposed method by Fichot<br />

[2004] only captures part of the variability of the contribution of CDOM to the total light<br />

absorption. There are some evi<strong>de</strong>nces suggesting limitations in Fichot’s method for the<br />

191


estimation of the <strong>de</strong>pth-integrated photoproduction of CO (or DIC). First, in large sub-<br />

tropical ultraoligotrophic gyre, CDOM is almost impossible to mea<strong>sur</strong>e and pure water is<br />

likely completely dominating the full UV-visible [Morel et al., in press in L&O, 2006].<br />

Thus the <strong>de</strong>pth-integrated CO photoproduction should be much lower in those area.<br />

Similarly, the <strong>de</strong>pth-integrated CO photoproduction should be reduced in coastal waters<br />

rich in particulate matter, and higher in river plumes where CDOM is concentrated at the<br />

<strong>sur</strong>face (e.g. Amazone and Oricono Rivers). Those general trends were not visible on the<br />

maps presented by Fichot [2004], likely because of the assumptions ma<strong>de</strong> on the ratio<br />

Kd(323)/aCDOM(323). Therefore, if one would like to adopt this method to estimate the<br />

<strong>de</strong>pth-integrated CO (or DIC) photoproduction, the validity of the above assumptions on<br />

the ratio Kd(323)/aCDOM(323) adopted by Fichot [2004] need to be evaluated in more<br />

<strong>de</strong>tails.<br />

We proposed an alternative method to use Ocean Color data for photooxidation<br />

quantification in which <strong>de</strong>pth-integrated DIC photoproduction varies as function of<br />

[aCDOM/at]. Because assuming constant proportion [aCDOM/at], as for example 0.90, may<br />

lead to significant overestimation of the DIC photoproduction (Table VI.4), I recommend<br />

to use either in situ or satellite-<strong>de</strong>rived [aCDOM/at]. Note that a value of 0.9 may appear<br />

unrealistically high, but based on in situ mea<strong>sur</strong>ements of [aCDOM/at], this value is<br />

generally valid for wavelength < 350 nm, where maximum photooxidation occurs (see<br />

Fig. IV.7).<br />

The difference between total annual DIC photoproduction calculated in the<br />

Beaufort Sea using the satellite-<strong>de</strong>rived and in situ [aCDOM/at] is not significant (i.e. <strong>les</strong>s<br />

than the algorithm accuracy). Because in situ mea<strong>sur</strong>ements of [aCDOM/at] are not<br />

available for every season and region of the Arctic Ocean, then the use of ocean color<br />

data could be useful. If one would like to use ocean color data for DIC photoproduction<br />

calculation, a particular attention should be paid to the data processing. First, it is<br />

strongly recommen<strong>de</strong>d to process the satellite data on an image-by-image basis in or<strong>de</strong>r<br />

to check the consistency of the water-leaving reflectance retrieval in the blue part of the<br />

spectrum. Second, adjacency effect due to sea ice could be extremely important, in<br />

particular at the beginning of the summer when a significant snow cover is still present.<br />

192


VI.5. Conclusions<br />

I conclu<strong>de</strong> that the proposed method on the use of satellite-<strong>de</strong>rived [aCDOM/at]<br />

could improve the estimation of <strong>de</strong>pth-integrated DIC photoproduction in regions where<br />

in situ mea<strong>sur</strong>ements are lacking. Neverthe<strong>les</strong>s, good estimation can be obtained if<br />

realistic assumption on [aCDOM/at] are ma<strong>de</strong>. Such assumptions, however, should be based<br />

on a sufficient number in situ mea<strong>sur</strong>ements to cover the spatio-temporal variability in<br />

[aCDOM/at]. At the moment, in situ optical mea<strong>sur</strong>ements are lacking for most of the<br />

Arctic coastal area (e.g. Kara and Laptev seas).<br />

The proposed method was tested using SeaWiFS imagery, but it could be applied<br />

to any ocean color sensors having ~412, ~490 and ~555 nm channels (e.g., MERIS and<br />

MODIS). Future improvements in the radiometric calibration and in the atmospheric<br />

correction over turbid waters should simplify the application of the method to satellite<br />

imagery.<br />

One important parameter to take into account for the estimation of DIC<br />

photoproduction is the variability in the apparent quantum yield. Besi<strong>de</strong> providing<br />

[aCDOM/at], Ocean Color data may be useful to predict the variability of φDIC(λ). Here, a<br />

method was tested to account for the variability in φDIC(λ) based on the magnitu<strong>de</strong> of the<br />

CDOM absorption coefficient at 412 nm (Method 2). The lack of information on the<br />

φDIC(λ) preclu<strong><strong>de</strong>s</strong> any further attempt to mo<strong>de</strong>l, or predict, its spatio-temporal variability<br />

from other external environmental conditions (e.g. mixed layer vs photic layer <strong>de</strong>pth,<br />

temperature, turbidity, salinity, etc.). Therefore, the AQY needs to be extensively<br />

documented and its variations un<strong>de</strong>rstood. Besi<strong>de</strong> the origin of CDOM, its light history,<br />

including the role of vertical mixing, is certainly a major aspect of the problem.<br />

VI.6. References<br />

Bailey, S. W., and P. J. Wer<strong>de</strong>ll (2006), A multi-sensor approach for the on-orbit<br />

validation of ocean color satellite data products, Remote Sens. Environ., 102, 12-<br />

23.<br />

Cullen, J. J., R. F. Davis, J. S. Barlett, and W. L. Miller (1997), Toward remote sensing<br />

of UV attenuation, photochemical <strong>flux</strong>es, and biological effects of UV in <strong>sur</strong>face<br />

193


waters, in Current and Emerging Issues in Aquatic Science: Aquatic Science<br />

meeting, Am. Soc. Limnol. and Oceanogr., Santa Fe, N. M.<br />

Fichot, C. G. (2004), Marine photochemistry from space: Algorithms for the Retreival of<br />

Diffuse Attenuation and CDOM absorption Coefficients (320-490 nm) from<br />

Ocean Color and Estimation of Depth-resolved photoproduction Rates of Carbon<br />

Monoxi<strong>de</strong> (CO) at Global Sca<strong>les</strong> using SeaWiFS imagery, Master of Science<br />

thesis, 217 pp., Dalhousie, Halifax, Canada.<br />

Gordon, H. R. (1997), Atmospheric correction of ocean color imagery in the Earth<br />

Observation System era, J. Geophys. Res., 102(D14), 17081-17106.<br />

Gordon, H. R., and M. Wang (1994), Retrieval of water-leaving radiance and aerosol<br />

optical thickness over the oceans with SeaWiFS: a preliminary algorithm, App.<br />

Opt., 33(3), 443-452.<br />

Johannessen, S. C. (2000), A photochemical sink for dissolved organic carbon in the<br />

ocean,176 pp., Dalhousie University, Halifax.<br />

Johannessen, S. C., and W. L. Miller (2001), Quantum yield for the photochemical<br />

production of dissolved inorganic carbon in seawater, Mar. Chem., 76, 271-283.<br />

Johannessen, S. C., W. L. Miller, and J. J. Cullen (2003), Calculation of UV attenuation<br />

and colored dissolved organic matter absorption spectra from mea<strong>sur</strong>ements of<br />

ocean color, J. Geophys. Res., 108(C9), 3301, doi:10.1029/2000JC000514.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and R. A. Arnone (2002), Deriving inherent optical properties<br />

from water color: a multiband quasi-analytical algorithm for optically <strong>de</strong>ep waters,<br />

App. Opt., 41(27), 5755-5772.<br />

Morel, A., B. Gentili, H. Claustre, M. Babin, A. Bricaud, J. Ras, and F. Tièche (in press,<br />

2006), Optical properties of the "clearest" natural waters, Limnol. Oceanogr.<br />

Oh<strong>de</strong>, T., B. Sturm, and H. Siegel (2002), Derivation of SeaWiFS vicarious calibration<br />

coefficients using in situ mea<strong>sur</strong>ements in Case 2 water of the Baltic Sea, Remote<br />

Sens. Environ., 80, 248-255.<br />

Ruddick, K. G., V. De Cauwer, Y.-J. Park, and G. Moore (2006), Seaborne<br />

mea<strong>sur</strong>ements of near infrared water-leaving reflectance: The similarity spectrum<br />

for turbid waters, Limnol. Oceanogr., 51(2), 1167-1179.<br />

Ruddick, K. G., F. Ovidio, and M. Rijkeboer (2000), Atmospheric correction of SeaWiFS<br />

imagery for turbid coastal and inland waters, App. Opt., 39(6), 897-895.<br />

Siegel, D. A., M. Wang, S. Maritorena, and W. Robinson (2000), Atmospheric correction<br />

of satellite ocean color imagery: the black pixel assumption, App. Opt., 39(21),<br />

3582-3591.<br />

194


Zibordi, G., F. Mélin, and J.-F. Berthon (2006), Comparison of SeaWiFS, MODIS and<br />

MERIS radiometric products at a coastal site, Geophys. Res. Lett., 33, L06617,<br />

doi:10.1029/2006GL025778.<br />

195


Chapitre VII: Conclusion générale et perspectives<br />

196


Le couvert <strong>de</strong> glace pluriannuel <strong>de</strong> l’Océan Arctique se rétréci en étendu et en<br />

épaisseur, laissant <strong>de</strong>rrière <strong><strong>de</strong>s</strong> étendues d’eau ouverte <strong>de</strong> plus en plus gran<strong><strong>de</strong>s</strong>. En parallèle,<br />

au cours <strong><strong>de</strong>s</strong> <strong>de</strong>rnières décennies, on a assisté à une augmentation significative du<br />

rayonnement UV reçu à la <strong>sur</strong>face <strong>de</strong> la mer en réponse à la diminution <strong>de</strong> l’ozone<br />

stratosphérique. Le travail réalisé dans le cadre <strong>de</strong> cette thèse représente un effort afin<br />

d’appréhen<strong>de</strong>r l’impact <strong><strong>de</strong>s</strong> ces <strong>changements</strong> environnementaux <strong>sur</strong> le cycle du <strong>carbone</strong><br />

organique dissous (DOC) dans l’Océan Arctique. En particulier, je me suis intéressé au<br />

<strong>carbone</strong> d’origine terrigène (tDOC) car il représente une importante fraction du pool <strong>de</strong><br />

DOC dans <strong>les</strong> eaux côtières <strong>de</strong> l’Océan Arctiques et que la fonte du pergélisol risque<br />

d’augmenter gran<strong>de</strong>ment cet apport dans le futur. Afin <strong>de</strong> mieux comprendre le rôle relatif<br />

que joue la lumière dans le cycle du DOC mon objectif principal était d’estimer la quantité<br />

<strong>de</strong> <strong>carbone</strong> organique qui est minéralisée à la <strong>sur</strong>face <strong>de</strong> l’océan par <strong>les</strong> processus <strong>de</strong><br />

photooxydation. Grâce à un modèle couplé optique/photochimique simple nécessitant en<br />

entré <strong>de</strong> l’information dérivée d’observations satellita<strong>les</strong>, j’ai quantifié pour la première fois la<br />

production photochimique <strong>de</strong> <strong>carbone</strong> inorganique dissous (DIC) dans une région <strong>de</strong><br />

l’Arctique.<br />

Les conclusions majeures <strong>de</strong> ce travail sont que : 1) la photoproduction <strong>de</strong> DIC dans<br />

le Sud-est <strong>de</strong> la Mer <strong>de</strong> Beaufort a augmenté <strong>de</strong> ~15% au cours <strong>de</strong> la pério<strong>de</strong> allant <strong>de</strong> 1979 à<br />

2004, principalement en réponse à la diminution <strong>de</strong> l’étendue <strong>de</strong> glace <strong>de</strong> mer, et 2) <strong>les</strong> taux<br />

<strong>de</strong> minéralisation du tDOC sont pratiquement équivalents aux taux actuels d’enfouissement<br />

<strong>de</strong> <strong>carbone</strong> organique particulaire dans <strong>les</strong> sédiments marins. Cette <strong>de</strong>rnière conclusion<br />

confirme l’importance relative <strong>de</strong> ce <strong>flux</strong> pour la balance nette du Bilan <strong>de</strong> Carbone<br />

Organique Arctique. En raison <strong>de</strong> l’intensité <strong>de</strong> sa stratification, <strong>les</strong> eaux <strong>de</strong> <strong>sur</strong>face arctiques<br />

sont généralement riches en tDOC et pauvres en nutriments inorganiques, expliquant en<br />

partie sa faible production primaire (e.g., couche euphotique plus mince due à la présence <strong>de</strong><br />

CDOM, mélange verticale faible, etc.). Si la minéralisation <strong>de</strong> <strong>carbone</strong> organique allochtone<br />

(i.e., terrigène) par <strong>les</strong> processus <strong>de</strong> « photooxydation/respiration bactérienne » est plus<br />

gran<strong>de</strong> que la séquestration <strong>de</strong> <strong>carbone</strong> organique autochtone (i.e. marin) dans <strong>les</strong> sédiments<br />

profonds, alors l’Océan Arctique doit être considéré comme une source nette <strong>de</strong> CO 2 pour<br />

la biosphère (i.e. hétérotrophie nette). La question à savoir si l’Océan Arctique agit comme<br />

un puits ou une source <strong>de</strong> <strong>carbone</strong> reste entièrement ouverte à ce jour.<br />

197


Cette thèse représente un premier pas vers une meilleure compréhension du rôle que<br />

joue la photooxydation du CDOM dans le bilan <strong>de</strong> <strong>carbone</strong> organique <strong>de</strong> l’Arctique. Elle<br />

fournit aussi <strong><strong>de</strong>s</strong> pistes quant à l’utilisation simultanée <strong>de</strong> paramètres divers dérivés <strong><strong>de</strong>s</strong><br />

données acquises <strong>de</strong>puis <strong><strong>de</strong>s</strong> plateformes satellita<strong>les</strong> pour quantifier, à <strong>de</strong> larges échel<strong>les</strong><br />

spatia<strong>les</strong>, <strong><strong>de</strong>s</strong> processus physico-chimiques telle que la photooxydation du CDOM. Dans ce<br />

qui suit, sont discutés certains aspects liés à la télédétection <strong>de</strong> la couleur <strong>de</strong> l’Océan, et à la<br />

quantification <strong>de</strong> la photooxydation dans <strong>les</strong> eaux <strong>de</strong> l’Océan Arctique.<br />

Vers une estimation semi-analytique du coefficient d’absorption du CDOM à partir<br />

<strong>de</strong> la couleur <strong>de</strong> l’océan<br />

Puisque la modélisation <strong>de</strong> la photooxydation du CDOM nécessite la connaissance<br />

<strong>de</strong> certaines propriétés optiques <strong>de</strong> l’eau <strong>de</strong> mer, un effort particulier a été mis <strong>sur</strong><br />

l’extraction <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan <strong><strong>de</strong>s</strong> paramètres bio-optiques pertinents. Or<br />

dans la littérature, <strong>les</strong> algorithmes semi-analytiques d’inversion proposés ne permettent pas<br />

<strong>de</strong> distinguer <strong>les</strong> coefficients d’absorption du CDOM et <strong><strong>de</strong>s</strong> particu<strong>les</strong> non-alga<strong>les</strong> (NAP).<br />

Ceci s’explique par la similarité <strong>de</strong> leurs spectres d’absorption respectifs. J’ai donc proposé<br />

un simple algorithme empirique permettant <strong>de</strong> retrouver directement la contribution du<br />

CDOM au coefficient d’absorption totale ([a CDOM/a t]) directement à partir du spectre <strong>de</strong><br />

réflectance <strong>de</strong> l’eau.<br />

Cependant, <strong>les</strong> relations empiriques peuvent empêcher notre compréhension <strong><strong>de</strong>s</strong><br />

fon<strong>de</strong>ments physiques, et par conséquent, limiter <strong>les</strong> futurs avancements dans le domaine <strong>de</strong><br />

la télédétection <strong>de</strong> la couleur <strong>de</strong> l’océan (IOCCG, 2006). Pour cette raison, <strong>les</strong> approches<br />

semi-analytiques basées <strong>sur</strong> <strong>les</strong> fon<strong>de</strong>ments physiques <strong>de</strong> l’optique hydrologique <strong>de</strong>vraient<br />

êtres préférées aux métho<strong><strong>de</strong>s</strong> empiriques. Or pour estimer a CDOM à l’ai<strong>de</strong> d’algorithmes semi-<br />

analytiques, <strong>les</strong> modè<strong>les</strong> bio-optiques se doivent d’être formulés <strong>de</strong> manière différente <strong>de</strong><br />

celle présentement adoptée. Le fait que, contrairement au CDOM, <strong>les</strong> NAP diffusent la<br />

lumière offre un pouvoir discriminant lorsque ceux-ci sont considérés individuellement dans<br />

la formulation du modèle d’IOPs. Ainsi, pour que <strong>les</strong> algorithmes semi-analytiques soient<br />

performant dans la discrimination <strong>de</strong> a CDOM et a NAP, ils <strong>de</strong>vront se baser <strong>sur</strong> un modèle<br />

d’IOPs qui inclura un lien robuste entre <strong>les</strong> propriétés <strong>de</strong> rétrodiffusion et d’absorption <strong><strong>de</strong>s</strong><br />

NAP (e.g. Gallegos et Neale, 2002). Une difficulté majeure d’un tel paramétrage est qu’il<br />

n’est présentement pas possible <strong>de</strong> me<strong>sur</strong>er distinctement <strong>les</strong> propriétés <strong>de</strong> rétrodiffusion <strong><strong>de</strong>s</strong><br />

198


particu<strong>les</strong> alga<strong>les</strong> et non-alga<strong>les</strong>. Le succès <strong>de</strong> notre algorithme empirique suggère néanmoins<br />

que ce type <strong>de</strong> paramétrage est possible dans certain environnement (e.g. Mer Baltique et<br />

Mer du Nord).<br />

Un autre aspect du problème lié à l’utilisation <strong><strong>de</strong>s</strong> modè<strong>les</strong> semi-analytiques concerne<br />

la métho<strong>de</strong> avec laquelle l’information contenue dans le spectre <strong>de</strong> l’eau est extraite. Dans <strong>les</strong><br />

eaux côtières, <strong>les</strong> variations dans <strong>les</strong> spectres <strong>de</strong> réflectance sont principalement dues aux<br />

variations <strong>de</strong> la turbidité <strong>de</strong> l’eau (e.g., Sathyendranath et al., 1989). Les subti<strong>les</strong> variations<br />

pouvant résultées <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>de</strong> proportions entre le phytoplancton, le CDOM et <strong>les</strong><br />

NAP, sont « camouflés » par <strong>les</strong> <strong>changements</strong> d’amplitu<strong>de</strong> <strong>de</strong> la réflectance causés par <strong>les</strong><br />

variations dans la turbidité à toutes <strong>les</strong> longueurs d’on<strong>de</strong>. Par conséquent, lorsqu’une<br />

information bien spécifique se doit d’être dérivée d’un spectre <strong>de</strong> réflectance, comme par<br />

exemple [a CDOM/a t], seu<strong>les</strong> <strong>les</strong> parties du spectre fournissant l’information la moins ambiguë<br />

possible se doivent d’être considérées par le modèle d’inversion. La justification du choix<br />

<strong><strong>de</strong>s</strong> longueurs d’on<strong><strong>de</strong>s</strong> utilisées dans notre algorithme empirique est donc tout autant valable<br />

pour <strong>les</strong> algorithmes semi-analytiques.<br />

Est-ce que <strong>les</strong> effets d’environnement dus à la présence <strong>de</strong> glace <strong>de</strong> mer peuvent être<br />

corrigés?<br />

La télédétection spatiale <strong>de</strong> la couleur <strong>de</strong> l’océan peut fournir <strong><strong>de</strong>s</strong> informations uti<strong>les</strong><br />

pour la quantification <strong><strong>de</strong>s</strong> processus photochimiques dans <strong>les</strong> eaux <strong>de</strong> l’Arctique souvent<br />

inaccessib<strong>les</strong>. Mais l’utilisation <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan dans <strong>les</strong> hautes latitu<strong><strong>de</strong>s</strong> est<br />

compliquée par la présence <strong>de</strong> glace <strong>de</strong> mer. Au début <strong>de</strong> la pério<strong>de</strong> estivale, près <strong>de</strong> la moitié<br />

<strong><strong>de</strong>s</strong> eaux ouvertes observéee <strong>de</strong>puis l’espace sont contaminées par <strong>les</strong> effets d’environnement.<br />

Ce phénomène comprommets sévèrement l’utilisation <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan<br />

dans <strong>les</strong> mers polaires où la glace est omniprésente. Par conséquent, il serait souhaitable <strong>de</strong><br />

mettre au point une métho<strong>de</strong> permettant <strong>de</strong> corriger <strong>les</strong> effets d’environement. Ceci pourrait<br />

être fait en générant <strong><strong>de</strong>s</strong> tab<strong>les</strong> pré-calculées (LUTs) additionnel<strong>les</strong> pour <strong>les</strong>quel<strong>les</strong> <strong>les</strong> effets<br />

d’environnement seraient inclus. Dans ce cadre, une étu<strong>de</strong> plus approfondie <strong>de</strong> la variabilité<br />

<strong>de</strong> la réflectance <strong>de</strong> la glace mer, aux échel<strong>les</strong> correspondant aux me<strong>sur</strong>es <strong>de</strong> couleur <strong>de</strong><br />

l’océan (i.e. ~1 Km <strong>de</strong> résolution), serait utile pour définir <strong>les</strong> types <strong>de</strong> glace <strong>les</strong> plus<br />

fréquents dans l’Arctique. Pour chaque type <strong>de</strong> glace, une LUT pourrait être générée. La<br />

réflectance <strong>de</strong> la glace pourrait éventuellement être me<strong>sur</strong>ée à partir du même instrument, et<br />

199


ainsi le type <strong>de</strong> glace i<strong>de</strong>ntifié <strong>de</strong>puis l’espace. Finalement, ces LUTs pourraient être utilisées<br />

dans le traitement <strong><strong>de</strong>s</strong> données uniquement lorsque <strong><strong>de</strong>s</strong> pixels contaminés par <strong>les</strong> effets<br />

d’environnement sont détectés avec la métho<strong>de</strong> proposée dans cette thèse.<br />

À propos <strong>de</strong> la variabilité naturelle du ren<strong>de</strong>ment quantique apparent pour la<br />

photoproduction <strong>de</strong> DIC, et <strong>de</strong> sa détermination<br />

Le φ DIC est un paramètre fondamental pour la modélisation <strong>de</strong> la photooxydation du<br />

CDOM. Il est, à l’heure actuelle, <strong>de</strong> loin le paramètre du modèle <strong>de</strong> photooxydation le moins<br />

bien documenté. Quelle est la représentativité <strong><strong>de</strong>s</strong> spectres <strong>de</strong> φ DIC me<strong>sur</strong>és dans <strong>les</strong> eaux du<br />

Fleuves Mackenzie, du Plateau du Mackenzie, et du Golfe d’Amundsen, du pool <strong>de</strong> CDOM<br />

<strong>de</strong> l’Arctique? Pour étendre notre étu<strong>de</strong> régionale à l’échelle <strong>de</strong> l’Océan Arctique, il est<br />

nécessaire <strong>de</strong> déterminer φ DIC dans <strong>les</strong> autres régions côtières <strong>de</strong> l’Arctique. En général, on<br />

s’attend à ce que la composition <strong>de</strong> la matière organique dissoute (DOM) provenant <strong><strong>de</strong>s</strong><br />

quatre plus grands fleuves arctiques (i.e, Ob, Yenisei, Lena and Mackenzie) diffère<br />

significativement. D’une part, <strong>les</strong> structures phénoliques <strong>de</strong> la lignine (Benner et al., 2005), <strong>les</strong><br />

aci<strong><strong>de</strong>s</strong> aminées et <strong>les</strong> sucres neutres (Amon et Benner, 2003), <strong>les</strong> propriétés <strong>de</strong> fluorescence<br />

(Amon et al., 2003; Retamal et al., 2006), tous indiquent <strong><strong>de</strong>s</strong> différences dans la composition<br />

moléculaire <strong>de</strong> la DOM provenant ces différents systèmes. D’autre part, <strong>les</strong> caractéristiques<br />

physiques et chimiques du Fleuve Mackenzie semblent être particulièrement différentes<br />

comparées aux autres grands fleuves sibériens : <strong>les</strong> plus frappantes sont ces fortes<br />

concentrations en particu<strong>les</strong> en suspension (> 4.3 fois; Rachold et al., 2004), son alcalinité et<br />

son pH élevés (e.g., Millot et al., 2003). Sous <strong><strong>de</strong>s</strong> conditions aussi différentes, on peut<br />

s’attendre à <strong><strong>de</strong>s</strong> différences significatives quant à la photoréactivité du CDOM entre ces<br />

environnements côtiers. Comme il a été mentionné précé<strong>de</strong>mment, l’augmentation attendue<br />

<strong>de</strong> la photominéralisation du tDOC est importante pour déterminer si l’Océan Arctique agira<br />

comme un puits ou une source <strong>de</strong> CO 2 pour la biosphère dans le futur. Pour mieux évaluer<br />

cet impact, il est nécessaire d’estimer la proportion relative <strong>de</strong> CDOM d’origine terrigène<br />

versus marine, ainsi que <strong>de</strong> connaître leur photoréactivité respective. La proportion <strong>de</strong><br />

chacun <strong>de</strong> ces composants est actuellement inconnue. Dans cette étu<strong>de</strong>, on a me<strong>sur</strong>é φ DIC <strong>de</strong><br />

l’ensemble du pool <strong>de</strong> CDOM. Il nous a donc par été possible d’estimer avec précision<br />

combien <strong>de</strong> DIC fut réellement produit suite à la photooxydation du tDOC. Par conséquent,<br />

200


il semble que φ DIC doit être déterminé en parallèle avec une caractérisation moléculaire du<br />

pool <strong>de</strong> CDOM.<br />

Outre la composition moléculaire du CDOM, la dose <strong>de</strong> lumière absorbée par une<br />

molécule <strong>de</strong> CDOM précé<strong>de</strong>nt son échantillonnage (i.e. son « histoire lumineuse ») est<br />

certainement importante pour déterminer sa photoréactivité. Comme il est pratiquement<br />

impossible <strong>de</strong> connaître l’histoire lumineuse d’un échantillon récolté en mer à un temps t,<br />

l’effet <strong>de</strong> l’histoire lumineuse <strong>sur</strong> φ DIC se doit d’être évaluée en laboratoire. Les étu<strong><strong>de</strong>s</strong> <strong>sur</strong> <strong>les</strong><br />

ren<strong>de</strong>ment quantique apparent (AQYs) autres que pour le CO 2 fournissent <strong><strong>de</strong>s</strong> pistes quant à<br />

l’effet <strong>de</strong> l’histoire lumineuse <strong>sur</strong> la photoréactivité du CDOM. Par exemple, Andrews et al.<br />

(2000) rapportaient une diminution <strong><strong>de</strong>s</strong> AQYs pour la consommation photochimique d’O 2<br />

et la photoproduction <strong>de</strong> H 2O 2, alors que Zhang et Xie (2006) observaient <strong><strong>de</strong>s</strong> tendances<br />

similaires pour <strong>les</strong> AQYs pour la photoproduction <strong>de</strong> CO. Cependant, l’effet <strong>de</strong> l’histoire<br />

lumineuse <strong>sur</strong> φ DIC est inconnu, et <strong>les</strong> résultats trouvés dans la littérature sont conflictuels.<br />

Vähätälo et Wetzel (2004) concluaient que <strong>les</strong> AQYs pour la perte en DOC (i.e. ~φ DIC)<br />

diminuaient en fonction <strong>de</strong> la dose <strong>de</strong> lumière absorbée par le CDOM. À l’opposé,<br />

Johannessen et Miller (2001) ont trouvé que le φ DIC avait pratiquement doublé après le<br />

photo-blanchissement intensif <strong>de</strong> l’échantillon (i.e., diminution <strong>de</strong> a CDOM <strong>de</strong> > 50%). Puisse<br />

qu’il est impossible <strong>de</strong> tirer <strong><strong>de</strong>s</strong> conclusions claires <strong>de</strong> ces observations, plus d’étu<strong><strong>de</strong>s</strong> sont<br />

nécessaires afin <strong>de</strong> déterminer l’effet <strong>de</strong> l’histoire lumineuse <strong>sur</strong> φ DIC.<br />

Parmi <strong>les</strong> autres aspects qui <strong>de</strong>man<strong>de</strong>nt plus d’étu<strong>de</strong>, l’impact <strong>de</strong> l’acidification que<br />

doivent subir <strong>les</strong> échantillons avant leurs incubations semble être cruciale. Cette étape est<br />

essentiel pour éliminer le DIC initial contenu dans l’échantillon qui masquerait totalement la<br />

photoproduction durant <strong>les</strong> incubations. Une façon d’évaluer <strong>les</strong> effets <strong>de</strong> l’acidification est<br />

<strong>de</strong> regar<strong>de</strong>r si <strong><strong>de</strong>s</strong> <strong>changements</strong> dans le AQY pour la photoproduction <strong>de</strong> CO (i.e. un proxy<br />

pour φ DIC) <strong>sur</strong> <strong><strong>de</strong>s</strong> échantillons traités <strong>de</strong> la même manière que ceux pour la détermination <strong>de</strong><br />

φDIC. Si l’acidification induit <strong><strong>de</strong>s</strong> <strong>changements</strong> majeurs dans <strong>les</strong> valeurs <strong>de</strong> CO-AQY, on peut<br />

donc s’attendre à <strong><strong>de</strong>s</strong> <strong>changements</strong> similaires <strong>de</strong> φDIC. Une autre métho<strong>de</strong> pourrait être <strong>de</strong><br />

travailler avec <strong><strong>de</strong>s</strong> échantillons provenant <strong>de</strong> lacs où <strong>les</strong> concentrations en DIC sont<br />

naturellement faib<strong>les</strong>. Ainsi, l’impact <strong>de</strong> l’acidification pourrait être évalué directement <strong>sur</strong><br />

φDIC. À côté <strong>de</strong> l’acidification, <strong>les</strong> effets <strong>de</strong> l’intensité <strong>de</strong> la lumière utilisé durant <strong>les</strong><br />

incubations et le stockage <strong><strong>de</strong>s</strong> échantillons pendant <strong>de</strong> longue pério<strong>de</strong> <strong>de</strong> temps (i.e. > 4<br />

201


mois) peuvent aussi avoir un impact <strong>sur</strong> l’amplitu<strong>de</strong> <strong>de</strong> φ DIC. Il est clair que ces aspects<br />

méthodologiques <strong>sur</strong> la détermination <strong>de</strong> φ DIC méritent plus d’attention.<br />

Enfin, si on arrive à mieux comprendre quels sont <strong>les</strong> paramètres biologiques,<br />

chimiques et physiques qui contrôlent φ DIC, ainsi il sera possible d’estimer, ou d’approximer,<br />

sa variabilité à partir d’observations environnementa<strong>les</strong>.<br />

Sur la modélisation <strong>de</strong> φφφφ DIC à partir d’informations dérivées <strong><strong>de</strong>s</strong> données satellita<strong>les</strong><br />

Le plein potentiel <strong><strong>de</strong>s</strong> observations satellita<strong>les</strong> pour la quantification <strong>de</strong> la<br />

photoproduction <strong>de</strong> DIC n’a pas été totalement exploré. En plus <strong>de</strong> fournir le rapport entre<br />

le a CDOM et a t, <strong>les</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan peuvent servir à modéliser la variabilité <strong>de</strong><br />

φ DIC. Par exemple, il semble qu’une relation entre φ DIC et a CDOM(412) pourrait exister (Chap.<br />

VI). D’autres produits dérivés <strong><strong>de</strong>s</strong> données <strong>de</strong> couleur <strong>de</strong> l’océan, comme la chlorophylle et<br />

le coefficient d’atténuation diffus, peuvent certainement fournir <strong><strong>de</strong>s</strong> informations clés. Alors<br />

que la chlorophylle est un proxy <strong>de</strong> la productivité primaire, et donc <strong>de</strong> la production <strong>de</strong><br />

CDOM marin, l’atténuation diffuse peut servir à déterminer la profon<strong>de</strong>ur <strong>de</strong> la couche<br />

euphotique dans laquelle la photoproduction <strong>de</strong> DIC prend place.<br />

Un modèle prédictif <strong>de</strong> φ DIC, basé <strong>sur</strong> <strong><strong>de</strong>s</strong> observations satellita<strong>les</strong>, doit probablement<br />

être développé en parallèle avec un modèle <strong>de</strong> mélange vertical. En effet, <strong>les</strong> taux <strong>de</strong> mélange<br />

verticaux et la profon<strong>de</strong>ur <strong>de</strong> la couche euphotique doivent être connus pour évaluer<br />

l’histoire lumineuse du CDOM, et ainsi pour prédire <strong>les</strong> <strong>changements</strong> <strong>de</strong> φ DIC. Dans un<br />

premier temps, la profon<strong>de</strong>ur <strong>de</strong> la couche mélangée peut être dérivée <strong><strong>de</strong>s</strong> données<br />

climatologiques <strong>de</strong> température et salinité (e.g. Steele et al., 2001). Mais un modèle <strong>de</strong><br />

mélange vertical utilisant <strong>les</strong> forcages dérivés d’observations satellita<strong>les</strong> (vitesse <strong>de</strong> vent,<br />

température <strong>de</strong> <strong>sur</strong>face, éclairement solaire, glace <strong>de</strong> mer, etc.) pourait être développé dans le<br />

futur pour l’Océan Arctique. Encore une fois, le manque d’information <strong>sur</strong> φ DIC empêche,<br />

pour le moment, toute tentative <strong>de</strong> modélisation <strong>de</strong> sa variabilité spatio-temporelle à partir<br />

d’autres observations environnementa<strong>les</strong>.<br />

Implication <strong><strong>de</strong>s</strong> <strong>changements</strong> <strong>climatiques</strong> <strong>sur</strong> la minéralisation du Carbone<br />

Organique Terrigène dans l’Océan Arctique<br />

202


La couche mélangée polaire Arctique (PML) contient <strong>les</strong> plus fortes concentrations<br />

<strong>de</strong> tDOC par rapport aux autres océans. Ceci est en partie le résultat <strong>de</strong> la gran<strong>de</strong> quantité<br />

d’apport terrigène qui alimente la PML par rapport à son volume. Une autre explication est<br />

la présence <strong>de</strong> ce couvert <strong>de</strong> glace permanent qui réduit fortement la pénétration <strong><strong>de</strong>s</strong> UV<br />

dans la colonne d’eau, limitant ainsi la dégradation du tDOC par <strong>les</strong> processus <strong>de</strong><br />

photooxydation.<br />

Selon plusieurs étu<strong><strong>de</strong>s</strong> récentes (e.g., Opsahl et al., 1999; Amon et al., 2003; Hansell et<br />

al., 2004), environ la moitié <strong><strong>de</strong>s</strong> apports annuels en tDOC dans la PML sont exportés vers<br />

l’Océan Atlantique Nord via <strong>les</strong> Détroit <strong>de</strong> Fram et l’Archipel Canadien. D’autres évi<strong>de</strong>nces<br />

récentes suggèrent qu’une fraction significative <strong>de</strong> ce matériel est ensuite exportée hors <strong>de</strong> la<br />

<strong>sur</strong>face <strong>de</strong> l’océan durant la formation <strong>de</strong> la masse d’Eau Profon<strong>de</strong> Nord Atlantique<br />

(NADW) (e.g., Benner et al., 2005). Dans <strong>les</strong> profon<strong>de</strong>urs <strong>de</strong> l’océan, le temps <strong>de</strong> rési<strong>de</strong>nce<br />

du tDOC est <strong>de</strong> plusieurs centaines d’années (Hernes et Benner, 2006). À l’opposé, la<br />

minéralisation du tDOC dans la couche <strong>de</strong> <strong>sur</strong>face est très rapi<strong>de</strong> en raison <strong>de</strong> l’efficacité <strong><strong>de</strong>s</strong><br />

processus <strong>de</strong> photooxydation et <strong>de</strong> respiration microbienne (e.g. Kieber et al. 1990; Miller et<br />

Zepp 1995; Hernes et Benner 2003), et ainsi ce <strong>carbone</strong> peut en partie retourner en dans<br />

l’atmosphère sous forme <strong>de</strong> CO 2.<br />

En me basant <strong>sur</strong> ces observations, je suggère que la réduction du couvert <strong>de</strong> glace<br />

pluriannuel accélèrera la minéralisation du tDOC dans la couche <strong>de</strong> <strong>sur</strong>face Arctique,<br />

réduisant l’export <strong>de</strong> ce <strong>carbone</strong> dans <strong>les</strong> eaux profon<strong><strong>de</strong>s</strong> océaniques, et augmentant la<br />

quantité <strong>de</strong> CO 2 pouvant retourner dans l’atmosphère. Cependant, comme il a été suggéré<br />

par Gobeil et al. (2001), la réduction du couvert <strong>de</strong> glace peut en contre partie augmenter la<br />

productivité primaire et la séquestration <strong>de</strong> <strong>carbone</strong> organique dans l’Océan Arctique. Par<br />

conséquent, l’augmentation prévue <strong>de</strong> l’oxydation du tDOC en <strong>sur</strong>face suite à une<br />

diminution <strong>de</strong> la glace pourrait être compensée, et peut être même <strong>sur</strong>passée en terme <strong>de</strong><br />

<strong>flux</strong> <strong>de</strong> <strong>carbone</strong>, par l’augmentation <strong>de</strong> la production primaire. Dans mes travaux futurs, <strong>les</strong><br />

conditions <strong>de</strong> glace prévues par <strong>les</strong> modè<strong>les</strong> globaux <strong>de</strong> circulation générale Atmosphère-<br />

Ocean (ACIA, 2005) serviront à évaluer la balance nette entre <strong>les</strong> processus stimulés par la<br />

lumière, c’est-à-dire la photooxydation du tDOC et la production primaire.<br />

203


VII.A. Version anglaise: Overall Conclusion and Perspectives<br />

The permanent sea ice cover of the Arctic Ocean is currently experiencing an<br />

important reduction in its extent and thickness, leaving behind increasingly wi<strong>de</strong>r area of<br />

open water. In parallel, the UV radiation reaching the sea <strong>sur</strong>face has increased<br />

significantly during the last <strong>de</strong>ca<strong>de</strong> as a response of the stratospheric ozone <strong>de</strong>pletion.<br />

The work achieved in this thesis represents an effort to evaluate the consequence of these<br />

environmental changes on the cycling of dissolved organic carbon (DOC) in the Arctic<br />

Ocean. In or<strong>de</strong>r to better un<strong>de</strong>rstand relative role of the light into the DOC cycling, in<br />

particular the terrigenous component (tDOC) as it represent a significant fraction of the<br />

DOC pool in the coastal Arctic Ocean, my main objective was to estimate how much<br />

organic carbon is mineralized in the <strong>sur</strong>face waters through photooxidation. Using a<br />

simplified mo<strong>de</strong>l that uses as inputs information <strong>de</strong>rived from satellite remote sensing<br />

instruments, I quantified for the first time the photoproduction of dissolved inorganic<br />

carbon in a region of the Arctic.<br />

The major conclusions of the present work are that 1) the DIC photoproduction in<br />

the southeastern Beaufort Sea has increased over the period from 1979 to 2004 by ~15%,<br />

mainly in response to the <strong>de</strong>creasing sea ice extent, and 2) the estimated rate of tDOC<br />

mineralization is nearly equivalent to that of particulate marine organic carbon currently<br />

buried into the <strong>de</strong>ep sediments. The latter conclusion confirms the relative importance of<br />

this <strong>flux</strong> in the net balance of the Arctic Organic Carbon Budget. Due to their strong<br />

stratification, the <strong>sur</strong>face Arctic waters are generally rich in tDOC and poor in inorganic<br />

nutrients, which results in low primary production (e.g., thinner euphotic zone due to<br />

high CDOM, weak mixing, etc.). If the mineralization of allochthonus organic carbon (i.e.<br />

terrigenous) through photo-oxidation/bacterial respiration is higher than the sequestration<br />

of autochtonous organic carbon (i.e. marine) in the <strong>de</strong>ep sediment, then the Arctic Ocean<br />

should be consi<strong>de</strong>r as a net source of CO2 for the biosphere (i.e. net heterotrophic).<br />

Whether the Arctic acts as a net source or sink of carbon still remains an open question.<br />

This thesis represents a first step toward a better un<strong>de</strong>rstanding of the role of<br />

CDOM photooxidation in the Arctic Ocean Organic Carbon Budget. It also provi<strong><strong>de</strong>s</strong><br />

valuable insights on how various satellites remote sensing information can be merged to<br />

204


quantify large scale processes such as CDOM photooxidation. Below, a number of<br />

unresolved issues related to the ocean color remote sensing and to the quantification of<br />

CDOM photooxidation in the Arctic Ocean are discussed.<br />

Toward semi-analytical estimation of CDOM absorption coefficient from ocean<br />

color<br />

Because mo<strong>de</strong>ling CDOM photooxidation involved in-water optical properties, a<br />

particular effort was ma<strong>de</strong> to extract relevant bio-optical parameters from satellite Ocean<br />

Color data. The current semi-analytical ocean color inversion algorithms found in the<br />

literature do not distinguish the CDOM and non-algal partic<strong>les</strong> (NAP) absorption<br />

coefficients because of the similarity in their respective spectral shapes. I proposed a<br />

simple empirical algorithm to retrieve the contribution of CDOM to the total absorption<br />

coefficient ([aCDOM/at]) from the remote sensing reflectance.<br />

Simple empirical relationships may prevent un<strong>de</strong>rstanding of the basics and,<br />

therefore, limit advancement in ocean color remote sensing [IOGGC, 2005]. For this<br />

reason semi-analytical approaches based on fundamentals of hydrological optics should<br />

be preferred. To <strong>de</strong>rive aCDOM using semi-analytical algorithms, however, the bio-optical<br />

mo<strong>de</strong>l must be formulated in a way different from that presently adopted. The fact that<br />

NAP scatters light, and CDOM do not, provi<strong><strong>de</strong>s</strong> discrimination power when CDOM and<br />

NAP are distinguished in an IOP mo<strong>de</strong>l. So, for future semi-analytical algorithms to be<br />

successful in discriminating aCDOM from aNAP using water reflectance, they will need to be<br />

based on an IOP mo<strong>de</strong>l that inclu<strong><strong>de</strong>s</strong> a good link between aNAP and bNAP (or bbNAP) [e.g.,<br />

Gallegos and Neale, 2002]. One major difficulty is that the partic<strong>les</strong> backscattering<br />

mea<strong>sur</strong>ements do not make the difference between the algal and non-algal partic<strong>les</strong>. The<br />

success of our empirical algorithm suggests that regional parameterization of the NAP<br />

optical properties is possible in certain environmental conditions (e.g. Baltic Sea and<br />

North Sea).<br />

The necessary information contained in the reflectance spectrum to discriminate<br />

aCDOM from aNAP, however, must be extracted properly, and this is probably one weakness<br />

of current semi-analytical mo<strong>de</strong>ls. In coastal waters, variations in reflectance are<br />

primarily due to variations in turbidity [Sathyendranath et al., 1989]. Subtle changes in<br />

205


the shape of the reflectance spectrum that may result from changes in the proportions of<br />

phytoplankton, CDOM and NAP, are overwhelmed by the change in the magnitu<strong>de</strong> of<br />

reflectance at all wavelengths due to variations in turbidity. Also, when some specific<br />

information is nee<strong>de</strong>d from the reflectance spectrum, such as the ratio of aCDOM to at, only<br />

the spectral regions that provi<strong>de</strong> <strong>les</strong>s ambiguous information must be consi<strong>de</strong>red in the<br />

inversion. The rationa<strong>les</strong> consi<strong>de</strong>red in the <strong>de</strong>velopment of our empirical algorithm are<br />

valid for a semi-empirical algorithm.<br />

Is the correction of the adjacency effect due to sea ice possible?<br />

Ocean Color remote sensing can provi<strong>de</strong> valuable information for the<br />

quantification of photochemical processes in the remote Arctic waters. But the use of<br />

ocean color remote sensing at high latitu<strong>de</strong> is complicated by the presence of sea ice.<br />

During spring time, more than the half of the open water area observed from space is<br />

contaminated by the adjacency effect. This phenomenon severely compromises the use of<br />

ocean color data over ice covered seas. Therefore, an extension of the atmospheric<br />

correction scheme to inclu<strong>de</strong> a correction for adjacency effect is highly <strong><strong>de</strong>s</strong>irable. This<br />

might be done by generating additional look-up-tab<strong>les</strong> (LUTs) that would account for the<br />

adjacency effect. For this purpose, further study <strong>de</strong>dicated to the variability of the sea ice<br />

reflectance would be necessary to <strong>de</strong>fine a number of typical Arctic sea ice reflectance<br />

spectra. For each type of ice, a LUT could be generated. In turns the sea ice reflectance<br />

could be mea<strong>sur</strong>es using the same Ocean Color instrument, and thus the type of ice<br />

obtained from space for the correction. Those additional LUTs could be used in the<br />

processing only when contaminated pixels are <strong>de</strong>tected, for instance, using the flagging<br />

approach <strong>de</strong>veloped in this thesis.<br />

On the natural variability of the apparent quantum yield for the DIC<br />

photoproduction, and its <strong>de</strong>termination<br />

The φDIC is fundamental for mo<strong>de</strong>ling CDOM photooxidation and is <strong>de</strong>finitely the<br />

least documented parameter of photooxidation mo<strong>de</strong>ls. How representative are the φDIC<br />

spectra for the Mackenzie River and Shelf and from the Amundsen Gulf, of the whole<br />

Arctic CDOM pool? To extend our regional study to the whole Arctic Ocean, it is<br />

206


necessary to <strong>de</strong>termine φDIC in other regions of the coastal Arctic. In general, the<br />

composition of DOM from the four major Arctic river systems (i.e. the Ob, Yenisei, Lena<br />

and Mackenzie) is expected to differ to some extent. For example, the lignin phenols<br />

[Benner et al., 2005], amino acids and neutral sugars analysis [Amon and Benner, 2003]<br />

and excitation/emission fluorescence spectrometry [Amon et al., 2003; Retamal et al.,<br />

2006] indicate varying composition of the DOM in these river systems. In addition, the<br />

Mackenzie River is also very different in terms of its physical and chemical environments:<br />

the most striking differences are its high concentration of particulate matter [> 4.3-fold<br />

the Siberian Rivers; Rachold et al., 2004], high alkalinity and pH [e.g., Millot et al.,<br />

2003]. Un<strong>de</strong>r these different conditions, one should expect important differences in their<br />

CDOM photoreactivity. As mentioned above, the expected increase in tDOC<br />

photomineralization is important to <strong>de</strong>termine if the Arctic Ocean will act as a net source<br />

or sink of CO2 for the biosphere in the future. To better evaluate this impact, it is<br />

necessary to assess the relative proportion of terrigenous versus marine CDOM, as well<br />

as their respective photoreactivity. The exact proportion each component is actually<br />

unknown. In this study, we mea<strong>sur</strong>ed the φDIC of the whole CDOM pool. We could not<br />

assess precisely how much of the DIC photoproduced came actually from the terrigenous<br />

component. Therefore, the φDIC needs to be documented in parallel with molecular<br />

characterization of the CDOM pool.<br />

Besi<strong>de</strong> the molecular composition of CDOM itself, the light history of CDOM is<br />

certainly important for <strong>de</strong>termining its photoreactivity. The light history refers to the light<br />

dose received by the CDOM prior to sampling. Because it is almost impossible to know<br />

the light history of water sample at the time of sampling, role of light history in the φDIC<br />

variations should be studied in the laboratory. The studies of the Apparent Quantum<br />

Yield (AQY) other than CO2 provi<strong>de</strong>d some insights regarding the impact of the light<br />

history on the CDOM photoreactivity. For example, Andrews et al. [2000] reported<br />

<strong>de</strong>creasing photochemical O2 consumption AQY and H2O2 photoproduction with the<br />

light dose absorbed by CDOM. Recently, Zhang and Xie [2006, in press in Environ. Sci.<br />

Technol.] reported a similar trend for CO photoproduction AQY. However, the effect of<br />

the light history on φDIC is still unknown, and the results found in the literature are<br />

conflictual. Vähätälo and Wetzel [2004] conclu<strong>de</strong>d that DOC loss AQY (i.e. ~φDIC)<br />

207


<strong>de</strong>creases with increasing aborbed dose. In contrast, Johannessen and Miller [2001]<br />

found that the φDIC nearly doubled after the sample was extensively photobleached (i.e., ><br />

50 %). Therefore, no clear conclusions can be drawn from the above observations. Hence,<br />

further studies are required to clarify the effect of light history of CDOM on φDIC. Again,<br />

to reduce the uncertainty on variability of φDIC resulting from CDOM light history in the<br />

mo<strong>de</strong>ling DIC photoproduction, an extensive data set on φDIC is necessary.<br />

Among the other aspects that require more investigation is the impact of<br />

acidification applied to the sample prior the incubation for φDIC <strong>de</strong>termination. This step<br />

was essential to remove the background in DIC that would mostly mask the DIC<br />

photoproduction during the experiment. One way to assess the effects of acidification is<br />

to look at the changes in the carbon monoxi<strong>de</strong> (CO) AQY (i.e. as a proxy of φDIC) on<br />

samp<strong>les</strong> treated with the same manner as DIC ones. If acidification induces major<br />

changes in the CO AQY, then it likely affects the φDIC as well. Another method would be<br />

first to <strong>de</strong>termine φDIC of a fresh water sample that contains naturally small amount of<br />

DIC (i.e. in a lake), and then, after a 24-h acidification and resetting the original pH, re-<br />

assess φDIC. In addition to the acidification, the effect of light intensity used during the<br />

incubation and sample storage may also have an impact on the magnitu<strong>de</strong> of φDIC. It is<br />

clear that these experimental aspects for the φDIC <strong>de</strong>termination <strong><strong>de</strong>s</strong>erve more study.<br />

Finally, if the main biological, chemical and physical parameters controlling φDIC<br />

are better un<strong>de</strong>rstood, then its variability may be assessed, or approximated, from<br />

environmental observations.<br />

On the mo<strong>de</strong>ling of φφφφDIC from space-<strong>de</strong>rived information<br />

The potential of satellite remote sensing observations to quantify DIC<br />

photoproduction has not been fully explored. Besi<strong><strong>de</strong>s</strong> providing the contribution of<br />

CDOM to the total light absorption coefficient, ocean color data may be used to constrain<br />

the variability of φDIC. For example, it appears that a relationship between φDIC(λ) and the<br />

magnitu<strong>de</strong> of the CDOM absorption coefficient at 412 nm might exist (Chap. VI). Other<br />

ocean color products like the chlorophyll a concentration and the diffuse attenuation<br />

coefficient can certainly provi<strong>de</strong> key information. While chlorophyll is a proxy of<br />

208


primary productivity and thus the production of fresh marine CDOM, the diffuse<br />

attenuation can be used to assess the <strong>de</strong>pth of the photic layer within which DIC<br />

photoproduction takes place.<br />

But a predictive mo<strong>de</strong>l of φDIC based on satellite-<strong>de</strong>rived parameters probably<br />

needs to be <strong>de</strong>veloped in parallel with a vertical mixing mo<strong>de</strong>l. Both the rate of vertical<br />

mixing and the photic layer <strong>de</strong>pth must be known to assess the light history of CDOM,<br />

and thus to predict changes in the φDIC. At a first approximation, mixed-layer <strong>de</strong>pth may<br />

be <strong>de</strong>rived from temperature and salinity climatology [e.g., Steele et al., 2001]. But a<br />

vertical mixing mo<strong>de</strong>l using as input satellite observations (<strong>sur</strong>face winds, sea <strong>sur</strong>face<br />

temperature, solar radiation, sea ice, etc) may also be <strong>de</strong>veloped in the future for the<br />

Arctic Ocean. Again, the lack of information on the φDIC preclu<strong><strong>de</strong>s</strong> any attempt to mo<strong>de</strong>l<br />

or predicts its spatio-temporal variability from observations of the environmental<br />

conditions.<br />

Implication of Climate Change on the mineralization of the Terrigenous Dissolved<br />

Organic Carbon in the Arctic<br />

The Arctic Polar Mixed Layer (PML) contains the highest concentration of tDOC<br />

relative to the other ocean basins. This is mainly a result of the large amount of riverine<br />

input that enter the Arctic Ocean relative to its volume. Another explanation may be the<br />

presence of a permanent sea ice cover strongly reduces the UV radiation penetration in<br />

the water column, limiting the <strong>de</strong>gradation of tDOC by photooxidation processes.<br />

According to a number of recent studies [e.g. Opshal et al., 1999; Amon et al.,<br />

2003; Hansell et al. 2004], about one half of the annual input of tDOC into the PML is<br />

exported to the North Atlantic Ocean via Fram Strait and the Canadian Archipelago.<br />

Recent evi<strong>de</strong>nces further suggest that a significant fraction of this material is further<br />

exported from the ocean <strong>sur</strong>face during the formation of the North Atlantic Deep Water<br />

(NADW) [e.g., Benner et al., 2005]. Interestingly, the resi<strong>de</strong>nce time of the tDOC in the<br />

<strong>de</strong>ep ocean may be of or<strong>de</strong>r several centuries [Hernes and Benner, 2006]. In contrast, in<br />

the <strong>sur</strong>face layer, tDOC is rapidly mineralized due to photooxidation and microbial<br />

processes [e.g., Kieber et al. 1990; Miller and Zepp, 1995; Hernes and Benner, 2003] and<br />

thus partly released to the atmosphere as CO2.<br />

209


Based on these observations, I conclu<strong>de</strong> that the reduction in the sea ice cover will<br />

accelerate the mineralization of tDOC within the <strong>sur</strong>face layer of the Arctic Ocean,<br />

reducing the export of organic carbon to the <strong>de</strong>ep ocean, and increasing the amount of<br />

CO2 potentially released to the atmosphere. However, as suggested by Gobeil et al.<br />

[2001], the reduction in sea ice cover could also increase primary productivity and carbon<br />

sequestration in the Arctic Ocean. Therefore, the expected increase in tDOC oxidation<br />

resulting from the <strong>de</strong>cline in sea ice could be compensated for, and perhaps even<br />

overtaken, by the increase in primary production in terms of carbon <strong>flux</strong>es. In my future<br />

work, the sea cover conditions predicted by the global coupled Atmosphere-Ocean<br />

General Circulation Mo<strong>de</strong>ls (AOGCMs) [ACIA, 2005] will be used to assess the net<br />

balance between light-related carbon <strong>flux</strong>es, i.e. tDOC photooxidation and primary<br />

production.<br />

210


Bibliographie générale<br />

211


Aagaard, K., and E. C. Carmack (1989), The role of sea ice and other fresh water in the<br />

Arctic circulation, J. Geophys. Res., 94(C10), 14485-14498.<br />

Aagaard, K., L. K. Coachman, and E. C. Carmack (1981), On the halocline of the Arctic<br />

Ocean, Deep Sea Res., 28A(6), 529-545.<br />

ACIA (2005), Arctic Climate Impact Assessment, 1042 pp., Cambridge Press University.<br />

Amon, R. M. W. (2004), The role of dissolved organic matter for the organic carbon cycle<br />

in the Arctic Ocean, in The organic carbon cycle in the Arctic Ocean, edited by R. Stein, and<br />

R. W. MacDonald, pp. 83-99, Springer, Berlin.<br />

Amon, R. M. W., and R. Benner (1996), Photochemical and microbial consumption of<br />

dissolved organic carbon and dissolved oxygen in the Amazon River System, Geochim.<br />

Cosmochim. Acta, 60(10), 1783-1792.<br />

Amon, R. M. W., and R. Benner (2003), Combined neutral sugars as indicators of the<br />

diagenetic state of dissolved organic matter in the Arctic Ocean, Deep Sea Res., 50(1),<br />

151-169.<br />

Amon, R. M. W., G. Budéus, and B. Meon (2003), Dissolved organic carbon distribution<br />

and origin in the Nordic Seas: Exchanges with the Arctic Ocean and North Atlantic, J.<br />

Geophys. Res., 108, 3221, doi: 10.1029/2002JC001594.<br />

Amon, R. M. W., and B. Meon (2004), The biogeochemistry of dissolved organic matter<br />

and nutrients in two large Arctic estuaries and potential implications for our<br />

un<strong>de</strong>rstanding of the Arctic Ocean system, Mar. Chem., 92, 311-330.<br />

An<strong>de</strong>rson, L. G. (2002), DOC in the Arctic Ocean, in Biogeochemistry of Marine Dissolved<br />

Organic Matter, edited by D. A. Hansell, and C. A. Carlson, pp. 665-683, Aca<strong>de</strong>mic<br />

Press, San Diego.<br />

An<strong>de</strong>rson, L. G., E. P. Jones, and J. H. Swift (2003), Export production in the central<br />

Arctic Ocean evaluated from phosphate <strong>de</strong>ficits, J. Geophys. Res., 108(C6).<br />

An<strong>de</strong>rson, L. G., K. Olsson, and M. Chierici (1998), A carbon budget for the Arctic Ocean,<br />

Global Biogeochem. Cyc<strong>les</strong>, 12(3), 455-465.<br />

Andrews, S. S., S. Caron, and O. C. Zafiriou (2000), Photochemical oxygen consumption<br />

in marine waters: A major sink for colored dissolved organic matter?, Limnol. Oceanogr.,<br />

45(2), 267-277.<br />

Antoine, D., and A. Morel (1999), A multiple scattering algorithm for atmospheric<br />

correction of remotely sensed ocean colour (MERIS instrument): principle and<br />

implementation for atmospheres carrying various aerosols including absorbing ones,<br />

Int. J. of Remote Sens., 20(9), 1875-1916.<br />

Antoine, D., and A. Morel (1996), Oceanic primary production : I. Adaptation of a spectral<br />

light-photosynthesis mo<strong>de</strong>l in view of application to satellite chlorophyll observations,<br />

Global Biogeochem. Cyc<strong>les</strong>, 10(1), 43-55<br />

Antoine, D., A. Morel, and J. M. André (1996), Oceanic primary production : II.<br />

Estimation at global scale from satellite (Coastal Zone Color Scanner) chlorophyll,<br />

Global Biogeochem. Cyc<strong>les</strong>, 10(1), 57-69.<br />

212


Arnell, N. W. (2005), Implications of climate change for freshwater inflows to the Arctic<br />

Ocean, J. Geophys. Res., 110(D7).<br />

Arrigo, K. R., and G. L. van Dijken (2004), Annual cyc<strong>les</strong> of sea ice and phytoplankton in<br />

Cape Bathurst polynya, southeastern Beaufort Sea, Canadian Arctic, Geophys. Res. Lett.,<br />

31, L088304.<br />

Arrigo, K. R., D. Worthen, A. Schnell, and M. P. Lizotte (1998), Primary production in<br />

Southern Ocean waters, J. Geophys. Res., 103(C8), 15587-15600.<br />

Arrigo, K. R., D. L. Worthen, and D. H. Robinson (2003), A coupled ocean-ecosystem<br />

mo<strong>de</strong>l of the Ross Sea : 2. Iron regulation of phytoplankton taxonomic variability and<br />

primary production, J. Geophys. Res., 108(C7), 3231, doi:10.1029/2001JC000856.<br />

Babin, M., and D. Stramski (2002), Light absorption by aquatic partic<strong>les</strong> in the nearinfrared<br />

spectral region, Limnol. Oceanogr., 47(3), 911-915.<br />

Babin, M., A. Morel, V. Fournier-Sicre, F. Fell, and D. Stramski (2003a), Light scattering<br />

properties of marine partic<strong>les</strong> in coastal and open ocean waters as related to the<br />

particle mass concentration, Limnol. Oceanogr., 48(2), 843-859.<br />

Babin, M., D. Stramski, G. M. Ferrari, H. Claustre, A. Bricaud, G. Obolensky, and N.<br />

Hoepffner (2003b), Variations in the light absorption coefficients of phytoplankton,<br />

nonalgal partic<strong>les</strong>, and dissolved organic matter in coastal waters around Europe, J.<br />

Geophys. Res., 108(C7), 3211, doi: 10.1029/2001JC000882.<br />

Bailey, S. W., and P. J. Wer<strong>de</strong>ll (2006), A multi-sensor approach for the on-orbit validation<br />

of ocean color satellite data products, Remote Sens. Environ., 102, 12-23.<br />

Barber, D. G., and J. M. Hanesiak (2004), Meteorological forcing of sea ice concentrations<br />

in the southern Beaufort Sea over the period 1979 to 2000, J. Geophys. Res., 109(C6),<br />

C06014, doi:10.1029/2003JC002027.<br />

Barnard, A. H., W. S. Pegau, and J. R. V. Zaneveld (1998), Global relationships of the<br />

inherent optical properties of the oceans, J. Geophys. Res., 103(C11), 24955-24968.<br />

Bauch, D., P. Schlosser, and R. G. Fairbanks (1995), Freshwater balance and the sources<br />

of <strong>de</strong>ep and bottom waters in the Arctic Ocean inferred from the distribution of H2 Prog. Oceanogr., 35, 53-80.<br />

Behrenfeld, M. J., E. Boss, D. A. Siegel, and D. S. Shea (2005), Carbon-based ocean<br />

productivity and phytoplankton physiology from space, Global Biogeochem. Cyc<strong>les</strong>, 19,<br />

GB1006, doi:10.1029/2004GB002299.<br />

Behrenfeld, M. J., and P. G. Falkowski (1997), Photosynthetic rates <strong>de</strong>rived from satellitebased<br />

chlorophyll concentration, Limnol. Oceanogr., 42(1), 1-20.<br />

Bélanger, S., H. Xie, N. Krotkov, P. Larouche, W. F. Vincent, and M. Babin (2006),<br />

Photomineralization of terrigenous dissolved organic matter in Arctic coastal waters<br />

from 1979 to 2003: Interannual variability and implications of climate change, Global<br />

Biogeochem. Cyc<strong>les</strong>, 20, GB4005, doi:10.1029/2006GB002708.<br />

Benner, R., B. Benitez-Nelson, K. Kaiser, and R. M. W. Amon (2004), Export of young<br />

terrigenous dissolved organic carbon from rivers to the Arctic Ocean, Geophys. Res.<br />

Lett., 31, L05305.<br />

213<br />

18 O,


Benner, R., P. Louchouarn, and R. M. W. Amon (2005), Terrigenous dissolved organic<br />

matter in the Arctic Ocean and its transport to <strong>sur</strong>face and <strong>de</strong>ep waters of the North<br />

Atlantic, Global Biogeochem. Cyc<strong>les</strong>, 19, GB2025, doi:10.1029/2004GB002398.<br />

Benner, R., and S. Opsahl (2001), Molecular indicators of the sources and transformations<br />

of dissolved organic matter in the Mississippi river plume, Org. Geochem., 32(4), 597-611.<br />

Billett, M. F., S. M. Palmer, D. Hope, C. Deacon, R. Storeton-West, K. J. Hargreaves, C.<br />

Flechard, and D. Fowler (2004), Linking land-atmosphere-stream carbon <strong>flux</strong>es in a<br />

lowland peatland system, Global Biogeochem. Cyc<strong>les</strong>, 18(1).<br />

Blough, N. V., and R. Del Vecchio (2002), Chromophoric DOM in the coastal<br />

environment, in Biogeochemistry of Marine Dissolved Organic Matter, edited by D. A. Hansell,<br />

and C. A. Carlson, pp. 509-546, Aca<strong>de</strong>mic Press, San Diego.<br />

Bricaud, A., M. Babin, A. Morel, and H. Claustre (1995), Variability in the chlorophyllspecific<br />

absorption coefficients of natural phytoplankton: Analysis and<br />

parameterization, J. Geophys. Res., 100(C7), 13321-13332.<br />

Bricaud, A., A. Morel, and L. Prieur (1981), Absorption by dissolved organic matter of the<br />

sea (yellow substance) in the UV and visible domains, Limnol. Oceanogr., 26(1), 43-53.<br />

Budac, D., and P. Wan (1992), Photo<strong>de</strong>carboxylation: mecanism and synthetic utility, J. of<br />

Photochem. Photobiology, 67, 135-166.<br />

Bussmann, I., and G. Kattner (2000), Distribution of dissolved organic carbon in the<br />

central Arctic Ocean: the influence of physical and biological properties, J. Mar. Syst.,<br />

27, 209-219.<br />

Camill, P. (2005), Permafrost thaw accelerates in boreal peatlands during late-20th century<br />

climate warming, Climatic Change, 68(1-2), 135-152.<br />

Car<strong>de</strong>r, Hawes, Baker, Smith, Steward, and Mitchell (1991), Reflectance mo<strong>de</strong>l for<br />

quantifying chlorophyll a in the presence of productivity <strong>de</strong>gradation products, J.<br />

Geophys. Res., 96(C11), 20599.<br />

Car<strong>de</strong>r, K. L., R. G. Steward, G. R. Harvey, and P. B. Ortner (1989), Marine humic and<br />

fulvic acids: Their effects on remote sensing of ocean chlorophyll, Limnol. Oceanogr., 34,<br />

86-81.<br />

Carlson, C. A. (2002), Production and Removal Processes, in Biogeochemistry of Marine<br />

Dissolved Organic Matter, edited by D. A. Hansell, and C. A. Carlson, pp. 91-151,<br />

Aca<strong>de</strong>mic Press, San Diego.<br />

Carmack, E. C., and R. W. MacDonald (2002), Oceanography of the Canadian Shelf of the<br />

Beaufort Sea: A setting for marine life, Arctic, 55(1), 29-45.<br />

Carmack, E. C., R. W. MacDonald, and S. Jasper (2004), Phytoplankton productivity on<br />

the Canadian Shelf of the Beaufort Sea, Mar. Ecol. Prog. Ser., 277, 37-50.<br />

Carmack, E. C., R. W. MacDonald, and J. E. Papadakis (1989), Water mass structure and<br />

boundaries in the Mackenzie Shelf estuary, J. Geophys. Res., 94(C12), 18043-18055.<br />

Cauwet, G., and I. Sidorov (1996), The biogeochemistry of Lena River: organic carbon and<br />

nutrients distribution, Mar. Chem., 53(3-4), 211-227.<br />

214


Cavalieri, D. J., P. Gloersen, and J. Zwally (1990), DMSP SSMI/I daily polar grid<strong>de</strong>d sea<br />

ice concentrations, edited by J. A. Maslanik, and J. Stroeve, pp. Digital media, National<br />

Snow and Ice Data Center, Boul<strong>de</strong>r, CO.<br />

Cavalieri, D. J., C. L. Parkinson, and K. Y. Vinnikov (2003), 30-Year satellite record reveals<br />

contrasting Arctic and Antarctic <strong>de</strong>cadal sea ice variability, Geophys. Res. Lett., 30(18).<br />

Chami, M., E. B. Shybanov, T. Y. Churilova, G. A. Khomenko, M. E. G. Lee, O. V.<br />

Martynov, G. A. Berseneva, and G. K. Korotaev (2005), Optical properties of the<br />

partic<strong>les</strong> in the Crimea coastal waters (Black Sea), J. Geophys. Res.,<br />

doi:10.1029/2005JC003008, 110(C11), C11020.<br />

Clarke, G. L., G. C. Ewing, and C. J. Lorenzen (1970), Spectra of backscattered light from<br />

the sea obtained from aircraft as a mea<strong>sur</strong>e of chlorophyll concentration, Science, 167,<br />

1119-1121.<br />

Comiso, J. C., N. G. Maynard, W. O. Smith Jr., and C. W. Sullivan (1990), Satellite ocean<br />

color studies of Antarctic ice edges in summer and autumn, J. Geophys. Res., 95(C6),<br />

9481-9486.<br />

Comiso, J. C., J. Yang, H. Susumo, and R. A. Krishfield (2003), Detection change in the<br />

Arctic using satellite and in situ data, J. Geophys. Res., 108(C12), 3384.<br />

Cota, G. F., J. Wang, and J. C. Comiso (2004), Transformation of global satellite<br />

chlorophyll retreivals with a regionally tuned algorithm, Remote Sens. Environ., 90, 373-<br />

377.<br />

Cox, C., and W. H. Munk (1954), Statistics of the sea <strong>sur</strong>face <strong>de</strong>rived from sun glitter, J. of<br />

Marine Res., 13, 198-227.<br />

Cullen, J. J., R. F. Davis, J. S. Barlett, and W. L. Miller (1997), Toward remote sensing of<br />

UV attenuation, photochemical <strong>flux</strong>es, and biological effects of UV in <strong>sur</strong>face waters,<br />

in Current and Emerging Issues in Aquatic Science: Aquatic Science meeting, Am. Soc. Limnol.<br />

and Oceanogr., Santa Fe, N. M.<br />

Dagg, M., R. Benner, S. Lohrenz, and D. Lawrence (2004), Transformation of dissolved<br />

and particulate materials on continental shelves influenced by large rivers: plume<br />

processes, Cont. Shelf Res., 24, 833-858.<br />

Davidson, E. A., and I. A. Janssens (2006), Temperature sensitivity of soil carbon<br />

<strong>de</strong>composition and feedbacks to climate change, Nature, 440, 165-173.<br />

Davis, J., and R. Benner (2005), Seasonal trends in the aboundance, composition and<br />

bioavailability of particulate and dissolved organic matter in the Chukchi/Beaufort<br />

Seas and western Canada Basin, Deep Sea Res. II, 52, 3396-3410.<br />

Del Vecchio, R., and N. V. Blough (2002), Photobleaching of chromophoric dissolved<br />

organic matter in natural waters: kinetics and mo<strong>de</strong>lling, Mar. Chem., 78, 231-253.<br />

Dittmar, T. (2004), Evi<strong>de</strong>nce for terrigenous dissolved organic nitrogen in the Arctic <strong>de</strong>ep<br />

sea, Limnol. Oceanogr., 49(1), 148-156.<br />

Dittmar, T., and G. Kattner (2003), The biogeochemistry of the river and shelf ecosystem<br />

of the Arctic Ocean: a review, Mar. Chem., 83, 105-120.<br />

215


Dixon, R. K., S. Brown, R. A. Houghton, A. M. Solomon, M. C. Trexler, and J. Wisniewski<br />

(1994), Carbon pools and <strong>flux</strong> of global forest ecosystems., Science, 263(5144), 185-190.<br />

Droppo, I. G., D. Jeffries, C. Jaskot, and S. Backus (1998), The prevalence of freshwater<br />

flocculation in cold regions: A case study from the Mackenzie River Delta, Northwest<br />

Territories, Canada, Arctic, 51(2), 155-164.<br />

D'Sa, E., and R. L. Miller (2003), Bio-optical properties in waters influenced by the<br />

Mississippi River during low flow conditions, Remote Sens. Environ., 84, 538-549.<br />

Dugdale, R. C., and J. J. Goering (1967), Uptake of new and regenerated forms of nitrogen<br />

in primary production, Limnol. Oceanogr., 12, 196-206.<br />

Ebuchi, N., and S. Kizu (2002), Probability distribution of <strong>sur</strong>face wave slope <strong>de</strong>rived<br />

using sun glitter images from geostastionay meteorological satellite and <strong>sur</strong>face vector<br />

winds from scatterometers, J. Oceanogr., 58, 477-486.<br />

Ferrari, G. M. (2000), The relationship between chromophoric dissolved organic matter<br />

and dissolved organic carbon in the European Atlantic coastal area and in the West<br />

Mediterranean Sea (Gulf of Lions), Mar. Chem., 70, 339-357.<br />

Fichot, C. G. (2004), Marine photochemistry from space: Algorithms for the retreival of<br />

diffuse attenuation and CDOM absorption coefficients (320-490 nm) from Ocean<br />

Color and estimation of <strong>de</strong>pth-resolved photoproduction rates of carbon monoxi<strong>de</strong><br />

(CO) at global sca<strong>les</strong> using SeaWiFS imagery, Master of Science thesis,217 pp.,<br />

Dalhousie, Halifax, Canada.<br />

Fioletov, V. E., M. G. Kimlin, N. Krotkov, L. J. B. McArthur, J. B. Kerr, D. I. Wardle, J. R.<br />

Herman, R. Meltzer, T. W. Mathews, and J. Kaurola (2004), UV in<strong>de</strong>x climatology<br />

over the United States and Canada from ground-based and satellite estimates,<br />

doi:10.1029/2004JD004820, J. Geophys. Res., 109(22).<br />

Franz, B. A., R. E. Eplee, S. W. Bailey, and M. Wang (2003), Changes to the atmospheric<br />

correction algorithm and retrieval of oceanic optical properties, Chap. 5, in NASA<br />

Tech. Memo. 2003-206892, vol. 22, edited by S. B. Hooker, and E. R. Firestone, pp. 29-<br />

33, NASA Goddard Space Flight Center, Greenbelt, Md.<br />

Freeman, C., C. D. Evans, D. T. Monteith, B. Reynolds, and N. Fenner (2001), Export of<br />

organic carbon from peat soils, Nature, 412(6849), 785-785.<br />

Freeman, C., N. Fenner, N. J. Ostle, H. Kang, D. J. Dowrick, B. Reynolds, M. A. Lock, D.<br />

Sleep, S. Hughes, and J. Hudson (2004), Export of dissolved organic carbon from<br />

peatlands un<strong>de</strong>r elevated carbon dioxi<strong>de</strong> levels, Nature, 430(6996), 195-198.<br />

Frey, K. E., and L. C. Smith (2005), Amplified carbon release from vast West Siberian<br />

peatlands by 2100, Geophys. Res. Lett., 32(9).<br />

Frouin, R., M. Schwindling, and P. Y. Deschamps (1996), Spectral reflectance of sea foam<br />

in the visible and near-infrared: In situ mea<strong>sur</strong>ements and remote sensing implications,<br />

J. Geophys. Res., 101(C6), 14361-14371.<br />

Fry, E. S. (2000), Visible and near-ultraviolet absorption spectrum of liquid water:<br />

comment, Appl. Optics., 39(16), 2743-2744.<br />

216


Gallegos, C. L., and P. J. Neale (2002), Partitioning spectral absorption in case 2 waters:<br />

discrimination of dissolved and particulates components, App. Opt., 41(21), 4220-4233.<br />

Gao, H., and R. G. Zepp (1998), Factors influencing photoreactions of dissolved organic<br />

matter in a coastal river of southeastern United States, Environ. Sci. Techn., 32(19),<br />

2940-2946.<br />

Garneau, M.-È., W. F. Vincent, L. Alonso-Saez, Y. Gratton, and C. Lovejoy (2006),<br />

Prokaryotic community structure and hetereotrophic production in a river-influence<br />

arctic ecosystem, Aquat. Microb. Ecol., 42, 27-40.<br />

Garver, D. L., and D. A. Siegel (1997), Inherent optical property inversion of ocean color<br />

spectra and its Biogeochem. interpretation, 1. Time series from the Sargasso Sea, J.<br />

Geophys. Res., 102(C8), 18607-18625.<br />

Gibson, J. A. E., W. F. Vincent, B. Nieke, and R. Pienitz (2000), Control of biological<br />

expo<strong>sur</strong>e to UV radiation in the Arctic Ocean: Comparison of the ro<strong>les</strong> of ozone and<br />

riverine dissolved organic matter, Arctic, 53(5), 372-382.<br />

Gobeil, C., B. Sundby, R. W. Macdonald, and J. N. Smith (2001), Recent change in organic<br />

carbon <strong>flux</strong> to Arctic Ocean <strong>de</strong>ep basins: Evi<strong>de</strong>nce from acid volatile silfi<strong>de</strong>,<br />

manganese and rhenium discord in sediments, Geophys. Res. Lett., 28(9), 1743-1746.<br />

Goñi, M. A., M. B. Yunker, R. W. Macdonald, and T. I. Eglinton (2005), The supply and<br />

preservation of ancient and mo<strong>de</strong>m components of organic carbon in the Canadian<br />

Beaufort Shelf of the Arctic Ocean, Mar. Chem., 93(1), 53-73.<br />

Gordon, H. R. (1992), Diffuse reflectance of the ocean: influence of nonuniform<br />

phytoplankton pigment profile, App. Opt., 31(12), 2116-2129.<br />

Gordon, H. R. (1997), Atmospheric correction of ocean color imagery in the Earth<br />

Observation System era, J. Geophys. Res., 102(D14), 17081-17106.<br />

Gordon, H. R., J. W. Brown, and R. H. Evans (1988b), Exact Rayleigh scattering<br />

calculations for use with the Nimbus-7 Coastal Zone Color Scanner, Appl. Optics., 27,<br />

862-871.<br />

Gordon, H. R., O. B. Brown, R. H. Evans, J. W. Brown, R. C. Smith, K. S. Baker, and D.<br />

K. Clark (1988a), A semianalytic radiance mo<strong>de</strong>l of ocean color, J. Geophys. Res., 93(D9),<br />

10909-10924.<br />

Gordon, H. R., O. B. Brown, and M. M. Jacobs (1975), Computed relationships between<br />

the inherent and apparent optical properties of a flat homogenous ocean, Appl. Optics.,<br />

14(2), 417-427.<br />

Gordon, H. R., and D. K. Clark (1981), Clear water radiances for atmospheric correction<br />

of coastal zone color scanner imagery, App. Opt., 20(24), 4175-4180.<br />

Gordon, H. R., D. K. Clark, J. L. Mueller, and W. A. Hovis (1980), Phytoplankton<br />

pigments from the Nimbus-7 Coastal Zone Color Scanner: comparisons with <strong>sur</strong>face<br />

mea<strong>sur</strong>ements, Science, 210, 63-66.<br />

Gordon, H. R., and K. Ding (1992), Self-shading of in-water optical instruments, Limnol.<br />

Oceanogr., 37(3), 491-500.<br />

217


Gordon, H. R., and M. Wang (1992a), Surface-roughness consi<strong>de</strong>rations for atmospheric<br />

correction of ocean color sensors. I : the Rayleigh- scattering component, Appl. Optics.,<br />

31, 4247-4260.<br />

Gordon, H. R., and M. Wang (1992b), Surface-roughness consi<strong>de</strong>rations for atmospheric<br />

correction of ocean color sensors. II : error in the retrieved water-leaving radiance,<br />

Appl. Optics., 31, 4261-4267.<br />

Gordon, H. R., and M. Wang (1994), Retrieval of water-leaving radiance and aerosol<br />

optical thickness over the oceans with SeaWiFS: a preliminary algorithm, App. Opt.,<br />

33(3), 443-452.<br />

Gorham, E. (1991), Northern peatlands: role in the carbon cycle and probable responses<br />

to climatic warming, Ecological Applications, 1(2), 182-195.<br />

Gosselin, M., M. Levasseur, P. A. Wheeler, R. A. Horner, and B. C. Booth (1997), New<br />

mea<strong>sur</strong>ements of phytoplankton and ice algal production in the Arctic Ocean, Deep-Sea<br />

Res. II, 44(8), 1623-1644.<br />

Gregg, W. W., and N. W. Casey (2004), Global and regional evaluation of the SeaWiFS<br />

chlorophyll data set, Remote Sens. Environ., 93, 463-479.<br />

Grenfell, T. C., and G. A. Maykut (1977), The optical properties of ice and snow in the<br />

Arctic Basin, J. of Glaciol., 18, 445-463.<br />

Guay, C. K., G. P. Klinkhammer, K. K. Falkner, R. Benner, P. G. Coble, T. E. Whitledge,<br />

B. Black, F. J. Bussell, and T. A. Wagner (1999), High-resolution mea<strong>sur</strong>ements of<br />

dissolved organic carbon in the Arctic Ocean by in situ fiber-optic spectrometry,<br />

Geophys. Res. Lett., 26(8), 1007-1010.<br />

Guéguen, C., L. Guo, and N. Tanaka (2005), Distributions and characteristics of colored<br />

dissolved organic matter in the Western Arctic Ocean, Cont. Shelf Res., 25(10), 1195-<br />

1207.<br />

Guo, L., and R. W. Macdonald (2006), Source and transport of terrigenous organic matter<br />

in the upper Yukon River: Evi<strong>de</strong>nce from isotope (δ 13 C, Δ 14 C, and δ �5 N) composition<br />

of dissolved colloïdal, and particulate phases, Global Biogeochem. Cyc<strong>les</strong>, 20, GB2011,<br />

doi:10.1029/2005GB002593.<br />

Hansell, D. A., and C. A. Carlson (1998), Deep-ocean gradients in the concentration of<br />

dissolved organic carbon, Nature, 395(6699), 263-266.<br />

Hansell, D. A., D. Kadko, and N. R. Bates (2004), Degradation of terrigenous dissolved<br />

organic carbon in the western Arctic Ocean, Science, 304, 858-861.<br />

Hedges, J. I., R. G. Keil, and R. Benner (1997), What happens to terrestrial organic matter<br />

in the ocean?, Org. Geochem., 27(5/6), 196-212.<br />

Herman, J. R., and E. A. Celarier (1997), Earth <strong>sur</strong>face reflectivity climatology at 340-380<br />

nm from TOMS data, J. Geophys. Res., 102(D23), 28003-28011.<br />

Herman, J. R., N. Krotkov, E. Celarier, D. Larko, and G. Labow (1999), Distribution of<br />

UV radiation at the Earth's <strong>sur</strong>face from TOMS-mea<strong>sur</strong>ed UV-backscattered radiances,<br />

J. Geophys. Res., 104(D10), 12059-12076.<br />

218


Hernes, P. J., and R. Benner (2002), Transport and diagenesis of dissolved and particulate<br />

terrigenous organic matter in the North Pacific Ocean, Deep-Sea Res., 49(12), 2119-2132.<br />

Hernes, P. J., and R. Benner (2003), Photochemical and microbial <strong>de</strong>gradation of dissolved<br />

lignin phenols: Implications for the fate of terrigenous dissolved organic matter in<br />

marine environments, J. Geophys. Res., 108(C9), 3291, doi: 1029/2002JC001421, 2003.<br />

Hernes, P. J., and R. Benner (2006), Terrigenous organic matter sources and reactivity in<br />

the North Atlantic Ocean and a comparison to the Arctic and Pacific oceans, Mar.<br />

Chem., 100(1-2), 66-79.<br />

Hinzman, L., N. D. Bettez, W. R. Bolton, F. S. Chapin, III, M. Dyurgerov, C. L. Fastie, B.<br />

Griffith, R. D. Hollister, A. Hope, H. P. Huntington, A. M. Jensen, G. J. Jia, T.<br />

Jorgenson, D. L. Kane, D. R. Klein, G. Kofinas, A. H. Lynch, A. H. Lloyd, A. D.<br />

McGuire, F. E. Nelson, W. C. Oechel, T. Osterkamp, C. H. Racine, V. E.<br />

Romankevich, R. S. Stone, D. A. Stow, M. Sturm, G. E. Tweedie, G. L. Vourlitis, M. D.<br />

Walker, D. A. Walker, P. J. Webber, J. M. Welker, K. S. Winker, and K. Yoshikawa<br />

(2005), Evi<strong>de</strong>nce and implications of recent climate change in northern Alaska and<br />

other Arctic regions, Climatic Change, 72, 251-298.<br />

Hoge, F. E., and P. E. Lyon (1996), Satellite retreival of inherent optical properties by<br />

linear matrix inversion of oceananic radiance mo<strong>de</strong>ls: An analysis of mo<strong>de</strong>l and<br />

radiance mea<strong>sur</strong>ement errors, J. Geophys. Res., 101, 16631-16648.<br />

Hoge, F. E., M. E. Williams, R. N. Swift, J. K. Yungel, and A. Vodacek (1995), Satellite<br />

retrieval of the absorption coefficient of chromophoric dissolved organic matter in<br />

continental margins, J. Geophys. Res., 100(C12), 24847-24854.<br />

Hooker, S. B., and C. R. McClain (2000), The calibration and validation of SeaWiFS data,<br />

Prog. Oceanogr., 45, 427-465.<br />

Hovis, W. A., D. K. Clark, F. An<strong>de</strong>rson, R. W. Austin, W. H. Wilson, E. T. Baker, D. Ball,<br />

H. R. Gordon, J. L. Mueller, S. Z. El-Sayed, B. Sturm, R. C. Wrigley, and C. S. Yentsch<br />

(1980), Nimbus-7 Coastal Zone Color Scanner: System <strong><strong>de</strong>s</strong>cription and initial imagery,<br />

Science, 210, 60-63.<br />

Hu, C., F. E. Muller-Karger, and R. G. Zepp (2002), Absorbance, absorption coefficient,<br />

and apparent quantum yield: A comment on common ambiguity in the use of these<br />

optical concepts, Limnol. Oceanogr., 47(4), 1261-1267.<br />

Huot, Y., C. A. Brown, and J. J. Cullen (2005), New algorithms for MODIS sun-induced<br />

chlorophyll fluorescence and a comparison with present data products, Limnol. Oceanogr.<br />

Meth., 3, 108-130.<br />

Ingvaldsen, R. B., L. Asplin, and H. Loeng (2004), The seasonal cycle in the Atlantic<br />

transport to the Barents Sea during the years 1997-2001, Cont. Shelf Res., 24(9), 1015-<br />

1032.<br />

IOCCG (2006), Remote sensing of Inherent Optical Properties: Fundamentals, Tests of algorithms, and<br />

Applications, edited by Z.-P. Lee, pp. 126, Reports of the International Ocean-Colour<br />

Coordinating Group, No. 5, IOCCG, Dartmouth, Canada.<br />

219


IPCC (2001), Climate Change 2001: The scientific basis. Contribution of Working Group I to the<br />

Third Assessment Report of the Intergovernmental Panel on Climate Change, 881 pp., Cambridge<br />

Press University.<br />

Jakobsson, M., A. Grantz, Y. Kristoffersen, and R. Macnab (2004), Bathymetry and<br />

physiography of the Arctic Ocean ans its constituent seas, in The organic carbon cycle in the<br />

Arctic Ocean, edited by R. Stein, and R. W. Macdonald, pp. 1-6, Springer, Berlin.<br />

Johannessen, S. C. (2000), A photochemical sink for dissolved organic carbon in the ocean,<br />

176 pp., Dalhousie University, Halifax.<br />

Johannessen, S. C., and W. L. Miller (2001), Quantum yield for the photochemical<br />

production of dissolved inorganic carbon in seawater, Mar. Chem., 76, 271-283.<br />

Johannessen, S. C., W. L. Miller, and J. J. Cullen (2003), Calculation of UV attenuation and<br />

colored dissolved organic matter absorption spectra from mea<strong>sur</strong>ements of ocean<br />

color, J. Geophys. Res., 108(C9), 3301, doi:10.1029/2000JC000514.<br />

Kattner, G., J. M. Lobbes, H. P. Fitznar, R. Engbrodt, E.-M. Nöthig, and R. J. Lara (1999),<br />

Tracing dissolved organic substances and nutrients from the Lena River through<br />

Laptev Sea (Arctic), Mar. Chem., 65, 25-39.<br />

Keeling, C. D., Atmospheric CO 2 concentration - Mauna Loa Observatory, Hawaii 1956-<br />

1986, NDP-001/R1 Carbon Dioxi<strong>de</strong> Information Center, Oak Ridge National<br />

Laboratory, Oak Ridge Tennessee, 1986.<br />

Kharu, M., and G. B. Mitchell (2001), Seasonal and nonseasonal variability of satellite<strong>de</strong>rived<br />

chlorophyll and colored dissolved organic matter concentration in the<br />

California Current, J. Geophys. Res., 106(C2), 2517-2529.<br />

Kieber, R. J., X. Zhou, and K. Mopper (1990), Formation of carbonyl compounds from<br />

UV-induced photo<strong>de</strong>gradation of humic substances in natural waters: Fate of riverine<br />

carbon in the sea, Limnol. Oceanogr., 35(7), 1503-1515.<br />

Kieber, D. J., J. McDaniel, and K. Mopper (1989), Photochemical source of biological<br />

substrates in sea water: implications for carbon cycling, Nature, 341, 637-639.<br />

Kinney, P. J., T. C. Lo<strong>de</strong>r, and J. Groves (1971), Particulate and dissolved organic matter<br />

in the Amerasian Basin of the Arctic Ocean, Limnol. Oceanogr., 16(1), 132-137.<br />

Kishino, M., M. Takahashi, N. Okami, and S. Ichimura (1985), Estimation of the spectral<br />

absorption coefficients of phytoplankton in the sea, Bull. Mar. Sci., 37(2), 634-642.<br />

Klein, B., B. LeBlanc, Z. P. Mei, R. Beret, J. Michaud, C. J. Mundy, C. H. von Quillfeldt, M.<br />

E. Garneau, S. Roy, Y. Gratton, J. K. Cochran, S. Bélanger, P. Larouche, J. D. Pakulski,<br />

R. B. Rivkin, and L. Legendre (2002), Phytoplankton biomass, production and<br />

potential export in the North Water, Deep-Sea Res., 49(22-23), 4983-5002.<br />

Köhler, H., B. Meon, V. V. Gor<strong>de</strong>ev, A. Spitzy, and R. M. W. Amon (2003), Dissolved<br />

organic matter (DOM) in the estuaries of Ob and Yenisei and the adjacent Kara Sea,<br />

Russia, in Siberian river runoff in the Kara Sea: Characterization, Quantification, Variability and<br />

Environmental Significance, Proceedings in Marine Science, edited by R. Stein, K. Fahl, D. K.<br />

Fütterer, E. M. Galimov, and O. V. Stepanets, pp. 281-308, Elsevier, Amsterdam.<br />

220


Kowalczuk, P., J. Olszewski, M. Darecki, and S. Kaczmarek (2005), Empirical relationships<br />

between coloured dissolved organic matter (CDOM) absorption and apparent optical<br />

properties in Baltic Sea waters, Int. J. Remote Sens., 26(2), 345-370.<br />

Krotkov, N., J. Herman, P. K. Bhartia, C. Seftor, A. Arola, S. Kalliskota, P. Taalas, and I.<br />

V. Geogdzhaev (2002), Version 2 total ozone mapping spectrometer ultraviolet<br />

algorithm: problems and enhancements, Opt. Engin., 41(12), 3028-3039.<br />

Krotkov, N. A., P. K. Bhartia, J. R. Herman, V. Fioletov, and J. Kerr (1998), Satellite<br />

estimation of spectral <strong>sur</strong>face UV irradiance in the presence of tropospheric aerosols -<br />

1. Cloud-free case, J. Geophys. Res., 103(D8), 8779-8793.<br />

Krotkov, N. A., J. R. Herman, P. K. Bhartia, V. Fioletov, and Z. Ahmad (2001), Satellite<br />

estimation of spectral <strong>sur</strong>face UV irradiance 2. Effects of homogeneous clouds and<br />

snow, J. Geophys. Res., 106(D11), 11743-11759.<br />

Laurion, I., F. Blouin, and S. Roy (2003), The quantitative filter technique for mea<strong>sur</strong>ing<br />

phytoplankton absorption: Interference by MAAs in the UV waveband, Limnol.<br />

Oceanogr.: Methods, 1(1), 1-9.<br />

Lee, S. H., and T. R. Whitledge (2005), Primary and new production in the <strong>de</strong>ep Canada<br />

Basin during summer 2002, Polar Biology, 28(3), 190-197.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and R. A. Arnone (2002), Deriving inherent optical properties<br />

from water color : a multiband quasi-analytical algorithm for optically <strong>de</strong>ep waters, App.<br />

Opt., 41(27), 5755-5772.<br />

Lee, Z. P., K. L. Car<strong>de</strong>r, R. F. Chen, and T. G. Peacock (2001), Properties of water column<br />

and bottom <strong>de</strong>rived from Airborne Visible Infrared Imaging Spectrometer (AVIRIS)<br />

data, J. Geophys. Res., 106(C6), 11639-11651.<br />

Lee, Z. P., K. L. Car<strong>de</strong>r, C. D. Mobley, R. G. Steward, and J. S. Patch (1999),<br />

Hyperspectral remote sensing for shallow waters: 2. Deriving bottom <strong>de</strong>pths and water<br />

properties by optimization, App. Opt., 38(18), 3831-3843.<br />

Lee, Z. P., K. L. Car<strong>de</strong>r, T. G. Peacock, C. O. Davis, and J. L. Mueller (1996), Method to<br />

<strong>de</strong>rive ocean absorption coefficients from remote-sensing reflectance, Appl. Optics.,<br />

35(3), 453-462.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and K. Du (2004), Effects of molecular and particle scatterings on<br />

the mo<strong>de</strong>l parameter for remote-sensing reflectance, App. Opt., 43(25), 4957-4964.<br />

Loisel, H., and A. Morel (1998), Light scattering and chlorophyll concentration in case 1<br />

waters: A reexamination, Limnol. Oceanogr., 43(5), 847-858.<br />

Loisel, H., and A. Morel (2001), Non-isotropy of the upward radiance field in typical<br />

coastal (Case 2) waters, Int. J. Remote Sens., 22(2&3), 275-295.<br />

Macdonald, R. W. (2000), Arctic estuaries and ice: a positive-negative estuarine couple, in<br />

The Freshwater Budget of the Arctic Ocean, edited by E. L. Lewis, pp. 383-407, Kluwer<br />

Aca<strong>de</strong>mic Publishers, Dordrecht, Netherlands.<br />

Macdonald, R. W., E. C. Carmack, F. A. McLaughlin, K. K. Falkner, and J. H. Swift (1999),<br />

Connections among ice, runoff and atmospheric forcing in the Beaufort Gyre, Geophys.<br />

Res. Lett., 26(15), 2223-2226.<br />

221


Macdonald, R. W., E. C. Carmack, F. A. McLaughlin, K. Iseki, D. M. MacDonald, and M.<br />

C. O'Brien (1989), Composition and modification of water mmasses in the Mackenzie<br />

Shelf Estuary, J. Geophys. Res., 94(C12), 18057-18070.<br />

Macdonald, R. W., F. A. McLaughlin, and E. C. Carmack (2002), Fresh water and its<br />

sources during the SHEBA drift in the Canada Basin of the Arctic Ocean, Deep-Sea Res.<br />

I, 49, 1769-1785.<br />

Macdonald, R. W., A. S. Naidu, M. B. Yunker, and C. Gobeil (2004b), The Beaufort Sea:<br />

distribution, sources, <strong>flux</strong>es and burial of organic carbon, in The organic carbon cycle in the<br />

Arctic Ocean, edited by R. Stein, and R. W. Macdonald, pp. 167-193, Springer, Berlin.<br />

Macdonald, R. W., D. W. Paton, E. C. Carmack, and A. Omstedt (1995), The freshwater<br />

budget and un<strong>de</strong>r-ice spreading of Mackenzie River Water in the Canadian Beaufort<br />

Sea based on salinity and 18 O/ 16 O mea<strong>sur</strong>ements in water and ice., J. Geophys. Res.,<br />

100(C1), 895-919.<br />

Macdonald, R. W., E. Sakshaug, and R. Stein (2004a), The Arctic Ocean: mo<strong>de</strong>rn status<br />

and recent climate change, in The organic carbon cycle in the Arctic Ocean, edited by R. Stein,<br />

and R. W. MacDonald, pp. 6-21, Springer, Berlin.<br />

Macdonald, R. W., S. M. Solomon, R. E. Cranston, H. E. Welsh, M. B. Yunker, and C.<br />

Gobeil (1998), A sediment and organic carbon budget for the Canadian Beaufort Shelf,<br />

Mar. Geol., 144, 255-273.<br />

Macdonald, R. W., C. S. Wong, and P. E. Erickson (1987), The distribution of nutrients in<br />

the Southeastern Beaufort Sea : Implications for the water circulation and primary<br />

production, J. Geophys. Res., 92(C3), 2939-2952.<br />

Mack, M. C., E. A. G. Schuur, M. S. Bret-Harte, G. R. Shaver, and F. S. Chapin, III (2004),<br />

Ecosystem carbon storage in arctic tundra reduced by long-term nutrient fertilization,<br />

Nature, 431(7007), 440-443.<br />

Madronich, S., and S. Flocke (1999), The role of solar radiation in atmospheric chemistry,<br />

in Handbook of Environmental Chemistry, edited by P. Boule, pp. 1-26, Springer-Verlag,<br />

Hei<strong>de</strong>lberg, Germany.<br />

Maritorena, S., D. A. Siegel, and A. R. Peterson (2002), Optimization of a semianalitical<br />

ocean color mo<strong>de</strong>l for global-scale applications, App. Opt., 41(15), 2705-2714.<br />

Mc Clatchey, R. A., R. W. Fenn, J. E. A. Selby, F. E. Volz, and J. S. Garing, Optical<br />

properties of the atmosphere, AFCRL-TR-71-079, L.G. Hancom Fiel, Bedford, Ms.,<br />

1971.<br />

McClelland, J. W., S. J. Dery, B. J. Peterson, R. M. Holmes, and E. F. Wood (2006), A panarctic<br />

evaluation of changes in river discharge during the latter half of the 20th century,<br />

Geophys. Res. Lett., 33(6), L06715.<br />

McKee, B. A., R. C. Aller, M. A. Allison, T. S. Bianchi, and G. C. Kineke (2004), Transport<br />

and transformation of dissolved and particulate materials on continental margins<br />

influenced by major rivers: benthic boundary layer and seabed processes, Cont. Shelf<br />

Res., 24(7-8), 899-926.<br />

222


Meon, B., and R. M. W. Amon (2004), Heterotrophic bacterial activity and <strong>flux</strong>es of<br />

dissolved free amino acids and glucose in the Arctic rivers Ob, Yenisei and the<br />

adjacent Kara Sea, Aquat. Microb. Ecol., 37(2), 121-135.<br />

Meyers-Schulte, K. J., and J. I. Hedges (1986), Molecular evi<strong>de</strong>nce for a terrestrial<br />

component of organic matter dissolved in ocean water, Nature, 321, 61-63.<br />

Mi<strong>les</strong>, C. J., and P. L. Brezonik (1981), Oxygen consumption in humic-colored waters by<br />

photochemical ferrous-ferric catalytic cycle, Environ. Sci. Technol., 15(9), 1089-1095.<br />

Miller, W. L., and M. A. Moran (1997), Interaction of photochemical and microbial<br />

processes in the <strong>de</strong>gradation of refractory dissolved organic matter from coastal marine<br />

environment, Limnol. Oceanogr., 42(6), 1317-1324.<br />

Miller, W. L., and R. G. Zepp (1995), Photochemical production of dissolved inorganic<br />

carbon from terrestrial organic matter: Significance to the oceanic carbon cycle, Geophys.<br />

Res. Lett., 22(4), 417-420.<br />

Millot, R., J. Gaillar<strong>de</strong>t, B. Dupre, and C. J. Allegre (2003), Northern latitu<strong>de</strong> chemical<br />

weathering rates: Clues from the Mackenzie River Basin, Canada, Geochim. Cosmochim.<br />

Acta, 67(7), 1305-1329.<br />

Mitchell, B. G. (1992), Predictive bio-optical relationships for polar oceans and marginal<br />

ice zones, J. Mar. Syst., 3, 91-105.<br />

Mobley, C. D. (1994), Light and water: Radiative Transfer in Natural Waters, 1st ed., 592 pp.,<br />

Aca<strong>de</strong>mic Press, San Diego, California.<br />

Mobley, C. D. (1999), Estimation of the remote-sensing réflectances from above-water<br />

mea<strong>sur</strong>ements, App. Opt., 38(36), 7442-7455.<br />

Mopper, K., and D. J. Kieber (2002), Photochemistry and the cycling of carbon, sulfur,<br />

nitrogen and phosphorus, in Biogeochemistry of Marine Dissolved Organic Matter, edited by<br />

D. A. Hansell, and C. A. Carlson, pp. 455-507, Aca<strong>de</strong>mic Press, San Diego.<br />

Mopper, K., X. Zhou, R. J. Kieber, D. J. Kieber, R. J. Sikorski, and R. D. Jones (1991),<br />

Photochemical <strong>de</strong>gradation of dissolved organic carbon and its impact on the oceanic<br />

carbon cycle, Nature, 353, 60-62.<br />

Morel, A. (1974), Optical properties of pure water and pure sea water, in Optical aspects of<br />

Oceanography, edited by N. G. Jerlov, and E. S. Nielsen, pp. 1-24, Aca<strong>de</strong>mic, New York.<br />

Morel, A. (1988), Optical mo<strong>de</strong>ling of the upper ocean in relation to its biogenous matter<br />

content (Case 1 Waters), J. Geophys. Res., 93(C9), 10749-10768.<br />

Morel, A., D. Antoine, and B. Gentili (2002), Bidirectional reflectance of oceanic waters:<br />

accounting for Raman emission and varying particle scattering phase function, App.<br />

Opt., 41(30), 6289-6306.<br />

Morel, A., and S. Bélanger (2006), Improved <strong>de</strong>tection of turbid waters from ocean color<br />

sensors information, Remote Sens. Environ., 102, 237-249.<br />

Morel, A., and B. Gentili (1991), Diffuse reflectance of oceanic waters: its <strong>de</strong>pendance on<br />

Sun angle as influenced the mollecular scattering contribution, App. Opt., 30(30), 4427-<br />

4438.<br />

223


Morel, A., and B. Gentili (1993), Diffuse reflectance of oceanic waters. II. Bidirectional<br />

aspects, App. Opt., 32(33), 6864-6879.<br />

Morel, A., and B. Gentili (1996), Diffuse reflectance of oceanic waters. III. Implication of<br />

bidirectionality for the remote-sensing problem, App. Opt., 35(24), 4850-4862.<br />

Morel, A., B. Gentili, H. Claustre, M. Babin, A. Bricaud, J. Ras, and F. Tièche (in press,<br />

2006), Optical properties of the "clearest" natural waters, Limnol. Oceanogr.<br />

Morel, A., and S. Maritorena (2001), Bio-optical properties of oceanic waters: A reappraisal,<br />

J. Geophys. Res., 106(C4), 7163-7180.<br />

Morel, A., and J. L. Mueller (2002), Normalized Water-Leaving Radiance and Remote<br />

Sensing Reflectance: Bidirectional Reflectance and Other Factors, in Ocean Optics<br />

Protocols For Satellite Ocean Color Sensor Validation, Revision 4, Volume III, edited by J. L.<br />

Mueller, G. S. Fargion, and C. R. McClain, pp. 32-59, NASA Goddard Space Flight<br />

Space Center, Greebelt.<br />

Morel, A., and L. Prieur (1977), Analysis of variations in ocean color, Limnol. Oceanogr.,<br />

22(4), 709-722.<br />

Morrison, J. R., and N. B. Nelson (2004), Seasonal cycle of phytoplankton UV absorption<br />

at Bermuda Atlantic Time-series Study (BATS) site, Limnol. Oceanogr., 49(1), 215-224.<br />

Mueller, J. L., A. Morel, R. Frouin, C. Davis, R. A. Arnone, K. L. Car<strong>de</strong>r, Z.-P. Lee, R. G.<br />

Stewart, S. B. Hooker, B. Holben, C. D. Mobley, S. McLean, M. Miller, C. Pietras, G. S.<br />

Fargion, K. D. Knobelspiesse, and K. Voss (2003), Radiometric mea<strong>sur</strong>ements and<br />

data analysis protocols, in Ocean Optics Protocols For Satellite Ocean Color Sensor Validation,<br />

Revision 4, Volume III, edited by J. L. Mueller, G. S. Fargion, and C. R. McClain, pp. 1-<br />

78, NASA Goddard Space Flight Center, Greenbelt.<br />

O'Brien, M. C., R. W. MacDonald, H. Melling, and K. Iseki (2006), Partic<strong>les</strong> <strong>flux</strong>es and<br />

geochemistry on the Canadian Beaufort Shelf: Implications for sediment transport and<br />

<strong>de</strong>position, Cont. Shelf Res., 26, 41-81.<br />

Oechel, W. C., G. L. Vourlitis, S. J. Hastings, R. C. Zulueta, L. Hinzman, and D. Kane<br />

(2000), Acclimation of ecosystem CO 2 exchange in the Alaskan Arctic in response to<br />

<strong>de</strong>cadal climate warming, Nature, 406(6799), 978-981.<br />

Oh<strong>de</strong>, T., B. Sturm, and H. Siegel (2002), Derivation of SeaWiFS vicarious calibration<br />

coefficients using in situ mea<strong>sur</strong>ements in Case 2 water of the Baltic Sea, Remote Sens.<br />

Environ., 80, 248-255.<br />

Opsahl, S., and R. Benner (1997), Distribution and cycling of terrigenous dissolved organic<br />

matter in the ocean, Nature, 386, 480-482.<br />

Opsahl, S., and R. Benner (1998), Photochemical reactivity of dissolved lignin and ocean<br />

waters, Limnol. Oceanogr., 43(6), 1297-1304.<br />

Opsahl, S., R. Benner, and R. M. W. Amon (1999), Major <strong>flux</strong> of terrigenous dissolved<br />

organic matter throught the Arctic Ocean, Limnol. Oceanogr., 44(8), 2017-2023.<br />

O'Reilly, J. E., S. Maritorena, G. Mitchell, D. A. Siegel, K. L. Car<strong>de</strong>r, D. L. Garver, M.<br />

Kahru, and C. R. McClain (1998), Ocean color chlorophyll algorithms for SeaWiFS, J.<br />

Geophys. Res., 103(C11), 24937-24950.<br />

224


O'Reilly, J. E., S. Maritorena, D. Siegel, M. C. O’Brien, D. Toole, B. G. Mitchell, M. Kahru,<br />

F. P. Chavez, P. Strutton, G. Cota, S. B. Hooker, C. R. McClain, K. L. Car<strong>de</strong>r, F.<br />

Muller-Karger, L. Harding, A. Magnuson, D. Phinney, G. F. Moore, J. Aiken, K. R.<br />

Arrigo, R. Letelier, and M. Culver (2000), Ocean color chlorophyll a algorithms for<br />

SeaWiFS, OC2, and OC4: version 4., in SeaWiFS Postlaunch Technical Report Series, vol.11.<br />

SeaWiFS postlaunch calibration and validation analyses: part 3, edited by S. B. Hooker, and E.<br />

R. Firestone, pp. 9-23, NASA Goddard Space Flight Center., Greenbelt, MD:.<br />

Overland, J. E., M. C. Spillane, and N. N. Sorei<strong>de</strong> (2004), Integrated analysis of physical<br />

and biological Pan-Arctic change, Climatic Change, 63(3), 291-322.<br />

Park, Y.-J., and K. G. Ruddick (2005), Mo<strong>de</strong>l of remote-sensing reflectance including<br />

bidirectional effects for case 1 and case 2 waters, App. Opt., 44(7), 1236-1249.<br />

Pegau, W. S. (2002), Inherent optical properties of the central Arctic <strong>sur</strong>face waters, J.<br />

Geophys. Res., 107(C10), doi: 10.1029/2000JC000382.<br />

Perovich, D. K., T. C. Grenfell, B. Light, and P. V. Hobbs (2002), Seasonal evolution of<br />

the albedo of multiyear Arctic sea ice, J. Geophys. Res., 107(C10).<br />

Perovich, D. K., C. S. Roesler, and W. S. Pegau (1998), Variability in Arctic sea ice optical<br />

properties, J. Geophys. Res., 103(C1), 1193-1208.<br />

Peterson, B. J., R. M. Holmes, J. W. McClelland, C. J. Vorosmarty, R. B. Lammers, A. I.<br />

Shiklomanov, and S. Rahmstorf (2002), Increasing river discharge to the Arctic Ocean,<br />

Science, 298, 2171-2173.<br />

Peterson, B. J., McClelland, J. W., Curry, R., Holmes, R. M., Walsh, J. E., & Aagaard, K.<br />

(2006). Trajectory shifts in the Arctic and Subarctic freshwater cycle. Science, 313, 1061-<br />

1066.<br />

Petzold, T. J., Volume scattering functions for selected ocean waters, Tech. Rep. SIO, 72-<br />

78 pp., Scripps Institution of Oceanography, San Diego, La Jolla, Ca., 1972.<br />

Pickart, R. S. (2004), Shelfbreak circulation in the Alaskan Beaufort Sea: Mean structure<br />

and variability, J. Geophys. Res., 109(C4), C04024.<br />

Pope, R. M., and E. S. Fry (1997), Absorption spectrum (380-700nm) of pure water. II.<br />

Integrating cavity mea<strong>sur</strong>ements, Appl. Optics., 36, 8710-8723.<br />

Preisendorfer, R. W. (1961), Application of radiative transfert theory to light<br />

mea<strong>sur</strong>ements in the sea, Union Géod. Géophys. Ints., Monogr., 10, 11-30.<br />

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery (1992), Numerical recipes<br />

in Fortran 77: The art of scientific computing, Chap 9 pp., Cambridge U. Press, New York.<br />

Proshutinsky, A., R. H. Bourke, and F. A. McLaughlin (2002), The role of the Beaufort<br />

Gyre in Arctic climate variability: Seasonal to <strong>de</strong>cadal climate sca<strong>les</strong>, Geophys. Res. Lett.,<br />

29(23), 2100, doi:10.1029/2002GL015847.<br />

Quicken<strong>de</strong>n, T. I., and J. A. Irvin (1980), The ultraviolet absorption spectrum of liquid<br />

water, J. Chem. Phys., 72(8), 4416-4428.<br />

Rachold, V., H. Eicken, V. V. Gor<strong>de</strong>ev, M. N. Grigoriev, H.-W. Hubberten, A. P. Lisitzin,<br />

V. P. Shevchenko, and L. Schirrmeister (2004), Mo<strong>de</strong>rn terrigenous organic carbon<br />

225


input to the Arctic Ocean, in The organic carbon cycle in the Arctic Ocean, edited by R. Stein,<br />

and R. W. MacDonald, pp. 33-41, Springer, Berlin.<br />

Rahmstorf, S. (2002), Ocean circulation and climate during the past 120,000 years, Nature,<br />

419(6903), 207-214.<br />

Retamal, L., W. F. Vincent, C. Martineau, and C. L. Osburn (in review, 2006), Comparison<br />

of optical properties of dissolved organic matter in two river-influence coastal regions<br />

of the Canadian Arctic, Estuarine Coast. Shelf Sci.<br />

Rex, M., R. J. Salawitch, P. von <strong>de</strong>r Gathen, N. R. P. Harris, M. P. Chipperfield, and B.<br />

Naujokat (2004), Arctic ozone loss and climate change, Geophys. Res. Lett., 31(4).<br />

Reynolds, R. A., D. Stramski, and B. G. Mitchell (2001), A chlorophyll-<strong>de</strong>pen<strong>de</strong>nt<br />

semianalytical reflectance mo<strong>de</strong>l <strong>de</strong>rived from field mea<strong>sur</strong>ements of absorption and<br />

backscattering coefficients within the Southern Ocean, J. Geophys. Res., 106(C4), 7125-<br />

7138.<br />

Rigor, I. G., J. M. Wallace, and R. L. Colony (2002), Response of sea ice to the Arctic<br />

Oscillation, J. Climate, 15(18), 2648-2668.<br />

Roesler, C. S., and M. J. Perry (1995), In situ phytoplankton absorption, fluorescence<br />

emission, and particulate backscattering spectra <strong>de</strong>termined from reflectance, J. Geophys.<br />

Res., 100(C7), 13279-13294.<br />

Roesler, C. S., M. J. Perry, and K. L. Car<strong>de</strong>r (1989), Mo<strong>de</strong>ling in situ phytoplankton<br />

absorption from total absorption spectra in productive inland marine waters, Limnol.<br />

Oceanogr., 34, 1510-1523.<br />

Rothrock, D. A., and J. Zhang (2005), Arctic Ocean sea ice volume: what explains its<br />

recent <strong>de</strong>pletion?, J. Geophys. Res., 110, C01002, doi:10.1029/2004JC002282.<br />

Ruddick, K. G., V. De Cauwer, Y.-J. Park, and G. Moore (2006), Seaborne mea<strong>sur</strong>ements<br />

of near infrared water-leaving reflectance: the similarity spectrum for turbid waters,<br />

Limnol. Oceanogr., 51(2), 1167-1179.<br />

Ruddick, K. G., F. Ovidio, and M. Rijkeboer (2000), Atmospheric correction of SeaWiFS<br />

imagery for turbid coastal and inland waters, App. Opt., 39(6), 897-895.<br />

Ru<strong>de</strong>ls, B., L. G. An<strong>de</strong>rson, and E. P. Jones (1996), Formation and evolution of the<br />

<strong>sur</strong>face mixed layer and halocline of the Arctic Ocean, J. Geophys. Res., 101(C4), 8807-<br />

8821.<br />

Ru<strong>de</strong>ls, B., G. Bjork, J. Nilsson, P. Winsor, I. Lake, and C. Nohr (2005), The interaction<br />

between waters from the Arctic Ocean and the Nordic Seas north of Fram Strait and<br />

along the East Greenland Current: results from the Arctic Ocean-02 O<strong>de</strong>n expedition,<br />

J. Mar. Syst., 55(1-2), 1-30.<br />

Sakshaug, E. (2004), Primary and secondary production in the Arctic seas, in The organic<br />

carbon cycle in the Arctic Ocean, edited by R. Stein, and R. W. MacDonald, pp. 57-81,<br />

Springer, Berlin.<br />

Sakshaug, E., and H. R. Skjoldal (1989), Life at the ice edge, Ambio., 18(1), 60-67.<br />

Saliot, A., G. Cauwet, G. Cahet, D. Mazaudier, and R. Daumas (1996), Microbial activities<br />

in the Lena River <strong>de</strong>lta and Laptev Sea, Mar. Chem., 53, 247-254.<br />

226


Sathyendranath, S., L. Prieur, and A. Morel (1989), A three-component mo<strong>de</strong>l of ocean<br />

colour and its application to remote sensing of phytoplankton pigments in coastal<br />

waters, Int. J. of Remote Sensing, 10(8), 1373-1394.<br />

Sch<strong>les</strong>inger, W. H. (1991), Biogeochemistry: an analysis of global change, 443 pp., Aca<strong>de</strong>mic press,<br />

San Diego.<br />

Scully, N. M., and W. L. Miller (2000), Spatial and temporal dynamics of colored dissolved<br />

organic matter in the North Water Polynya, Geophys. Res. Lett., 27(7), 1009-1011.<br />

Serreze, M. C., J. A. Maslanik, T. A. Scambos, F. Fetterer, J. Stroeve, K. Know<strong>les</strong>, C.<br />

Fowler, S. Drobot, R. G. Barry, and T. M. Haran (2003), A record minimum arctic sea<br />

ice extent and area in 2002, Geophys. Res. Lett., 30(3), 1110,<br />

doi:10.1029/2002GL016406.<br />

Serreze, M. C., J. E. Walsh, F. S. Chapin, T. Osterkamp, M. Dyurgerov, V. Romanovsky,<br />

W. C. Oechel, J. Morison, T. Zhang, and R. G. Barry (2000), Observational evi<strong>de</strong>nce<br />

of recent change in the northern high-latitu<strong>de</strong> environment, Climatic Change, 46(1-2),<br />

159-207.<br />

Siegel, D. A., S. Maritorena, and N. B. Nelson (2005), In<strong>de</strong>pen<strong>de</strong>nce and<br />

inter<strong>de</strong>pen<strong>de</strong>ncies among global ocean color properties: Reassessing the bio-optical<br />

assumption, J. Geophys. Res., 110(C7).<br />

Siegel, D. A., S. Maritorena, N. B. Nelson, D. A. Hansell, and M. Lorenzi-Kayser (2002),<br />

Global distribution and dynamics of colored dissolved and <strong>de</strong>trital organic materials, J.<br />

Geophys. Res., 107(C12), 3228, doi:10.1029/2001JC000965.<br />

Siegel, D. A., M. Wang, S. Maritorena, and W. Robinson (2000), Atmospheric correction<br />

of satellite ocean color imagery: the black pixel assumption, App. Opt., 39(21), 3582-<br />

3591.<br />

Smith, L. C., G. M. MacDonald, A. A. Velichko, D. W. Beilman, O. K. Borisova, K. E.<br />

Frey, K. V. Kremenetski, and Y. Sheng (2004), Siberian peatlands a net carbon sink<br />

and global methane source since the early Holocene, Science, 303(5656), 353-356.<br />

Smith, R. C. (1973), Optical properties of the Arctic upper water, Arctic, 26(4), 303-313.<br />

Smith, S. D., R. D. Muench, and C. H. Pease (1990), Polynyas and leads: An overview of<br />

physical processes and environment, J. Geophys. Res., 95(C6), 9461-9479.<br />

Smith, W. O., and D. M. Nelson (1986), The importance of ice-edge phytoplankton<br />

production in the Southern Ocean, BioScience, 36, 251-257.<br />

Sokal, R. R., and F. J. Rohlf (1995), Biometry: princip<strong>les</strong> and practice of statistic in biological research,<br />

Third ed., 887 pp., Freeman, New York.<br />

Steele, M., R. Morley, and W. Ermold (2001), A global ocean hydrography with a high<br />

quality Arctic Ocean, J. Climate, 14, 2079-2087.<br />

Stein, R., and R. W. MacDonald (2004), Organic carbon budget: Arctic Ocean vs. Global<br />

Ocean, in The Organic Carbon Cycle in the Arctic Ocean, edited by R. Stein, and R. W.<br />

MacDonald, pp. 315-322, Springer, Berlin.<br />

227


Stramska, M., D. Stramski, R. Hapter, S. Kaczmarek, and J. Ston (2003), Bio-optical<br />

relationships and ocean color algorithms for the north polar region of the Atlantic, J.<br />

Geophys. Res., 108(C5), 3143, doi: 10.1029/2001JC001195.<br />

Stirling, I. (1997), The importance of polynyas, ice edges, and leads to marine mammals<br />

and birds, J. Mar. Syst., 10(1-4), 9-21.<br />

Stroeve, J. C., M. C. Serreze, F. Fetterer, T. Arbetter, W. Meier, J. Maslanik, and K.<br />

Know<strong>les</strong> (2005), Tracking the Arctic's shrinking ice cover: Another extreme September<br />

minimum in 2004, Geophys. Res. Lett., 32(L04501), doi:10.1029/2004GL021810.<br />

Stumpf, R. P., R. A. Arnone, R. W. Gould, R. M. Martinolich, and V. Ransibrahmanakul<br />

(2003), A partially coupled ocean-atmosphere mo<strong>de</strong>l for retrieval of water-leaving<br />

radiance from SeaWiFS in coastal waters, Chap. 9, in NASA/TM-2003-206892, edited<br />

by S. B. Hooker, and E. R. Firestone, pp. 51-59, Greenbelt, MD.<br />

Tanré, D., M. Herman, and P. Y. Deschamps (1981), Influence of the background<br />

contribution upon space mea<strong>sur</strong>ements of ground reflectance, App. Opt., 20(20), 3676-<br />

3684.<br />

Tanré, D., M. Herman, P. Y. Deschamps, and A. <strong>de</strong> Leffe (1979), Atmospheric mo<strong>de</strong>ling<br />

for space mea<strong>sur</strong>ements of ground reflectances, including bidirectional properties, App.<br />

Opt., 18(21), 3587-3594.<br />

Tassan, S., and G. M. Ferrari (1995), An alternative approach to absorption mea<strong>sur</strong>ements<br />

of aquatic partic<strong>les</strong> retained on filters, Limnol. Oceanogr., 40(8), 1358-1368.<br />

Tassan, S., and G. M. Ferrari (2002), A sensitivity analysis of the 'transmittance-reflectance'<br />

method for mea<strong>sur</strong>ing light absorption by aquatic partic<strong>les</strong>, J. Plankton Res., 24(8), 757-<br />

776.<br />

Telang, S. A., R. Pocklington, A. S. Naidu, E. A. Romankevich, I. I. Gitelson, and<br />

Glacyshev (1991), Carbon and mineral transport in major North American, Russian<br />

Arctic, and Siberian Rivers: The St Lawrence, the Mackenzie, the Yukon, the Arctic<br />

Alaskan Rivers, the Arctic Basin Rivers in the Soviet Union, and the Yenisei, in<br />

Biogeochemistry of Major World Rivers. SCOPE 42, edited by E. T. Degens, S. Kempe, and<br />

J. E. Richey, pp. 74-104, Wiley, New York.<br />

Topliss, B. J., J. R. Miller, and P. W. Horne (1989), Ocean optical mea<strong>sur</strong>ements-II.<br />

Statistical analysis of data from Canadian eastern Arctic waters, Cont. Shelf Res., 9(2),<br />

135-142.<br />

Twardowski, M. S., E. Boss, J. B. Macdonald, W. S. Pegau, A. H. Barnard, and J. R. V.<br />

Zaneveld (2001), A mo<strong>de</strong>l for estimating bulk refractive in<strong>de</strong>x from the optical<br />

backscattering ratio and the implication for un<strong>de</strong>rstanding particle composition in case<br />

I and case II waters, J. Geophys. Res., 106(C7), 14129-14142.<br />

Twardowski, M. S., E. Boss, J. M. Sullivan, and Donaghay, P. L. (2004). Mo<strong>de</strong>ling the<br />

spectral shape of absorption by chromophoric dissolved organic matter. Marine<br />

Chemistry, 89, 69-88.<br />

Vähätalo, A. V., M. Salkinoja-Salonen, P. Taalas, and K. Salonen (2000), Spectrum of the<br />

quantum yield for the photochemical mineralization of dissolved organic carbon in a<br />

humic lake, Limnol. Oceanogr., 45(3), 664-676.<br />

228


Vähätalo, A. V., K. Salonen, M. Salkinoja-Salonen, and A. Hatakka (1999), Photochemical<br />

mineralization of synthetic lignin in lake water indicates enhanced of aromatic organic<br />

matter un<strong>de</strong>r solar radiation, Bio<strong>de</strong>gradation, 10(6), 415-420.<br />

Vähätalo, A. V., and R. G. Wetzel (2004), Photochemical and microbial <strong>de</strong>composition of<br />

chromophoric dissolved organic matter during long (months-years) expo<strong>sur</strong>es, Mar.<br />

Chem., 89, 313-326.<br />

Vermote, E., D. Tanré, J. L. Deuzé, M. Herman, and J. J. Morcrette, 6s User Gui<strong>de</strong><br />

Version 2, 218 pp., Laboratoire d'Optique Atmosphérique, Lille, France, 1997.<br />

Vincent, W. F., and C. Belzile (2003), Biological UV expo<strong>sur</strong>e in the polar oceans: Arctic-<br />

Antarctic comparisons, in Antarctic Biology in a Global Context, edited by A.H.L. Huiskes,<br />

W.W.C. Gieskes, J. Rozema, R.M.L. Schorno, S. M. v. d. Vies, and W. J. Wolff, pp.<br />

176-181., Backhuys Publishers, Lei<strong>de</strong>n, Netherlands.<br />

Vodacek, A., N. V. Blough, M. D. DeGrandpre, E. T. Peltzer, and R. K. Nelson (1997),<br />

Seasonal variation of CDOM and DOC in the Middle Atlantic Bight: Terrestrial inputs<br />

and photooxidation, Limnol. Oceanogr., 42(4), 674-686.<br />

Wadhams, P., and N. R. Davis (2000), Further evi<strong>de</strong>nce of ice thinning in the Arctic<br />

Ocean, Geophys. Res. Lett., 27, 3973-3975.<br />

Wang, J., G. Cota, and D. A. Ruble (2005a), Absorption and backscattering in the Beaufort<br />

and Chukchi Seas, J. Geophys. Res., 110(C4), C04014, doi: 10.1029/2002JC001653.<br />

Wang, J., and G. F. Cota (2003), Remote-sensing reflectance in the Beaufort and Chukchi<br />

seas: observations and mo<strong>de</strong>ls, App. Opt., 42(15), 2754-2765.<br />

Wang, J., G. F. Cota, and J. C. Comiso (2005b), Phytoplankton in the Beaufort and<br />

Chukchi Seas: Distribution dynamics, and environmental forcing, Deep Sea Res. II, 52,<br />

3355-3368.<br />

Wer<strong>de</strong>ll, P. J., and S. W. Bailey (2005), An improved in-situ bio-optical data set for ocean<br />

color algorithm <strong>de</strong>velopment and satellite product validation, Remote Sens. Environ., 98,<br />

122-140.<br />

Wheeler, P. A., M. Gosselin, E. Sherr, D. Thibault, D. L. Kirchman, R. Benner, and T. E.<br />

Whitledge (1996), Active cycling of organic carbon in the central Arctic Ocean, Nature,<br />

380(6576), 697-699.<br />

Wheeler, P. A., J. M. Watkins, and R. L. Hansing (1997), Nutrients, organic carbon and<br />

organic nitrogen in the upper water column of the Arctic Ocean: implications for<br />

sources of dissolved organic carbon, Deep Sea Res. II, 44(8), 1571-1592.<br />

Woodgate, R. A., K. Aagaard, J. H. Swift, K. K. Falkner, and W. M. Smethie (2005), Pacific<br />

ventilation of the Arctic Ocean's lower halocline by upwelling and diapycnal mixing<br />

over the continental margin, Geophys. Res. Lett., 32(18).<br />

Woodgate, R. A., K. Aagaard, and T. J. Weingartner (2005), Monthly temperature, salinity,<br />

and transport variability of the Bering Strait through flow, Geophys. Res. Lett., 32(4).<br />

Woodwell, G. M., F. T. Mackenzie, R. A. Houghton, M. Apps, E. Gorham, and E.<br />

Davidson (1998), Biotic feedbacks in the warming of the earth, Climatic Change, 40,<br />

495-518.<br />

229


Xie, H., O. C. Zafiriou, W.-J. Cai, R. G. Zepp, and Y. Wang (2004), Photooxidation and its<br />

effects on the carboxyl content of dissolved organic matter in two coastal rivers in the<br />

Southeastern United States, Environ. Sci. Techn., 38(15), 4113-4119.<br />

Xie, H. X., and M. Gosselin (2005), Photoproduction of carbon monoxi<strong>de</strong> in first-year sea<br />

ice in Franklin Bay, southeastern Beaufort Sea, Geophys. Res. Lett., 32(12).<br />

Zafiriou, O. C. (1977), Marine organic photochemistry previewed, Mar. Chem., 5, 497-522.<br />

Zaneveld, J. R. V., J. C. Kitchen, and C. Moore (1994), The scattering error correction of<br />

reflecting-tube absorption meters, in Ocean Optics XII, edited by J. S. Jaffe, pp. 44-55,<br />

SPIE, Bergen.<br />

Zepp, R. G., T. V. Callaghan, and D. Erickson III, J. (2003), Interactive effects of ozone<br />

<strong>de</strong>pletion and climate change on biogeochemical cyc<strong>les</strong>, Photochem. Photobiol. Sci., 2, 51-<br />

61.<br />

Zhang, Y., and H. X. Xie (in press 2006), Factors affecting the efficiency of carbon<br />

monoxi<strong>de</strong> photoproduction in the St Lawrence estuarine system (Canada), Environ. Sci.<br />

Tech.<br />

Zibordi, G., and G. M. Ferrari (1995), Instrument Self-Shading in Un<strong>de</strong>rwater Optical<br />

Mea<strong>sur</strong>ements - Experimental-Data, App. Opt., 34(15), 2750-2754.<br />

Zibordi, G., F. Mélin, and J.-F. Berthon (2006), Comparison of SeaWiFS, MODIS and<br />

MERIS radiometric products at a coastal site, Geophys. Res. Lett., 33, L06617,<br />

doi:10.1029/2006GL025778.<br />

230


Annexe 1 : Modification <strong>de</strong> l’algorithme Quasi-<br />

Analytique pour l’application à la Mer <strong>de</strong> Beaufort<br />

231


Annex A1. Modification of the Quasi-Analytical Algorithm for an application to<br />

the Beaufort Sea<br />

A1.1. Introduction and motivation<br />

In Chapter IV, we presented a method to <strong>de</strong>rived the contribution of CDOM to the<br />

total absorption coefficient ([aCDOM/at]) at 412 nm. To extrapolate [aCDOM/at] from 412 to<br />

the full spectrum between 300 and 600 nm, the magnitu<strong>de</strong> of at(412) is necessary (see<br />

section VI.3.3). It has been shown that the total absorption coefficient could be <strong>de</strong>rived<br />

with a relatively good precision from remote sensing reflectance [IOCCG, 2006].<br />

In addition, little is known on the spatial variability of the different <strong>sur</strong>face water<br />

masses in the southeastern Beaufort Sea. In particular, the distribution of river runoff,<br />

which constitutes a major part of the freshwater budget in the western part of the Arctic<br />

[Macdonald et al., 2002], remains poorly documented. Spatial-temporal variability of<br />

satellite-<strong>de</strong>rived inherent optical properties (IOP) may provi<strong>de</strong> new insights on river<br />

runoff distribution in the southeastern part of the Beaufort Sea.<br />

For those reasons, here I present the modifications to the method proposed by Lee<br />

et al. [2002] to <strong>de</strong>rive IOPs in the southern Beaufort Sea from SeaWiFS imagery. The<br />

objectives were 1) to inclu<strong>de</strong> recent <strong>de</strong>velopments achieved the field of Ocean Color, and<br />

2) to tune the empirical relationships involved in the algorithm with the CASES data set.<br />

The results of the application of the algorithm to the CASES data set are presented and<br />

discussed. The application to SeaWiFS imagery is shown and discussed in Chapter VI<br />

(Figs. VI.6 and VI.7).<br />

A1.2. Quasi-Analytical Algorithm <strong><strong>de</strong>s</strong>cription<br />

Lee et al. [2002] recently proposed a quasi-analytical algorithm (hereafter <strong>de</strong>noted<br />

as QAA02) that calculates the bulk IOPs, i.e. the total absorption (at) and backscattering<br />

(bb) coefficients, from the remote-sensing reflectance spectrum (Rrs). The later is <strong>de</strong>fined<br />

as the ratio of water-leaving radiance to the above-<strong>sur</strong>face downwelling irradiance. Our<br />

algorithm differs from the QAA02 with respect to 1) the formulation of Rrs versus the<br />

IOP, 2) the empirical formulations for the absorption coefficient at reference wavelength<br />

232


(λ0), and 3) the use of a single value for the exponent of the power law used to mo<strong>de</strong>l the<br />

particulate backscattering spectrum (Y).<br />

First, a semi-analytical remote-sensing reflectance mo<strong>de</strong>l <strong><strong>de</strong>s</strong>igned to account for<br />

the bidirectional effects [Park and Ruddick, 2005] is used instead of the Gordon et al.<br />

[1988] formulation. Since the Park and Ruddick [2005] mo<strong>de</strong>l required a priori<br />

information on backscattering nature of the medium (i.e. molecular versus particulate) to<br />

operate, additional steps are nee<strong>de</strong>d in the implementation of our algorithm with respect<br />

to the QAA02. Those are illustrated in Fig. A1, which shows the inputs/outputs data flow<br />

through each step of the processing chain. Table A1 provi<strong><strong>de</strong>s</strong> a summary of all the<br />

equations used in the different steps. Briefly, the correction factors for both air-water<br />

interface reflection and bidirectional effects for the observation geometry are expressed<br />

by a single fourth-polynomial mo<strong>de</strong>l as<br />

4<br />

Rrs ( λ,<br />

θ s , θ v , Δ φ,<br />

W , γ b ( λ))<br />

= ∑ gi<br />

( θ s , θ v , Δφ,<br />

W , γ b ( λ))<br />

ωb<br />

( λ)<br />

, (A1)<br />

i=<br />

1<br />

where ωb is the ratio bb/(at+bb), and gi are coefficients tabulated in look-up table (LUT)<br />

for given values of the Sun and sensor zenith ang<strong>les</strong> (θs and θv), the azimuth difference<br />

between the Sun-pixel and pixel-sensor half vertical plane (Δφ), the wind speed (W) and<br />

the phase function parameter, γb(λ). The latter is <strong>de</strong>fined as the relative contribution of<br />

partic<strong>les</strong> to the total backscattering (i.e. bbp(λ)/bb(λ)). Thus the first three steps of the<br />

algorithm (eqs. no. 1-4 in Table A1 and Fig. A1), similar to the QAA02 (steps 0-3 and 5<br />

in Lee et al. 2002), provi<strong>de</strong> an initial estimate of γb at a reference wavelength. With the<br />

initial value of γb(λ0) and Rrs(λ0, θs, θv, Δφ, W), ωb(λ0) is estimated by inverting the<br />

fourth-or<strong>de</strong>r polynomial using a numerical solution (no. 5) [Laguerre's method, or the<br />

zroots function in Chap 9.5 of Press et al., 1992]. Two iterations are sufficient to<br />

obtain a difference between two consecutives estimate of γb(λ0)


chlorophyll concentration in case 1 water [e.g., O'Reilly et al., 1998] that uses the<br />

maximum reflectance band ratio between 443, 490 and 510 nm upon 555 nm (r 2 = 0.785,<br />

N = 35; for at-w(555)0.5aw(555)):<br />

234<br />

⎡ Rrs<br />

( 650)<br />

⎤<br />

−0.<br />

69+<br />

1.<br />

41log10<br />

⎢ ⎥<br />

⎣ Rrs<br />

( 555)<br />

⎦<br />

a ( 650)<br />

= a ( 650)<br />

+ 10<br />

(8)<br />

t<br />

w<br />

where aw(650)=0.34 [Pope and Fry, 1997]. For continuity in the at retrieval, when at-<br />

w(555) falls between 0.03 and 0.06 m -1 the a linear combination of the two results<br />

obtained with λ0 = 555 and 650 nm is performed following equation 20 in Lee et al.<br />

[2002].<br />

Third, the spectral partic<strong>les</strong> backscattering coefficient (bbp) is calculated using λ −Y<br />

law with Y = 1. Note that in the QAA02, an empirical relationship based on the<br />

reflectance ratio Rrs(443)/Rrs(555) is used to estimate Y (<strong>de</strong>veloped with data from the<br />

Gulf of Mexico).<br />

In Chapter IV, an empirical algorithm was proposed to estimate the ratio between<br />

the CDOM and the total absorption coefficients at 412 nm ([aCDOM/at](412)). Multiplying


the QAA-<strong>de</strong>rived at(412) by the empirically <strong>de</strong>rived [aCDOM/at](412) yield the CDOM<br />

absorption coefficient.<br />

Table A1. List of equations find in Fig. A1.<br />

Equation<br />

number 1<br />

1<br />

2<br />

3<br />

4<br />

If λ0=555 then<br />

a<br />

t<br />

( 555)<br />

Equations Ref.<br />

a<br />

= w<br />

( 555)<br />

235<br />

2<br />

−1.<br />

3 − 2.<br />

53ρ<br />

+ 0.<br />

46ρ<br />

+ 10<br />

⎛ max[<br />

Rrs<br />

( 443,<br />

490,<br />

510)<br />

] ⎞<br />

with ρ = log ⎜<br />

⎟ 10<br />

.<br />

⎝ Rrs<br />

( 555)<br />

⎠<br />

If λ0=650 then<br />

a<br />

t<br />

( 650)<br />

= a<br />

w<br />

( 650)<br />

+ 10<br />

⎡ Rrs<br />

( 650)<br />

⎤<br />

−0.<br />

69+<br />

1.<br />

41log10<br />

⎢ ⎥<br />

⎣ Rrs<br />

( 555)<br />

⎦<br />

with Rrs(650)= 1.082*Rrs(665) or 1.101* Rrs(670)<br />

r<br />

rs<br />

( λ)<br />

=<br />

1<br />

0.<br />

52<br />

2<br />

g 0 + g 0 + 4g1rrs<br />

R<br />

rs<br />

( λ)<br />

+ 1.<br />

7R<br />

rs<br />

( λ)<br />

−<br />

( λ)<br />

ωb<br />

( λ)<br />

= , g0=0.0945 and g1=0.0794<br />

2g<br />

ωb<br />

( λ)<br />

at<br />

( λ)<br />

bb<br />

( λ)<br />

= and bbp(λ)=bb(λ) - bbw(λ)<br />

1−<br />

ω ( λ)<br />

5 ωb(λ)=zroots(Rrs(λ), gi; i=1..4)<br />

6<br />

b<br />

Y<br />

⎛ λ0<br />

⎞<br />

bbp ( λ)<br />

= bbp<br />

( λ0<br />

) ⎜ ⎟<br />

⎝ λ ⎠<br />

7 ( 1 − ω ( λ)<br />

)<br />

b bb<br />

( λ)<br />

at<br />

( λ)<br />

=<br />

ωb<br />

( λ)<br />

1<br />

Numbers correspond to the upper subscript in Fig. A1.<br />

This study<br />

Lee et al.<br />

[2002]<br />

Gordon et al.<br />

[1988]<br />

Press et al.<br />

[1992]<br />

e.g., Gordon<br />

et al. [1988]


Figure A1. Schematic representation of the quasi-analytical algorithm<br />

<strong>de</strong>veloped to <strong>de</strong>rive the partic<strong>les</strong> backscattering coefficient at 555 nm and<br />

the spectral total absorption coefficient. The data (parallelogram)<br />

necessary for each steps of the algorithm (rectangular boxes) are i<strong>de</strong>ntified<br />

by the dotted arrows, while the outputs are i<strong>de</strong>ntified by the thin line with<br />

filled arrows. The final outputs, i.e. bbp(555) and at(λ), are i<strong>de</strong>ntified by<br />

the think line with filled arrows at bottom right of the diagram. The<br />

equations (i<strong>de</strong>ntified with upper script number) are listed in Table A1.<br />

236


A1.3. Results and discussion<br />

A1.3.1. Application of QAA to in situ data set<br />

The QAA02 and its modified version <strong><strong>de</strong>s</strong>cribed above (QAA-CASES) were<br />

applied to the remote-sensing reflectance mea<strong>sur</strong>ements obtained during the CASES field<br />

campaign (Fig. A2). Table A2 provi<strong><strong>de</strong>s</strong> the statistics on the performance to retrieve at of<br />

both algorithms. The criteria presented inclu<strong><strong>de</strong>s</strong> the 95% confi<strong>de</strong>nce interval range for<br />

intercept and the slope of the linear regression (Type 2) between retrieved and mea<strong>sur</strong>ed<br />

at [Sokal and Rohlf, 1995], the geometric standard <strong>de</strong>viation (GSD 95% ), and the mean<br />

relative difference (AD). A perfect performance of the inverse algorithm would give an<br />

intercept of 0 and a slope of 1. The GSD 95% is calculated as 10 1.966*σ , where σ is the<br />

standard <strong>de</strong>viation of the difference between mea<strong>sur</strong>ed and retrieved estimate computed<br />

on the log10-transformed data. Note that GSD 95% is almost equivalent to 10 1.966*RMSE<br />

where RMSE is the root-mean-square-error which is commonly used is algorithm<br />

evaluation [IOCCG, 2006]. The interpretation of GSD 95% is easier than the RMSE since it<br />

means that a retrieved value of x (x retrieved ) falls within the range between x true / GSD 95%<br />

and x true * GSD 95% at 95% confi<strong>de</strong>nce interval.<br />

As expected, the tuned algorithm (QAA-CASES) provi<strong><strong>de</strong>s</strong> more accurate estimate<br />

of the total absorption than the QAA02, as revealed by the lower GSD 95% and AD at all<br />

wavelengths (Table A2). For the QAA02, both the intercept and slope parameter <strong>de</strong>viated<br />

significantly from 0 and 1 respectively, indicating a bias in the retrieval of at. The low<br />

slope of the regression obtained using the QAA02 is largely explained by the use of an<br />

inappropriate relationship for Y in the inversion. In fact, when the QAA02’s empirical<br />

formulation for Y is used in instead of 1.0 in the QAA-CASES, the slope of the regression<br />

between retrieved and mea<strong>sur</strong>ed at(412) was significantly lower than 1. In addition, the<br />

performance of the QAA-CASES <strong>de</strong>creases with <strong>de</strong>creasing wavelength as indicated by<br />

the higher values for GSD 95% and AD. These results also suggest that the value of Y is not<br />

constant in this area. While studies carried out in the Arctic found empirical relationships<br />

between Y and the magnitu<strong>de</strong> of bb(555) [Stramska et al., 2003; Wang et al., 2005], no<br />

significant relationships between neither Rrs ratios nor the magnitu<strong>de</strong> of bp(555) and the<br />

mea<strong>sur</strong>ed Y values were find in this study (not shown). Hence the highest Y values (~1.4),<br />

as <strong>de</strong>rived from the bp(λ) mea<strong>sur</strong>ed with the ac-9, were observed near the Mackenzie<br />

237


River mouth, while offshore the values ranged between 0.3 to 1.0. This trend is opposite<br />

to the one proposed by Lee et al. [2002] and Morel and Maritorena [2001] where Y<br />

<strong>de</strong>creases with increasing turbidity. Here a value of 1.0 was found to give the best results<br />

in the retrieval of at at 412 nm.<br />

Figure A2. Comparison between retrieved and mea<strong>sur</strong>ed at with the ac9 at<br />

A) 412 nm and B) 440 nm. Total absorption coefficients were retrieved<br />

using the optimized version of QAA proposed by Lee et al. [2002] that use<br />

λ0 555 and 640 nm (open circ<strong>les</strong>) and the tuned QAA with the CASES<br />

dataset (inverted triang<strong>les</strong>).<br />

238


Table A2. Performance of the QAA to retrieved at(λ) with empirical relationships of Lee<br />

et al. [2002] versus QAA tuned with CASES dataset (n=46).<br />

Wavelength Approach Intercept<br />

(nm)<br />

a Slope a 95% b<br />

GSD AD c<br />

(%)<br />

412 QAA02 [-0.138, -0.043] [0.809, 0.945] 1.51 16.2<br />

QAA-CASES [-0.019, 0.054] [0.965, 1.069] 1.38 12.9<br />

440 QAA02 [-0.163, -0.062] [0.811, 0.928] 1.42 14.3<br />

QAA-CASES [-0.030, 0.042] [0.963, 1.046] 1.28 10.1<br />

490 QAA02 [-0.179, -0.059] [0.799, 0.909] 1.35 18.9<br />

QAA-CASES [-0.044, 0.053] [0.930, 1.019] 1.29 12.9<br />

510 QAA02 [0.736, 0.839] [0.736, 0.839] 1.27 9.9<br />

QAA-CASES [-0.028, 0.075] [0.997, 1.094] 1.25 10.1<br />

555 QAA02 [-0.174, -0.026] [0.818, 0.962] 1.25 9.4<br />

QAA-CASES [-0.006, 0.124] [0.992, 1.119] 1.22 7.4<br />

a<br />

95% Confi<strong>de</strong>nce interval calculated on the log10-transformed data using Type II regression<br />

mo<strong>de</strong>l with the major axis (MA method based on Sokal and Rohlf [1995],<strong><strong>de</strong>s</strong>cribed by Pierre<br />

Legendre, 2001: http://www.bio.umontreal.ca/casgrain/fr/labo/mo<strong>de</strong>l-ii.html).<br />

b 95%<br />

1.<br />

966*<br />

σ log<br />

Geometric Standard Deviation (95% confi<strong>de</strong>nce interval): GSD = 10 , where σlog is the<br />

standard <strong>de</strong>viation calculated on the log10-transformed difference between retrieved and mea<strong>sur</strong>ed<br />

values.<br />

c Mean absolute difference in %,<br />

retrieved mea<strong>sur</strong>ed<br />

n ai<br />

− ai<br />

AD = ∑<br />

mea<strong>sur</strong>ed<br />

n a<br />

1 .<br />

i<br />

i<br />

The inaccuracy of the QAA-CASES could also become from the empirical<br />

relationships used to obtain the total absorption at the reference wavelength (Eqs. 7 and<br />

8). The mean absolute error ma<strong>de</strong> on at(555) and at(650) was only 5.1% and 2.5%<br />

respectively. Those are lower than the error ma<strong>de</strong> using the QAA02’s empirical<br />

relationships (~10%).<br />

Unlike most semi-analytical mo<strong>de</strong>ls currently available [see review by IOCCG,<br />

2006], the proposed version of the QAA accounts for the non-isotropic nature of the<br />

radiant light field that emerged from the ocean [Loisel and Morel, 2001; Morel et al.,<br />

2002; Park and Ruddick, 2005]. This <strong>de</strong>pen<strong>de</strong>nce is largely explained by the nature of<br />

light scattering (i.e. molecular vs particulate) [Lee et al., 2004; Morel et al., 2002] which<br />

is inclu<strong>de</strong>d in the mo<strong>de</strong>l form adopted here. According to Park and Ruddick [2005],<br />

bidirectional effects (BRDF) due to illumination-viewing geometry and the nature of the<br />

scattering can explain ~10% of the variability in the relationship between Rrs and the ratio<br />

bb/(at+bb). When satellite data acquired in various illumination conditions are averaged to<br />

239


produce level 3 data, the proposed algorithm should minimize the un<strong><strong>de</strong>s</strong>irable variability<br />

resulting from the ocean BRDF.<br />

Unfortunately, no in situ mea<strong>sur</strong>ements of bbp were performed during CASES. So<br />

I compared the retrieved bbp(555) values versus the mea<strong>sur</strong>ed bp(555) values (Fig. A3). A<br />

relatively strong correlation was obtained for the linear regression between the logarithm-<br />

transformed data (r²=0.86; N=46). Interestingly, the average b bp<br />

~ calculated using<br />

retrieved bbp(555) and mea<strong>sur</strong>ed bp(555) was 1.4% with a standard <strong>de</strong>viation of 0.7%.<br />

Those are within the expected range for partic<strong>les</strong> found in coastal waters [e.g., Chami et<br />

al., 2005; Twardowski et al., 2001; Wang et al., 2005].<br />

Figure A3. Comparison between QAA-<strong>de</strong>rived bbp(555) and mea<strong>sur</strong>ed<br />

bp(555).<br />

A1.3.2. Estimation of aCDOM<br />

To estimate aCDOM, several investigators have proposed empirical relationships<br />

based on a single water-leaving radiance or reflectance ratio [D'Sa and Miller, 2003;<br />

Johannessen et al., 2003; Kahru and Mitchell, 2001; Kowalczuk et al., 2005]. We fitted<br />

all possible ratios from the 13 Rrs SPMR channels with the aCDOM values at 412 nm. The<br />

240


egression between the ratio of Rrs at 412 nm to Rrs at 665 nm with aCDOM at 412 nm gives<br />

the highest correlation (aCDOM(412)=10 -.301-.656*[Rrs(412)/Rrs(665)] ; r 2 =0.924, GSD 95% =1.63,<br />

AD=26.8%). For comparison, we estimate aCDOM at 412 nm multiplying the QAA-<strong>de</strong>rived<br />

at(412) by the empirically-<strong>de</strong>rived [aCDOM/at](412) (Chapter IV). Figure A1.4 compares<br />

the retrieved aCDOM(412) with its mea<strong>sur</strong>ed value. Combining the results from the QAA<br />

and the empirical algorithm provi<strong><strong>de</strong>s</strong> a more robust estimate of the CDOM absorption<br />

coefficient than an empirical algorithm based on a single Rrs ratio in this region<br />

(GSD 95% =1.45, AD=14.8%).<br />

Figure A4. Retrieved versus mea<strong>sur</strong>ed aCDOM(412) in the Beaufort Sea.<br />

Statistical parameters of the Type II regression are also shown.<br />

A1.4. Conclusion<br />

In view of <strong>de</strong>riving IOPs from SeaWiFS imagery in the Beaufort Sea, a semi-<br />

analytical algorithm was modified 1) to account for the anisotropic nature of the oceanic<br />

upwelling light field and 2) for the regionally distinct relationships between absorption<br />

coefficients at reference wavelength (i.e. 555 or 650 nm) and various reflectance ratios.<br />

The method provi<strong><strong>de</strong>s</strong> an estimation of the total absorption coefficient at 412 and 443 nm<br />

within a precision of about 38 and 28%, respectively, at confi<strong>de</strong>nce level of 95%.<br />

Combined with the [aCDOM/at](412) algorithm (Chap. IV) the CDOM absorption<br />

coefficient could be estimated within a precision of 45% at confi<strong>de</strong>nce level of 95%. For<br />

241


the Beaufort Sea, the latter performance is better than any other single band ratio found in<br />

the literature.<br />

Satellite-<strong>de</strong>rived IOPs (bbp and at) could be further used to study the spatio-<br />

temporal variability of the Makcenzie River plume and to quantify the partic<strong>les</strong><br />

distribution in the <strong>sur</strong>face waters (Forest et al., Particulate organic carbon <strong>flux</strong>es on the<br />

Mackenzie Shelf slope: physical and biological forcing shelf basin exchanges, submitted<br />

manuscript to Journal of Marine Systems, 2006). Finally, the magnitu<strong>de</strong> of the CDOM<br />

absorption coefficient can also be use as an indicator of the DOM photoreactivity (see<br />

Chap. VI).<br />

A1.4. References<br />

Chami, M., E. B. Shybanov, T. Y. Churilova, G. A. Khomenko, M. E. G. Lee, O. V.<br />

Martynov, G. A. Berseneva, and G. K. Korotaev (2005), Optical properties of the<br />

partic<strong>les</strong> in the Crimea coastal waters (Black Sea), J. Geophys. Res., 110(C11),<br />

C11020, doi:10.1029/2005JC003008.<br />

D'Sa, E., and R. L. Miller (2003), Bio-optical properties in waters influenced by the<br />

Mississippi River during low flow conditions, Remote Sens. Environ., 84, 538-<br />

549.<br />

Huot, Y., C. A. Brown, and J. J. Cullen (2005), New algorithms for MODIS sun-induced<br />

chlorophyll fluorescence and a comparison with present data products, Limnol.<br />

Oceanogr.: Methods, 3, 108-130.<br />

IOCCG (2006), Remote sensing of inherent optical properties: Fundamentals, tests of<br />

algorithms, and applications, edited by Z.-P. Lee, pp. 126, Reports of the<br />

International Ocean-Colour Coordinating Group, No. 5, IOCCG, Dartmouth,<br />

Canada.<br />

Johannessen, S. C., W. L. Miller, and J. J. Cullen (2003), Calculation of UV attenuation<br />

and colored dissolved organic matter absorption spectra from mea<strong>sur</strong>ements of<br />

ocean color, J. Geophys. Res., 108(C9), 3301, doi:10.1029/2000JC000514.<br />

Kahru, M., and G. B. Mitchell (2001), Seasonal and nonseasonal variability of satellite<strong>de</strong>rived<br />

chlorophyll and colored dissolved organic matter concentration in the<br />

California Current, J. Geophys. Res., 106(C2), 2517-2529.<br />

Kowalczuk, P., J. Olszewski, M. Darecki, and S. Kaczmarek (2005), Empirical<br />

relationships between coloured dissolved organic matter (CDOM) absorption and<br />

apparent optical properties in Baltic Sea waters, Int. J. Remote Sens., 26(2), 345-<br />

370.<br />

Lee, Z.-P., K. L. Car<strong>de</strong>r, and R. A. Arnone (2002), Deriving inherent optical properties<br />

from water color : a multiband quasi-analytical algorithm for optically <strong>de</strong>ep<br />

waters, Appl. Opt., 41(27), 5755-5772.<br />

242


Lee, Z.-P., K. L. Car<strong>de</strong>r, and K. Du (2004), Effects of molecular and particle scatterings<br />

on the mo<strong>de</strong>l parameter for remote-sensing reflectance, Appl. Opt., 43(25), 4957-<br />

4964.<br />

Loisel, H., and A. Morel (2001), Non-isotropy of the upward radiance field in typical<br />

coastal (Case 2) waters, Int. J. Remote Sens., 22(2&3), 275-295.<br />

Macdonald, R. W., F. A. McLaughlin, and E. C. Carmack (2002), Fresh water and its<br />

sources during the SHEBA drift in the Canada Basin of the Arctic Ocean, Deep-<br />

Sea Research I, 49, 1769-1785.<br />

Morel, A., D. Antoine, and B. Gentili (2002), Bidirectional reflectance of oceanic waters:<br />

accounting for Raman emission and varying particle scattering phase function,<br />

Appl. Opt., 41(30), 6289-6306.<br />

O'Reilly, J. E., S. Maritorena, G. Mitchell, D. A. Siegel, K. L. Car<strong>de</strong>r, D. L. Garver, M.<br />

Kahru, and C. R. McClain (1998), Ocean color chlorophyll algorithms for<br />

SeaWiFS, J. Geophys. Res., 103(C11), 24937-24950.<br />

Park, Y.-J., and K. G. Ruddick (2005), Mo<strong>de</strong>l of remote-sensing reflectance including<br />

bidirectional effects for case 1 and case 2 waters, Appl. Opt., 44(7), 1236-1249.<br />

Pope, R. M., and E. S. Fry (1997), Absorption spectrum (380-700 nm) of pure water. II.<br />

Integrating cavity mea<strong>sur</strong>ements, Appl. Opt., 36, 8710-8723.<br />

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery (1992), Numerical<br />

recipes in Fortran 77: The art of scientific computing, Chap 9 pp., Cambridge U.<br />

Press, New York.<br />

Sokal, R. R., and F. J. Rohlf (1995), Biometry: princip<strong>les</strong> and practice of statistic in<br />

biological research, Third ed., 887 pp., Freeman, New York.<br />

Stramska, M., D. Stramski, R. Hapter, S. Kaczmarek, and J. Ston (2003), Bio-optical<br />

relationships and ocean color algorithms for the north polar region of the Atlantic,<br />

J. Geophys. Res., 108(C5), 3143, doi: 10.1029/2001JC001195.<br />

Twardowski, M. S., E. Boss, J. B. Macdonald, W. S. Pegau, A. H. Barnard, and J. R. V.<br />

Zaneveld (2001), A mo<strong>de</strong>l for estimating bulk refractive in<strong>de</strong>x from the optical<br />

backscattering ratio and the implication for un<strong>de</strong>rstanding particle composition in<br />

case I and case II waters, J. Geophys. Res., 106(C7), 14129-14142.<br />

Wang, J., G. Cota, and D. A. Ruble (2005), Absorption and backscattering in the<br />

Beaufort and Chukchi Seas, J. Geophys. Res., 110(C4), C04014, doi:<br />

10.1029/2002JC001653.<br />

243

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!