29.03.2013 Views

THESE EN COTUTELLE - paleoclimat ephe

THESE EN COTUTELLE - paleoclimat ephe

THESE EN COTUTELLE - paleoclimat ephe

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

N° d’ordre: 3344<br />

<strong>THESE</strong> <strong>EN</strong> <strong>COTUTELLE</strong><br />

Universidade de Lisboa Université Bordeaux<br />

Faculdade de Ciências Ecole Doctorale des<br />

Departamento de Geologia Sciences du vivant, Géosciences<br />

Science de l’environnement<br />

Présentée à<br />

L’UNIVERSITÉ de LISBONNE<br />

Par Filipa Naughton<br />

POUR OBT<strong>EN</strong>IR LE GRADE DE<br />

DOCTEUR<br />

. en Géologie spécialité Paléontologie et Stratigraphie (Portugal)<br />

. en Océanographie, Paléo-Océanographie (France)<br />

As variações climáticas dos últimos 30 000 anos e sua influência<br />

na evolução dos sistemas costeiros do norte de Portugal<br />

Les variations climatiques des derniers 30 000 ans et leur influence sur<br />

l’évolution des systèmes côtiers du nord de Portugal<br />

Dirigée par :Mme Maria Fernanda SÁNCHEZ GOÑi et Mme Maria da Conceição FREITAS<br />

Devant la commission d’examen formée de :<br />

Soutenue le: 18 Janvier 2007<br />

Président du Jury :<br />

M. J.-L. TURON, Directeur de Recherche CNRS (EPOC-Université Bordeaux 1)<br />

Rapporteurs :<br />

M. H. SEPPÄ, Maître de Conférence (Dép. de Géologie-Université Helsinki)<br />

Mme F. ABRANTES, Directeur de Recherche (Dép. de Géologie Marine-INETI, Lisbonne)<br />

Examinateur :<br />

Mme T. DRAGO, Chercheur Auxiliaire (IPIMAR, Olhão)<br />

Directeurs de thèse :<br />

Mme M.F. SÁNCHEZ GOÑi, Maître de Conférence EPHE (EPOC-Université Bordeaux1)<br />

Mme M.C. FREITAS, Professeur Associé-FCUL (Université de Lisbonne)


Principalmente para ti....................<br />

Maria da Luz Naughton


Agradecimentos, Remerciements<br />

A execução do presente trabalho não teria sido possível sem a contribuição<br />

de diversas pessoas e entidades, a quem desejo expressar os mais sinceros<br />

agradecimentos, nomeadamente:<br />

A Maria Fernanda Sánchez Goñi pour m’avoir contaminée avec sa passion<br />

pour la recherche et soutenue au cours de ces derniers quatre ans. Je te remercie,<br />

Maria, pour toute ta gentillesse et ton amitié, pour avoir toujours été disponible et<br />

aidée à n’importe quelle heure, chaque jour de la semaine voire les week-ends.<br />

Merci, Maria, pour m’avoir appris non seulement les pollens mais aussi beaucoup de<br />

choses sur la paléoclimatologie. Merci, Maria, pour m’avoir fait voir le vrai travail en<br />

équipe, avec toutes ces discussions scientifiques partagées entre autres dans la salle<br />

à café.<br />

À Teresa Drago a quem gostaria de expressar toda a minha gratidão e sem a<br />

qual este trabalho não teria sido possível de ser realizado. Obrigada Teresa por teres<br />

apostado na minha formação em palinologia, pela confiança em mim depositada,<br />

pelo interesse e dedicação com que tens acompanhado o meu trabalho e pelo<br />

teu encorajamento.<br />

A Stéphanie Desprat, pour m’avoir toujours encouragée et principalement<br />

pour m’avoir soutenue au cours de ces 4 ans. Je te remercie « Stephy » pour avoir<br />

parfois été ma tête et pour avoir retrouver tous mes fichiers perdus dans l’immensité<br />

de mes dossiers sur l’ordinateur. Je te remercie pour le temps passé à m’apprendre<br />

les logiciels compliqués et à partager tes connaissances et ton enthousiasme pour la<br />

recherche. Surtout, je te remercie pour cette complicité et notre amitié, nos<br />

discussions parfois « trop » scientifiques dans les cafés de la place Camille Julian …<br />

heureusement qu’on ne peut pas discuter dans la salle du cinéma UTOPIA …<br />

À Professora Conceição Freitas pela sua disponibilidade e co-orientação,<br />

pelas suas críticas e sugestões e pela sua simpatia sempre presente.<br />

A Josette Duprat, pour avoir été toujours là pour partager ses résultats<br />

magnifiques de foraminifères planctoniques ainsi toute sa connaissance sur les<br />

changements des ces associations. Merci Josette pour votre dynamisme et votre<br />

sens pratique afin de répondre à toutes les questions scientifiques posées sur le<br />

moment. Je vous remercie aussi pour toute votre gentillesse …<br />

A Marie-Hélène Castera, pour les préparations des nombreuses lames<br />

palynologiques, son dynamisme et sa bonne humeur. Je te remercie pour le temps<br />

que tu as passé sur mes lames incluant le montage d’une deuxième, troisième voir<br />

sixième lame sur un même niveau. Il faut le dire : « quel courage ! ». Pour la dernière<br />

année, ta nouvelle partenaire de préparations d’échantillons, Muriel, a<br />

heureusement beaucoup participé à finaliser quelques séries d’échantillons pour<br />

moi et je la remercie aussi vivement.


A Jean-Marie Jouanneau, pour son soutien et son assistance quotidiens. J’ai<br />

beaucoup apprécié votre intérêt pour mon travail et je vous remercie pour toutes les<br />

discussions et les échanges sur de nombreux sujets scientifiques, vous avez contribué<br />

par vos relectures et critiques à l’aboutissement des articles scientifiques.<br />

A Jean-Louis Turon pour m’avoir reçu en toute gentillesse au sein de votre<br />

équipe « Biopal », pour m’avoir donné la possibilité de travailler sur des carottes<br />

Bordelaises et pour m’avoir passé votre magnifique microscope. Je vous remercie<br />

aussi pour les donnés sur les assemblages de dinokyste de la carotte Française.<br />

A l’équipe des traceurs isotopiques, en particulier Karine Charlier et Bruno<br />

Malaisé pour les donnés isotopiques. A Francis Grousset pour les discussions sur les<br />

événements d’Heinrich et son étudiante Elsa Jullien pour avoir toujours été disponible<br />

pour partager ses connaissances acquises pendant cette dernière année et pour les<br />

discussions de « paléoclimapoésie ». Je tiens aussi à remercier aussi Philippe Martinez<br />

pour m’avoir proposé une vacation qui a contribué à financer partiellement ma<br />

dernière année de thèse.<br />

A l’équipe d’environnement côtiers, en particulier à Monsieur Castaing pour<br />

m’avoir aidé sur le sujet des estuaires et pour m’avoir permis de participer à<br />

quelques sorties de terrains ainsi qu’à quelques cours de DEA. Je tiens aussi à<br />

remercier Sébastien Zaragosi pour avoir toujours été disponible à m’aider et à<br />

discuter sur les images radiographiques des carottes marines et aussi sur les données<br />

obtenues de l’étude sédimentologique des lames minces. A Michel Cremer pour la<br />

gentillesse de m’aider sur les images radiographiques des carottes marines.<br />

Je tiens aussi à remercier tout le couloir BIOPAL pour ces dernières années<br />

passées dans la bonne humeur et pour m’avoir reçue et intégrée au sein de leur<br />

équipe.<br />

A Jean-François Bourillet pour m’avoir donné la possibilité de travailler sur une<br />

de ses magnifiques carottes (IFREMER-Brest), dans le cadre d’un contrat de travail de<br />

3 mois lequel à contribuer partiellement à financer ma dernière année de thèse. Je<br />

vous remercie aussi pour m’avoir intégré dans l’équipe d’étude de cette carotte et<br />

de m’avoir permis l’écriture d’un article scientifique.<br />

A Elsa Cortijo et Elisabeth Michel du Laboratoire des Sciences du Climat et de<br />

l’Environnement (LSCE), Gif-sur-Yvette, pour les analyses isotopiques et pour toujours<br />

avoir été disponibles à résoudre des questions scientifiques du moment.<br />

A Frauke Rostek et Edouard Bard (Cerege) pour les données des alkénones.<br />

A Vincent Marieu pour avoir perdu pas mal de temps à m’apprendre<br />

l’analyse spectrale et d’autres choses comme corriger mes fautes d’orthographe,<br />

etc. Je te remercie aussi pour toute ta gentillesse et ton amitié. Et je m’excuse<br />

d’avoir fumée quelques cigarettes dans notre bureau.


A Delphine Denis et Elodie Marchès pour avoir toujours été disponibles pour<br />

m’aider à corriger mon français catastrophique ainsi que pour m’avoir montré qu’il<br />

existe encore des étudiants passionnés et dédiés à la recherche tout en restant<br />

simples. Je vous remercie aussi pour votre amitié et pour m’avoir supporté dans les<br />

moments plus difficiles. Je te remercie aussi Delphine pour m’avoir beaucoup aidé<br />

les derniers jours de cauchemar………..et pour la nuit blanche.<br />

A Anne-Laure Danieu pour avoir aussi toujours été disponibles pour m’aider à<br />

corriger mon français principalement dans les derniers jours de cauchemar…. Je te<br />

remercie aussi pour les conversations partagées à la salle à café et pour les<br />

moments d’amitié partagés.<br />

A William Fletcher for helping me to improve my catastrophic and terrible<br />

English.<br />

Je tiens aussi à remercier mes deux autres collègues de Bureau : Marc,<br />

Vincent et Jonathan pour m’avoir supporté pendant quelques années et aussi pour<br />

leur amitié.<br />

Je tiens aussi à remercier Alexandra Coynel pour m’avoir écouté et partagé<br />

quelques moments très agréables.<br />

À Pauliana Valente por ter passado algumas horas a corrigir o meu português<br />

de actual emigrante em França.<br />

À Àurea Narciso pela sua disponibilidade e ajuda em relação aos problemas<br />

administrativos relacionados com a tese.<br />

Tenho ainda necessidade de agradecer a uma série de pessoas que me<br />

ajudaram, apoiaram há cerca de 4 anos (durante o meu mestrado) as quais<br />

contribuíram em muito para a decisão de iniciar este trabalho de Doutoramento,<br />

nomeadamente:<br />

A Luisa Santos da Universidade da Coruña, aos colegas do grupo DISEPLA:<br />

Rui Taborda, à Anabela Oliveira e João Cascalho e Francisco Fatela.<br />

Gostaria ainda de agradecer:<br />

Au UMR CNRS 5805 EPOC (Environnements et Paléoenvironnements<br />

Océaniques- Univérsité Bordeaux 1) pour m’avoir reçue pendant 4 ans et de m’avoir<br />

donné toutes les conditions favorables à la réalisation de mon travail de thèse.<br />

À Universidade de Lisboa por ter aceite a minha candidatura a<br />

Doutoramento em regime de co-tutela.


Ao IPIMAR- Instituto de Investigação das Pescas e do Mar por me ter acolhido<br />

desde o meu mestrado e início do meu Doutoramento.<br />

Gostaria de expressar ainda a minha maior gratidão às entidades e projectos<br />

que financiaram este trabalho de Doutoramento, nomeadamente:<br />

O suporte financeiro durante 3 anos do projecto -“Envi-Changes”. Late<br />

Quaternary Environmental Changes From Estuary and Shelf Sedimentary Record<br />

inserido no Programa Dinamizador das Ciências e Tecnologias do Mar<br />

(PDCTM/PP/MAR/15251/99) e no Programa de Apoio à Reforma dos Laboratórios de<br />

Estado da Fundação para a Ciência e a tecnologia (FCT), o qual foi coordenado<br />

pela Doutora Teresa Drago.<br />

À cooperação Luso-Françesa (ICCTI- IFREMER) por me ter financiado algumas<br />

missões entre Portugal e França.<br />

Ao programa “PESSOA” integrado no programa, de acções integradas<br />

franco-português por me ter financiado algumas missões entre Portugal e França.<br />

À embaixada de França em Portugal por me ter apoiado financeiramente 2<br />

meses no início do meu 4° ano de Doutoramento.<br />

O projecto CNRS-ECLIPSE “La variabilité climatique d’ordre millénaire de la<br />

dernière période glaciaire dans les moyennes latitudes de l’Atlantique Nord, d’après<br />

l’analyse pollinique d’une carotte marine profonde prélevée sur la marge ibérique »<br />

o qual me proporcionou uma « vacation » de um mês.<br />

O programa EuroCLIMATE-European Science Foundation.<br />

L’Agence National de la Recherche (ANR)<br />

Finalmente gostaria ainda de agradeçer:<br />

À minha avó Maria da Luz Naughton pelo seu apoio incondicional e pela<br />

força que toda a vida me deu.<br />

E a:<br />

Heikki Seppä and Fátima Abrantes for their presence as the principal jury<br />

members during my PhD defense and for their valuable comments which greatly<br />

improved this manuscript.


Resumo<br />

O principal objectivo deste trabalho é caracterizar a variabilidade climática<br />

que ocorreu nos últimos 30 000 anos, nas médias latitudes do Atlântico Norte. Este<br />

trabalho focaliza-se ainda no impacto da variabilidade climática Holocénica na<br />

evolução dos sistemas costeiros do noroeste de Portugal. Desta forma, foi<br />

efectuado um estudo multidisciplinar (ex: pólens, associações de foraminíferos<br />

planctónicos, δ 18 O de foraminíferos planctónicos e bentónicos e alcanonas) em<br />

duas sondagens marinhas profundas recolhidas no noroeste da margem Ibérica. Foi<br />

ainda efectuado um estudo “multi-proxy” em duas sondagens estuarinas do<br />

noroeste de Portugal assim como numa sondagem marinha pouco profunda<br />

recolhida na plataforma continental noroeste Francesa. Foi igualmente efectuada<br />

uma calibração do sinal polínico ao longo da margem Ibérica.<br />

Este estudo mostra que os eventos associados a fortes descargas de<br />

icebergues no Atlântico Norte, designados por eventos de Heinrich, são complexos<br />

e compostos essencialmente por duas fases climáticas distintas no noroeste da<br />

margem Ibérica e continente adjacente. A primeira fase é marcada por um<br />

arrefecimento extremo das condições atmosféricas no continente (declínio da<br />

floresta de Pinus) e oceânicas de superfície. Esta fase é ainda caracterizada no<br />

continente, por uma certa humidade a qual é representada pelo desenvolvimento<br />

de Ericaceae incluindo Calluna e pelo aumento da concentração polínica total. A<br />

segunda fase é marcada por condições menos frias tanto no oceano como no<br />

continente assim como por um aumento da aridez continental (desenvolvimento de<br />

plantas semi-desérticas). Foram propostos dois mecanismos, um oceânico<br />

(circulação oceânica) e um outro atmosférico (relacionado com a Oscilação Norte<br />

Atlântica), para explicar o tal padrão complexo deixado pelos eventos de Heinrich<br />

no noroeste da margem ibérica.<br />

O último máximo glaciar apresenta um sinal oposto entre as condições<br />

oceânicas de superfície (temperaturas relativamente elevadas) e a vegetação<br />

continental a qual foi dominada por formações abertas (temperaturas<br />

relativamente baixas). Este sinal oposto é explicado, por um lado, pela<br />

intensificação da “Meridional overturning circulation” (MOC) no Atlântico Norte<br />

(MOC), e por outro, pelo aumento do albedo, pelo forte contraste sazonal existente<br />

nas altas latitudes do Atlântico Norte, e pela diminuição da concentração de CO2


na atmosfera. A intensificação da MOC facilitou o aumento da temperatura<br />

oceânica, enquanto que os outros mecanismos propostos contribuíram para a<br />

manutenção das temperaturas frias no continente.<br />

Após o evento H1, o aumento da insolação de verão nas médias latitudes do<br />

Hemisfério Norte e a intensificação da MOC provocaram um aquecimento<br />

oceânico e continental (expansão da floresta decídua) o qual caracteriza o evento<br />

Bölling-Alleröd no noroeste da Península Ibérica. Há cerca de 14 000 anos cal BP,<br />

ocorreu um incidente extremamente quente o qual é marcado pela máxima<br />

expansão da floresta decídua (Quercus). Este episódio é contemporâneo do<br />

importante evento relacionado com uma subida rápida do nível do mar (Meltwater<br />

pulse 1A) e do pico máximo observado na temperatura da Gronelândia durante o<br />

Bölling-Alleröd. Isto sugere que este aquecimento extremo do Hemisfério Norte terá<br />

sido provavelmente o impulsionador da fusão drástica da calote glaciar canadiana<br />

“Laurentide ice sheet”.<br />

A diminuição da floresta decídua de Quercus e a expansão de plantas semi-<br />

desérticas caracterizam o evento frio e seco, Dryas recente, no noroeste da<br />

Península Ibérica. Estas condições foram favorecidas pela redução da intensidade<br />

da MOC e pelo predomínio de condições semelhantes ao modo positivo da NAO-<br />

like.<br />

O início do Holocénico (11 600-8 200 anos cal BP) é marcado por um<br />

aquecimento máximo no noroeste da margem Ibérica e continente adjacente e<br />

sugere um episódio máximo térmico (HTM) nesta região.<br />

Após o HTM, a diminuição gradual da floresta temperada sugere um<br />

arrefecimento progressivo o qual segue o padrão geral de diminuição da insolação<br />

de verão das médias latitudes do Hemisfério Norte.<br />

Superimposta a esta variabilidade climática orbital, um episódio<br />

caracterizado pela expansão de Corylus indica um aumento do contraste sazonal<br />

no noroeste de França e marca o início do evento multi-secular “8.6-8.0 ka”. No<br />

interior deste evento, o súbito declínio da floresta de Corylus marca o evento frio<br />

designado “8.2 Ka”. O forte contraste sazonal resulta dos episódios terminais de<br />

expulsão dos lagos de “Agassiz” e de “Ojibway” e da consequente redução<br />

gradual da MOC, enquanto que o súbito arrefecimento reflecte o momento de


máxima diminuição na intensidade da MOC a qual terá provocado um<br />

arrefecimento suplementar da temperatura na Europa e na Gronelândia.<br />

A melhoria climática característica do início do Holocénico permitiu a<br />

expansão da floresta de Quercus na bacia do Douro e favoreceu a subida gradual<br />

do nível do mar. Após 6 535 anos cal BP, o desenvolvimento de uma barreira<br />

cascalhenta na parte sul do estuário resultou da atenuação da subida do nível do<br />

mar e do aumento do hidrodinamismo do rio. A presença desta barreira contribuiu<br />

para a migração do canal principal do rio para norte e para o alargamento deste<br />

estuário.<br />

Résumé<br />

Le principal objectif de cette thèse est de caractériser la variabilité climatique<br />

des moyennes latitudes de l’Atlantique Nord qui a eu lieu durant les derniers 30 000<br />

ans. Cette thèse traite aussi de l’impact de la variabilité climatique sur l’évolution<br />

des systèmes côtiers au cours de l’Holocène. Pour cela, une étude multiproxy<br />

(pollen, assemblages de foraminifères planctoniques, δ 18 O benthique et<br />

planctonique, alcénones) à été réalisée à partir de deux carottes marines profondes<br />

et une estuarienne prélevées dans le nord-ouest de la marge Ibérique et d’une<br />

carotte marine de la plateforme continentale nord-ouest Française. Egalement, une<br />

calibration du signal pollinique des sédiments de la marge occidentale ibérique a<br />

été préalablement effectuée.<br />

Cette étude montre que les événements associés à la forte décharge<br />

d’icebergs dans l’Atlantique nord, événements d’Heinrich, sont complexes,<br />

composés de deux phases climatiques distinctes dans la marge Ibérique et sur le<br />

continent adjacent. La première phase est marquée par des températures marines<br />

et continentales (chute de la forêt de Pinus) particulièrement froides et une certaine<br />

humidité (développement des Ericaceae et augmentation de la concentration<br />

pollinique); la deuxième, par des conditions moins froides mais une sécheresse<br />

accrue (fréquences maximales des plantes semi-désertiques). Des mécanismes de<br />

forçages marins (circulation océanique) et atmosphériques (l’oscillation Nord<br />

Atlantique) ont été proposés pour expliquer ce scénario.


Le Dernier Maximum Glaciaire montre un découplage entre les températures<br />

des eaux de surface, qui sont relativement élevées dans cette région, et la réponse<br />

de la végétation qui est dominée par les formations ouvertes. Ce découplage est<br />

expliqué d’une part par l’intensification de la circulation méridienne Atlantique de<br />

renversement (MOC) observée par des études précédentes et, d’autre part, par<br />

l’augmentation de l’albédo, le fort contraste saisonnier et la chute de la<br />

concentration de CO2 atmosphérique, ces derniers permettant le maintien de<br />

températures froides sur le continent.<br />

L’augmentation de l’insolation d’été des moyennes latitudes de l’Hémisphère<br />

Nord et l’intensification de la MOC ont produit le réchauffement climatique du<br />

Bölling-Alleröd. Un incident extrêmement chaud dans les moyennes et hautes<br />

latitudes de l’Atlantique Nord vers 14 000 cal ans BP est synchrone d’un épisode<br />

important de montée du niveau marin global (Meltwater Pulse 1A). Cela suggère<br />

que ce réchauffement de l’Atlantique Nord pourrait être le principal responsable de<br />

la fonte drastique et soudaine de la calotte glaciaire de Laurentide.<br />

Ensuite, l’événement froid du Dryas récent, associé paradoxalement au<br />

maximum d’insolation estivale dans l’Hémisphère Nord, serait le résultat d’une<br />

réduction de l’intensité de la MOC et de la prédominance du mode positif de la<br />

NAO-like.<br />

Le début de l’Holocène est marqué par un fort réchauffement climatique<br />

jusqu’à 8 200 cal ans BP dans la marge Ibérique. Cette période caractérise le<br />

Maximum Thermique de l’Holocène (HTM) sur cette région.<br />

Après le HTM, un refroidissement atmosphérique à long-terme suit la<br />

diminution graduel de l’insolation d’été des moyennes latitudes.<br />

Surimposé à cette variabilité climatique orbitale, un épisode caractérisé par le<br />

développement de la forêt de Corylus indiquant un fort contraste saisonnier dans le<br />

nord-ouest de la France marque l’événement global pluriséculaire "8.6-8.0 ka". A<br />

l’intérieur de cet événement, la chute soudaine de cette forêt, marque le<br />

refroidissement du "8.2 ka". Le fort contraste saisonnier serait le résultat des épisodes<br />

terminaux de purges des lacs d’Agassiz et d’Ojibway et de la réduction graduelle de<br />

la MOC tandis que le refroidissement brusque serait lié à la phase ultime de<br />

réduction de la MOC provoquant une diminution supplémentaire des températures<br />

sur l’Europe.


Le réchauffement climatique du début de l’Holocène a permis le<br />

développement d’une chênaie caducifoliée dans le bassin du Douro au même<br />

temps que nous avons détecté l’augmentation graduelle du niveau marin dans son<br />

estuaire. Vers 6535 ans cal BP, le développement d’une barrière de gravier dans la<br />

partie sud de celui-ci serait le résultat de l’atténuation de la montée du niveau marin<br />

et l’augmentation de l’hydrodynamisme de la rivière. La présence de cette barrière<br />

a contribué à la migration vers le nord du chenal principal de la rivière, identifié<br />

précédemment par des études sismiques, et par conséquence, à l’élargissement de<br />

cet estuaire. Nous avons pu donner une date maximale pour cette migration qui a<br />

eu lieu après 6535 ans cal BP.


Agradecimentos<br />

Resumo<br />

Résumé<br />

ÍNDICE GERAL<br />

Índice de Figuras<br />

Índice de Tabelas<br />

Índice de Anexos<br />

Capítulo 1 | Introdução 1<br />

Introduction 6<br />

1. 1 Paleoclimatologia do quaternário recente, contexto e objectivos principais deste trabalho 12<br />

1. 1. 1 Variabilidade climática orbital 12<br />

1. 1. 2 Variabilidade climática milenar 15<br />

1. 1. 2. 1 Os eventos de Heinrich e os eventos de D-O 15<br />

1. 1. 2. 2 O início da deglaciação 31<br />

1. 1. 2. 3 Holocénico 36<br />

1. 2 Calibração da assinatura polínica marinha ao longo da Península Ibérica 40<br />

1. 3 Impacto da variabilidade climática na evolução dos sistemas costeiros 44<br />

1. 3. 1 Variações do nível médio do mar 44<br />

1. 3. 2 Evolução dos sistemas costeiros 45<br />

1. 4 Zona de estudo 47<br />

1. 4. 1 Margem Ibérica 47<br />

1. 4. 1. 1 Clima e vegetação actual 47<br />

1. 4. 1. 2 Oceanografia 49<br />

1. 4. 1. 3 Geomorfologia e dinâmica sedimentar actual 52<br />

1. 4. 2 Plataforma continental noroeste Francesa 57<br />

1. 4. 2. 1 Geomorfologia, oceanografia e sedimentação actual 57<br />

1. 4. 2. 2 Clima e vegetação 58<br />

1. 5 Material e Métodos 59<br />

1. 5. 1 Amostras superficiais 59<br />

1. 5. 2 Sondagens 62<br />

1. 5. 2. 1 Sondagens marinhas profundas 62<br />

1. 5. 2. 2 Sondagem marinha pouco profunda VK03-58Bis 64<br />

1. 5. 2. 3 Sondagens estuarinas 64<br />

1. 5. 3 Cronologia e datações 14C 66<br />

1. 5. 4 Indicadores paleoclimáticos 68<br />

1. 5. 4. 1 Variação do coberto vegetal e do clima continental 68<br />

1. 5. 4. 2 Indicadores paleoclimáticos marinhos 73<br />

1. 5. 4. 3 Variações do volume de gelo acumulado nos pólos 77<br />

Referências 77


Capítulo 2| Present-day and past (last 25 000 years) marine pollen signal off western Iberia 105<br />

Resumo 106<br />

Résumé 107<br />

Abstract 109<br />

2. 1 Introduction 111<br />

2. 2 Environmental Setting 112<br />

2. 2. 1 Study area and present-day vegetation and climate 112<br />

2. 2. 2 Oceanography 114<br />

2. 2. 3 Morphology and recent sedimentation 115<br />

2. 2. 3. 1 North-western Iberian margin 117<br />

2. 2. 3. 2 South-western Iberian margin 118<br />

2. 3 Material and methods 119<br />

2. 3. 1 Deep-sea cores: MD99-2331 and MD03-2697 119<br />

2. 3. 1. 1 Radiometric dating 119<br />

2. 3. 1. 2 Marine proxy analyses 121<br />

2. 3. 1. 3 Pollen analysis 122<br />

2. 3. 2 Modern pollen samples 122<br />

2. 4 Results and Discussion 124<br />

2. 4. 1 Present day pollen signature 124<br />

2. 4. 1. 1 Western Iberian terrestrial sites 124<br />

2. 4. 1. 2 Western Iberian estuarine and margin sites 126<br />

2. 4. 2 Present-day pollen transport patterns 128<br />

2. 4. 3 Climatic and vegetational response in western Iberia to North Atlantic climatic events over the<br />

last 25 000 years 129<br />

2. 4. 3. 1 Marine Isotopic Stage 2 132<br />

2. 4. 3. 1. 1 Heinrich events (H2 and H1) 132<br />

2. 4. 3. 1. 2 The LGM 137<br />

2. 4. 3. 2 Marine Isotopic Stage 1 137<br />

2. 4. 3. 2. 1 The Bölling-Allerød 137<br />

2. 4. 3. 2. 2 The Younger Dryas cold event 139<br />

2. 4. 3. 2. 3 The Holocene 139<br />

2. 5 Conclusions 142<br />

References 143<br />

Capítulo 3| New insights on the impact of Heinrich events and LGM in the mid-latitudes of the eastern<br />

North Atlantic and in the adjacent continent 153<br />

Resumo 154<br />

Résumé 155<br />

Abstract 157<br />

3. 1 Introduction 159<br />

3. 2 Environmental Setting 161


3. 3 Material and methods 162<br />

3. 3. 1 Chronostratigraphy 162<br />

3. 3. 2 Pollen analysis 165<br />

3. 3. 3 Marine proxy analysis 166<br />

3. 3. 3. 1 Isotopic analyses 166<br />

3. 3. 3. 2 Ice rafted detritus (IRD) 166<br />

3. 3. 3. 3 Planktonic foraminifer-derived SST<br />

166<br />

3. 3. 3. 4 Alkenone-derived SST 167<br />

3. 4 Results and discussion 168<br />

3. 4. 1 Long-term climate variability and the LGM period 178<br />

3. 4. 2 Heinrich events 173<br />

3. 4. 3 Possible mechanisms triggering the complex pattern signal of Heinrich events in and off northwestern<br />

Iberia 175<br />

3. 5 Conclusions 179<br />

References 180<br />

Capítulo 4| Climate variability during the last deglaciation in north-western Iberian margin and<br />

adjacent continent 189<br />

Resumo 190<br />

Résumé 192<br />

Abstract 193<br />

4. 1 Introduction 197<br />

4. 2 Environmental Setting 198<br />

4. 3 Material and methods 199<br />

4. 3. 1 Stratigraphy and age model 199<br />

4. 3. 2 Pollen analysis 201<br />

4. 3. 3 Marine proxy analyses 201<br />

4. 3. 3. 1 Ice rafted detritus (IRD) and planktonic foraminiferal assemblages<br />

201<br />

4. 3. 3. 2 Isotopic analyses 202<br />

4. 4 Vegetation and climate changes in north-western Iberia and adjacent margin during the last<br />

deglaciation 203<br />

4. 4. 1 The end of the Last Glacial maximum (LGM) 205<br />

4. 4. 2 The Heinrich 1 (H1) 205<br />

4. 4. 3 The Bölling-Alleröd (B-A) 208<br />

4. 4. 4 The Younger Dryas (YD) 209<br />

4. 4. 5 The Holocene 211<br />

4. 4. 5. 1 The 8.2 k yr event 212<br />

4. 5 Conclusion 213<br />

References 214


Capítulo 5| Long-term and millennial-scale climate variability in north-western France during the last 8<br />

850 years 221<br />

Resumo 222<br />

Résumé 223<br />

Abstract 224<br />

5. 1 Introduction 227<br />

5. 2 Environmental Setting 228<br />

5. 3 Material and methods 230<br />

5. 3. 1 Radiometric dating 230<br />

5. 3. 2 Pollen and dinocyst analyses 231<br />

5. 3. 3 Pollen-based quantitative climate reconstruction 232<br />

5. 4 Results 233<br />

5. 4. 1 Lithostratigraphy and age model 233<br />

5. 4. 2 Evolution of dinocyst assemblages 234<br />

5. 4. 3 Vegetation succession and quantitative climate reconstruction 234<br />

5. 5 Climate variability in north-western France 240<br />

5. 5. 1 Long-term cooling pattern and the Holocene thermal maximum 240<br />

5. 5. 2 Sub-orbital climate variability 243<br />

5. 5. 2. 1 The multi-centennial-scale climate cooling and the 8.2 ka event 244<br />

5. 5. 2. 2 Other possible millennial scale cooling episodes 247<br />

5. 6 Conclusions 247<br />

References 249<br />

Capítulo 6| HOLOC<strong>EN</strong>E CHANGES IN THE DOURO ESTUARY (NORTHWESTERN IBERIA) 257<br />

Resumo 258<br />

Résumé 258<br />

Abstract 259<br />

6. 1 Introduction 261<br />

6. 2 Environmental Setting 262<br />

6. 3 Material and methods 264<br />

6. 3. 1 Radiometric dating 265<br />

6. 3. 2 Sedimentological analyses 265<br />

6. 3. 3 Micropalaeontological analyses 266<br />

6. 4 Results and Discussion 267<br />

6. 4. 1 Chronology 267<br />

6. 4. 2 Holocene sedimentary processes in the Douro estuary 267<br />

6. 4. 3 Vegetation changes versus variations in pollen catchment area during the Holocene 270<br />

6. 4. 4 Geomorphological changes in the Douro estuary during the Holocene 275<br />

6. 5 Conclusions 277<br />

References 277


Capítulo 7| Conclusions and perspectives 283<br />

Conclusão 293<br />

Índice de Figuras<br />

Capítulo 1|<br />

Fig. I.1 | Variação dos parâmetros astronómicos da terra (excentricidade, obliquidade e precessão dos equinócios)<br />

e da sua resultante, a qual é designada por insolação (W.m -2), durante os últimos 400 000 anos (Berger, 1978). A<br />

insolação representa a quantidade de energia recebida no verão pela terra a 65° N. As listas cinzentas representam<br />

os ciclos interglaciários, os quais são intercalados por períodos glaciários, representados pelas listas brancas.<br />

Fig. I.2 | Parâmetros astronómicos da terra: a) excentricidade, b) obliquidade e c) precessão dos equinócios.<br />

Fig. I.3 | Variabilidade climática orbital nos registos marinhos, continentais e de gelo (adaptado de Tzedakis et al.,<br />

2003). A: curva contínua a negro representa a percentagem de árvores excluindo Juniperus e Pinus e a curva a<br />

tracejado representa a totalidade percentual de árvores; B: curva isotópica do oxigénio contido nas carapaças de<br />

foraminíferos planctónicos e bentónicos da sequência ODP 980 (McManus et al., 1999); C: representa o volume de<br />

gelo versus nível do mar, obtido a partir do δ 18O de foraminíferos bentónicos da sequência ODP 980 (McManus et<br />

al., 1999); D: teor em CO2 atmosférico contido na sondagem de gelo de Vostok (Petit et al., 1999); E: curva de<br />

insolação a 40°N e 65°N (Berger, 1978).<br />

Fig. I.4 | Curva de variação da composição isotópica do oxigénio contido na sondagem de gelo GRIP (Dansgaard<br />

et al., 1993). O valor Isotópico do oxigénio (δ 18O) contido no gelo representa indirectamente a temperatura<br />

atmosférica do momento no qual ocorreu a acumulação de gelo no pólo Norte (Johnsen et al., 1992). Os números<br />

de 1 a 19 representam os episódios quentes interestadiais. Na parte superior da figura estão representados os<br />

estádios isotópicos marinhos (MIS-Marine isotopic stage) os quais foram definidos a partir da curva de variação do<br />

δ 18O contido nas carapaças de foraminíferos bentónicos (Shackleton & Opdyke, 1973). O MIS 3 terminou há cerca<br />

de: a) 24 000 anos segundo Shackleton & Opdyke (1973) e Martinson et al. (1987) ou b) 29 000 anos segundo<br />

Voelker et al. (1998) e van Kreveld et al. (2000). Ao longo do último período glaciário, a variabilidade de D-O é mais<br />

proeminente durante o MIS 3 do que durante o MIS 2 (Voelker et al., 2002) e o MIS 4.<br />

Fig. I.5 | Localização de alguns dos registos paleoclimáticos citados no texto: GRIP (Dansgaard et al., 1993); Vostok<br />

(Blunier et al., 1998; Petit et al., 1999); Byrd (Blunier et al., 1998); V23-81 (Bond et al., 1993; Bond & Lotti, 1995);<br />

<strong>EN</strong>AM93-21 (Rasmussen et al., 1996; 1997); SU90-24 e SU90-16 (Elliot et al., 1998); SO82-5 (van Kreveld et al., 2000);<br />

NOAMP (Heinrich, 1988) (a sondagem marinha ODP609 localiza-se exactamente na posição da NOAMP, Bond et al.,<br />

1993); HU75-55, 56 e HU87-09 (Andrews & Tedesco, 1992); MD95-2002 (Grousset et al., 2000); PS2644 (Voelker et al.,<br />

1998); <strong>EN</strong>AM97-09 (Richter et al., 2001); SU90-38 (Cortijo et al., 1995); SU90-03 (Chapman & Shackleton,1998;<br />

Chapman & Maslin,1999; Chapman et al., 2000); A (margem Ibérica): D11957P; SO75-26KL; PO 28-1; PO 8-2; MD95-<br />

2042; SU81-18; MD95-2041; MD95-2040 e MD95-2039) (Lebreiro et al.,1996; Baas et al., 1997; 1998; Zahn et al., 1997;<br />

Abrantes et al., 1998; Bard et al., 2000; Thouveny et al., 2000; de Abreu et al., 2003); RC11-83 (Charles et al., 1996);<br />

TTN057-10/13/21 (Kanfoush et al., 2000). O rectângulo cinza claro representa a localização da cintura de Ruddiman.<br />

Fig. I.6 | Identificação dos eventos de Heinrich na sondagem ODP609 (Bond et al., 1993). Curva de variação: da %<br />

de foraminíferos planctónicos de origem polar (N. pachyderma sin.); de detritos provenientes das calotes glaciárias<br />

do Hemisfério Norte (IRD); de δ 18O contido nas carapaças de N. pachyderma sin. e cujos picos representam uma<br />

grande quantidade de fluxo de água fundida (Hemming, 2004).<br />

Fig. I.7 | Calotes glaciárias (Ruddiman, 2001).<br />

Fig. I.8 | Representação esquemática do padrão geral da circulação termohalina (Rahmstorf, 2002). MOC- Atlantic<br />

Meridional Overturning Circulation.<br />

Fig. I.9 | Distribuição global dos registos de D-O durante o MIS 3 (ver Voelker et al., 2002).<br />

Fig. I.10 | Localização das sondagens onde foi efectuado uma correlação directa oceano-continente: 8057B<br />

(Hooghiemstra et al., 1992), ODP 976 (Combourieu Nebout, et al., 1999; 2002), MD95-2042 (Sánchez Goñi et al., 2000;<br />

2002), SO75-6KL (Boessenkool et al., 2001), MD95-2039 (Roucoux et al., 2001; 2005), MD95-2043 (Sánchez Goñi et al.,<br />

2002) e SU81-18 (Turon et al., 2003). Delimitação das zonas biogeográficas (adaptado de Peinado & Rivas-Martinez,<br />

1987).<br />

Fig. I.11 | Alguns registos da variabilidade climática milenar durante o LGIT (adaptado de Goslar et al., 2000): Lago<br />

de Gosciaz na Polónia (Goslar et al., 1992); Bacia de Caríaco (Hughen et al., 1996); GRIP δ 18O (Dansgaard et al.,<br />

1993).


Fig. I.12 | Comparação dos registos de GRIP δ 18O (Gronelândia) (Dansgaard et al., 1993) com δ 18O das sondagens<br />

gelo da Antárctica: Taylor Dome e Byrd e com a variação de Deutério na sondagem de gelo Vostok (adaptado de<br />

Blunier et al., 1998). Esta correlação foi efectuada utilizando a curva de metano de cada um dos registos.<br />

Fig. I.13 | Localização geográfica de alguns dos registos polínicos continentais. a) quadrado A localiza as<br />

sequências de 1 a 5: 1- Laguna de la Roya (Allen et al., 1996), 2- Sanabria March (Allen et al., 1996), 3-Laguna de las<br />

Sanguijuelas (Muñoz Sobrino et al., 2004), 4-Lleguna (Muñoz Sobrino et al., 2004), 5- Pozo do Carballal (Muñoz<br />

Sobrino et al., 1997); b) os pontos 6 a 13 correspondem a: 6- Laguna Lucenza (Santos et al., 2000); 7- Lagoa Lucenza<br />

(Muñoz Sobrino et al., 2001); 8-Lago de Ajo (Allen et al., 1996); 9- Los Tornos (Peñalba, 1994); 10-Saldropo (Peñalba,<br />

1994); 11-Belate (Peñalba, 1994); 12- Atxuri (Peñalba, 1994); 13- Banyoles (Pérez-Obiol & Julià, 1994); c) quadrado B<br />

inclui as sequências 14 a 19: 14- Quintanar de la Sierra (Peñalba et al., 1997); 15- Sierra de Neila - Quintanar de la<br />

Sierra (Ruiz Zapata et al., 2002); 16- Hoyos de Iregua (Gil-Garcia et al., 2002); 17- Laguna Masegosa (Von<br />

Engelbrechten, 1998); 18- Laguna Negra (Von Engelbrechten, 1998); 19- Las Pardillas lake (Sánchez Goñi & Hannon,<br />

1999); 20- Padul (Pons & Reille, 1988); 21- Mougás (Gómez-Orellana et al., 1998); 22- Charco da Candieira (Van der<br />

Knaap & Van Leeuwen, 1995).<br />

Fig. I.14 | Curva de variação do nível global do mar (adaptado de Lambeck et al., 2002).<br />

Fig. I.15 | Localização das zonas biogeográficas (adaptado de Blanco Castro et al., 1997).<br />

Fig. I.16 | Esquema detalhado das principais correntes de superfície do Atlântico Norte: EG-corrente Este da<br />

Gronelândia, Ei-corrente este da Islândia, Gu-Gulf Stream, Ir-corrente de Irminger, La-corrente do Labrador, Nacorrente<br />

Norte Atlântica, Nc-corrente do Cap Norte, Ng-corrente da Noroega, Ni-corrente do Norte da Islândia, Pocorrente<br />

de Portugal, Sb-corrente de Spitsbergen, Wg-corrente Oeste da Gronelândia. Linhas negras representam as<br />

correntes relativamente quentes enquanto as linhas a tracejado representam correntes relativamente frias<br />

(adaptado de Dietrich et al., 1980).<br />

Fig. I.17 | Representação da localização dos centros de altas e baixas pressões durante o verão e o Inverno ao<br />

longo do Hemisfério Norte (adaptado de Hurrell & Dickson, 2004). As setas representam a direcção dos ventos<br />

dominantes.<br />

Fig. I.18 | Esquema das principais correntes oceânicas que circulam ao longo da margem Ibérica. PCS: Portugal<br />

Current system; <strong>EN</strong>ACWsp: Eastern North Atlantic Central Water de origem sub-polar; <strong>EN</strong>ACWst: Eastern North<br />

Atlantic Central Water de origem sub-tropical; MSW: Mediterranean Sea Water; LSW: Labrador Sea water; NADW:<br />

North Atlantic Deep Water (adaptado de Sprangers et al., 2004).<br />

Fig. I.19 | a) Morfologia da margem continental Ibérica: Canhões submarinos de Mugia (MC), do Porto (PC), de<br />

Aveiro (AC), da Nazaré (NC), de Cascais (CC), de Lisboa (LC), de Setúbal (SC) e de São Vicente (S.VC); montanhas<br />

submarinas de Tore (TS), do Porto (PS), Vasco da Gama (VDGS) e de Vigo (VS).<br />

Fig. I.20 | Morfologia da plataforma continental do noroeste da Península Ibérica (adaptado de Dias et al., 2002).<br />

Fig. I.21 | Morfologia da plataforma continental sudoeste portuguesa (adaptado de Araujo et al., 2002).<br />

Fig. I.22 | Localização do corpo lodoso “Grande Vasière” na plataforma continental Francesa.<br />

Fig. I.23 | Amostras sedimentares de superfície analisadas neste estudo (VIR-18, Ría de Vigo; Laquasup, Estuário do<br />

Douro; PO287-13-2G, Complexo silto-argiloso do Douro; CG11, Complexo silto-argiloso do Minho; MD99-2331, Talude<br />

continental ao largo de Vigo; MD04-2814 CQ, Talude continental ao largo do Porto; Barreiro, Estuário do Tejo; MD99-<br />

2332, Complexo silto-argiloso de Lisboa; FP8-1, Talude continental ao largo de Sines; e MD95-2042, Talude-planície<br />

abissal ao largo de Sines).<br />

Fig. I.24 | Localização das amostras de superfície e das sondagens estudadas ao longo do estuário do Douro. Os<br />

círculos brancos com pinta negra representam as amostras de superfície e os círculos pretos representam as<br />

sondagens.<br />

Fig. I.25 | Localização das sondagens utilizadas neste trabalho. Sondagens marinhas profundas: MD99-2331 e MD03-<br />

2697; sondagem marinha pouco profunda: VK03-58Bis; sondagens estuarinas: Core1 e Core1B.<br />

Capítulo 2|<br />

Fig. II.1 | Fig. 1- Study area. Dashed line divides the Atlantic and Mediterranean biogeographical zones (Blanco<br />

Castro et al., 1997). White circles with a dark point represent the top samples analysed in this study; white circles<br />

represent the modern samples from the European Pollen Database; white circles with a cross represent the studied


cores sites (MD03-2697 and MD99-2331); dark circles represent marine and terrestrial core sites used for comparison<br />

with our study. Continental sequences: a) Square A locates sequences 1 to 5: 1- Laguna de la Roya (Allen et al.,<br />

1996), 2- Sanabria March (Allen et al., 1996), 3-Laguna de las Sanguijuelas (Muñoz Sobrino et al., 2004), 4-Lleguna<br />

(Muñoz Sobrino et al., 2004), 5- Pozo do Carballal (Muñoz Sobrino et al., 1997); b) Sites 6 to 13 correspond to: 6-<br />

Laguna Lucenza (Santos et al., 2000); 7- Lagoa Lucenza (Muñoz Sobrino et al., 2001); 8-Lago de Ajo (Allen et al.,<br />

1996); 9- Los Tornos (Peñalba, 1994); 10-Saldropo (Peñalba, 1994); 11-Belate (Peñalba, 1994); 12- Atxuri (Peñalba,<br />

1994); 13- Banyoles (Pérez-Obiol and Julià, 1994); c) Square B includes sequences 14 to 19: 14- Quintanar de la Sierra<br />

(Peñalba et al., 1997); 15- Sierra de Neila - Quintanar de la Sierra (Ruiz Zapata et al., 2002); 16- Hoyos de Iregua (Gil-<br />

Garcia et al., 2002); 17- Laguna Masegosa (Von Engelbrechten, 1998); 18- Laguna Negra (Von Engelbrechten, 1998);<br />

19- Las Pardillas lake (Sánchez Goñi and Hannon, 1999); 20- Padul (Pons and Reille, 1988); 21- Mougás (Gómez-<br />

Orellana et al., 1998); 22- Charco da Candieira (Van der Knaap and Van Leeuwen, 1995). The marine cores<br />

represented on the map are: 8057 B (Hooghiemstra et al., 1992), SO75-6KL (Boessenkool et al., 2001), SU81-18 (Turon<br />

et al., 2003) and ODP 976 (Combourieu Nebout, et al., 1999; 2002) and MD95-2039 (Roucoux et al., 2001; 2005).<br />

Fig. II.2 | West to east scheme of the different water masses from the western Iberian margin (adapted from<br />

Sprangers et al., 2004). White circles with a dark point represent southward water flow and white circle with a cross<br />

represent northward water flow. PCS-Portugal Current System; <strong>EN</strong>ACW st-Eastern North Atlantic Central Water of<br />

subtropical origin; <strong>EN</strong>ACW sp-Eastern North Atlantic Central Water of subpolar origin; MSW-Mediterranean Sea Water;<br />

LSW-Labrador Sea Water; NADW-North Atlantic Deep Water.<br />

Fig. II.3 | a) Morphology of the Iberian margin. Location of the surface samples from b) north-western Iberian margin<br />

and c) south-western Iberian margin. White arrows indicate the present-day pattern of pollen dispersion in the<br />

western Iberian margin.<br />

Fig. II.4 | Pollen spectra from western Iberian modern samples. Total temperate and humid (Tot. Temp./Hum.) trees<br />

includes: Alnus, Betula, Corylus, deciduous Quercus and other temperate and humid species (Acer, Fagus, Fraxinus,<br />

Salix, Tilia, Ulmus, Hedera helix, Myrica and Vitis). Total mediterranean (Tot. Mediter.) plants includes: evergreen<br />

Quercus, Olea and Cistus. Taraxacum-type, Asteraceae, Poaceae, Ericaceae and Calluna represent the ubiquist<br />

group. Semi-desert plants include Ephedra, Chenopodiaceae and Artemisia. Climate parameters: Alt: Altitude; PP:<br />

Precipitation; MTCO: Mean temperature of the coldest month; MTWA: Mean temperature of the warmest month;<br />

TANN: Annual temperature.<br />

Fig. II.5 | Pollen assemblages of top samples from coastal and marine western Iberian sites (see also caption of Fig.<br />

II.4). TPC: total pollen concentration.<br />

Fig. II.6 | Galician margin composite record (MD99-2331 and MD03-2697 deep-sea cores). From the left to the right:<br />

corrected radiocarbon ages; marine proxies: δ 18O of G. bulloides, % N. pachyderma (s.), ice-rafted detritus (IRD),<br />

Marine and Greenland climatic events; % pollen taxa; pollen zones and chronostratigraphy. Pollen zones were<br />

established using qualitative and quantitative fluctuations of a minimum of 2 curves of ecologically important taxa<br />

(Pons and Reille, 1986). They are defined by the abbreviated name of the core (MD31 or MD97) followed by the<br />

number of the marine isotopic stage (1 or 2) and numbered from the bottom to the top (MD31-2-1 to MD31-2-5 and<br />

MD97-1-1 to MD97-1-6).<br />

Fig. II.7 | - Comparison between continental (Quintanar de la Sierra; Peñalba et al., 1997) and marine (MD99-2331<br />

and MD03-2697) pollen sequences.<br />

Capítulo 3|<br />

Fig. III.1 | Map showing MD99-2331 location and sites of the cores referred in the text: 1: MD95-2040 (Pailler and Bard,<br />

2002; de Abreu et al., 2003; Schönfeld et al., 2003; Narciso et al., 2006), 2: MD95-2039 (Thouveny et al., 2000; Roucoux<br />

et al., 2001; 2005; Schönfeld et al., 2003), 3: PO 28-1 (Abrantes et al., 1998), 4: D11957P (Lebreiro et al., 1996; 1997), 5:<br />

SO75-26KL (Zahn et al., 1997; Boessenkool et al., 2001), 6: PO 8-2 (Abrantes et al., 1998), 7: MD95-2042 (Cayre et al.,<br />

1999; Sánchez Goñi et al., 2000; 2002; Thouveny et al., 2000; Pailler and Bard, 2002), 8: SU81-18 (Bard et al., 2000;<br />

Turon et al., 2003); 9: ODP 976 (Combourieu et al., 2002), 10: SU90-03 (Chapman et al., 2000), 11: ESSCAMP-KS02<br />

(Loncaric et al., 1998; Zaragosi et al., 2001), 12: MD95-2002 (Grousset et al., 2000; Zaragosi et al., 2001), 13: AKS01<br />

(Zaragosi et al., 2001), 14: VM 23-81 (Bond and Lotti, 1995), 15: MD04-2845 (work in progress), 16: SU90-11 (Jullien et<br />

al., in press.), 17: MD03-2705 (Jullien et al., submitted), 18: OCE326-GGC5 (McManus et al., 2004).<br />

Fig. III.2 | Chronostratigraphy of the MD99-2331 record. AMS radiocarbon dates are represented by triangles while<br />

the calibrated ones are represented by squares. North Atlantic Heinrich events are delimited by both the age limits<br />

(not calibrated) proposed by Elliot et al. (2002) and by the results obtained from the multi-proxy study of the MD99-<br />

2331 record (see below). White triangles and squares reflect the rejected levels for the model age while the dark<br />

ones represent the accepted ages.<br />

Fig. III.3 | Comparison between long term trends of MD99-2331 record and Greenland temperatures (Sánchez Goñi<br />

et al., in prep.) during the Late MIS 3 and MIS 2 against age (cal yr BP). From the bottom to the top: MD99-2331


enthic δ 13C; percentages of Pinus, temperate and humid trees and Poaceae and Greenland temperatures. Dashed<br />

line represents the long term trend of each signature.<br />

Fig. III.4 | Multi-proxy results of MD99-2331 record. From bottom to top: ice-rafted detritus (IRD) concentrations,<br />

percentages of planktonic foraminifera associations (polar, sub-polar and warm), planktonic foraminifera-based<br />

winter and summer SST estimates, alkenone-based annual SST reconstruction, δ 18O of G. bulloides, percentages of<br />

temperate and humid trees, Pinus percentages and Greenland temperatures (Sánchez Goñi et al., in prep.). Grey<br />

lines represent the Heinrich events. Note: the age limits of H4 that are based on GISP2 chronology are slightly different<br />

from those estimated from the calibration using NGRIP.<br />

Fig. III.5 | Response of the north-western Iberia vegetation to the complex pattern of Heinrich events in the Iberian<br />

margin. From bottom to top: Heinrich events, percentages of Pinus, percentages of Calluna representing wet<br />

conditions, percentages of semi-desert plants reflecting continental dryness and total of pollen concentration.<br />

Dashed line separated wet from dryness conditions during H events.<br />

Fig. III.6 | Prevailing NAO negative conditions scheme (adapted from Wanner et al., 2001).<br />

Fig. III.7 | Prevailing NAO positive conditions scheme (adapted from Wanner et al., 2001).<br />

Capítulo 4|<br />

Fig. IV.1 | Study area. Location of deep-sea cores referred in the text: OCE326-GGC5 (McManus et al., 2004); MD95-<br />

2002 (Zaragosi et al., 2001; Auffret et al., 2002; Ménot et al., 2006) and MD99-2331 (Naughton et al., 2006; in prep.).<br />

Fig. IV.2 | Pollen diagram of MD03-2697 deep-sea core against depth. From left to right: calibrated ages and<br />

percentages of selected pollen taxa. The stratigraphy is based on previous work by (Naughton et al., 2006) where the<br />

Oldest Dryas represents the continental counterpart of Heinrich 1 event in the ocean.<br />

Fig. IV.3 | Multi-proxy record of MD03-2697 against calibrated ages. From bottom to top: percentages of selected<br />

pollen taxa (trees: Betula, Corylus, deciduous Quercus, Pinus; Calluna and semi-desert plants: Artemisia,<br />

Chenopodiaceae and Ephedra); Ice-rafted detritus (IRD); planktonic foraminifera associations; Sea Surface<br />

Temperature (SST) estimates; δ 18O of planktonic and benthic foraminifera and Greenland temperatures (Sánchez<br />

Goñi et al., in prep.).<br />

Fig. IV.4 | Long-term and small-scalle pattern of vegetation changes in north-western Iberia. From bottom to top:<br />

benthic foraminifera δ 18O; percentages of temperate trees includes (Acer, Alnus, Betula, Corylus, Cupressaceae,<br />

deciduous Quercus, Fagus, Fraxinus excelsior-type, Pinus, Quercus ilex-type, Salix, Tilia and Ulmus); percentages of<br />

Pinus, percentages of deciduous Quercus and summer insolation at 45° N (after Berger, 1978).<br />

Capítulo 5|<br />

Fig. V.1 | Location of shelf core VK03-58Bis; and deep-sea core MD99-2551 (Ellison et al., 2006).<br />

Fig. V.2 | Lithology and synthetic pollen diagram against depth (cm). From left to right: radiocarbon and calibrated<br />

ages; lithology (after Folliot, 2004) including T. communis level (represented by small shells); dinocyst percentages<br />

(Operculodnium centrocarpum; Total of Spiniferites and Lingulodinium machaerophorum); pollen diagram and<br />

pollen zones.<br />

Fig. V.3 | Pollen diagram and quantitative pollen-based climate estimates against depth (cm). From left to right:<br />

calibrated ages; selected pollen taxa from the synthetic pollen diagram (other deciduous trees include: Fraxinus<br />

excelsior-type, Tilia and Ulmus); climate parameters: PANN (mean annual precipitation); difference between the<br />

temperature of the warmest (MTWA) and the coldest (MTCO) months (seasonality) and TANN (mean annual<br />

temperatures). Dashed lines represent maxima (bold) and minima values and the dark line represents mean values.<br />

Grey dashed lines represent the tendency of each curve; pollen zones.<br />

Fig. V.4 | Correlation between vegetation changes, quantitative climate estimates, summer insolation at 45° N and<br />

precessional signal (after Berger, 1978) and δ 18O-isotope composition of the NorthGRIP ice-core (Johnsen et al.,<br />

2001) during the Holocene. Temperate and humid trees include: Acer, Alnus, Betula, Corylus, Cupressaceae,<br />

deciduous Quercus, Fagus, Fraxinus excelsior-type, Pinus, Quercus ilex-type, Salix, Tilia and Ulmus while<br />

Brassicaceae, Caryophyllaceae, Asteraceae (including Aster- and Anthemis- types) and Taraxacum-type,<br />

Cyperaceae, Ericaceae and Calluna, Plantago, Poaceae and semi-desert plants (including Chenopodiaceae,<br />

Artemisia and Ephedra) are integrated in the herbaceous plants association.<br />

Fig. V.5 | Correlation between selected pollen taxa, quantitative climate estimates (PANN, TANN, MTCO, MTWA and<br />

seasonality) and δ 18O-isotope composition of the NorthGRIP ice-core (Johnsen et al., 2001) during the Holocene. The


8.2 kyr event is represented by the dark grey bar which is superimposed to 8.6-8.0 kyr event represented by the light<br />

grey bar. Dark arrows indicate possible millennial-scale cooling events during the Holocene.<br />

Fig. V.6 | Present day and past marine biogeographical zones in the North-East Atlantic (adapted from Funder et al.,<br />

2002). Bold dashed lines represent the limits of the present-day marine biogeographical zones in the North-East<br />

Atlantic; Grey dashed lines represent: a) the northward displacement of the boreal southern limit during the early<br />

Eemian (Funder et al., 2002) and b) the southward displacement of the boreal southern limit during the during 8.6-8.0<br />

kyr event (this work).<br />

Capítulo 6|<br />

Fig. VI.1 | a) Douro estuary localisation in the Iberian Península. b) Core (1, 1B and 2) and surface sampling sites.<br />

Dark circles represent core sites and white circles surface samples sites. Dark lines represent the palaeoisobathic<br />

curves defined by Carvalho and Rosa (1988). Dashed line represents the ancient direction of the river main channel<br />

flow and bold dark line the present day river main channel flow. This palaeobathymetric map shows the<br />

palaeovalley of the Douro river.<br />

Fig. VI.2 | Lithlogy of the Douro sequence. Depth is represented in Ordnance Datum which is a coordinate system for<br />

heights above mean sea level (optometric heights). Five calibrated ages are also represented along the two cores.<br />

Fig. VI.3 | Lithology of the Douro sequence. Depth is represented in Ordnance Datum which is a coordinate system<br />

for heights above mean sea level (optometric heights). Five calibrated ages are also represented along the two<br />

cores.<br />

Fig. VI.4 | a) Grain roundness and b) effective sphericity of gravel pebbles plotted against particle size. Two black<br />

lines delimit the different roundness and sphericity averages defined by Dobbkins and Folk (1970), representing the<br />

average limit of grain characterising river and/or low and high-energy beach environments. Dark squares represent<br />

all the measures obtained and Circle represents the value means of all measures.<br />

Fig. VI.5| Pollen diagram. From the left to the right: lithology, calibrated ages, arboreal pollen, AP (total of arboreal<br />

pollen), Pinus, NAP (total of non-arboreal pollen), pollen of herbaceous plants and pollen zones.<br />

Fig. VI.6| LAQUASUP: surface samples sites and pollen percentages.<br />

Fig. VI.7| Conceptual model of the Holocene geomorphological evolution of the Douro estuary: a) between 10720<br />

and 6530 cal yr BP, b) from 6530 to 1500 cal yr BP and c) the last 1500 years.<br />

Índice de Tabelas<br />

Capítulo 1 |<br />

Tab. I.1 | Localização das amostras de superfície. Da esquerda para a direita encontra-se representado o nome das<br />

amostras, a profundidade na coluna sedimentar, a latitude, a longitude, a profundidade da coluna de água, o ano<br />

da colheita das amostras, o nome dos projectos científicos ou o nome das missões oceanográficas ou o nome das<br />

Instituições que forneceram as amostras.<br />

Capítulo 2|<br />

Tab. II.1| Radiocarbon ages of MD99-2331 and MD03-2697 deep-sea cores. a Not acceptable dating (bioturbated<br />

layers); b Radiocarbon dates too old (not used); c dates calibrated by matching conventional AMS 14C with calendar<br />

ages estimated for MD95-2042 deep-sea core by Bard et al. (2004).<br />

Tab. II.2| Location, water depth and year of sample sampling from coastal, shelf and slope sequences of the Iberian<br />

margin.<br />

Tab. II.3| Description of the pollen zones in the Galician margin composite core and respective chronostratigraphy.


Tab. II.4| Description of pollen zones from the well-dated reference sites of Quintanar de la Sierra (Peñalba 1994,<br />

Peñalba et al., 1997), Laguna de la Roya (Allen et al., 1996) and Padul (Pons and Reille, 1988).<br />

Tab. II.5| Holocene tree succession in north-western Iberia.<br />

Capítulo 3|<br />

Tab. III.1| Radiocarbon ages of MD99-2331 deep sea cores. Bold levels represent the accepted ages while not bold<br />

ones represent the rejected ages for the age model.<br />

Capítulo 4|<br />

Tab. IV.1| Radiocarbon ages of MD03-2697 deep-sea core and one level from the twin core (MD99-2331). a<br />

Radiocarbon dates too young or too old and b Not acceptable dating (bioturbated layers).<br />

Capítulo 5|<br />

Tab. V.1| Radiocarbon ages from VK03-58Bis and VK03-58 and VK03-59Bis shelf cores.<br />

Capítulo 6|<br />

Tab. VI.1 | Radiocarbon and calibrated dates from the site under study.<br />

Índice de Anexos<br />

Anexo A<br />

Climate variability of the last five isotopic interglacials: direct land-sea-ice correlation from the<br />

multiproxy analysis of north western Iberian margin deep-sea cores.<br />

S. Desprat, M.F., Sánchez Goñi, F., Naughton, J.-L., Turon, J. Duprat, B. Malaizé, E. Cortijo and J.-P.<br />

Peypouquet. in press in The climate of past interglacials. Elsevier publications.<br />

Paleoenvironmental evolution of estuarine systems during the last 14000 years – the case of Douro<br />

estuary (NW Portugal).<br />

T. Drago, C.Freitas, F.Rocha, M.Cachão, J.Moreno, F.Naughton, C.Fradique, F.Araújo, T.Silveira, A.Oliveira,<br />

J.Cascalho, F.Fatela. (in press). In press in Journal of Coastal Research, SI 39, 7 pp.


Capítulo 1 | Introdução<br />

Face à crescente inquietude relativa ao aquecimento actual do<br />

planeta, torna-se necessário compreender a dinâmica natural do clima no<br />

passado, assim como identificar as variações ambientais que dela<br />

resultaram.<br />

As variações climáticas são um fenómeno global as quais resultam da<br />

interacção entre a atmosfera, hidrosfera, criosfera, litosfera e biosfera. Estes<br />

diferentes componentes interagem por sua vez com o clima produzindo<br />

retroacções negativas ou positivas ou, pelo contrário, amplificam ou<br />

reduzem um dado sinal climático.<br />

Os mecanismos forçadores externos (os quais se encontram<br />

intimamente ligados à posição da terra em relação sol e à constante solar) e<br />

internos (circulação termohalina, mecanismo interno ao gelo, erupções<br />

vulcânicas, concentração de CO2, albedo, etc.....) são responsáveis pela<br />

frequência, duração e amplitude das variações climáticas.<br />

De forma a detectar e compreender a frequência, duração e<br />

amplitude, assim como os mecanismos responsáveis pela variabilidade<br />

climática natural, torna-se necessário reconstituir, para o passado, o impacto<br />

desta variabilidade nos cinco sub-sistemas climáticos, no passado, utilizando<br />

uma cronologia única. Esta reconstituição pode ser realizada através da<br />

correlação directa entre registos paleoclimáticos marinhos, continentais e de<br />

gelo.<br />

A correlação indirecta entre sequências sedimentares terrestres,<br />

marinhas e de gelo é no entanto difícil devido à falta de precisão existente<br />

entre a conexão dos diferentes modelos de idade. A maioria dos registos<br />

climáticos encontram-se dispersos geograficamente pelo mundo e<br />

apresentam modelos de idades distintos os quais se baseiam em diversos<br />

tipos de metodologias tais como: datações radiométricas ( 14 C, U/Th, etc),<br />

contagem das camadas de gelo anuais nas sondagens da Gronelândia, e<br />

ainda com base em modelos físicos de acumulação de gelo na Antárctica.<br />

De forma a contornar esta problemática, o estudo de sondagens<br />

marinhas ricas em pólen e esporos permite-nos obter informações sobre a<br />

história da vegetação, assim como do clima no continente adjacente. Este


tipo de sondagens englobam ainda uma série de indicadores<br />

paleoclimáticos marinhos que permitem estimar a temperatura e salinidade<br />

da massa oceânica de superfície (foraminíferos planctónicos,<br />

dinoflagelados, nanoplancton, diatomáceas, etc), as condições da massa<br />

oceânica de fundo (Mg/Ca, ostracodos, foraminíferos bentónicos, δ 13 C), a<br />

dinâmica dos icebergues e a instabilidade das calotes polares (sedimentos<br />

grosseiros) assim como o volume de gelo acumulado nos pólos (δ 18 O de<br />

foraminíferos bentónicos). A comparação directa entre os diferentes tipos de<br />

registos paleoclimáticos permite-nos identificar as variações climáticas que<br />

afectaram o continente e correlacioná-las directamente com a resposta dos<br />

outros componentes do sistema climático (oceano, gelo e atmosfera). É-nos<br />

possível ainda documentar eventuais não contemporaneidades entre a<br />

resposta dos vários sub-sistemas climáticos a uma dada variação climática,<br />

assim como compreender a frequência e a natureza dessas variações. A<br />

comparação entre dados e resultados obtidos a partir de simulações<br />

numéricas permite-nos ainda testar a validade das últimas assim como<br />

confirmar os mecanismos utilizados nessas mesmas estimativas.<br />

Nos últimos milhões de anos, o clima foi sujeito a alternâncias entre<br />

períodos glaciários e interglaciários, segundo ciclos de cerca de 100 000<br />

anos, induzidos por variações na insolação de verão. Super imposta a esta<br />

ciclicidade orbital, outras oscilações da ordem do milénio foram detectadas<br />

em sondagens de gelo da Gronelândia. Estas oscilações, ditas Dansgaard-<br />

Oeschger (D-O), são marcadas por uma alternância entre aquecimentos<br />

súbitos e arrefecimentos progressivos. Alguns dos episódios frios são<br />

considerados eventos extremos e resultam da forte descarga de icebergues<br />

no Atlântico Norte.<br />

Estes eventos, ditos Heinrich, foram identificados em várias sequências<br />

marinhas do Atlântico Norte, nomeadamente na cintura de Ruddiman entre<br />

40° e 55° N, e são caracterizados pela forte presença de grãos detríticos de<br />

dimensões superiores a 150 μm (designados por IRD - Ice Rafted Detritus),<br />

assim como por um sinal magnético bastante importante. Estas camadas de<br />

IRDs foram ainda detectadas fora da sua zona preferencial de acumulação,<br />

2<br />

F. Naughton, 2007


nomeadamente nas médias latitudes do Atlântico Norte, incluindo na<br />

margem Ibérica.<br />

Nos últimos 40 000 anos, quatro eventos de Heinrich (H),<br />

nomeadamente H4 (34.9-33.9 k anos 14 C BP), H3 (27.4-26.1 k anos 14 C BP), H2<br />

(22.1-20.4 k anos 14 C BP) e H1 (15.1-13.4 k 14 C anos BP) foram registados no<br />

Atlântico Norte. No entanto, estes eventos apresentam um padrão<br />

complexo, composto por duas fases distintas, no sudoeste da margem<br />

ibérica (Bard et al., 2000; Sánchez Goñi et al., 2000). A primeira fase é<br />

caracterizada por uma diminuição abrupta da temperatura das águas de<br />

superfície (SST-Sea Superfície Temperatura), uma fraca presença de IRDs e<br />

um clima relativamente húmido no continente adjacente, enquanto que a<br />

segunda fase é caracterizada pela deposição máxima de IRDs e por um<br />

clima árido no continente. Este padrão complexo não foi até à data<br />

documentado para o noroeste da Península Ibérica. Várias hipóteses foram<br />

propostas na tentativa de explicar este padrão complexo associado aos<br />

eventos de Heinrich (Bard et al., 2000; Sánchez Goñi et al., 2000; Abrantes et<br />

al., 1998). No entanto, várias questões permanecem ainda sem resposta.<br />

Por essas razões um dos principais objectivos deste trabalho seria<br />

descrever com precisão o impacto destes eventos extremos no noroeste da<br />

margem Ibérica e continente adjacente, assim como discutir sobre os<br />

eventuais mecanismos responsáveis pelo padrão complexo deixado pelos<br />

eventos de Heinrich nesta região.<br />

Neste trabalho, pretendemos ainda compreender as interacções<br />

existentes entre os diferentes sub-sistemas climáticos durante o último<br />

máximo glaciar (LGM - Last Glacial Maximum). Sabe-se que a vegetação<br />

respondeu contemporaneamente às variações de SST que caracterizam as<br />

oscilações de D-O durante a fase tardia do estádio isotópico marinho 3 (late<br />

MIS3) (Sánchez Goñi et al., 2000; 2002). A expansão da floresta temperada é<br />

geralmente associada a um aquecimento oceânico superficial, enquanto<br />

que a regressão desta floresta corresponde a um arrefecimento das<br />

condições oceânicas de superfície. Contudo, durante o período de máxima<br />

extensão das calotes glaciárias no Hemisfério Norte, observa-se uma<br />

assincronia entre a SST (elevada nas médias latitudes do Atlântico Norte<br />

oriental) (e.g Chapman et al., 2000; de Abreu et al., 2003; Morey et al., 2005),<br />

3<br />

F. Naughton, 2007


e a biosfera continental adjacente (dominada por uma paisagem aberta<br />

indicando temperaturas frias no continente) (Peyron et al., 1998). Por essas<br />

razões, o segundo objectivo deste trabalho foi então, de tentar<br />

compreender esta situação paradoxal.<br />

Pretendemos ainda identificar e compreender a resposta da<br />

vegetação das médias latitudes do Atlântico Norte aos eventos sucessivos<br />

que caracterizam a última deglaciação, nomeadamente: o final do LGM, o<br />

H1, o evento quente Bölling-Alleröd (interestadial 1 nas sondagens de gelo<br />

da Gronelândia) e o evento frio Dryas recente (estadial 1 nas sondagens de<br />

gelo da Gronelândia), assim como o aquecimento que caracteriza o início<br />

do Holocénico. A comparação detalhada entre sequências polínicas<br />

marinhas e continentais permitiu-nos correlacionar claramente a clássica<br />

estratigrafia continental com os eventos oceânicos detectados no Atlântico<br />

Norte e os eventos registados nas sondagens de gelo da Gronelândia.<br />

Finalmente, interessamo-nos em identificar e compreender a resposta<br />

da vegetação face à variabilidade climática orbital e sub-orbital que<br />

caracteriza o actual interglaciário (Holocénico).<br />

Pretendeu-se ainda detectar e definir o intervalo de tempo associado<br />

ao Máximo Térmico Holocénico (HTM) na Europa centroeste e sudoeste.<br />

De forma a alcançar os objectivos mencionados foram utilizadas três<br />

sondagens às quais foi aplicado um estudo multidisciplinar (indicadores<br />

paleoclimáticos continentais, marinhos e de volume de gelo acumulado nos<br />

pólos) de alta resolução temporal (~100 anos).<br />

Previamente a efectuar a correlação directa oceano-continente-gelo<br />

dessas sondagens, foi efectuada uma calibração do sinal polínico marinho<br />

ao longo da margem Ibérica. De facto, a comparação do sinal polínico<br />

actual de amostras colhidas na margem Ibérica (num perfil de sul para<br />

Norte) com amostras continentais (representantes das duas regiões<br />

biogeográficas principais: Mediterrânica a sul e Atlântica a norte)<br />

provenientes da base de dados Europeia (http:/www.imep-<br />

cnrs.com/pages/EPD.htm; Peyron et al., 1998; Barboni et al., 2004) permite-<br />

nos verificar:<br />

- se assinatura polínica marinha actual representa uma imagem<br />

integral da vegetação regional do continente adjacente, e;<br />

4<br />

F. Naughton, 2007


- se a assinatura polínica das amostras de superfície costeiras e<br />

marinhas do noroeste e sudoeste da Península Ibérica é semelhante aos<br />

espectros polínicos actuais representantes das regiões biogeográficas<br />

Atlântica e Mediterrânica, respectivamente.<br />

De forma a compreender melhor o sinal polínico ao longo desta<br />

margem foram ainda determinados os padrões de transporte e dispersão<br />

polínica, do continente para o mar aberto.<br />

As sondagens marinhas utilizadas ao longo deste trabalho, a MD99-<br />

2331 (42° 09’ 00 N, 09° 40’ 90 W; 2110 m de profundidade) e a MD03-2697 (42°<br />

09’ 59 N, 59° 42’ 10 W; 2164 m de profundidade), foram recolhidas no<br />

noroeste da Península Ibérica pelo navio oceanográfico Marion Dufresne<br />

durante as campanhas GINNA e GEOSCI<strong>EN</strong>CES. No entanto, a resolução<br />

temporal destas sondagens para o Holocénico é bastante fraca pelo que,<br />

foi utilizada uma terceira sondagem, a Vk03-58Bis, recolhida na plataforma<br />

continental noroeste francesa, na “Grande Vasière” (47°36’ N, 4°08’ W; 98 m<br />

de profundidade), a qual é constituída por 2.72 m de sedimento. Os<br />

principais indicadores climáticos utilizados neste trabalho foram: os grãos de<br />

pólen, os sedimentos grosseiros provenientes da fusão dos icebergues, os<br />

foraminíferos planctónicos e os isótopos de oxigénio de foraminíferos<br />

bentónicos.<br />

O último objectivo deste trabalho seria documentar e compreender o<br />

impacto da variabilidade climática Holocénica na evolução dos sistemas<br />

costeiros do norte de Portugal. Sabe-se que o clima tem uma forte influência<br />

na evolução dos sistemas costeiros. De facto, a variabilidade climática<br />

orbital induz oscilações no volume de gelo acumulado nos pólos que, por<br />

sua vez provocam modificações do nível do mar as quais vão agir na<br />

evolução dos sistemas costeiros. A relação entre as variações do nível do<br />

mar e a evolução dos sistemas costeiros foi aprofundada para a zona sul<br />

portuguesa (Freitas et al., 2002 ; 2003).<br />

De forma a compreender o impacto das variações climáticas e<br />

consequentes modificações do nível do mar na evolução geomorfológica<br />

do estuário do Douro (Norte de Portugal) durante o Holocénico, foi<br />

efectuado um estudo polínico e sedimentar em duas sondagens estuarinas<br />

5<br />

F. Naughton, 2007


colhidas no estuário do Douro. Neste trabalho foi ainda efectuada a<br />

distinção dos diferentes mecanismos responsáveis pela evolução deste<br />

sistema costeiro.<br />

INTRODUCTION<br />

Face à l’inquiétude croissante sur le réchauffement actuel de la<br />

planète dû sans doute au moins en partie aux activités humaines, il est<br />

nécessaire plus que jamais de connaître la dynamique naturelle du climat<br />

dans le passé et les modifications environnementales qui en découlent. La<br />

compréhension du système climatique naturel peut permettre de dissocier ce<br />

qui dans le changement climatique actuel est imputable à l’homme de ce<br />

qui est propre à la cyclicité naturelle.<br />

Le changement climatique est un phénomène global qui agît sur<br />

l’atmosphère, l’hydrosphère, notamment les océans, la cryosphère, en<br />

particulier les calottes de glace polaires, la lithosphère et la biosphère. Ces<br />

différents réservoirs agissent, à leur tour, sur le climat en produisant des<br />

rétroactions négatives ou, au contraire, amplifiant le signal climatique. Le<br />

forçage externe lié à la position de la Terre par rapport au Soleil et à la<br />

constante solaire ainsi que plusieurs forçages internes (circulation<br />

thermohaline, mécanique interne de la glace, éruptions volcaniques,<br />

concentration du CO2, albédo…) sont les responsables de la fréquence, la<br />

durée et l’amplitude du changement climatique. Pour comprendre ces trois<br />

paramètres et les mécanismes associés au changement climatique naturel il<br />

est donc nécessaire de reconstituer son impact sur ces cinq réservoirs dans le<br />

passé avec une chronologie commune.<br />

Cette reconstitution peut être réalisée par la corrélation des<br />

enregistrements paléoclimatiques marins, de glace et continentaux. Toutefois<br />

la corrélation des séquences terrestres avec les archives marines et de glace<br />

est difficile en raison du manque de précision des différents calages<br />

chronologiques d’une séquence à l’autre. En effet, ces séquences dispersées<br />

géographiquement, ont de plus des modèles d’âge différents, certains basés<br />

sur des datations radiométriques ( 14 C, U/Th…), d’autres sur le comptage des<br />

6<br />

F. Naughton, 2007


couches de glace accumulées annuellement sur le Groenland et encore<br />

d’autres sur des modèles physiques d’accumulation de glace en<br />

Antarctique. Une façon de contourner ce problème est de travailler sur des<br />

carottes marines à sédimentation continue et non perturbée riches en pollen<br />

et spores qui nous renseignent sur l’histoire de la végétation et, en<br />

conséquence, du climat du proche continent. Ces carottes ont de plus<br />

l’avantage de renfermer des indicateurs paléoclimatiques provenant<br />

d’autres réservoirs permettant d’estimer les températures et la salinité des<br />

eaux de surface (foraminifères planctoniques, kystes de dinoflagellés,<br />

alcénones, nannoplancton, diatomées), les conditions des eaux du fond<br />

(Mg/Ca, ostracodes, foraminifères benthiques, δ 13 C), la dynamique des<br />

icebergs et l’instabilité des calottes polaires (sédiments grossiers) et le volume<br />

de glace stocké aux pôles (isotopes des foraminifères benthiques). La<br />

corrélation directe de ces différents types d’enregistrements<br />

paléoclimatiques permet d’une part d’identifier les changements climatiques<br />

qui ont affecté le continent et de les corréler directement avec la réponse<br />

d’autres composants du système climatique (océan, glace, atmosphère). On<br />

pourra ensuite documenter d’éventuels déphasages dans la réponse de ces<br />

réservoirs à un même changement climatique et enfin discuter de la<br />

fréquence et nature de ces changements. Une comparaison entre données<br />

et sorties des modèles permet en plus de tester la validité de ces dernières et,<br />

donc, la robustesse des mécanismes impliqués dans les différents modèles<br />

pour reproduire le changement climatique. En particulier, il est possible de<br />

tester le rôle possible des changements de la végétation sur le climat.<br />

Au cours du dernier million d’années, le climat a été rythmé par des<br />

alternances entre périodes glaciaires et interglaciaires, sur des cycles de<br />

100 000 ans environ, induites par les variations d’insolation. Surimposée à<br />

cette cyclicité orbitale de long terme, d’autres oscillations d’ordre millénaire<br />

ont été détectées, au cours de la dernière période glaciaire, dans des<br />

carottes de glace du Groenland. Ces oscillations, dites de Dansgaard-<br />

Oeschger (D-O), sont marquées par une alternance entre des<br />

réchauffements importants très rapides suivis de refroidissements progressifs.<br />

Certains de ces événements sont associés à des refroidissements extrêmes<br />

des eaux de surface de l’Atlantique Nord induites par l’introduction de<br />

7<br />

F. Naughton, 2007


grandes quantités d’eau de fonte provenant des importantes débâcles<br />

d’icebergs dans l’hémisphère Nord. Ces événements, dits d’Heinrich, sont<br />

identifiés dans les séquences marines de l’Atlantique Nord, et notamment<br />

dans la ceinture de Ruddiman (45°-55° N). Ils sont caractérisés par une forte<br />

accumulation des grains détritiques de dimension supérieure à 150 μm<br />

(désignés IRD – Ice Rafted Detritus), ainsi que par un signal magnétique<br />

robuste. Des couches d’IRD ont été détectées aussi en dehors de cette zone<br />

préférentielle de déposition, et en particulier dans des séquences marines<br />

des moyennes latitudes incluant celles de la marge ibérique.<br />

Pour la période qui nous intéresse, les derniers 40 000 ans, quatre<br />

événements d’Heinrich H4 (34.9-33.9 14 C k ans BP), H3 (27.4-26.1 14 C k ans BP),<br />

H2 (22.1-20.4 14 C k ans BP) et H1 (15.1-13.4 14 C k ans BP) on été enregistrés<br />

dans cette région.<br />

Les événements d’Heinrich montrent des scénarii complexes, avec<br />

deux phases distinctes enregistrées dans les carottes marines de la marge<br />

sud-ouest de l’Ibérie (Bard et al., 2000 ; Sánchez Goñi et al., 2000). La<br />

première phase correspond à une forte diminution de la température des<br />

eaux de surface (SST-Sea Surface Temperature), la presque absence d’IRD et<br />

un climat relativement humide sur le continent. La deuxième phase est<br />

caractérisée par le dépôt maximal d’IRD, un climat aride sur le continent<br />

adjacent tout en conservant des SST froides. Les changements associées à<br />

ces phases dans le nord-ouest de la Péninsule Ibérique n’ont pas été,<br />

toutefois, documentés. Quelques hypothèses ont été proposées pour<br />

expliquer ces scénarii complexes associés aux événements d’Heinrich.<br />

Néanmoins des questions restent en suspend.<br />

C’est pour cela qu’un des principaux objectifs de ce travail a été,<br />

dans un premier temps, de décrire avec précision l’impact que ces<br />

événements ont produit sur le nord-ouest de la marge ibérique et le continent<br />

adjacent. Ensuite, nous avons discuté des mécanismes associés à ces<br />

scénarii.<br />

D’autre part, ce travail s’est aussi intéressé à la compréhension des<br />

interactions entre les différents réservoirs terrestres pendant le dernier<br />

maximum glaciaire (LGM – Last Glacial Maximum). Généralement, la<br />

végétation a répondu de façon synchrone aux variations D-O des SST<br />

8<br />

F. Naughton, 2007


(Sánchez Goñi et al., 2000 ; 2002). L’expansion de la forêt caducifoliée est<br />

associée à des réchauffements océaniques importants, inversement le retrait<br />

de cette forêt correspond à un refroidissement de la surface marine.<br />

Cependant, durant la période d’extension maximale des calottes glaciaires<br />

dans l’hémisphère nord, nous avons observé un découplage entre les SST, qui<br />

sont élevées dans les moyennes latitudes de l’Atlantique Nord oriental (e.g<br />

Chapman et al., 2000 ; de Abreu et al., 2003 ; Morey et al., 2005), et la<br />

biosphère du continent adjacent où les paysages ouverts indiquent des<br />

températures froides (Peyron et al., 1998). Comprendre cette situation<br />

paradoxale a été le deuxième objectif de cette thèse.<br />

De plus, nous nous sommes aussi attachés à identifier et comprendre la<br />

réponse de la végétation des moyennes latitudes aux événements successifs<br />

qui caractérisent la dernière déglaciation : la fin du LGM, l’H1, l’événement<br />

chaud du Bölling-Alleröd (interstade 1 des carottes Groenlandaises), et<br />

l’événement froid du Dryas Récent (stade 1 des carottes Groenlandaises)<br />

ainsi que le réchauffement qui caractérise le début de l’Holocène. Une<br />

comparaison détaillée entre séquences polliniques marines et continentales<br />

nous a permis de corréler sans ambigüité la stratigraphie continentale<br />

classique avec les événements océaniques et groenlandais.<br />

Enfin, nous avons aussi considéré la variabilité climatique orbitale et<br />

sub-orbitale qui caractérise l’actuel interglaciaire (l’Holocène). Comment la<br />

végétation des moyennes latitudes a réagi t’elle aux forçages<br />

astronomiques ?, Quelle est sa réponse aux événements froids de l’Atlantique<br />

Nord (l’événement multi-centenaire 8.6-8.0 k ans et l’événement de 8.2 k<br />

ans), associés aux phases finales de purge des lacs d’Agassiz et d’Ojibway ?<br />

Pour atteindre les objectifs mentionnés ci-dessus, nécessitant d’une<br />

chronologie commune et fiable entre les différents réservoirs terrestres, nous<br />

avons utilisés des carottes marines riches en pollen. Forts de cette approche<br />

de corrélation directe entre indicateurs climatiques marins, terrestres et de<br />

glace, nous avons appliqué une analyse à très haute résolution (~100 ans)<br />

afin de détecter le plus grand nombre d’événements climatiques des<br />

derniers 40 000 ans.<br />

Avant d’effectuer la corrélation directe océan-continent-glace, nous<br />

avons tout d’abord calibré le signal pollinique marin de la zone d’étude.<br />

9<br />

F. Naughton, 2007


Nous avons confronté le signal pollinique actuel des échantillons de surface<br />

prélevée sur la marge (du Sud vers le Nord) avec ceux de la Péninsule<br />

Ibérique (représentant les deux principales régions biogéographiques<br />

Méditerranéenne au sud et Atlantique au nord) provenant de la base de<br />

données polliniques européenne (Peyron et al., 1998; Barboni et al., 2004;<br />

http:/www.imep-cnrs.com/pages/EPD.htm ), de façon à montrer que :<br />

- les assemblages polliniques de la marge Ibérique représentent une<br />

image intégrée de la végétation régionale qui colonise le continent<br />

adjacent ;<br />

- les communautés forestières présentes sur les deux zones<br />

biogéographiques pré-citées sont bien discriminées par les spectres<br />

polliniques marins du sud et du nord de la marge Ibérique, respectivement.<br />

Pour améliorer la compréhension du signal pollinique sur cette marge,<br />

les scénarii de dispersion pollinique actuelle et les principaux mécanismes<br />

responsables du transport sur cette région ont été approfondis.<br />

Les carottes marines utilisées, MD99-2331 (42° 09’ 00 N, 09° 40’ 90 W; 2110 m<br />

profondeur) et MD03-2697 (42° 09’ 59 N, 59° 42’ 10 W; 2164 m profondeur), ont<br />

été prélevées par le bateau océanographique Marion Dufresne en face de<br />

la Galice (nord-ouest de la Péninsule Ibérique) lors des campagnes GINNA et<br />

GEOSCI<strong>EN</strong>CES. Toutefois, la résolution faible de ces carottes pour étudier la<br />

variabilité climatique des derniers 10 000 ans (Holocène) nous a amené à<br />

travailler sur une autre carotte marine, cette fois-ci moins profonde, prélevée<br />

en face de Brest dans la Grande Vasière (nord-ouest de la France, 47°36’ N<br />

et 4°08’ W) et constituée de 2.72 m de sédiment.<br />

Les indicateurs climatiques (proxies) utilisés ont été principalement : le<br />

pollen, la fraction de sédiment grossier provenant de la fonte des icebergs,<br />

les foraminifères planctoniques et les isotopes de l’oxygène des foraminifères<br />

benthiques. Ces proxies vont documenter les variations dans les<br />

communautés végétales et des foraminifères indiquant la variabilité<br />

climatique sur le continent et l’océan, respectivement ; les périodes des<br />

débâcles et fonte d’icebergs et la variation du volume de la calotte polaire<br />

de l’Hémisphère Nord.<br />

10<br />

F. Naughton, 2007


Un but supplémentaire de notre thèse était de déceler l’impact des<br />

variations climatiques sur l’évolution des systèmes côtiers du nord-ouest de la<br />

marge Ibérique au cours des derniers 10 000 ans. Nous savons que le climat a<br />

une très forte influence sur l’évolution de ces systèmes. En particulier, la<br />

variabilité orbitale du climat est traduite par des variations du volume de<br />

glace accumulé aux pôles qui induisent des modifications du niveau marin<br />

avec un impact important sur l’évolution des systèmes côtiers. La relation<br />

entre les variations du niveau marin et l’évolution des systèmes côtiers de la<br />

marge occidentale ibérique a été étudiée pour le sud de la Péninsule<br />

ibérique (Freitas et al., 2002 ; 2003). Cependant, la réponse des systèmes<br />

côtiers du nord-ouest ibérique reste mal connue. Afin de comprendre<br />

l’évolution géomorphologique de cet environnement pendant l’Holocène<br />

nous avons réalisé les analyses pollinique et sédimentaire d’une carotte<br />

estuarienne prélevée dans l’estuaire du Douro. Dans ce travail, les<br />

mécanismes de forçage global lié aux variations du niveau marin sont<br />

distingués des mécanismes locaux qui ont agît sur l’évolution de ce système<br />

côtier.<br />

11<br />

F. Naughton, 2007


1. 1 Paleoclimatologia do quaternário recente, contexto e objectivos<br />

principais deste trabalho<br />

1. 1. 1 Variabilidade climática orbital<br />

Nos últimos milhões de anos o clima terrestre sofreu variações de<br />

longo-termo, entre condições glaciárias e interglaciárias, como resposta a<br />

modificações físicas que ocorreram entre o sol e a terra (Imbrie et al., 1992).<br />

Este mecanismo forçador externo, resulta da associação dos parâmetros<br />

astronómicos (excentricidade, obliquidade e precessão dos equinócios), que<br />

controlam a distribuição sazonal e latitudinal da energia proveniente do sol<br />

(Insolação) (Fig. I.1) (Berger & Loutre, 2004).<br />

Fig. I.1 | Variação dos parâmetros astronómicos da terra (excentricidade, obliquidade e precessão dos<br />

equinócios) e da sua resultante, a qual é designada por insolação (W.m -2), durante os últimos 400 000<br />

anos (Berger, 1978). A insolação representa a quantidade de energia recebida no verão pela terra a 65°<br />

N. As listas cinzentas representam os ciclos interglaciários, os quais são intercalados por períodos<br />

glaciários, representados pelas listas brancas.<br />

A órbita da terra à volta do sol varia de circular a elíptica, segundo<br />

ciclos próximos de 100 000 e 400 000 anos (Fig. I.2.a). O grau de<br />

achatamento da elipse em relação ao círculo caracteriza a excentricidade.<br />

O valor da excentricidade varia entre 0 a 0.05 (Berger & Loutre, 2004). Este<br />

pequeno valor da excentricidade da órbita terrestre afecta fracamente a<br />

quantidade de energia solar recebida pela terra anualmente.<br />

A obliquidade representa a relação entre o ângulo de inclinação do<br />

eixo da terra e a perpendicular ao plano da sua órbita (plano da eclíptica),<br />

a qual vai afectar a quantidade de energia recebida pela terra durante as<br />

estações do ano (Fig. I.2.b). O valor da obliquidade varia entre 22° e 25°<br />

segundo um ciclo de 41 000 anos (Berger & Loutre, 2004). O aumento da<br />

12<br />

F. Naughton, 2007


obliquidade provoca um aumento do contraste sazonal nas altas latitudes e<br />

em particular, invernos muito frios e verões muito quentes em ambos os<br />

hemisférios. Quando o valor da obliquidade decresce, o contraste sazonal<br />

diminui em ambos os hemisférios e, a presença de verões amenos e de<br />

invernos húmidos favorece o crescimento das calotes glaciárias nos pólos.<br />

A precessão dos equinócios resulta da combinação de dois<br />

movimentos de precessão: axial e de elipse (Fig. I.2.c). A precessão axial<br />

resulta da modificação da orientação do eixo de rotação da terra<br />

relativamente ao periélio e ao afélio, descrevendo uma figura cónica em<br />

redor de uma recta perpendicular ao plano da eclíptica, a qual é<br />

provocada pela força de atracção exercida pelo sol e lua à terra, a nível do<br />

equador (Ruddiman, 2001). A precessão da elipse resulta do movimento da<br />

rotação da terra sobre a órbita terrestre.<br />

A precessão dos equinócios ocorre segundo ciclos de 23 000 a 19 000<br />

anos (Ruddiman, 2001) sendo em média de 21 000 anos (Berger & Loutre,<br />

2004). A variação da precessão dos equinócios produz um forte contraste<br />

sazonal em ambos os hemisférios, onde verões quentes e invernos frios no<br />

Hemisfério Norte contrastam com verões frios e invernos quentes no<br />

Hemisfério Sul.<br />

Fig. I.2 | Parâmetros astronómicos da terra: a) excentricidade, b) obliquidade e c) precessão dos<br />

equinócios.<br />

Nos últimos 450 000 anos, a alternância entre ciclos glaciários e<br />

interglaciários foi dominada principalmente por variações da excentricidade<br />

da órbita terrestre (Fig. I.1). As variações climáticas de longo-termo,<br />

associadas à alternância entre os referidos ciclos glaciários e interglaciários,<br />

foram detectadas pela primeira vez por Shackleton & Opdyke em 1973, no<br />

registo de δ 18 O presente nas carapaças de foraminíferos bentónicos, numa<br />

13<br />

F. Naughton, 2007


sequência sedimentar marinha, confirmando pela primeira vez a teoria<br />

astronómica. Posteriormente, esta variabilidade orbital foi observada noutros<br />

registos marinhos, nomeadamente na sondagem ODP 980 (McManus et al.,<br />

1999), assim como noutro tipo de registos tais como: no CO2 incluso nas<br />

bolhas de ar da sondagem de gelo recolhida em Vostok, na Antárctica (Petit<br />

et al., 1999) (Fig. I.3 B, C, D). A variabilidade orbital foi ainda detectada pela<br />

primeira vez em sequências polínicas continentais nomeadamente, na<br />

Grécia (Van der Hammen et al., 1971). Por exemplo, a sequência de<br />

Tenaghi Philippon, evidência uma forte regressão da floresta durante os<br />

períodos glaciários e uma forte expansão da mesma ao longo dos períodos<br />

interglaciários (Fig. I.3 A). Existem ainda outros episódios de regressão da<br />

floresta durante os períodos interglaciários, associados a ciclicidades<br />

astronómicas de 20 000 a 40 000 anos (Fig. I.3 A, E).<br />

Fig. I.3 | Variabilidade climática orbital nos registos marinhos, continentais e de gelo (adaptado de<br />

Tzedakis et al., 2003). A: curva contínua a negro representa a percentagem de árvores excluindo<br />

Juniperus e Pinus e a curva a tracejado representa a totalidade percentual de árvores; B: curva isotópica<br />

do oxigénio contido nas carapaças de foraminíferos planctónicos e bentónicos da sequência ODP 980<br />

(McManus et al., 1999); C: representa o volume de gelo versus nível do mar, obtido a partir do δ 18O de<br />

foraminíferos bentónicos da sequência ODP 980 (McManus et al., 1999); D: teor em CO2 atmosférico<br />

contido na sondagem de gelo de Vostok (Petit et al., 1999); E: curva de insolação a 40°N e 65°N (Berger,<br />

1978).<br />

14<br />

F. Naughton, 2007


1. 1. 2 Variabilidade climática milenar<br />

1. 1. 2. 1 Os eventos de Heinrich e os eventos de D-O<br />

Sobreposta à variabilidade climática orbital, ocorreram uma série de<br />

flutuações rápidas, durante o último período glaciário (70 000-15 000 anos<br />

calendário BP), cuja periodicidade não pode ser explicada pela teoria<br />

orbital de Milankovitch. Esta variabilidade climática sub-orbital, ocorreu de<br />

forma cíclica todos os 1470-1500 anos (Bond et al., 1993; Bond et al., 1997;<br />

Mayeswski et al., 1997; Schulz et al., 2004) e é representada por uma<br />

alternância entre episódios de aquecimento abrupto, designados por<br />

interestadiais e episódios de arrefecimento gradual (estadiais), que<br />

ocorreram ao longo do último período glaciário (Dansgaard et al., 1993). Esta<br />

alternância entre episódios interestadiais (GIS-Greenland interstadials) e<br />

estadiais (GS-Greenland stadials) é designada por oscilações de Dansgaard-<br />

Oeschger (D-O). As oscilações de D-O foram detectadas pela primeira vez<br />

no registo isotópico do oxigénio (δ 18 O) do gelo, na sondagem GRIP<br />

(European Greenland Ice-core project) (Dansgaard et al., 1993) (Fig. I.4) (Fig.<br />

I.5). A variação da temperatura associada a esta oscilação de D-O chegou<br />

a atingir 16°C na Gronelândia (Severinghaus & Brook, 1999).<br />

Fig. I.4 | Curva de variação da composição isotópica do oxigénio contido na sondagem de gelo GRIP<br />

(Dansgaard et al., 1993). O valor Isotópico do oxigénio (δ 18O) contido no gelo representa indirectamente a<br />

temperatura atmosférica do momento no qual ocorreu a acumulação de gelo no pólo Norte (Johnsen et<br />

al., 1992). Os números de 1 a 19 representam os episódios quentes interestadiais. Na parte superior da<br />

figura estão representados os estádios isotópicos marinhos (MIS-Marine isotopic stage) os quais foram<br />

definidos a partir da curva de variação do δ 18O contido nas carapaças de foraminíferos bentónicos<br />

(Shackleton & Opdyke, 1973). O MIS 3 terminou há cerca de: a) 24 000 anos segundo Shackleton &<br />

Opdyke (1973) e Martinson et al. (1987) ou: b) 29 000 anos segundo Voelker et al. (1998) e van Kreveld et<br />

al. (2000). Durante o último período glaciário, a variabilidade de D-O é mais proeminente durante o MIS 3<br />

do que durante o MIS 2 (Voelker et al., 2002) e o MIS 4.<br />

15<br />

F. Naughton, 2007


Fig. I.5 | Localização de alguns dos registos paleoclimáticos citados no texto: GRIP (Dansgaard et al.,<br />

1993); Vostok (Blunier et al., 1998; Petit et al., 1999); Byrd (Blunier et al., 1998); V23-81 (Bond et al., 1993;<br />

Bond & Lotti, 1995); <strong>EN</strong>AM93-21 (Rasmussen et al., 1996; 1997); SU90-24 e SU90-16 (Elliot et al., 1998); SO82-<br />

5 (van Kreveld et al., 2000); NOAMP (Heinrich, 1988) (a sondagem marinha ODP609 localiza-se<br />

exactamente na posição da NOAMP, Bond et al., 1993); HU75-55, 56 e HU87-09 (Andrews & Tedesco,<br />

1992); MD95-2002 (Grousset et al., 2000); PS2644 (Voelker et al., 1998); <strong>EN</strong>AM97-09 (Richter et al., 2001);<br />

SU90-38 (Cortijo et al., 1995); SU90-03 (Chapman & Shackleton,1998; Chapman & Maslin,1999; Chapman et<br />

al., 2000); A (margem Ibérica): D11957P; SO75-26KL; PO 28-1; PO 8-2; MD95-2042; SU81-18; MD95-2041;<br />

MD95-2040 e MD95-2039) (Lebreiro et al.,1996; Baas et al., 1997; 1998; Zahn et al., 1997; Abrantes et al.,<br />

1998; Bard et al., 2000; Thouveny et al., 2000; de Abreu et al., 2003); RC11-83 (Charles et al., 1996); TTN057-<br />

10/13/21 (Kanfoush et al., 2000). O rectângulo cinza claro representa a localização da cintura de<br />

Ruddiman.<br />

A alternância entre fases de aquecimento abrupto e de<br />

arrefecimento gradual dos valores de temperatura da massa oceânica<br />

superficial foi detectada ainda, em várias sondagens marinhas colhidas no<br />

oceano Atlântico Norte, nomeadamente na V23-81 e ODP 609 (Fig. I.5)<br />

(Bond et al., 1993; Bond & Lotti, 1995).<br />

16<br />

F. Naughton, 2007


Os episódios frios, ditos estadiais (GS), estão especialmente bem<br />

representados nos registos obtidos em sondagens marinhas das altas<br />

latitudes do Atlântico Norte, pela presença de níveis sedimentares pouco<br />

espessos, ricos em material detrítico grosseiro, designado por IRD (Ice rafted<br />

detritus), proveniente da descarga de icebergues (Bond & Lotti., 1995; Elliot<br />

et al., 1998; van Kreveld et al., 2000).<br />

Alguns destes episódios estadiais são considerados como episódios<br />

extremos, e são conhecidos por eventos de Heinrich (H) (Broecker, 1994). Os<br />

eventos de H ocorreram ciclicamente todos os 5 000-10 000 anos (Elliot et al.,<br />

1998), e foram identificados pela primeira vez por Heinrich (1988), em<br />

sondagens marinhas colhidas entre 45 e 50° N, na chamada “cintura de<br />

Ruddiman” (Fig. I.5). Os níveis sedimentares representantes dos eventos de H<br />

são caracterizados pela abundância anómala de IRDs (Fig. I.6) provenientes<br />

das calotes glaciárias da “Laurentide”, Fenoescandinávia, Islândia e<br />

Britânico-Irlandesa (Fig. I.7) (Heinrich, 1988; Andrews & Tedesco, 1992;<br />

Broecker, 1994; Bond et al., 1992; Grousset et al., 1993, 2000; Bond & Lotti,<br />

1995; Elliot et al., 1998; Hemming et al., 1998; Scourse et al., 2000; Richter et al.,<br />

2001), assim como por um aumento dos valores de susceptibilidade<br />

magnética (MS) (Grousset et al., 1993). A presença de IRDs foi detectada,<br />

ainda, fora da cintura de Ruddiman, a norte de 50°N (Cortijo et al., 1995 ;<br />

Fronval et al., 1995; Rasmussen et al., 1996; 1997; Revel et al., 1996 ; Andrews<br />

et al., 1998; Elliot et al., 1998; Voelker et al., 1998; Van Kreveld et al., 2000),<br />

assim como nas médias latitudes do Atlântico Norte, a sul de 45° N (Lebreiro<br />

et al.,1996; Baas et al., 1997; 1998; Zahn et al., 1997; Abrantes et al., 1998; Bard<br />

et al., 2000; Thouveny et al., 2000; de Abreu et al., 2003; Chapman &<br />

Shackleton, 1998; Chapman & Maslin, 1999; Chapman et al., 2000) (Fig. I.5). A<br />

espessura das camadas de IRDs (IRD-layers) detectada nas sondagens<br />

marinhas das médias latitudes é, contudo, mais fina do que a espessura das<br />

camadas que caracterizam os eventos de Heinrich ao longo da cintura de<br />

Ruddiman. O mesmo se passa relativamente aos valores de MS, durante um<br />

evento de Heinrich: estes são elevados na cintura de Ruddiman e<br />

relativamente baixos nas latitudes médias do Atlântico Norte (Thouveny et<br />

al., 2000).<br />

17<br />

F. Naughton, 2007


Fig. I.6 | Identificação dos eventos de Heinrich na sondagem ODP609 (Bond et al., 1993). Curva de<br />

variação: da % de foraminíferos planctónicos de origem polar (N. pachyderma sin.); de detritos<br />

provenientes das calotes glaciárias do Hemisfério Norte (IRD); de δ 18O contido nas carapaças de N.<br />

pachyderma sin. e cujos picos representam uma grande quantidade de fluxo de água fundida (Hemming,<br />

2004).<br />

Fig. I.7 | Calotes glaciárias (Ruddiman, 2001).<br />

18<br />

F. Naughton, 2007


Apesar da espessura das camadas de IRDs e do sinal magnético<br />

serem menos importantes nas médias do que nas altas latitudes do Atlântico<br />

Norte, o impacto destes eventos extremos é evidente, principalmente no que<br />

se refere à diminuição brutal dos valores da temperatura da massa de água<br />

superficial (SST-Sea surface temperature) em ambas as regiões. Este drástico<br />

arrefecimento da massa de água superficial, resultou da introdução de<br />

grandes quantidades de água, proveniente da fusão dos icebergues no<br />

Atlântico Norte (Rosell-Melé et al., 1997; Bard et al., 2000; van Kreveld et al.,<br />

2000; Pailler & Bard, 2002). A diminuição da SST é particularmente bem<br />

representada no sinal isotópico do oxigénio incluso nas carapaças de<br />

foraminíferos planctónicos do tipo Globigerina bulloides (Maslin et al., 1995;<br />

Bond & Lotti, 1995; Bond et al., 1993; Cortijo et al., 1997; Chapman &<br />

Shackleton, 1998; Chapman & Maslin, 1999; Chapman et al., 2000;<br />

Shackleton et al., 2000a; Schönfeld et al., 2003), assim como, no aumento da<br />

abundância de foraminíferos planctónicos polares do tipo<br />

Neogloboquadrina pachyderma sinistrógira (Bond et al., 1992; Fronval et al.,<br />

1995; Cortijo et al., 1997; Lebreiro et al., 1997; Rasmussen et al., 1997; Cayre et<br />

al., 1999; de Abreu et al., 2003) (Fig. I.6).<br />

Para além da diminuição drástica da SST, a introdução de grandes<br />

quantidades de água doce, provocou uma diminuição dos valores de<br />

salinidade na massa de água oceânica de superfície (SSS-Sea surface<br />

salinity) (Vidal et al., 1997; 1999), produzindo mudanças paleoceanográficas<br />

muito importantes (Lehman & Keigwin, 1992; Maslin et al., 1995; Rahmstorf,<br />

1995; Cortijo et al., 1997; Vidal et al., 1997; Zahn et al., 1997; Chapman &<br />

Schackelton, 1998). A variabilidade dos parâmetros SST e SSS observada<br />

durante os eventos de Heinrich induziu a uma forte estratificação da coluna<br />

de água, afectando o padrão natural da circulação termohalina<br />

(circulação global oceânica, também conhecida por “conveyor belt”;<br />

Broecker, 1991) (Fig. I.8) e, em particular, da “Atlantic Meridional Overturning<br />

Circulation” (MOC) impedindo a transferência de calor das baixas para as<br />

altas latitudes, provocando um forte arrefecimento do Atlântico Norte<br />

(Fawcett et al., 1997).<br />

Para além da redução ou interrupção da MOC, a introdução de<br />

grandes quantidades de água doce impediu, chegando mesmo a bloquear<br />

19<br />

F. Naughton, 2007


totalmente, a formação da massa de água profunda, no Atlântico Norte<br />

(NADW-North Atlantic Deep Water) (Broecker, 1994; Keigwin & Lehman, 1994;<br />

Maslin et al., 1995; Oppo & Lehman, 1995; Vidal et al., 1997; Zahn et al., 1997;<br />

Seidov & Maslin, 1999; Ganopolski & Rahmstorf, 2001; Elliot et al., 2002). A<br />

formação da NADW pode ser ainda perturbada por variações na posição<br />

da frente polar, a qual impede o transporte da corrente superficial Norte<br />

Atlântica para as altas latitudes do Atlântico Norte, o consequente<br />

arrefecimento da massa de água superficial, o mergulho da mesma para o<br />

oceano profundo e finalmente o retorno para sul dessa massa de água<br />

profunda (Broecker et al., 1990).<br />

Fig. I.8 | Representação esquemática do padrão geral da circulação termohalina (Rahmstorf, 2002).<br />

MOC- Atlantic Meridional Overturning Circulation.<br />

Nas últimas duas décadas, vários mecanismos têm sido propostos<br />

para explicar a ocorrência dos eventos de Heinrich (ver Hemming, 2004) no<br />

Atlântico Norte, nomeadamente:<br />

1- Instabilidade interna da calote glaciar da “Laurentide” (“binge-<br />

purge model”, MacAyeal, 1993; Verbitsky & Saltzaman, 1995; Clark et al.,<br />

1996; Papa et al., 2006). Este mecanismo pressupõe que o calor geotermal é<br />

o principal responsável pela descarga de icebergues no Atlântico Norte;<br />

2- Actividade repetitiva de fenómenos de “jökulhaup” no lago da Baía<br />

de Hudson (Johnson & Lauritzen, 1995). Esta actividade jökulhaups representa<br />

20<br />

F. Naughton, 2007


a ocorrência de sucessivos episódios drásticos de expulsão de água os quais<br />

estão directamente relacionados com variações do nível do lago;<br />

3- Crescimento de plataformas de gelo e consequente colapso no<br />

Mar de Labrador (Hulbe, 1997; Hulbe et al., 2004; Alley et al., 2006).<br />

A diminuição da SST, detectada nos vários registos do Atlântico Norte,<br />

durante os episódios estadiais de D-O é de menor amplitude do que aquela<br />

observada durante os episódios estadiais extremos (eventos de Heinrich).<br />

Contudo, estes arrefecimentos são, tal como durante os eventos de Heinrich,<br />

principalmente induzidos pela introdução de grandes quantidades de água<br />

doce no Atlântico Norte, as quais perturbaram a circulação termohalina<br />

durante o último período glaciário (van Kreveld et al., 2000; Boyle, 2000;<br />

Ganopolski & Rahmstorf, 2001; Elliot et al., 2002; Knutti et al., 2004). Para além<br />

da alteração no modo de funcionamento da circulação termohalina (THC),<br />

a formação da NADW foi igualmente afectada: sendo esta reduzida durante<br />

os episódios estadiais e semelhante à situação actual, durante os episódios<br />

interestadiais.<br />

O efeito produzido por tais modificações hidrológicas, que ocorreram<br />

no Atlântico Norte, foi transmitido globalmente de forma súbita e está bem<br />

representado em varios registos paleoclimáticos (Leuschner & Siroko, 2000;<br />

ver Voelker et al., 2002) (Fig. I.9). No entanto, esta interpretação associada a<br />

uma reorganização rápida entre o oceano e a atmosfera foi recentemente<br />

posta em questão por Wunsch (2006). Este autor propõe que mudanças do<br />

volume de gelo no pólo Norte teram sido responsáveis por variações da<br />

direcção dos ventos as quais afectaram por consequência o oceano.<br />

21<br />

F. Naughton, 2007


Fig. I.9 | Distribuição global dos registos de D-O durante o MIS 3 (ver Voelker et al., 2002).<br />

Algumas zonas, tais como o Oceano Atlântico Sul (Charles et al., 1996;<br />

Vidal et al., 1999; Kanfoush et al., 2000) e a Antárctica (cores de gelo Byrd e<br />

Vostok) (Blunier et al., 1998), apresentam um registo climático de ordem<br />

temporal milenar em anti-fase, quando comparado com os registos obtidos<br />

na Gronelândia e Atlântico Norte (Bender et al., 1994; Blunier et al., 1998;<br />

Blunier & Brook, 2001) (Fig. I.5). Para além deste registo em anti-fase, a<br />

variabilidade climática observada no Hemisfério Sul é bastante mais fraca<br />

em amplitude do que aquela registada na Gronelândia.<br />

A correlação anti-fásica entre os dois hemisférios é explicada pelo<br />

designado efeito de “thermal bipolar sea-saw”, o qual sugere que variações<br />

abruptas na intensidade da THC, causadas por modificações no “input” de<br />

água doce, afecta o clima nos pólos através de modificações associadas à<br />

transferência de calor meridional (Ganopolski & Rahmstorf, 2001; Knutti et al.,<br />

2004).<br />

Este efeito de “sea-saw” é um dos muitos mecanismos propostos para<br />

explicar a passagem abrupta de um evento frio (estadial ou de Heinrich) a<br />

um evento interestadial de D-O, no Hemisfério Norte (Knutti et al., 2004). Para<br />

além desta teoria, vários têm sido os mecanismos propostos na tentativa de<br />

explicar este fenómeno, nomeadamente:<br />

22<br />

F. Naughton, 2007


- concentração de sal durante a formação de gelo marinho nas zonas<br />

de convecção do Atlântico Norte (van Kreveld et al., 2000). Durante a<br />

acumulação de gelo marinho, o sal é rejeitado e introduzido no oceano<br />

provocando o aumento da intensidade da MOC (Atlantic Meridional<br />

Overturning). Este mecanismo foi confirmado recentemente através de um<br />

modelo numérico de complexidade intermédia (Earth system Model of<br />

Intermediate Complexity) efectuado por Wang et al. (2006).<br />

- “feedback” oceânico o qual envolve a corrente oceânica<br />

mediterrânea (MOW-Mediterranean Outflow Water) (Voelker et al., 2006).<br />

Este mecanismo sugere que o aumento na intensidade da MOW que<br />

ocorreu durante episódios de fraca intensidade da circulação termohalina<br />

(THC) levou à introdução de calor e sal, proveniente da MOW, nas massas<br />

de água intermédias durante os episódios frios o que terá levado a THC a<br />

mudar o seu modo de funcionamento;<br />

- mudanças de sazonalidade, nomeadamente variações severas na<br />

temperatura de inverno as quais são provocadas por variações na extensão<br />

de gelo marinho no Atlântico Norte (Denton et al., 2005). No inverno, durante<br />

a formação de grandes camadas de gelo marinho, o sal incorporado na<br />

água do mar é rejeitado e a sua incorporação nas massas de água<br />

intermédias favorece o restabelecimento da MOC.<br />

Para além desta problemática associada à passagem abrupta de<br />

condições estadiais a condições interestadiais, outras questões têm sido<br />

levantadas nomeadamente em relação ao tipo de mecanismos que agiram<br />

de forma a amplificar o sinal climático, numa dada região, durante os<br />

eventos de Heinrich, nomeadamente:<br />

- a subida do nível do mar, causada pela introdução de grandes<br />

quantidades de água doce no Atlântico Norte, durante os eventos de<br />

Heinrich, contribuíram para a destabilização das plataformas de gelo e das<br />

calotes glaciárias do Hemisfério Norte induzindo uma contínua descarga de<br />

23<br />

F. Naughton, 2007


icebergs amplificando o sinal destes eventos nas altas latitudes do Atlântico<br />

Norte (Flückiger et al., 2006);<br />

- mecanismo atmosférico: a presença de um elevado índice na<br />

oscilação Norte Atlântica (NAO-North Atlantic Oscillation) provocaria a<br />

migração dos centros de alta pressão móveis polares Escandinavo e<br />

Atlântico, para a Península Ibérica, contribuindo para uma amplificação das<br />

condições extremamente frias e áridas, durante os eventos de Heinrich, nas<br />

regiões situadas mais a oeste da zona mediterrânica as quais não têm<br />

ligação directa com o Atlântico Norte (Sánchez Goñi et al., 2002).<br />

Todos estes mecanismos citados anteriormente sugerem que as<br />

oscilações de D-O resultam de modificações que ocorreram inicialmente no<br />

Atlântico Norte e cuja resultante foi posteriormente transmitida globalmente.<br />

Contudo, nos últimos anos, outras teorias têm sido enunciadas na tentativa<br />

de explicar a variabilidade de D-O, as quais sugerem que esta variabilidade<br />

global envolveu outro tipo de mecanismos que ocorreram inicialmente na<br />

zona tropical, tais como: monções indianas e asiáticas (Leuschner & Sirocco,<br />

2000; Kudrass et al., 2001) e as modificações que ocorreram na zona do<br />

Pacífico equatorial associadas ao El Niño Southern Oscilation (<strong>EN</strong>SO) (Cane &<br />

Clement, 1999).<br />

Recentemente, Wunsch (2006), sugeriu ainda uma alternativa à teoria<br />

do “freshwater pulse mechanism” (associada à introdução de grandes<br />

quantidades de água no oceano), a qual sugere que as modificações da<br />

circulação oceânica resultam de variações na intensidade dos ventos e<br />

que, os eventos de D-O resultam de uma interacção entre o vento e a<br />

topografia das calotes glaciárias permitindo assim a transferência rápida do<br />

sinal climático a nível global.<br />

Contudo, até à data, os vários mecanismos propostos na tentativa de<br />

explicar a natureza quase cíclica dos eventos de D-O e de Heinrich não são<br />

conclusivos.<br />

24<br />

F. Naughton, 2007


Para além da impressão digital deixada por esta variabilidade<br />

climática nas sondagens de gelo da Gronelândia e nas sondagens marinhas<br />

do Atlântico Norte, estes eventos tiveram também um forte impacto nas<br />

variações do coberto vegetal, nomeadamente nas médias latitudes, uma<br />

vez que no passado esta zona nunca terá sido coberta por gelo.<br />

A correlação efectuada entre os vários registos continentais e os<br />

registos paleoclimáticos obtidos tanto em sequências marinhas como em<br />

sondagens de gelo não é directa, uma vez que cada um destes locais<br />

apresenta um modelo cronológico próprio. Por estas razões, o estudo de<br />

sondagens marinhas profundas ricas em conteúdo polínico, situadas nas<br />

proximidades do continente, é muito importante na compreensão da<br />

resposta da vegetação à variabilidade climática detectada no Atlântico<br />

Norte, uma vez que permite correlacionar directamente ambos os registos<br />

marinho e continental utilizando uma cronologia única (Groot & Groot, 1966;<br />

Balsam & Heusser, 1976; Heusser & Shackleton, 1979).<br />

De facto, a correlação directa dos diferentes tipos de registos<br />

paleoclimáticos permite-nos não só identificar a variabilidade climática que<br />

afectou o continente mas também correlacioná-la com os outros<br />

componentes do sistema climático (oceano-atmosfera-gelo).<br />

A margem ibérica (situada a latitudes médias do Atlântico Norte) é<br />

directamente influenciada pela variabilidade climática detectada no<br />

Atlântico Norte e Gronelândia, sendo por isso considerada uma zona<br />

preferencial para efectuar este tipo de correlação directa.<br />

Até à data, foram efectuadas várias correlações directas (oceano-<br />

continente ou oceano-continente-gelo) ao longo da margem ibérica, na<br />

tentativa de compreender como é que a vegetação respondeu à<br />

variabilidade climática que caracteriza o último período glaciário (o qual<br />

engloba os estádios isotópicos marinhos MIS 4, MIS 3 e MIS 2), nas latitudes<br />

médias do Atlântico nordeste, nomeadamente: durante o MIS 3 -<br />

Combourieu Nebout et al., 1999, 2002; Sánchez Goñi et al., 2000, 2002;<br />

Roucoux et al., 2001, 2005, e durante o MIS 2 - Hooghiemstra et al., 1992;<br />

Combourieu Nebout et al., 1999, 2002; Boessenkool et al., 2001; Roucoux et<br />

al., 2001, 2005; Turon et al., 2003 (Fig. I.10). Contudo, a resolução temporal<br />

25<br />

F. Naughton, 2007


utilizada para o estudo do MIS 2 nas várias sequências sedimentares é<br />

bastante baixa.<br />

Fig. I.10 | Localização das sondagens onde foi efectuado uma correlação directa oceano-continente:<br />

8057B (Hooghiemstra et al., 1992), ODP 976 (Combourieu Nebout, et al., 1999; 2002), MD95-2042 (Sánchez<br />

Goñi et al., 2000; 2002), SO75-6KL (Boessenkool et al., 2001), MD95-2039 (Roucoux et al., 2001; 2005), MD95-<br />

2043 (Sánchez Goñi et al., 2002) e SU81-18 (Turon et al., 2003). Delimitação das zonas biogeográficas<br />

(adaptado de Peinado & Rivas-Martinez, 1987).<br />

Para além da baixa resolução temporal, a maioria dos estudos foram<br />

efectuados em áreas adjacentes à zona biogeográfica mediterrânica<br />

(Hooghiemstra et al., 1992; Combourieu Nebout et al., 1999, 2002;<br />

Boessenkool et al., 2001; Turon et al., 2003) a qual é caracterizada por um<br />

período de seca estival (Fig. I.10). A única correlação directa oceano-<br />

continente-gelo elaborada nas proximidades da zona biogeográfica<br />

Eurosiberiana (numa zona de transição entre esta e a zona biogeográfica<br />

Mediterrânica), caracterizada por um clima mais húmido, foi efectuada por<br />

Roucoux et al. (2005), (Fig. I.10).<br />

A maioria destas correlações sugere que a vegetação respondeu de<br />

forma síncrona à variabilidade da SST ou seja, a diminuição da temperatura<br />

da massa de água superficial ocorreu contemporaneamente à diminuição<br />

26<br />

F. Naughton, 2007


da floresta decídua/mediterrânica e à forte expansão da vegetação semi-<br />

desértica no continente, enquanto que, durante os episódios quentes,<br />

ocorreu uma forte expansão da floresta mediterrânica e uma forte redução<br />

da vegetação semi-desértica.<br />

Alguns destes trabalhos (Boessenkool et al., 2001; Roucoux et al., 2005)<br />

delimitam os eventos de Heinrich baseados apenas na espessura da<br />

camada de IRD, omitindo o registo dos indicadores de SST o qual permite<br />

definir o intervalo completo destes episódios.<br />

Outros estudos, efectuados ao longo da margem Ibérica, detectaram<br />

um padrão sedimentar complexo durante os eventos de Heinrich<br />

nomeadamente, nas sondagens marinhas: PO 28-1 e a PO 8-2 (Abrantes et<br />

al., 1998), MD95-2039 (Thouveny et al., 2000; Schönfeld et al., 2003), MD95-<br />

2042 (Sánchez Goñi et al., 2000; Thouveny et al., 2000), SU81-18 (Bard et al.,<br />

2000), MD95-2040 (Schönfeld et al., 2003; Narciso et al., 2006).<br />

A conjugação dos dois indicadores de presença de icebergues ao<br />

longo da margem ibérica, IRD e susceptibilidade magnética, efectuada nas<br />

sondagens marinhas SU81-18, MD95-2039, MD95-2042, permitiu detectar a<br />

ocorrência de dois sub-episódios associados aos eventos de H2 e H1 (Bard et<br />

al., 2000; Thouveny et al., 2000). O sub-episódio mais antigo apresenta uma<br />

fraca quantidade de IRDs e um grande pico no sinal magnético, enquanto<br />

que o mais recente, é representado por uma grande quantidade de IRDs e<br />

valores elevados em MS.<br />

Contemporaneamente aos episódios complexos de deposição de<br />

IRDs, ocorreu uma forte diminuição da SST, sugerindo que, apesar da<br />

ausência deste material grosseiro, o impacto dos eventos de Heinrich é bem<br />

evidente ao longo da margem Ibérica (Lebreiro et al., 1997; Cayre et al.,<br />

1999; Bard et al., 2000; Pailler & Bard, 2002; Chapman et al., 2000; Shackleton<br />

et al., 2000a; de Abreu et al., 2003; Schönfeld et al., 2003). Apesar dos<br />

indicadores marinhos apresentarem um padrão complexo relativamente aos<br />

eventos de H2 e H1, na sondagem SU81-18, o continente adjacente é<br />

caracterizado por um sinal homogéneo representante de condições áridas<br />

(Turon et al., 2003). No entanto, a correlação directa continente-oceano de<br />

alta resolução temporal, efectuada na sondagem MD95-2042 (situada nas<br />

proximidades da SU81-18), sugere a presença de um padrão tri-fásico<br />

27<br />

F. Naughton, 2007


associado aos eventos de Heinrich durante o MIS 3 (Sánchez Goñi et al.,<br />

2000), evidente no sinal obtido tanto a partir dos indicadores marinhos como<br />

dos indicadores continentais. Cada um dos eventos de H5, H4 e H3 é<br />

inicialmente marcado por condições húmidas no continente, seguido por<br />

um aumento de aridez durante o episódio de máxima deposição de IRDs<br />

(provenientes das calotes glaciárias Canadianas) e finalmente, por uma<br />

nova fase húmida no final de cada um desses eventos.<br />

Várias hipóteses foram enunciadas na tentativa de explicar a<br />

ocorrência deste padrão complexo ao longo da margem Ibérica, durante os<br />

eventos de Heinrich, nomeadamente:<br />

a) a proveniência de IRD resulta de diferentes fontes (Bard et al., 2000,<br />

Thouveny et al., 2000);<br />

b) resulta de múltiplas descargas de icebergues das calotes glaciárias<br />

canadianas (Abrantes et al., 1998); ou<br />

c) migração da frente polar (Chapman et al., 2000).<br />

Contudo, nenhum destes mecanismos propostos permite explicar<br />

variações entre humidade e aridez detectadas no continente.<br />

Objectivo 1<br />

Por estas razões, um dos principais objectivos deste trabalho é,<br />

identificar e compreender a resposta da vegetação ao complexo sinal<br />

deixado pelos eventos de Heinrich, no noroeste da margem Ibérica, assim<br />

como de discutir sobre os eventuais mecanismos responsáveis por esta<br />

variabilidade, através da comparação entre dados e modelos numéricos.<br />

Especial atenção será dada aos eventos de Heinrich que ocorreram durante<br />

o MIS 2 (26 000-15 500 anos cal BP) (intervalo cronológico definido para o MIS<br />

2 por Shackleton & Opdyke. 1973), nomeadamente H2 e H1. De forma a<br />

alcançar estes objectivos, foi efectuada uma correlação directa (oceano-<br />

continente-gelo) de altíssima resolução temporal (< 200 anos) numa região<br />

directamente influenciada pelo clima das latitudes temperadas que se<br />

encontre fora da zona sub-tropical (Margem Ibérica).<br />

28<br />

F. Naughton, 2007


Os episódios de H2 e H1, são separados por um período associado a<br />

condições glaciárias plenas (máxima extensão de gelo nos pólos), entre<br />

24 300 e 18 500 anos cal BP, designado por último máximo glaciar (LGM-Last<br />

Glacial Maximum). Apesar do nível do mar ter atingido o seu valor mínimo,<br />

entre 30 000 e 20 000 anos cal BP, o LGM foi definido preferencialmente num<br />

período caracterizado por uma certa estabilidade climática, na qual não<br />

existem variações bruscas tais como eventos de Heinrich e de D-O (Mix et al.,<br />

2001).<br />

O LGM é considerado como um período chave na compreensão da<br />

sensibilidade a mudanças dos vários sistemas ambientais globais uma vez<br />

que o clima permaneceu relativamente estável, embora bastante diferente<br />

e oposto ao estado actual (condições interglaciárias) (Mix et al., 2001).<br />

Recentemente foram efectuadas, no âmbito do projecto MARGO<br />

(Multiproxy Approach for the Reconstruction of the Glacial Ocean surface),<br />

uma série de reconstruções das condições oceânicas superficiais (SST e<br />

extensão de gelo marinho) para o LGM, nomeadamente na zona do<br />

Atlântico e Mares do Norte. Contrariamente às reconstruções prévias,<br />

obtidas no âmbito do projecto CLIMAP (Climate Long-range Investigation,<br />

Mapping, And Prediction) (CLIMAP project members, 1981), as quais<br />

sugeriam a presença de uma grande extensão de coberto de gelo perene a<br />

norte do Atlântico Norte e Mares do Norte durante este período, o grupo<br />

Margo mostra que esta cobertura de gelo de mar seria bastante mais<br />

variável (Kucera et al., 2005; Meland et al., 2005). Esta variabilidade resulta<br />

essencialmente do forte contraste sazonal que caracteriza este período, ou<br />

seja: invernos frios que favorecem uma forte expansão de gelo marinho<br />

enquanto que verões quentes reduzem fortemente a área coberta por gelo<br />

no Atlântico e Mares do Norte (de Vernal et al., 2005a). Esta diferença entre<br />

condições perenes e sazonais tem implicações muito importantes no clima,<br />

nomeadamente no que se refere à quantidade de humidade fornecida às<br />

altas latitudes do Hemisfério Norte assim como na localização da zona<br />

preferencial de formação da NADW e ainda, na Intensidade e extensão da<br />

MOC (Meland et al., 2005).<br />

29<br />

F. Naughton, 2007


Nos últimos anos foram efectuadas várias reconstruções de SST ao<br />

longo da margem Ibérica as quais incluem o LGM. Estas reconstruções foram<br />

essencialmente baseadas na composição química de alcanonas<br />

“alkenones” (Bard et al., 2000; Pailler & Bard, 2002) e na aplicação de<br />

funções de transferência às associações de foraminíferos planctónicos<br />

(Lebreiro et al., 1997; Cayre et al., 1999; de Abreu et al., 2003). Os resultados<br />

obtidos nessas reconstituições mostram que a temperatura média anual seria<br />

igual ou superior a 13°C; a de verão seria igual ou superior a 15°C e a<br />

temperatura de inverno situada entre 12 a 14°C, ao longo do LGM.<br />

Para além do aquecimento da massa de água superficial durante o<br />

LGM, algumas das reconstruções obtidas pelo grupo MARGO, mostram uma<br />

gradual diminuição da SST anual segundo um perfil latitudinal de sul a norte<br />

nomeadamente: cerca de 18°C na extremidade sudoeste Portuguesa, de<br />

16°C à latitude de Lisboa, de 14°C à latitude do Porto e finalmente de 12°C<br />

na extremidade noroeste da Península Ibérica (Morey et al., 2005).<br />

Baseado no trabalho desenvolvido por de Van Campo (1984),<br />

Sánchez Goñi (2006) sugeriu que o desenvolvimento da floresta temperada,<br />

durante os interestadiais de D-O que ocorreram ao longo do MIS 3, está<br />

intimamente associada a valores de SST de verão iguais ou superiores a 12°C.<br />

De facto, o aumento da SST de verão durante o LGM favoreceu o<br />

restabelecimento da “Meridional Overturnig Circulation” (MOC) permitindo a<br />

transferência de calor e humidade para o continente europeu, incluindo a<br />

margem Ibérica.<br />

Desta forma, e como resposta à variação da SST de verão estimada<br />

pelos trabalhos referidos anteriormente, onde os valores da SST de verão são<br />

da ordem dos 15°C, deveríamos esperar encontrar uma forte expansão da<br />

floresta temperada no noroeste da Península Ibérica.<br />

No entanto, o diagrama polínico obtido para a sondagem MD95-2039<br />

(Roucoux et al., 2005), mostra uma maior expansão de árvores temperadas<br />

durante os episódios interestadiais de D-O especialmente entre 65 000 e 30<br />

000 anos cal BP do que durante o LGM, apesar dos valores de SST serem<br />

semelhantes ao longo dos referidos episódios.<br />

30<br />

F. Naughton, 2007


A maioria dos trabalhos de correlação directa (oceano-continente e<br />

oceano-continente-gelo) efectuados ao longo da margem ibérica tem<br />

dado pouca importância ao período que caracteriza o último máximo<br />

glaciar (LGM).<br />

Objectivo 2<br />

Desta forma, pretende-se identificar a resposta da vegetação ao<br />

longo do período de máxima extensão de gelo nos pólos e relacioná-la com<br />

as variações das condições oceânicas de superfície que caracterizam o<br />

LGM. Para melhor compreender a relação entre a redução da floresta<br />

temperada do noroeste da Península Ibérica e as condições de SST da<br />

margem adjacente pretende-se ainda comparar os dados obtidos para o<br />

LGM com outros referentes ao interestadiais de D-O do tardi-MIS 3.<br />

Ambiciona-se ainda discutir sobre potenciais mecanismos responsáveis pela<br />

amplificação ou redução do sinal climático continental, nomeadamente o<br />

impacto do volume de gelo acumulado nos pólos e ainda as variações<br />

sazonais de expansão de gelo marinho nas altas latitudes do Atlântico Norte,<br />

nas latitudes médias.<br />

A variabilidade climática milenar, para além de ter sido detectada ao<br />

longo do último período glaciário (MIS 4, MIS 3 e MIS 2), foi ainda observada<br />

durante o MIS 1, o qual engloba o período de transição entre o último<br />

episódio glaciário e o actual interglaciário (LGIT-Last Glacial Interglacial<br />

transition) (15 500 – 11 500 cal anos BP) e o Holocénico (últimos 11 500 cal<br />

anos BP).<br />

1. 1. 2. 2 O início da deglaciação<br />

A fase de transição entre o último período glaciário e o actual<br />

interglaciário (LGIT: Last Glacial-Interglacial Transition) é caracterizada por<br />

um valor máximo da insolação de verão a 65° N e por uma forte redução do<br />

volume de gelo acumulado nos pólos. Apesar das condições orbitais<br />

favorecerem o início de um episódio interglaciário durante o LGIT, este<br />

período é, contudo marcado por uma série de oscilações climáticas<br />

complexas e abruptas.<br />

31<br />

F. Naughton, 2007


Esta variabilidade climática foi detectada pela primeira vez numa<br />

série de sequências polínicas continentais norte europeias e, os seus<br />

sucessivos eventos foram designados por: Oldest Dryas (frio), Bølling (quente),<br />

Older Dryas (frio), Allerød (quente) e Younger Dryas (frio) (Iversen, 1954;<br />

Mangerud et al., 1974). Na Península Ibérica, a qual se situa nas latitudes<br />

médias do Atlântico Norte, apenas três destes episódios foram detectados: o<br />

Oldest Dryas (Dryas Antigo), o Bølling-Allerød e o Younger Dryas (Dryas<br />

Recente) (Pons & Reille, 1988; de Beaulieu et al., 1994; Peñalba et al., 1997;<br />

Von Engelbrechten, 1998).<br />

Posteriormente, a variabilidade climática associada ao LGIT foi<br />

observada nas sondagens de gelo da Gronelândia e os seus eventos<br />

designados por: GS 2 (Greenland stadial 2); GIS 1 (Greenland interstadial 1) e<br />

GS 1 (Greenland stadial 1) (Alley et al., 1993; Dansgaard et al., 1993; Johnsen<br />

et al., 2001) e ainda numa série de sequências marinhas recolhidas no<br />

Atlântico Norte (Ruddiman & McIntyre, 1981; Lehman & Keigwin, 1992; Bond<br />

et al., 1993; Rasmussen et al., 1996). Estes registos marinhos do Atlântico Norte<br />

mostram um aumento abrupto na temperatura da massa de água<br />

superficial após o evento de H1 (Severinghaus & Brook, 1999), o qual<br />

permaneceu mais ou menos estável durante 2000 anos (Broecker, 2000). Este<br />

fenómeno está intimamente correlacionado com o evento “meltwater pulse<br />

1A” detectado no registo dos corais de Barbados (Bard et al., 1990a). O<br />

episódio quente oceânico associado ao evento continental Bølling-Allerød<br />

(B-A), foi seguido por um arrefecimento rápido e intenso (Ruddiman &<br />

McIntyre, 1981; Lehman & Keigwin, 1992; Bond et al., 1993; Rasmussen et al.,<br />

1996) que caracteriza o Dryas recente (Younger Dryas) no oceano. O Dryas<br />

recente é considerado como o incidente climático mais abrupto que<br />

ocorreu após o início da deglaciação (Broecker, 1994; 2000). O retorno a<br />

condições glaciárias plenas ocorreu em simultaneo com a diminuição das<br />

calotes glaciárias do Hemisfério Norte e, consequente introdução de<br />

grandes quantidades de água doce no oceano, a qual perturbou o padrão<br />

geral da circulação termohalina (Teller et al., 2002) assim como a formação<br />

de NADW.<br />

A mudança da direcção do fluxo de água doce, proveniente das<br />

fases finais da fusão da calote glaciar da “Laurentide”, do rio Mississippi para<br />

32<br />

F. Naughton, 2007


o rio St. Laurence, é tida como o mecanismo principal responsável para a<br />

ocorrência do Dryas recente no Atlântico Norte (Clark et al., 2001). Para<br />

além da introdução de água doce proveniente da calote glaciar<br />

canadiana no Atlântico Norte, a incorporação de água doce do lago de<br />

gelo Báltico na Escandinávia terá contribuído também para a destabilização<br />

da MOC durante este período (Nesje et al., 2004).<br />

Para além do Atlântico Norte e continente Europeu, o Dryas recente<br />

deixou indícios da sua presença nos mais variadíssimos registos climáticos<br />

mundiais, nomeadamente: na América do Norte (Mott et al., 1986; Peteet et<br />

al., 1990; 1993; Levesque et al., 1993; Peteet & Man, 1994), na América<br />

Central e Caraíbas (Hughen et al., 1996; 2000; Peterson., 2000), na América<br />

do Sul (van der Hammen & Hooghiemstra, 1995), em Africa (deMenocal.,<br />

2000), no Mar Sulu situado no Pacífico tropical (Linsley & Thunell, 1990;<br />

Rosenthal et al., 2003), no Pacífico noroeste (Kotilainen & Shackleton, 1995), a<br />

norte do Mar Arábico (Schulz et al., 1998) e ainda no sudeste Atlântico<br />

(situado a este da Bacia Angolana ~ a 17° S) (Kim et al., 2002) (Fig. I.11).<br />

Fig. I.11 | Alguns registos da variabilidade climática milenar durante o LGIT (adaptado de Goslar et al.,<br />

2000): Lago de Gosciaz na Polónia (Goslar et al., 1993); Bacia de Caríaco (Hughen et al., 1996); GRIP δ 18O<br />

(Dansgaard et al., 1993).<br />

33<br />

F. Naughton, 2007


Contudo e tal como acontece durante os eventos estadiais de D-O, o<br />

arrefecimento da Antártida (ACR-Antarctic Cold Reversal) e das altas<br />

latitudes do Hemisfério Sul ocorreu anteriormente ao arrefecimento que<br />

caracteriza o Dryas recente na zona tropical e Atlântico Norte (Bender et al.,<br />

1994; Sowers & Bender, 1995; Blunier et al., 1998; Blunier & Brook, 2001; Bianchi<br />

& Gersonde, 2004) (Fig. I.12). Este sinal anti-fásico é, tal como para os eventos<br />

de D-O, explicado pela teoria do “sea-saw”.<br />

Fig. I.12 | Comparação dos registos de GRIP δ 18O (Gronelândia) (Dansgaard et al., 1993) com δ 18O das<br />

sondagens gelo da Antárctica: Taylor Dome e Byrd e com a variação de Deutério na sondagem de gelo<br />

Vostok (adaptado de Blunier et al., 1998). Esta correlação foi efectuada utilizando a curva de metano de<br />

Vostok.<br />

Durante a última deglaciação, a fusão terminal dos glaciares<br />

continentais favoreceu a formação de numerosos lagos. Desta forma, os<br />

sedimentos lacustres foram considerados, durante muito tempo, como locais<br />

propícios ao estudo da variabilidade climática que caracteriza este período.<br />

Infelizmente as sequências lacustres raramente apresentam registos mais<br />

antigos que 15 000 anos.<br />

Ao longo de vários anos foram efectuados uma série de estudos<br />

polínicos em sequências lacustres de forma a detectar modificações no<br />

coberto vegetal durante este período e, em particular, na Península Ibérica<br />

(Fig. I.13).<br />

34<br />

F. Naughton, 2007


Fig. I.13 | Localização geográfica de alguns dos registos polínicos continentais. a) quadrado A localiza as<br />

sequências de 1 a 5: 1- Laguna de la Roya (Allen et al., 1996), 2- Sanabria March (Allen et al., 1996), 3-<br />

Laguna de las Sanguijuelas (Muñoz Sobrino et al., 2004), 4-Lleguna (Muñoz Sobrino et al., 2004), 5- Pozo do<br />

Carballal (Muñoz Sobrino et al., 1997); b) os pontos 6 a 13 correspondem a: 6- Laguna Lucenza (Santos et<br />

al., 2000); 7- Lagoa Lucenza (Muñoz Sobrino et al., 2001); 8-Lago de Ajo (Allen et al., 1996); 9- Los Tornos<br />

(Peñalba, 1994); 10-Saldropo (Peñalba, 1994); 11-Belate (Peñalba, 1994); 12- Atxuri (Peñalba, 1994); 13-<br />

Banyoles (Pérez-Obiol & Julià, 1994); c) quadrado B inclui as sequências 14 a 19: 14- Quintanar de la Sierra<br />

(Peñalba et al., 1997); 15- Sierra de Neila - Quintanar de la Sierra (Ruiz Zapata et al., 2002); 16- Hoyos de<br />

Iregua (Gil-Garcia et al., 2002); 17- Laguna Masegosa (Von Engelbrechten, 1998); 18- Laguna Negra (Von<br />

Engelbrechten, 1998); 19- Las Pardillas lake (Sánchez Goñi & Hannon, 1999); 20- Padul (Pons & Reille, 1988);<br />

21- Mougás (Gómez-Orellana et al., 1998); 22- Charco da Candieira (Van der Knaap & Van Leeuwen,<br />

1995).<br />

Contudo, e tal como foi referido anteriormente, a correlação<br />

efectuada entre os vários registos continentais e os registos paleoclimáticos<br />

obtidos tanto em sequências marinhas como em sondagens de gelo, é<br />

indirecta. Para além disso, algumas sequências continentais apresentam por<br />

vezes hiatos sedimentares importantes, os quais impedem a realização de<br />

um estudo detalhado sobre a deglaciação, sendo por isso, necessário<br />

recorrer a sondagens marinhas situadas nas proximidades do continente.<br />

Até à data, foram efectuados apenas dois estudos de correlação<br />

directa (oceano-continente) ao longo da margem ibérica, para o MIS 1<br />

35<br />

F. Naughton, 2007


(Boessenkool et al., 2001; Turon et al., 2003). Contudo, os resultados obtidos<br />

não foram discutidos em pormenor para este período de transição.<br />

Objectivo 3<br />

Por estas razões pretende-se detectar a variabilidade climática que<br />

ocorreu durante o LGIT no noroeste da Península Ibérica, nomeadamente:<br />

documentar as modificações do coberto vegetal contemporâneas da<br />

variabilidade climática detectada no Atlântico Norte e Gronelândia: fase<br />

final do evento de Heinrich 1, o Interestadial GIS 1 (episódio quente B-A),<br />

estádio GS 1 (Dryas recente) e o início do Holocénico.<br />

1. 1. 2. 3 Holocénico<br />

Durante muitos anos o actual interglaciário foi considerado como um<br />

período onde o clima permaneceu relativamente estável.<br />

No entanto, nos últimos anos, vários estudos efectuados em registos<br />

marinhos, continentais e de gelo, assim como outros dados de modelos<br />

numéricos, mostraram que o Holocénico foi afectado por uma variabilidade<br />

climática de longo termo induzida por modificações orbitais (Kutzbach &<br />

Gallimore, 1988; Koç & Jansen, 1994; Duplessy et al., 2001; Crucifix et al., 2002;<br />

Marchal et al., 2002; Weber & Oerlemans, 2003; Andersen et al., 2004a; Moros<br />

et al., 2004; Solignac et al., 2004; Keigwin et al., 2005; Renssen et al., 2005;<br />

Lorenz et al., 2006) e por uma variabilidade milenar (sub-orbital),<br />

superimposta à primeira, que ocorreu de forma cíclica todos os 1 000 a 1 500<br />

anos (Denton & Karlén, 1973; Koç et al., 1993; O’Brien et al., 1995; Bond et al.,<br />

1997; Campbell, et al., 1998; Bianchi & McCave; 1999; Calvo et al., 2002;<br />

Risebrobakken et al., 2003; Andersen et al., 2004b; Magny, 2004; Mayewski et<br />

al., 2004).<br />

A variabilidade climática holocénica, induzida por modificações dos<br />

parâmetros astronómicos da terra, é geralmente representada por uma subtil<br />

e gradual diminuição da temperatura nos mais variadíssimos registos<br />

sedimentares e de gelo do Hemisfério Norte (Johnsen et al., 2001, Marchal et<br />

al., 2002; Andersen et al., 2004a; Moros et al., 2004). Este padrão gradual de<br />

arrefecimento, foi ainda simulado a partir de modelos numéricos (Kutzbach<br />

& Gallimore, 1988; Crucifix et al., 2002; Weber & Oerlemans, 2003; Renssen et<br />

36<br />

F. Naughton, 2007


al., 2005), nos quais foram introduzidos os valores da insolação de verão a 65<br />

°N definidos para o Holocénico (Berger, 1978).<br />

A data de início deste arrefecimento é, contudo variável, e depende<br />

essencialmente da localização geográfica das zonas estudadas (Kaufman et<br />

al., 2004). De facto, o padrão de arrefecimento gradual tem início durante<br />

ou, após, a ocorrência do episódio térmico máximo do Holocénico (HTM-<br />

Holocene thermal maximum). Vários registos paleoclimáticos do Atlântico<br />

Norte, detectaram valores de temperatura máxima no início do Holocénico<br />

(11 500 - 8 000 anos cal BP) (Andrews & Giraudeau, 2002; Marchal et al; 2002;<br />

Duplessy et al., 2001; Kaufman et al., 2004; Knudsen et al., 2004; de Vernal et<br />

al., 2005b) enquanto outros, da mesma região e da Gronelândia,<br />

detectaram-no durante ou após 8 000 anos cal BP (no início do Holocénico<br />

médio) (Dahl-Jensen et al., 1998; Bauch et al., 2001; Johnsen et al., 2001;<br />

Levac et al., 2001; Kaplan et al., 2002; Solignac et al., 2004; Kaufman et al.,<br />

2004; Keigwin et al., 2005).<br />

A reconstrução das anomalias da temperatura atmosférica anual<br />

(TANN-annual mean temperature), de inverno (MTCO- mean temperature of<br />

the coldest month) e de verão (MTWA-mean temperature of the warmest<br />

month), efectuada em cerca de 500 sequências polínicas europeias,<br />

permitiram detectar igualmente a ocorrência do HTM, na Europa em geral,<br />

entre 8 000 e 4 000 anos cal BP (Davies et al., 2003). Este autor sugeriu que os<br />

valores da MTWA seriam superiores aos actuais, durante o HTM. O sudoeste<br />

da Europa, incluindo a Península Ibérica, apresenta contudo, valores de<br />

MTWA e MTCO bastante inferiores aos actuais durante o período situado<br />

entre 8 000 e 4 000 anos cal BP (Davies et al., 2003). O limite cronológico<br />

definido por Davies et al. (2003) para o HTM europeu é semelhante ao<br />

intervalo definido a partir de algumas sequências polínicas norte europeias<br />

tais como: nos lagos Raigastvere, Viitna, Ruila (Estónia) e Flarken (Suécia)<br />

(Seppä & Poska, 2004; Seppä et al., 2005).<br />

Para além do gradual decréscimo da temperatura induzido por<br />

modificações dos parâmetros orbitais, a variabilidade climática que ocorreu<br />

à escala milenar está bem expressa em variadíssimos registos climáticos tais<br />

como por exemplo (ver Mayewski et al., 2004):<br />

37<br />

F. Naughton, 2007


-O avanço e recuo de glaciares do norte da Europa e América<br />

(Denton & Karlén, 1973);<br />

- A presença de IRDs e diminuição da SST (de 1 a 3° C) durante os<br />

episódios frios em sondagens marinhas do Atlântico Norte (Bond et al., 1997;<br />

Solignac et al., 2004);<br />

2003);<br />

-A subida e descida dos níveis lacustres (ex: Alpes Suissos, Magny et al.,<br />

-As variações de K + na sondagem de gelo GISP2, as quais representam<br />

um aumento do fornecimento de poeiras à Gronelândia durante os<br />

episódios mais frios da variabilidade milenar (Mayewski et al., 1997);<br />

-As variações de cor nos sedimentos em sondagens marinhas<br />

recolhidas em “Caríaco Basin” (Hughen et al., 1996).<br />

Dentro desta variabilidade climática milenar, ocorreu um drástico<br />

episódio de arrefecimento, há cerca de 8 200 anos cal BP, o qual é<br />

designado por evento de 8.2 (8.2 kyr event). Este evento foi detectado em<br />

vários registos paleoclimáticos tais como: nas sondagens de gelo da<br />

Gronelândia (O’Brien et al., 1995; Alley et al., 1997; Muscheler et al., 2004),<br />

nas sondagens marinhas do Atlântico Norte (Bond et al., 1997; 2001; Bianchi<br />

& McCave; 1999) e no continente Europeu (Von Grafenstein et al., 1998;<br />

Klitgaard-Kristensen et al., 1998; Nesje & Dahl, 2001; Tinner & Lotter, 2001; 2006;<br />

Baldini et al., 2002; Spurk et al., 2002; Heiri et al., 2003; Magny et al., 2003;<br />

Seppä & Poska, 2004; Veski et al., 2004; Seppä et al., 2005).<br />

Os factores que provocaram este evento têm sido largamente<br />

debatidos nos últimos anos. Grande número de investigadores sugeriu que<br />

este evento teria sido provocado pelo rápido colapso da calote glaciar da<br />

“Laurentide”, que ocorreu há cerca de 8 470 anos (Barber et al., 1999). Este<br />

colapso favoreceu a introdução de grandes quantidades de água doce e<br />

fria no Atlântico Norte, destabilizando o padrão geral da circulação<br />

termohalina e a redução da formação de NADW (Alley et al., 1997; Clark et<br />

al., 2001). A diminuição da intensidade da circulação termohalina impediu o<br />

38<br />

F. Naughton, 2007


transporte de calor para as altas latitudes (Fawcett et al., 1997) provocando<br />

um forte arrefecimento no Atlântico Norte (Barber et al., 1999; Rahmstorf,<br />

2002; Renssen et al., 2001) e, como consequência, terá favorecido a<br />

formação de gelo marinho durante o inverno nas altas latitudes do Atlântico<br />

Norte (Denton et al., 2005). A expansão de gelo marinho amplificou o sinal<br />

de diminuição da temperatura durante o inverno, facilitando assim o<br />

aumento do contraste sazonal anual.<br />

Outros autores sugeriram que este evento foi provocado<br />

essencialmente por variações relacionadas com a actividade solar (Denton<br />

& Karlén, 1973; Bond et al., 2001; Van Geel et al., 2003). Esta hipótese<br />

pressupõe que as variações da actividade solar produzem flutuações no<br />

vento solar, o qual controla a intensidade dos raios cósmicos e a produção<br />

de 14 C na atmosfera e como tal nos valores de temperatura atmosférica.<br />

Contudo, o facto deste evento ser mais proeminente nos registos climáticos<br />

do Atlântico Norte do que noutras regiões, de ocorrer logo após o colapso<br />

da calote glaciária da “Laurentide”, e do facto de que a reconstrução<br />

baseada em modelos numéricos da anomalia que caracteriza este evento<br />

apresentar grandes semelhanças com o sinal emitido pelos vários<br />

indicadores paleoclimáticos (marinhos, continentais e de gelo), vai assim<br />

favorecer a primeira teoria a qual se baseia no designado “freshwater pulse<br />

mechanism” (Alley & Ágústsdóttir, 2005).<br />

Duas publicações recentes, Rohling & Pälike (2005) e Ellison et al.<br />

(2006), sugerem que o evento drástico e abrupto “8.2 kyr event”, ocorreu no<br />

interior de uma longa anomalia fria, de escala secular, entre 8 600 e 8 000<br />

anos cal BP. Este episódio de relativa longa duração (~ 600 anos) foi<br />

observado previamente em alguns trabalhos científicos, nomeadamente:<br />

no registo de K + da sondagem de gelo GISP2 (Mayeski et al., 1997); nos<br />

resultados de SST obtidos em algumas sondagens marinhas colhidas no<br />

Atlântico Norte (Risebrobakken et al., 2003; Knudsen et al., 2004; Keigwin et<br />

al., 2005) e ainda em sequências polínicas norte europeias (Seppä & Poska,<br />

2004). Contudo este evento foi correlacionado com o evento de 8.2 k anos.<br />

Até à data, poucos estudos foram efectuados de forma a demonstrar<br />

a resposta da vegetação às variações orbitais e sub-orbitais, no sul da<br />

Europa, sendo um exemplo o estudo efectuado por Magri (1995).<br />

39<br />

F. Naughton, 2007


Objectivos 4<br />

Por estas razões, um dos objectivos deste trabalho é testar se a<br />

vegetação das latitudes médias da Europa ocidental respondeu à<br />

variabilidade climática orbital e sub-orbital, e em particular aos eventos de<br />

8.2 k anos e multi-secular (8.6 -8.0 k anos). Pretendeu-se ainda definir o HTM<br />

nesta região e finalmente integrar os resultados e interpretações num<br />

contexto global.<br />

1. 2 Calibração da assinatura polínica marinha ao longo da Península<br />

Ibérica<br />

Tal como foi referido anteriormente, a correlação directa entre os<br />

vários indicadores paleoclimáticos: marinhos, continentais e de volume de<br />

gelo preservados em sedimentos marinhos, permite-nos avaliar o tempo e a<br />

natureza de resposta da vegetação às variações climáticas detectadas no<br />

Atlântico Norte.<br />

Contudo, antes de efectuar este tipo de correlações, deve-se verificar<br />

se os grãos de pólen preservados nas sequências marinhas representam uma<br />

imagem integral da vegetação continental adjacente, numa dada região<br />

em estudo.<br />

Nas últimas décadas, foram elaborados uma série de estudos<br />

palinológicos em amostras sedimentares de superfície, de forma a<br />

determinar os tipos de transporte polínico (do continente para o oceano), os<br />

padrões de dispersão polínica no oceano assim como o tipo de relação<br />

existente entre a assinatura polínica marinha e a sua fonte original, em vários<br />

locais do mundo, nomeadamente: noroeste Pacífico (Heusser & Florer, 1973;<br />

Heusser & Balsam, 1977); no Mar de Okhotsk (noroeste da Russia) (Koreneva,<br />

1966); no Mar do Japão (Koreneva, 1966); no Oceano ĺndico e nos Mares da<br />

Indonésia (van der Kaars, 2001; van der Kaars & de Deckker, 2003); no Mar<br />

mediterrânico (Rossignol, 1969; Koreneva, 1971); no Mar Negro (Koreneva,<br />

1966); no Mar da Noruega (Combaz et al., 1977); a Este do Canadá (Mudie,<br />

40<br />

F. Naughton, 2007


1982); na margem sudoeste francesa (Turon, 1984); no Rio Mississipi (Chmura<br />

et al., 1999); no noroeste da costa Africana (Rossignol-Strick & Duzer, 1979;<br />

Leroy & Dupont, 1994; Hooghiemstra et al., 2006); no Golfo da Guiné (Lézine<br />

& Vergnaud-Grazzini, 1993) e no sudoeste da costa Africana (Dupont &<br />

Wyputta, 2003). A maioria destes estudos mostram que os grãos de pólen<br />

contidos nos sedimentos marinhos dessas regiões representam uma imagem<br />

integral da vegetação regional do continente adjacente.<br />

Sabe-se que os grãos de pólen, após serem produzidos pelas plantas,<br />

são primeiramente dispersos pelo vento, depositados no solo, lagos e rios ou<br />

mantidos a pairar na atmosfera durante um período curto de tempo (Muller,<br />

1959; Cour et al., 1999). O meio de transporte polínico desde as zonas<br />

costeiras para o mar aberto depende essencialmente das condições<br />

ambientais de cada zona (Groot & Groot, 1966; Koreneva, 1966; Dupont et<br />

al., 2000):<br />

- em zonas áridas, com sistemas hidrológicos pouco desenvolvidos, tais<br />

como o noroeste da costa Africana (Leroy & Dupont, 1994; Hooghiemstra et<br />

al., 2006) e o Canadá oriental (Mudie & McCarthy, 1994), o transporte<br />

polínico do continente para o oceano, é essencialmente efectuado pelo<br />

vento;<br />

- em zonas áridas a semi-áridas, atravessadas por alguns sistemas<br />

fluviais tais como o Golfo da Guiné (Lézine & Vergnaud-Grazzini, 1993) e o<br />

Mar Alboran (Moreno et al., 2002), o tipo de transporte polínico é misto<br />

(fluvial e eólico);<br />

- em zonas costeiras com sistemas fluviais complexos e bacias<br />

hidrográficas bem desenvolvidas, os grãos de pólen são essencialmente<br />

transportados para o mar pelos rios (Muller, 1959; Cross et al., 1966; Bottema<br />

& Van Straaten, 1966; Heusser & Balsam, 1977; Peck, 1973).<br />

Vários trabalhos experimentais, efectuados desde a década de 60,<br />

mostraram que os grãos de pólen e esporos são considerados como<br />

partículas sedimentares finas, que quando suspensos na água comportam-se<br />

41<br />

F. Naughton, 2007


e obedecem às mesmas leis físicas dos sedimentos finos em suspensão<br />

(Muller, 1959; Koreneva, 1966; Traverse & Ginsburg, 1966; Stanley, 1966; Groot<br />

& Groot, 1966; Chmura & Eisma, 1995). Os grãos de pólen, tal como as<br />

partículas sedimentares finas, são ainda sujeitos a processos de<br />

aglomeração, floculação e pelitização, que favorecem o aumento da<br />

velocidade de sedimentação dos mesmos em meios marinhos (Chmura &<br />

Eisma, 1995; Chmura et al., 1999). Vários autores estimaram que a velocidade<br />

de sedimentação na coluna de água é cerca de 100 m/dia (Hogghiemstra<br />

et al., 1992).<br />

Posteriormente a serem libertados no mar pelos rios ou pelo vento, os<br />

grãos de pólen podem ser depositados na plataforma continental ou ser<br />

transportados através de correntes oceânicas para o mar profundo (Muller,<br />

1959; Heusser & Shackleton, 1979; Muller, 1959; Groot & Groot, 1966; Dupont<br />

et al., 2000). Estes podem ser transportados para distâncias longínquas à<br />

linha de costa tais como 1500 a 2500 Km (Muller, 1959). Contudo, alguns<br />

autores sugerem que, distâncias à linha de costa superiores a 500 Km<br />

apresentam geralmente um fraco conteúdo esporopolínico, e que o<br />

espectro polínico dessas sequências pode, por vezes, não representar uma<br />

imagem integral da vegetação do continente adjacente (Koreneva, 1966;<br />

Stanley, 1966).<br />

Em meio marinho, o conteúdo polínico diminui gradualmente à<br />

medida que nos distanciamos da actual linha de costa (em direcção ao<br />

largo), seguindo geralmente o mesmo padrão de dispersão das partículas<br />

sedimentares finas (Muller, 1959; Cross et al., 1966; Groot & Groot, 1966;<br />

Bottema & Van Straaten, 1966; Koreneva, 1966; Stanley, 1966; van der Kaars<br />

& de Deckker, 2003).<br />

Durante a última década, vários trabalhos de palinologia foram<br />

realizados em sondagens marinhas profundas ao longo da margem Ibérica<br />

de forma a compreender a relação entre as modificações do coberto<br />

vegetal a as variações climáticas detectadas no Atlântico Norte e<br />

Gronelândia (Hooghiemstra et al., 1992; Sánchez Goñi et al., 1999; 2000; 2002;<br />

2005; Boessenkool et al., 2001; Roucoux et al., 2001; 2005; Turon et al., 2003;<br />

Tzedakis et al., 2004; Desprat, 2005; Desprat et al., 2005; 2006; in press).<br />

42<br />

F. Naughton, 2007


Contudo, e até à data, não foram efectuados quaisquer tipo de<br />

estudos experimentais de forma a demonstrar que os grãos de pólen<br />

preservados nessas sequências marinhas representam uma imagem integral<br />

da vegetação da Península Ibérica, e estudos que permitam determinar os<br />

tipos de mecanismos envolvidos na transferência desses grãos, do<br />

continente para o oceano, nesta região.<br />

Objectivo 5 Pretende-se desta forma:<br />

- investigar se a assinatura polínica das amostras de superfície costeiras e<br />

marinhas do noroeste e sudoeste da Península Ibérica é semelhante aos<br />

espectros polínicos actuais representantes das regiões biogeográficas<br />

Atlântica e Mediterrânica, respectivamente;<br />

- mostrar que a assinatura polínica marinha actual representa uma imagem<br />

integral regional e não uma imagem local da vegetação do continente<br />

adjacente;<br />

- determinar os padrões de transporte e dispersão polínica, do continente<br />

para o mar aberto, a partir do estudo comparativo entre a concentração<br />

polínica total de cada amostra e modelos de dispersão sedimentar de<br />

partículas finas efectuados ao longo da margem ibérica.<br />

43<br />

F. Naughton, 2007


1. 3 Impacto da variabilidade climática na evolução dos sistemas<br />

costeiros<br />

As variações climáticas têm um papel muito importante no controlo<br />

das modificações do nível do mar global, as quais afectam profundamente<br />

os sistemas costeiros.<br />

1. 3. 1 Variações do nível médio do mar<br />

Tal como foi referido anteriormente, durante o Quaternário, o clima<br />

terrestre sofreu variações de grande e pequena escala induzidas por<br />

modificações internas e ou externas (insolação e a constante solar) que<br />

ocorreram no planeta. Estas variações estão associadas a mudanças mais<br />

ou menos importantes no volume de gelo acumulado nos pólos e<br />

continentes que, por consequência, provocaram alterações no nível médio<br />

do mar (Shackleton & Opdyke, 1973).<br />

As modificações do nível do mar são, contudo, espacialmente<br />

afectadas por oscilações gravitacionais potenciais do sistema terra-oceano<br />

e gelo, sendo por isso necessário efectuar correcções a nível glacio-hidro-<br />

isostático durante a elaboração de uma curva de variação do nível do mar<br />

global (Lambeck et al., 2002).<br />

Nos últimos anos, várias curvas de variação do nível do mar foram<br />

elaboradas na zona tropical, tendo em conta as variações de altura dos<br />

terraços coralíferos datados a partir do U/Th obtidos em: Barbados<br />

(Fairbanks, 1989; Bard et al., 1990a; 1990b), Tahiti (Bard et al., 1996), Península<br />

de Huon (Chappell & Polach, 1991; Edwards et al., 1993), noroeste<br />

Australiano (Yokoyama et al., 2000; 2001) etc. e utilizadas por Lambeck et al.<br />

(2002) para estimar a variabilidade do nível do mar global e respectivas<br />

modificações no volume de gelo desde o estádio isotópico marinho 3 (MIS 3)<br />

até à actualidade (Fig. I.14).<br />

O nível do mar desceu desde 60 000 a 30 000 anos cal BP onde atingiu<br />

o seu valor mínimo de – 120 m. Entre 30 000 e 19 500 anos, o nível do mar<br />

permaneceu relativamente estável e, no final deste período, subiu<br />

rapidamente (15 m) em cerca de 500 anos (Yokoyama et al., 2000).<br />

44<br />

F. Naughton, 2007


Fig. I.14 | Curva de variação do nível global do mar (adaptado de Lambeck et al., 2002).<br />

Durante o início da deglaciação (entre 19 000 e 16 000 anos) a fusão<br />

global dos gelos foi lenta provocando um aumento gradual do nível do mar<br />

de cerca de 3.3 mm/ano, enquanto que entre 16 000 e 12 500 anos cal BP a<br />

subida do nível do mar foi bastante mais rápida (16.7 mm/ano) (Lambeck et<br />

al., 2002). Há cerca de 14 000 anos, um hiato, marca o evento designado de<br />

“melwater pulse 1A” (Fairbanks, 1989; Bard et al., 1990a). A ocorrência de um<br />

episódio de curta duração, entre 12 500 e 11 500 anos cal BP, marcado por<br />

um “plateau” onde o nível do mar permaneceu relativamente estável,<br />

representa o evento Dryas recente (Bard et al., 1996; Lambeck et al., 2002).<br />

Desde 11 500 a 8 500 anos o nível do mar subiu gradualmente cerca de 15.2<br />

mm/ano (Lambeck et al., 2002). Finalmente o volume oceânico aproximou-<br />

se do seu actual valor há cerca de aproximadamente 7 000 anos (Lambeck,<br />

2000; Lambeck et al., 2002).<br />

1. 3. 2 Evolução dos sistemas costeiros<br />

As modificações do nível do mar que ocorreram durante a<br />

deglaciação, encontram-se registadas nas sequências sedimentares por<br />

uma alternância entre episódios de incisão fluvial ou de preenchimento<br />

sedimentar, associados a momentos de descida ou subida do nível do mar,<br />

respectivamente. Estas características foram detectadas pela primeira vez<br />

45<br />

F. Naughton, 2007


no Vale do Mississipi (Fisk & McFarlan, 1955) e posteriormente noutros sistemas<br />

costeiros mundiais, nomeadamente: na costa da Louisiana (Nichol et al.,<br />

1996), na costa de Delaware (Belknap & Kraft, 1985), no sul da Austrália (Roy<br />

et al., 1995) e no delta Yangtze na China (Li et al., 2002).<br />

Para além deste indicador sedimentar de modificações relativas do<br />

nível relativo do mar, existem também outro tipo de variações, relacionadas<br />

com a amplitude da influência fluvial relativamente à marinha e vice-versa,<br />

nos registos sedimentares costeiros, os quais podem fornecer informações<br />

importantes no que respeita à variabilidade do nível de base marinho<br />

(Dalrymple et al., 1994; Zaitlin et al., 1994).<br />

Este tipo de abordagem, que se baseia no estudo das mudanças<br />

ambientais ao longo de registos sedimentares costeiros (lagoas e estuários)<br />

durante a deglaciação, foi efectuado nos últimos anos ao longo da<br />

Península Ibérica nomeadamente na zona costeira sudoeste portuguesa<br />

(Freitas et al., 2002; 2003). Estes trabalhos sugerem a presença de um<br />

ambiente fluvial durante o LGIT seguido de um gradual aumento da<br />

influência marinha ao longo do Holocénico como resposta à subida gradual<br />

do nível do mar o qual atingiu a sua máxima influência no final da<br />

deglaciação. Contudo, no final da deglaciação, a formação de barreiras<br />

arenosas na embocadura destes estuários assim como a desaceleração da<br />

subida do nível do mar contribuíram, para que outros mecanismos locais<br />

tenham sido responsáveis pelas modificações geomorfológicas destas zonas<br />

(Freitas & Andrade, 1997; 2001; Freitas et al., 2002; 2003).<br />

Este tipo de estudos começou a ser efectuado para a região noroeste<br />

de Portugal nos últimos anos (Naughton, 2002; Guerreiro et al., 2005; Moreno<br />

et al., 2005; 2006; Drago et al., in press; Fradique et et al., in press).<br />

De forma a compreender a evolução geomorfológica do noroeste de<br />

Portugal e descriminar os mecanismos globais e locais responsáveis pela<br />

evolução do estuário do Douro, foi efectuado um estudo polínico e<br />

sedimentar, de alta resolução temporal, em duas sequências sedimentares<br />

estuarinas situadas no noroeste da Península Ibérica para o Holocénico<br />

(Naughton, 2002). Este trabalho foi sintetizado numa publicação científica, a<br />

qual é apresentada no capítulo 6.<br />

46<br />

F. Naughton, 2007


1. 4 Zona de estudo<br />

1. 4. 1 Margem Ibérica<br />

1. 4. 1. 1 Clima e vegetação actual<br />

A Península Ibérica é ocupada por uma série de biomas, climas e tipos<br />

de solos que variam de acordo com a topografia. Esta região é<br />

caracterizada por duas zonas biogeográficas principais: a zona atlântica<br />

(Blanco Castro et al., 1997) e a zona mediterrânica (Fig. I.15). A zona<br />

atlântica inclui, a norte da Península ibérica, a designada zona Eurosiberiana<br />

(Peinado & Rivas-Martinez, 1987) (Fig. I.10).<br />

Uma vez que os pólens e esporos inclusos nos sedimentos marinhos<br />

representam geralmente a vegetação que coloniza as bacias hidrográficas,<br />

deve-se ter em conta ambas as delimitações biogeográficas referenciadas<br />

nas figuras seguintes: Fig. I.10 e Fig. I.15.<br />

Fig. I.15 | Localização das zonas biogeográficas (adaptado de Blanco Castro et al., 1997).<br />

A região biogeográfica mediterrânica ocupa grande parte da<br />

Península enquanto que a zona atlântica é estreita e está essencialmente<br />

confinada ao norte e noroeste da mesma. A região mediterrânica é<br />

caracterizada por um forte contraste sazonal (verões quentes e secos e<br />

47<br />

F. Naughton, 2007


invernos frios e húmidos) enquanto a região atlântica, influenciada pelo<br />

oceano Atlântico, apresenta condições sazonais menos contrastantes,<br />

sendo por isso caracterizada por um clima temperado e húmido ao longo de<br />

todo o ano.<br />

O noroeste espanhol, incluindo a bacia hidrográfica do Minho, é<br />

caracterizado por temperaturas situadas entre -7 e +10°C e por uma<br />

precipitação média anual que varia entre 900 e 1400 mm. Esta zona é<br />

dominada pela floresta de carvalhos, nomeadamente por: Quercus robur,<br />

Q. pyrenaica e Q. petraea e por uma vegetação rasteira constituída<br />

essencialmente por herbáceas do tipo Ericaceae, Calluna e Ulex. Nesta<br />

zona, é ainda possível detectar localmente Betula pubescens subsp.<br />

celtiberica, Corylus avellana e Genista (Alcara Ariza et al., 1987).<br />

Ligeiramente a sul desta região, encontra-se uma zona de transição a<br />

qual inclui a bacia hidrográfica do Douro. Esta zona é caracterizada por<br />

valores de temperatura média anual de 12º C (Loureiro et al., 1986) e por<br />

temperaturas invernais que rondam os 4 e -4°C nas baixas e médias altitudes,<br />

podendo chegar aos -8°C nas altas altitudes (Polunin & Walters, 1985). A<br />

precipitação é elevada, sendo aproximadamente de 700 a 1000 mm/ano<br />

nas baixas e médias altitudes, e de 1600 mm/ano nas altas altitudes. A<br />

influência oceânica é particularmente importante no noroeste da bacia,<br />

favorecendo a predominância da associação Q. robur e Q. suber (Braun-<br />

Blanquet et al., 1956). A expansão de Pinus pinaster, Pinus sylvestris,<br />

Eucalyptus globulus, Castanea sativa e Juglans regia resulta do impacto<br />

antrópico (Valdès & Gil Sanchez, 2001). A vegetação é ainda dominada por<br />

Ulex e Ericaceae e as margens do rio são colonizadas por Alnus glutinosa,<br />

Fraxinus angustifolia, Ulmus spp., Salix spp. e Populus spp.. A zona oriental é<br />

afectada por condições ligeiramente mais continentais, favorecendo a<br />

expansão de Q. ilex, Q. suber, Juniperus spp. e a floresta de carvalhos (Q.<br />

pyrenaica, Q. faginea).<br />

O sudoeste da Península Ibérica, nomeadamente as bacias do Tejo e<br />

Sado, são influenciadas por um clima mediterrânico (precipitação média<br />

anual de 200 a 600 mm e temperaturas que variam entre 4 a 14°C) o qual<br />

favorece o desenvolvimento da floresta esclerófila. A zona oeste é<br />

representada essencialmente por uma floresta de Q. Ilex, Q. rotundifolia e Q.<br />

48<br />

F. Naughton, 2007


Suber, assim como pela presença de Phillyrea angustifolia e Pistacia<br />

terebinthus enquanto a parte oriental, é sobretudo colonizada por Q.<br />

rotundifolia e Q. coccifera associada a Juniperus communis e Pinus<br />

halepensis. As zonas representantes de altitudes médias (700-1000 m a.s.l.)<br />

são dominadas pela floresta de carvalhos de folha caduca (Q. pyrenaica e<br />

Q. faginea) associada a espécies norte Europeias tais como Taxus baccata.<br />

A degradação da floresta favorece a expansão de Cistaceaes nas zonas<br />

ligeiramente áridas e de Ericaceas nas zonas mais húmidas (Blanco Castro et<br />

al., 1997).<br />

1. 4. 1. 2 Oceanografia<br />

A margem Ibérica é dominada pelo sistema de correntes superficiais<br />

designado por Sistema de correntes de Portugal (PCS-Portugal Current<br />

System) o qual é composto por uma corrente lenta que se desloca em<br />

direcção ao equador ao longo do oceano aberto (Arhan et al., 1994) (Fig.<br />

I.16) e pela rápida corrente costeira a qual reverte sazonalmente a direcção<br />

do seu percurso (Ambar & Fiúza, 1994; Barton, 1998).<br />

Fig. I.16 | Esquema detalhado das principais correntes de superfície do Atlântico Norte: EG-corrente Este<br />

da Gronelândia, Ei-corrente este da Islândia, Gu-Gulf Stream, Ir-corrente de Irminger, La-corrente do<br />

Labrador, Na-corrente Norte Atlântica, Nc-corrente do Cap Norte, Ng-corrente da Noroega, Ni-corrente<br />

do Norte da Islândia, Po-corrente de Portugal, Sb-corrente de Spitsbergen, Wg-corrente Oeste da<br />

Gronelândia. Linhas negras representam as correntes relativamente quentes enquanto as linhas a<br />

tracejado representam correntes relativamente frias (adaptado de Dietrich et al., 1980).<br />

49<br />

F. Naughton, 2007


Durante o verão, a célula de altas pressões dos Açores encontra-se<br />

localizada na zona central do Atlântico Norte e o centro de baixas pressões<br />

da Gronelândia é fraco (Torres et al., 2003) (Fig. I.17).<br />

Fig. I.17 | Representação da localização dos centros de altas e baixas pressões durante o verão e o<br />

Inverno ao longo do Hemisfério Norte (adaptado de Hurrell & Dickson, 2004). As setas representam a<br />

direcção dos ventos dominantes.<br />

Esta situação gera ventos dominantes vindos de norte e do noroeste<br />

favorecendo a circulação para sul das correntes costeiras superficiais (Fiúza<br />

et al., 1982; Haynes & Barton, 1990), ao longo da zona externa da<br />

plataforma, nos primeiros 50 a 100 m de profundidade da coluna de água<br />

(Álvarez-Salgado et al., 2003), assim como a ocorrência de upwelling ao<br />

longo da margem portuguesa (Fiúza et al., 1982; Haynes & Barton, 1990;<br />

Torres et al., 2003). A massa de água fria rica em nutrientes (<strong>EN</strong>ACWsp-<br />

Eastern North Atlantic Central Water of subpolar sources) dirige-se para norte<br />

dos 45° N enquanto a massa de água salina e pobre em nutrientes<br />

(<strong>EN</strong>ACWst- Eastern North Atlantic Central Water of subtropical origin) dirige-<br />

se0 para sul de 40° N (Fiúza, 1984; Rios et al., 1992) (Fig. I.18).<br />

Durante o inverno, o centro de altas pressões dos Açores localiza-se<br />

na margem noroeste Africana, o centro de baixas pressões da Gronelândia<br />

é forte e encontra-se desviado para sudoeste da Gronelândia (Fig. I.17).<br />

50<br />

F. Naughton, 2007


Fig. I.18 | Esquema das principais correntes oceânicas que circulam ao longo da margem Ibérica. PCS:<br />

Portugal Current system; <strong>EN</strong>ACWsp: Eastern North Atlantic Central Water de origem sub-polar; <strong>EN</strong>ACWst:<br />

Eastern North Atlantic Central Water de origem sub-tropical; MSW: Mediterranean Sea Water; LSW: Labrador<br />

Sea water; NADW: North Atlantic Deep Water (adaptado de Sprangers et al., 2004).<br />

O gradiente de pressão entre estes dois sistemas depressionários<br />

favorece os ventos vindos de sul, ao longo da margem ibérica, provocando<br />

processos de downwelling e o transporte das correntes costeiras de<br />

superfície para norte (Frouin et al., 1990; Haynes & Barton, 1990). Esta inversão<br />

dos padrões hidrológicos ocorre entre Setembro-Outubro e Março-Abril e<br />

representa a designada Contra-corrente costeira portuguesa (PCCC-<br />

Portugal Coastal Counter Current) (Ambar et al., 1986). Esta circulação<br />

ocorre numa zona estreita (aproximadamente de 30 Km) e transporta uma<br />

massa de água quente e salina (<strong>EN</strong>ACWst) para norte, a profundidades<br />

compreendidas entre 200 a 300 m (Pingree & Le Cann, 1990).<br />

Entre 550 a 1500 m de profundidade da coluna de água, a massa de<br />

água altamente salina e relativamente quente (MSW- Mediterranean Sea<br />

Water) proveniente do estreito de Gibraltar migra para norte (Mazé et al.,<br />

1997) (Fig. I.18). Contudo, a salinidade desta massa de água diminui<br />

principalmente a latitudes superiores a 41° N onde se mistura com a massa<br />

de água de fraca salinidade proveniente do Mar de Labrador (LSW-<br />

Labrador Sea water) (McCave & Hall, 2002). A LSW é uma das três<br />

componentes que compõem a massa de água profunda norte Atlântica<br />

(NADW- North Atlantic Deep Water) (Huthnance et al., 2002).<br />

51<br />

F. Naughton, 2007


1. 4. 1. 3 Geomorfologia e dinâmica sedimentar actual<br />

A Margem Ibérica é caracterizada por uma plataforma continental<br />

muito estreita (30 a 50 Km) e por um talude irregular, bastante inclinado, o<br />

qual mergulha rapidamente para as grandes profundidades da planície<br />

abissal (Fig. I.19). Esta margem continental é interceptada parcialmente por<br />

canhões submarinos muito profundos, nomeadamente: Mugia, Porto, Aveiro,<br />

Cascais, Lisboa e São Vicente. Os grandes canhões submarinos de Setúbal e<br />

da Nazaré cortam a plataforma continental na sua quase totalidade<br />

facilitando a captura dos sedimentos que se encontram em suspensão<br />

permitindo assim a sua conduta directa para a planície abissal (Vanney &<br />

Mougenot, 1981). Pensa-se que os canhões submarinos Ibéricos devem ter<br />

sido bastante mais activos durante períodos de baixo nível do mar (Van<br />

Weering & McCave, 2002). O talude continental é ainda marcado pela<br />

presença de várias montanhas submarinas tais como: Vigo (VS), Vasco da<br />

Gama (VDGS), Porto (PS), Tore (TS), pelo Banco da Galiza e por algumas<br />

depressões tectónicas (Vanney & Mougenot, 1981).<br />

Fig. I.19 | a) Morfologia da margem continental Ibérica: Canhões submarinos de Mugia (MC), do Porto<br />

(PC), de Aveiro (AC), da Nazaré (NC), de Cascais (CC), de Lisboa (LC), de Setúbal (SC) e de São Vicente<br />

(S.VC); montanhas submarinas de Tore (TS), do Porto (PS), Vasco da Gama (VDGS) e de Vigo (VS).<br />

52<br />

F. Naughton, 2007


No noroeste da Península Ibérica, grandes quantidades de sedimentos<br />

são libertados pelos rios Douro, Ave, Cavado, Lima e Minho para a margem<br />

continental Ibérica. O principal fornecedor de sedimentos à plataforma<br />

noroeste portuguesa é o rio Douro seguido pelo rio Minho (~8.2 x 10 9 m 3<br />

descarga média anual) (Dias et al., 2002; Jouanneau et al., 2002; Oliveira et<br />

al., 2002) (Fig. I.20). A área das suas bacias hidrográficas é de 97 682 km 2 e 17<br />

081 km 2 e o seu comprimento total é de 927 km e 300 km, respectivamente<br />

(Loureiro et al., 1986). Contrariamente aos rios que desaguam no noroeste de<br />

Portugal, as rias situadas a norte de 42° N (Vigo, Pontevedra, Arousa e Muros)<br />

são consideradas como armadilhas sedimentares impedindo o fornecimento<br />

de sedimentos continentais à plataforma adjacente (Dias et al., 2002;<br />

Jouanneau et al., 2002).<br />

A plataforma continental noroeste Ibérica é composta por: a) uma<br />

zona interna, situada a menos de 30 m de profundidade, constituída<br />

essencialmente por areias finas; b) uma zona intermédia constituída<br />

principalmente por areias e cascalhos e uma zona externa rica em<br />

carbonatos e areias de dimensão média (Van Weering et al., 2002). Esta<br />

plataforma é ainda constituída por dois importantes complexos silto-argilosos<br />

(Douro e Galiza) separados por uma zona desprovida de lodo (Lopez-Jamar<br />

et al., 1992) (Fig. I.20).<br />

Fig. I.20 | Morfologia da plataforma continental do noroeste da Península Ibérica (adaptado de Dias et al.,<br />

2002).<br />

53<br />

F. Naughton, 2007


A evolução destes corpos lodosos depende essencialmente da<br />

quantidade de sedimentos fornecidos pelo continente adjacente, da acção<br />

das barreiras morfológicas e ainda, das condições hidrológicas do meio<br />

(Dias et al., 2002; Jouanneau et al., 2002).<br />

Os processos de sedimentação são bastante complexos no noroeste<br />

da margem Ibérica e ocorrem muitas vezes associados a eventos episódicos<br />

de cheias (Dias et al., 2002) e /ou a episódios associados a descargas fluviais<br />

importantes (Araújo et al., 1994; Drago et al., 1998).<br />

Depois de serem libertados pelos rios, os sedimentos finos são<br />

transportados essencialmente por camadas nefelóides: de fundo (BNL-<br />

Bottom N<strong>ephe</strong>loid Layers), intermédias (INL-Intermediate N<strong>ephe</strong>loid Layers) e<br />

de superfície (SNL-Surface N<strong>ephe</strong>loid Layers). Oliveira et al. (1999),<br />

detectaram um decréscimo da concentração de sedimentos finos ao longo<br />

das diferentes camadas nefelóides com o aumento da distância à linha de<br />

costa e que as correntes e a ondulação induzem a resuspensão dos<br />

sedimentos depositados nos complexos silto-argilosos do Douro e Galiza<br />

especialmente durante episódios de grandes tempestades. Durante alguns<br />

destes episódios extremos, tais como tempestades de inverno as quais estão<br />

associadas a condições de downwelling (ventos vindos do sul), os<br />

sedimentos contidos na BNL podem ficar retidos nos afloramentos rochosos<br />

(Fig. I.20) (Drago et al., 1998) os quais funcionam como barreiras à<br />

transferência de sedimentos da plataforma para o talude e planície abissal<br />

(Jouanneau et al., 2002; Van Weering et al., 2002), estimulando assim o<br />

transporte sedimentar para norte, ao longo da plataforma continental<br />

(Drago et al., 1998; Dias et al., 2002; Jouanneau et al., 2002; Van Weering et<br />

al., 2002).<br />

A ondulação associada a esses eventos extremos podem<br />

esporadicamente induzir a ressuspensão de partículas situadas na BNL<br />

(Vitorino et al., 2002) alimentado a camada subjacente INL (Oliveira et al.<br />

2002) e contribuir para a exportação dos mesmos para o largo (Vitorino et<br />

al., 2002).<br />

A inversão de correntes de superfície induzidas pela presença local de<br />

redemoinhos pode também contribuir para a transferência de partículas<br />

para o talude e planície abissal (Pingree & LeCann, 1992).<br />

54<br />

F. Naughton, 2007


Durante o verão e associado a condições de upwelling (ventos vindos<br />

do norte ou de noroeste) a exportação de sedimentos efectua-se<br />

essencialmente a nível do eixo da plataforma (McCave & Hall 2002; Van<br />

Weering et al., 2002). A transferência lateral de sedimentos da plataforma<br />

para o largo pode ocorrer por vezes, em locais onde existam filamentos<br />

transversais (Huthnance et al., 2002).<br />

A sudoeste da margem Ibérica, o rio Tejo é o principal fornecedor de<br />

sedimentos à plataforma continental, talude e planície abissal, seguido pelo<br />

rio Sado (Dias, 1987; Jouanneau et al., 1998) (Fig. I.21). O Rio Tejo apresenta<br />

cerca de 1110 km de comprimento e a área da sua bacia hidrográfica é de<br />

80 600 km 2 enquanto o rio Sado é bastante mais pequeno, com apenas 175<br />

km de comprimento e uma área da bacia hidrográfica de 7 640 km 2<br />

(Loureiro et al., 1986).<br />

Fig. I.21 | Morfologia da plataforma continental sudoeste portuguesa (adaptado de Araújo et al., 2002).<br />

A diferença entre a descarga fluvial existente entre os dois rios assim<br />

como o papel desempenhado pelas correntes litorais, condicionam a<br />

dinâmica sedimentar desta zona (Jouanneau et al., 1998). O complexo silto-<br />

55<br />

F. Naughton, 2007


argiloso localiza-se logo após a embocadura do estuário do Tejo e cobre<br />

toda a área da plataforma continental (Araújo et al., 2002).<br />

Durante o verão, a concentração de matéria particulada em<br />

suspensão (SPM-Suspended Particulate Matter) é quatro vezes mais elevada<br />

na embocadura do estuário do Tejo do que no Sado e as camadas<br />

nefelóides estendem-se cerca de 30 Km para oeste da actual linha de costa<br />

(Jouanneau et al., 1998).<br />

O transporte de partículas finas para o largo é favorecido pelos<br />

canhões submarinos de Cascais, Lisboa e Setúbal (Jouanneau et al., 1998) e<br />

por filamentos transversais (Huthnance et al., 2002).<br />

56<br />

F. Naughton, 2007


1. 4. 2 Plataforma continental noroeste Francesa<br />

1. 4. 2. 1 Geomorfologia, oceanografia e sedimentação actual<br />

A plataforma continental Francesa localiza-se no Golfo da Gasconha<br />

(Golfe de Gascogne) entre 43° a 48° N (Fig. I.22). Morfologicamente a<br />

plataforma francesa é muito larga e pouco inclinada no noroeste (300 Km<br />

de largura) tornando-se dez vezes mais estreita e íngreme para sul (30 Km de<br />

largura).<br />

Fig. I.22 | Localização do corpo lodoso “Grande Vasière” na plataforma continental Francesa.<br />

Esta plataforma, é constituída por dois corpos lodosos principais: um<br />

situado próximo do estuário da Gironde (Gironde shelf mud field) e um outro,<br />

situado entre a Bretanha e a Charente (Grande Vasière) (Allen & Castaing,<br />

1977). Estes corpos lodosos são essencialmente alimentados por sedimentos<br />

finos libertados pelos rios Gironde e Loire e, em menores quantidades, pelos<br />

rios Adour, Vilaine e Charente (Castaing & Jouanneau, 1987; Jouanneau et<br />

al. 1999). A área das suas bacias hidrográficas é de: Gironde (que inclui as<br />

bacias hidrográficas dos rios Garonne e Dordogne) 79 180 km 2 , Loire 118 420<br />

km 2 , Adour 17 100 km 2 , Vilaine 10 420 km 2 e Charente 11 970 km 2 .<br />

O estuário da Gironde é um dos mais longos da Europa, apresentando<br />

cerca de 80 km de comprimento e uma superfície que ronda os 625 Km 2 .<br />

A plataforma Francesa é afectada fortemente pela ondulação,<br />

principalmente em períodos de fortes tempestades. Para além da<br />

57<br />

F. Naughton, 2007


ondulação, esta plataforma é ainda sujeita a um regime de maré semi-<br />

diurno, meso a macrotidal. Durante o verão a massa de água apresenta<br />

uma forte estratificação enquanto que, durante o inverno, a acção dos<br />

ventos que sopram de Oeste favorecem a não estratificação da coluna de<br />

água impedindo a exportação de sedimentos finos para o talude e planície<br />

abissal.<br />

O complexo silto-argiloso “Grande Vasière”, apresenta cerca de 250<br />

Km de comprimento (paralelamente à linha de costa actual), situa-se a<br />

cerca de 100 m de profundidade da coluna de água, e apresenta uma taxa<br />

de sedimentação de 0.1-0.2 cm/ano (Lesueur et al., 2001) (Fig. I.22).<br />

O fornecimento de sedimentos através das camadas nefelóides<br />

permite a manutenção deste corpo lodoso (Jouanneau et al:, 1999; Lesueur<br />

et al., 2001), principalmente durante períodos de cheia (Bourillet et al., 2005).<br />

Durante estes episódios cerca de 30% dos sedimentos em suspensão são<br />

libertados pelos rios e depositam-se a cerca de 20 Km da costa.<br />

1. 4. 2. 2 Clima e vegetação<br />

O noroeste francês é caracterizado por um clima teperado e húmido,<br />

com temperaturas médias anuais que rondam os 13°C e precipitação média<br />

anual de cerca de 1000 mm (dados obtidos na agência pública: “Meteo<br />

France”).<br />

À medida que nos afastamos do litoral, a precipitação média anual<br />

diminui até cerca 600 mm/ano, nas zonas de baixa altitude, enquanto que<br />

estes valores são bastante elevados em áreas situadas nas altas altitudes, tais<br />

como: o “Massif Central”, que apresenta valores superiores a 2200 mm/ano e<br />

nos Pirinéus onde variam entre 2000 mm/ano (oeste) e 1000 mm/ano (este).<br />

O clima temperado e húmido que caracteriza o noroeste francês,<br />

favorece a expansão da floresta decídua e da floresta mista quente,<br />

nomeadamente: deciduous Quercus (Quercus pedunculata, Q. pubescens e<br />

Q. sessiflora) e a presença local de Q. Ilex, Q. Suber, Ulmus e Fraxinus. As<br />

zonas litorais são essencialmente compostas por Pinus pinaster e Ulex. Nas<br />

zonas de maiores altitudes é possível encontrar ainda Fagus e Carpinus.<br />

58<br />

F. Naughton, 2007


1. 5 Material e Métodos<br />

1. 5. 1 Amostras superficiais<br />

De forma a calibrar a assinatura polínica marinha ao longo da<br />

Península Ibérica, foram recolhidos uma série de níveis sedimentares<br />

superficiais, em vários tipos de ambientes distintos, a oeste da Península<br />

Ibérica, tais como: estuários, plataforma e talude continental (Fig. I.23; Tab.<br />

I.1), nos quais foi efectuada a análise do conteúdo polínico. Algumas destas<br />

amostras foram recolhidas na área adjacente à zona biogeográfica<br />

mediterrânica, enquanto outras foram preferencialmente recolhidas na<br />

região adjacente à zona biogeográfica Atlântica. Cada amostra<br />

corresponde a um sinal polínico representativo dos últimos 200 anos.<br />

Fig. I.23 | Amostras sedimentares de superfície analisadas neste estudo (VIR-18, Ría de Vigo; Laquasup,<br />

Estuário do Douro; PO287-13-2G, Complexo silto-argiloso do Douro; CG11, Complexo silto-argiloso do<br />

Minho; MD99-2331, Talude continental ao largo de Vigo; MD04-2814 CQ, Talude continental ao largo do<br />

Porto; Barreiro, Estuário do Tejo; MD99-2332, Complexo silto-argiloso de Lisboa; FP8-1, Talude continental<br />

ao largo de Sines; e MD95-2042, Talude-planície abissal ao largo de Sines).<br />

59<br />

F. Naughton, 2007


Nome da<br />

amostra<br />

Prof.<br />

(cm)<br />

MD95-2042 Top<br />

(0-1)<br />

1FP8-1 Top<br />

(0-1)<br />

MD99-2332 Top<br />

(0-1)<br />

Barreiro Top<br />

(0-1)<br />

MD04-2814 CQ Top<br />

(0-1)<br />

Laquasup Top<br />

(0-5)<br />

PO287-13-2G Top<br />

(0-1)<br />

CG11 Top<br />

(0-1)<br />

Latitude Longitude Prof.<br />

coluna<br />

água<br />

(m)<br />

60<br />

Ano<br />

Projecto,<br />

missões<br />

ou instituições<br />

37°48’N 10°10’W 3148 1995 IMAGES V<br />

38°01’N 09°20’W 980 2003 FORAMPROX1<br />

38°33’N 09°22’W 97 1999 GINNA-IMAGESV<br />

38°40’N 09°07’W 0 1999 IPIMAR<br />

40°37’N 09°52’W 2449 2004 ALI<strong>EN</strong>OR<br />

41°09’N 08°38’W 0 2001 LAQUATEDE<br />

(PRAXIS/P/CTE/1<br />

1101/1998)<br />

<strong>EN</strong>VI-CHANGES<br />

(PDCTM/PP/MAR<br />

/15251/99)<br />

41°09’N 09°01’W 81 2002 <strong>EN</strong>VI-CHANGES<br />

PDCTM/PP/MAR/<br />

15251/99<br />

41°48’N 09°04’W 107 1992 GEOMAR92<br />

MD99-2331 3-4 42°09’N 09°42’W 2120 1999 GINNA-IMAGES<br />

V<br />

Vir-18 Top 42°14’N 08°47’W 45 1990 Departamento<br />

(0-1)<br />

de estratigrafia<br />

da Universidade<br />

de Vigo<br />

Tab. I.1 | Localização das amostras de superfície. Da esquerda para a direita encontra-se representado o<br />

nome das amostras, a profundidade na coluna sedimentar, a latitude, a longitude, a profundidade da<br />

coluna de água, o ano da colheita das amostras, o nome dos projectos científicos ou o nome das missões<br />

oceanográficas ou o nome das Instituições que forneceram as amostras.<br />

Os resultados obtidos foram posteriormente comparados com as<br />

assinaturas polínicas de várias amostras de superfície continentais<br />

(sedimentos lacustres, musgos e turfeiras) provenientes da base de dados<br />

“European pollen database” http:/www.imep-cnrs.com/pages/EPD.htm<br />

(Peyron et al., 1998; Barboni, et al., 2004). Estas amostras representam<br />

igualmente um sinal polínico dos últimos 200 anos.<br />

Na tentativa de compreender as variações do registo polínico obtido<br />

em duas sequências sedimentares colhidas no estuário do Douro (Core 1 e<br />

F. Naughton, 2007


Core 1B) (Capítulo 6), foi necessário, efectuar um estudo polínico em cinco<br />

amostras de superfície colhidas após um episódio de cheia que ocorreu em<br />

Novembro de 2000, de forma a poder descriminar os diferentes tipos de<br />

assinatura polínica: regional (associado à cheia) e local (associada a uma<br />

parcial exposição sub-aérea) (Fig. I.24).<br />

Fig. I.24 | Localização das amostras de superfície e das sondagens estudadas ao longo do estuário do<br />

Douro. Os círculos brancos com pinta negra representam as amostras de superfície e os círculos pretos<br />

representam as sondagens.<br />

61<br />

F. Naughton, 2007


1. 5. 2 Sondagens<br />

Duas sondagens marinhas profundas (MD99-2331 e MD03-2697), uma<br />

sondagem marinha pouco profunda (Vk03-58Bis) e duas sondagens<br />

estuarinas foram escolhidas para estudar a variabilidade climática dos<br />

últimos 30 000 anos nas latitudes médias do Atlântico Norte (Fig. I.25).<br />

Fig. I.25 | Localização das sondagens utilizadas neste trabalho. Sondagens marinhas profundas: MD99-<br />

2331 e MD03-2697; sondagem marinha pouco profunda: VK03-58Bis; sondagens estuarinas: Core1 e<br />

Core1B.<br />

1. 5. 2. 1 Sondagens marinhas profundas<br />

As sondagens MD99-2331 (42° 09’ N, 09° 40’ 90 W) e MD03-2697 (42° 09’<br />

59 N, 09° 42’ 10 W) foram colhidas na margem Galega, a 2110 m e 2164 m de<br />

profundidade, a bordo do navio oceanográfico Marion Dufresne, usando um<br />

“corer” CALYPSO, durante as missões oceanográficas GINNA (a qual se<br />

insere no programa IMAGES V) e PICABIA, respectivamente (Fig. I.25). Estas<br />

sondagens apresentam cerca de 37.2 m e 41.23 m sedimento hemipelágico<br />

e cobrem aproximadamente os últimos 225 000 e 425 000 anos.<br />

195 níveis sedimentares foram recolhidos em todos os 2 cm nos<br />

primeiros 3 m da sondagem MD99-2331 e, em todos os 5 cm nos restantes 4<br />

m de sedimento. Infelizmente, os resultados obtidos ao longo dos níveis<br />

sedimentares correspondentes ao MIS 1 (primeiros 2 metros de sondagem), o<br />

qual engloba o LGIT e o actual interglaciário (Holocénico), não serão<br />

62<br />

F. Naughton, 2007


apresentadas nesta dissertação uma vez que, posteriormente às análises<br />

efectuadas (pólen, δ 18 O de foraminíferos planctónicos e bentónicos,<br />

alcanonas, associações de foraminíferos planctónicos), verificou-se a<br />

existência de mistura de sedimentos, nos resultados obtidos pelos vários<br />

indicadores paleoclimáticos, assim como, nos resultados obtidos pelas<br />

idades 14 C. Esta mistura de sedimentos, foi igualmente confirmada a partir da<br />

análise radiográfica a qual foi efectuada utilizando um processador de<br />

Imagem (SCOPIX image-processing mode). Por essa razão, apenas 110 dos<br />

níveis analisados (a nível do seu conteúdo polínico) serão apresentados<br />

neste trabalho.<br />

De forma a complementar os resultados obtidos na sondagem MD99-<br />

2331 foi efectuada a análise polínica em 54 amostras, ao longo dos primeiros<br />

4.10 m de sedimento, na sondagem vizinha MD03-2697, com uma resolução<br />

entre amostras que variou de 1 a 10 cm. Dois dos níveis foram considerados<br />

estéreis uma vez que apresentavam uma fraca ou quase nula quantidade<br />

esporopolínica.<br />

Na sondagem MD99-2331, entre 2 m e 7 m de profundidade, foi<br />

efectuada ainda a análise semi-quantitativa de IRDs e das associações de<br />

foraminíferos planctónicos (em 78 níveis sedimentares) assim como a<br />

determinação do δ 18 O contido nas carapaças de foraminíferos planctónicos<br />

(em 58 níveis sedimentares) e bentónicos (em 29 níveis sedimentares) com<br />

uma resolução entre amostras que variou de 2 a 10 cm. Foram ainda<br />

efectuadas 40 datações por 14 C AMS (Accelerator Mass Spectrometry) em<br />

foraminíferos planctónicos do tipo Globigerina bulloides ou<br />

Neogloboquadrina pachyderma (s.), ao longo dos primeiros 7 m de<br />

sondagem.<br />

Nos primeiros 4 m da sondagem MD03-2697 foi ainda efectuada a<br />

análise semi-quantitativa de IRDs (em 60 níveis sedimentares) e das<br />

associações de foraminíferos planctónicos (em 71 níveis sedimentares), a<br />

determinação de δ 18 O de foraminíferos planctónicos (em 59 níveis<br />

sedimentares) e bentónicos (em 35 níveis sedimentares) com uma resolução<br />

entre amostras de 2 a 10 cm, e ainda 11 datações 14 C AMS.<br />

63<br />

F. Naughton, 2007


1. 5. 2. 2 Sondagem marinha pouco profunda VK03-58Bis<br />

A sondagem VK03-58Bis foi colhida na plataforma continental<br />

noroeste francesa, na Baía de Biscaia, no sector “Sud-Glénan” do complexo<br />

silto-argiloso designado por “La Grande Vasière” (47°36’ N e 4°08’ W) a 96.8<br />

m de profundidade) usando uma vibrosonda, durante a missão<br />

oceanográfica Vibarmor (Fig. I.25). Esta sondagem apresenta cerca de 2.72<br />

m de sedimento silto-argiloso e cobre aproximadamente os últimos 8 855<br />

anos (8 855 cal yr BP). Esta elevada taxa de sedimentação permite a<br />

obtenção de um registo climático de altíssima resolução temporal para o<br />

Holocénico.<br />

Nesta sondagem, foi efectuada a análise polínica em 42 níveis<br />

sedimentares, em todos os 2 a 8 cm de sedimento ao longo dos 2.72 m de<br />

sondagem. Foi ainda efectuada a análise das associações de<br />

dinoflagelados e o estudo das comunidades bentónicas (nomeadamente<br />

do gastrópode Turritella communis) em 15 amostras sedimentares e ainda 5<br />

datações 14 C AMS nas carapaças de gastrópodes do tipo T. communis.<br />

1. 5. 2. 3 Sondagens estuarinas<br />

De forma a compreender o Impacto da variabilidade climática na<br />

evolução dos sistemas costeiros do noroeste da Península Ibérica, durante a<br />

fase final da deglaciação (Holocénico), foram utilizadas 2 sondagens, as<br />

quais foram colhidas na zona do estuário do Douro (Fig. I.24; Fig. I.25).<br />

As sondagens “Core 1” e “Core 1B” (situadas à distância de 50 cm<br />

uma da outra) foram recolhidas numa zona temporariamente emersa da<br />

Baía de S. Paio (estuário do Douro; 41°09’N; 8°38’W), com recurso a uma<br />

sonda hidráulica (rotary perforation). A sondagem total, composta pelas<br />

sondagens “Core 1” e “Core 1B”, é designada por sequência do Douro. A<br />

sequência do Douro atingiu cerca de 20 m de profundidade e cobre<br />

essencialmente o período Holocénico.<br />

Foi efectuada a análise polínica em 9 níveis nos primeiros 7.90 m (-4.4<br />

m OD) da sondagem“Core1”. Esta sondagem contém importantes hiatos<br />

polínicos os quais estão associados à deposição de grandes quantidades de<br />

material grosseiro: areia e cascalho. (OD-Ordnance Datum, representa a<br />

profundidade corrigida em relação ao actual nível médio do mar). Na<br />

64<br />

F. Naughton, 2007


sondagem, “Core 1B”, foi efectuada a análise polínica em 68 níveis<br />

sedimentares, todos os 2 a 20 cm de sedimento entre, 10.50 a 17.40 m (-6.81<br />

a -13.90 m OD) de profundidade. Foi ainda efectuada uma análise<br />

sedimentológica nas sondagens “Core1 e Core 1B”, a qual compreende a<br />

determinação de teores em carbonatos, granulometria, determinação de<br />

teores em matéria orgânica e caracterização sedimentológica de 64<br />

amostras, com um espaçamento entre amostras que varia de 2 a 5 cm. Foi<br />

ainda efectuado a caracterização do nível de cascalho presente nas<br />

sondagens estuarinas, através do estudo morfométrico de 221 balastros, e<br />

finalmente 2 datações por 14 C AMS em material orgânico no “Core1” e 3 no<br />

“Core 1B”.<br />

A análise polínica de todas as sondagens referenciadas anteriormente<br />

(MD99-2331, MD03-2697, Vk03-58Bis, “Core1” e “Core1B”) foi efectuada, no<br />

laboratório UMR CNRS 5805 EPOC, da Universidade de Bordéus. A análise<br />

semi-quantitativa de IRD’s e associações de foraminíferos planctónicos das<br />

sondagens marinhas profundas, foi realizada no mesmo laboratório por<br />

Josette Duprat. A análise de alcanonas da sondagem MD99-2331 foi<br />

efectuada por Edouard Bard e Frauke Rostek, no laboratório CNRS UMR-6635<br />

Cerege da Universidade de Aix-MarseilleIII. A determinação do δ 18 O de<br />

foraminíferos planctónicos do tipo Globigerina bulloides, na sondagem<br />

MD99-2331, foi executada por Bruno Malaizé, no UMR CNRS 5805 EPOC da<br />

Universidade de Bordéus 1, enquanto que o mesmo tipo de análises, na<br />

sondagem MD03-2697, foi efectuado por Elsa Cortijo no “Laboratoire des<br />

Sciences du Climat et de l’Environnement (LSCE)” em Gif-sur-Yvette. Elsa<br />

Cortijo analisou ainda, o conteúdo em δ 18 O dos foraminíferos bentónicos do<br />

tipo Cibicides wuellestorfi, em ambas as sondagens marinhas profundas. A<br />

análise radiográfica destas sondagens marinas profundas foi efectuada no<br />

no laboratório UMR CNRS 5805 EPOC, da Universidade de Bordéus com a<br />

ajuda de Sébastien Zaragosi e Michel Cremer.<br />

As datações 14 C AMS destas sondagens marinhas profundas, foram<br />

efectuadas nos laboratórios “LMC-Laboratoire de Mesure du Carbone 14”<br />

em Saclay (França), “GifA-AMS laboratory” em Gif-sur-Yvette (França) e no<br />

“Beta Analytic Inc.” (USA) enquanto que, as datações 14 C AMS da<br />

65<br />

F. Naughton, 2007


sondagem marinha pouco profunda (Vk03-58Bis) foram realizadas no<br />

“Poznan Radiocarbon Laboratory” (Polónia). As sondagens estuarinas foram<br />

datadas no laboratório “Beta Analytic Inc.” (USA).<br />

A análise sedimentológica da sondagem Vk03-58Bis assim como a<br />

contagem de T. Communis foi efectuada por Folliot no IFREMER a Brest. A<br />

contagem de dinoflagelados foi por seu lado executada por Jean-Loius<br />

Turon no laboratório UMR CNRS 5805 EPOC, da Universidade de Bordéus.<br />

Uma parte da análise sedimentológica das sondagens “Core1 e Core<br />

1B”, nomeadamente a determinação de teores em carbonatos,<br />

granulometria, determinação de teores em matéria orgânica e<br />

caracterização do nível de cascalho presente nessas sondagens estuarinas<br />

foi efectuada pela candidata no IPIMAR-Instituto de Investigação das Pescas<br />

e do Mar (Portugal) enquanto que, a caracterização sedimentológica da<br />

fracção arenosa foi efectuada pela Anabela Oliveira no IPIMAR.<br />

1. 5. 3 Cronologia e datações 14 C<br />

A principal ferramenta necessária ao estudo comparativo e à<br />

correlação indirecta de diferentes registos paleoclimáticos do Quaternário<br />

tardio é, sem dúvida, a cronologia absoluta.<br />

À excepção das sequências constituídas por “varvas” (laminação<br />

anual), a cronologia obtida em sequências sedimentares continentais e<br />

marinhas baseia-se nas datações radiométricas.<br />

O cálculo da datação 14 C baseia-se no princípio de que a actividade<br />

do 14 C contido no CO2 atmosférico tem sido constante ao longo dos anos<br />

(Libby, 1952). Contudo, vários estudos mostram que a concentração de 14 C<br />

na atmosfera variou significativamente ao longo do tempo (Linick et al.,<br />

1986; Stuiver et al., 1991). Esta variabilidade ocorreu como resposta a<br />

modificações da intensidade no campo geomagnético terrestre, da<br />

actividade solar e ainda, da redistribuição de 14 C entre os diferentes tipos de<br />

sub-sistemas climáticos, nomeadamente em função da variação da<br />

circulação termohalina (Hughen et al., 2000). Por essas razões, torna-se<br />

necessário converter as idades 14 C convencionais (anos BP) em idades<br />

calibradas (anos cal BP) (BP-Before present, sendo o presente considerado<br />

como o ano de 1950).<br />

66<br />

F. Naughton, 2007


Recentemente, foram elaboradas uma série de curvas de calibração<br />

tais como: IntCal04, Marine04 e SHCal 04, as quais são recomendadas para<br />

efectuar a calibração das idades 14 C convencionais (Reimer et al., 2004). A<br />

série de dados estão inseridas no programa CALIB Rev 5.0<br />

(http://calib.qub.ac.uk) (Stuiver et al., 2005).<br />

Nas sondagens marinhas MD99-2331, MD03-2697 e Vk03-Bis, as quais se<br />

situam no Atlântico Norte utilizou-se a série de dados marine 04.14c inserida<br />

no programa CALIB Rev 5.0 para calibrar Idades mais jovens do que 21 786<br />

BP (Stuiver & Reimer, 1993; Hughen et al., 2004; Stuiver et al., 2005). Nesta<br />

calibração foi utilizado o intervalo de confiança de 2 sigma (95.4%), assim<br />

como as áreas relativas da curva de probabilidade e, ainda, a<br />

probabilidade mediana da distribuição de probabilidade (Telford et al.,<br />

2004), tal como sugerido por Stuiver et al. (2005). Este programa incorpora a<br />

correcção temporal do efeito de reservatório do oceano global, que é de<br />

400 anos. Este valor está perfeitamente de acordo com o valor da idade<br />

reservatório estipulada para a margem Ibérica por Bard et al. (2004).<br />

Contudo, a utilização deste programa de calibração é limitada a idades<br />

mais recentes do que 21 786 BP, sendo por isso necessário recorrer a outro<br />

tipo de curvas de calibração de forma a calibrar idades 14 C convencionais<br />

mais antigas. As datações mais antigas do que 21 786 BP foram calibradas<br />

utilizando o polinómio simples de segunda ordem, o qual foi elaborado por<br />

Bard et al. (2004), a partir da correlação entre curvas de variação obtidas<br />

por diferentes indicadores paleoclimáticos marinhos e continentais, na<br />

sondagem marinha MD95-2042 (margem Ibérica), com a curva isotópica do<br />

oxigénio contido na sondagem de gelo GISP2:<br />

Idade anos cal BP = -6.2724 x 10 -6 x [idade 14 C anos BP] 2 + 1.3818 x [idade 14 C anos BP] –1388<br />

As datações efectuadas nas sondagens “Core1” e “Core 1B”, do<br />

estuário do Douro, foram calibradas pela Beta Analytic Inc. (USA) utilizando o<br />

programa INTCAL 98 (Stuiver et al., 1998).<br />

67<br />

F. Naughton, 2007


1. 5. 4 Indicadores paleoclimáticos<br />

Os grãos de pólen representam a ferramenta principal utilizada no<br />

decorrer deste trabalho. Contudo, para além deste indicador<br />

paleoclimático terrestre, as sondagens marinhas profundas foram sujeitas a<br />

um estudo multidisciplinar, o qual inclui uma série de indicadores<br />

paleoclimáticos marinhos e um indicador do volume de gelo acumulado nos<br />

pólos.<br />

1. 5. 4. 1 Variação do coberto vegetal e do clima continental<br />

Tal como foi referido previamente, os grãos de pólen e esporos<br />

inclusos nos sedimentos marinhos, reflectem uma imagem integral da<br />

vegetação regional do continente adjacente (Muller, 1959; Koreneva, 1971;<br />

Cross et al., 1966; Manten, 1966; Groot & Groot, 1966; Turon, 1984; Heusser &<br />

Balsam, 1977; Heusser & Shackleton, 1979) e como tal são indicadores das<br />

condições atmosféricas do continente adjacente.<br />

De forma a identificar variações do coberto vegetal e do clima,<br />

procedeu-se ao tratamento laboratorial, à identificação e contagem dos<br />

esporomorfos, ao cálculo da percentagem e concentração polínica, à<br />

elaboração de diagramas polínicos e, finalmente, à reconstrução qualitativa<br />

e quantitativa dos parâmetros climáticos:<br />

a) Tratamento das amostras e montagem das lâminas<br />

Todas as amostras foram preparadas segundo o protocolo<br />

experimental descrito por de Vernal et al. (1996) ligeiramente modificado no<br />

laboratório “Environnements et Paléoenvironnements Océaniques” (UMR<br />

CNRS 5805 EPOC) da Universidade de Bordéus 1 (Desprat, 2005).<br />

Foram lavados cerca de 3 a 5 cm 3 de sedimento seco utilizando um<br />

crivo de forma a recuperar a fracção inferior a 150 μm. Esta fracção final foi<br />

decantada durante 48 horas. Eliminou-se a água em excesso com a ajuda<br />

de uma bomba de vácuo, recuperando-se o resíduo com a ajuda de água<br />

destilada, para um tubo de fundo cónico de 100 cm 3 . Este resíduo foi<br />

centrifugado a 2500 rpm (rotações por minuto) durante 7 minutos. A água<br />

em excesso foi retirada e colocaram-se, no tubo, duas pastilhas contendo<br />

um número conhecido de esporos exóticos de Lycopodium (25084 grãos no<br />

68<br />

F. Naughton, 2007


seu total), para poder estimar posteriormente a concentração de<br />

palinomorfos por cm 3 .<br />

De forma a eliminar totalmente o conteúdo em carbonatos das<br />

amostras, foram efectuados vários ataques químicos, a frio, utilizando ácido<br />

clorídrico: primeiro utilizando uma solução de HCl a 10% seguido de uma<br />

solução de HCl a 25% e posteriormente uma solução de HCl a 50%.<br />

De seguida, as amostras foram sujeitas a dois ataques químicos com<br />

ácido fluorídrico de forma a eliminar os silicatos contidos no sedimento. Na<br />

primeira manipulação, foi utilizado o HF a 40% deixando-o reagir durante três<br />

a quatro horas sobre um agitador e, na segunda, foi aplicado o HF a 70%, o<br />

qual permaneceu em repouso, durante 48 horas.<br />

De forma a eliminar os géis fluorissilicatados formados durante a<br />

reacção do resíduo com o ácido fluorídrico, procedeu-se de novo a dois<br />

ataques químicos com uma solução de HCl - 25%, a frio.<br />

O resíduo foi lavado com água destilada a fim de eliminar todos os<br />

excedentes deixados pelos ácidos e, finalmente, filtrado (filtro de 10 µm),<br />

recuperando-se a fracção compreendida entre 10 µm e 150 µm (Heusser &<br />

Stock, 1984).<br />

O resíduo final foi montado sobre lâmina e lamela, juntamente com<br />

algumas gotas de glicerol de forma a obter uma lâmina móvel, a qual<br />

permite a observação dos palinomorfos em todas as suas vistas (polar e<br />

equatorial), facilitando a sua identificação.<br />

b) Identificação e contagem dos palinomorfos<br />

A contagem dos palinomorfos foi efectuada utilizando um<br />

microscópio óptico Zeiss, com objectivas de X40 e X100 (óleo de imersão). A<br />

identificação e determinação polínica foi feita com base nas características<br />

morfológicas apresentadas pelos esporomorfos, que permitem distinguir os<br />

diferentes taxa polínicos. Esses taxa polínicos compreendem famílias, géneros<br />

e tipos polínicos (grupos de espécies ou géneros com características<br />

morfológicas semelhantes). Neste último caso foi utilizado o sufixo “tipo”<br />

(type). Esta determinação taxonómica dos pólens foi efectuada com a<br />

ajuda de uma colecção de referência, existente no UMR CNRS 5805 EPOC<br />

69<br />

F. Naughton, 2007


da Universidade Bordéus 1 e de dois atlas polínicos: Moore et al. (1991) e<br />

Reille (1992).<br />

Nas sondagens marinhas profundas e pouco profundas e amostras<br />

superficiais foi contado, em cada amostra, um valor mínimo de 100 grãos de<br />

pólen (excluindo os grãos de Pinus, plantas aquáticas e os esporos), pelo<br />

menos 20 taxa e mais de 100 grãos de Lycopodium (McAndrew & King, 1976;<br />

Maher, 1981).<br />

Nas sondagens e amostras superficiais estuarinas foram contados em<br />

cada uma das amostras cerca de 300 a 350 grãos de pólen, excluindo as<br />

plantas aquáticas e os esporos (Rull,1987), pelo menos 20 taxa e mais de 100<br />

grãos de Lycopodium (McAndrew & King, 1976; Maher, 1981).<br />

Os esporomorfos danificados (partidos, corroídos, dobrados e<br />

escondidos) foram assinalados como indeterminados. Existem ainda outros<br />

grãos que não foram identificados, os quais foram inseridos na categoria dos<br />

não identificados.<br />

c) Cálculo da percentagem e da concentração esporopolínica<br />

A percentagem de cada taxon, também designada por frequência<br />

polínica relativa (de acordo com Birks & Birks, 1980), foi calculada<br />

relativamente ao somatório de base:<br />

% taxa w = número de pólens contados do taxa w X 100/ somatório de base<br />

Geralmente, nas sequências polínicas continentais, o somatório de<br />

base corresponde à soma de todos os grãos de pólen de árvores, arbustos e<br />

herbáceas, enquanto o somatório total representa a adição do somatório<br />

de base com o somatório das plantas aquáticas, dos esporos, dos<br />

indeterminados e dos grãos não identificados.<br />

No entanto, os grãos de pólen do género Pinus são geralmente sobre-<br />

representados nos sedimentos marinhos (Heusser & Balsam, 1977; Turon, 1984)<br />

70<br />

F. Naughton, 2007


pelo que estes são excluídos da soma de base quando se efectua um<br />

estudo polínico de sondagens marinhas.<br />

Desta forma, nas sondagens marinhas profundas (MD99-2331 e MD03-<br />

2697) o Pinus não foi incluído no somatório de base e a sua percentagem foi<br />

determinada em relação ao somatório total a qual inclui todos os taxa<br />

polínicos, as plantas aquáticas, os esporos, os indeterminados e os grãos não<br />

identificados.<br />

A sondagem marinha pouco profunda Vk03-58Bis foi recolhida<br />

próxima da actual linha de costa e sabendo que normalmente a sobre-<br />

representação de Pinus aumenta à medida que nos afastamos da linha de<br />

costa (Muller, 1959; Groot & Groot, 1966; Bottema & Van Straaten, 1966;<br />

Koreneva, 1966; van der Kaars & de Deckker, 2003), e tendo verificado que<br />

este taxa não aparenta estar sobre-representado relativamente às<br />

sequências polínicas do continente adjacente, admitimos a não sobre-<br />

representação do mesmo na sondagem Vk03-58Bis, pelo que este taxa foi<br />

incluído no somatório de base.<br />

Nas sondagens estuarinas Core1 e Core1B o Pinus foi incluído também<br />

no somatório de base.<br />

A concentração esporopolínica contida nos sedimentos é definida<br />

pela quantidade de esporos e pólens presentes por unidade de volume ou<br />

de massa de sedimento. Tal concentração, é expressa em grãos/cm 3 ou em<br />

grãos/g. Com base na concentração polínica obtida em cada amostra, é-<br />

nos possível avaliar se as variações ocorridas, em termos de percentagens<br />

de taxa, são reais ou se resultam apenas de efeitos estatísticos. De forma a<br />

determinar a concentração de palinomorfos por unidade de volume<br />

(grãos/cm 3 ), utilizou-se a seguinte fórmula:<br />

[taxa w] = n° de pólens contados do taxa w X [exóticos] / n° de<br />

exóticos<br />

d) Diagrama Polínico<br />

A proporção de cada taxa expressa em frequências relativas para<br />

cada nível sedimentar constitui o chamado espectro polínico.<br />

71<br />

F. Naughton, 2007


O diagrama polínico é a representação sucessiva e vertical de todos<br />

os espectros polínicos, permitindo uma melhor visualização da variação<br />

desses espectros ao longo das sequências sedimentares. Para a elaboração<br />

dos diagramas polínicos recorreu-se ao programa PSIMPLOLL (Bennett, 1992).<br />

Estabeleceram-se vários tipos distintos de diagramas polínicos, de<br />

modo a compreender claramente as variações ocorridas ao longo da<br />

sondagem, nomeadamente:<br />

- diagramas polínicos detalhados de percentagens em função da<br />

profundidade, constituído por todos os taxa polínicos, esporos,<br />

indeterminados e não determinados;<br />

- diagramas sintéticos de percentagens em função da profundidade,<br />

constituído por uma selecção dos taxa que permitem decifrar as variações<br />

do coberto vegetal;<br />

-diagramas sintéticos representando uma série de grupos ecológicos distintos<br />

que permitem identificar as variações ecológicas que ocorreram ao longo<br />

do tempo;<br />

- diagramas de concentrações polínicas em função da profundidade.<br />

O diagrama polínico de uma sequência sedimentar permite-nos<br />

reconstituir associações ecológicas passadas quando comparadas com<br />

associações actuais (princípio do uniformitarismo - “o presente é a chave do<br />

passado”), uma vez que, na sua maioria, estas ainda persistem como<br />

géneros actuais (Reille, 1990). A distribuição actual das diferentes formações<br />

vegetais está directamente associada a um dado clima. Assim, conhecendo<br />

as variações que ocorrem nas associações polínicas ao longo do tempo, é-<br />

nos possível deduzir através de um diagrama polínico, as variações<br />

climáticas associadas de forma qualitativa, visto que essas associações<br />

estão em equilíbrio com o clima.<br />

O diagrama polínico é dividido em zonas polínicas de maneira a<br />

facilitar a interpretação das variações do conteúdo esporopolínico em<br />

termos de vegetação e clima. O estabelecimento de uma zona polínica<br />

baseia-se na flutuação qualitativa de pelo menos duas curvas de taxa<br />

72<br />

F. Naughton, 2007


ecologicamente importantes relativamente a uma zona sub ou suprajacente<br />

(Pons & Reille, 1986).<br />

e) Reconstrução quantitativa de parâmetros climáticos baseada nas<br />

associações polínicas<br />

A reconstrução climática quantitativa de uma dada zona pode ser<br />

efectuada em sequências polínicas utilizando a técnica dos melhores<br />

análogos actuais (MAT-Modern Analogue Technique)(Guiot et al., 1989;<br />

Guiot; 1990), com recurso ao programa 3Pbase (Guiot & Goeury, 1996). Esta<br />

técnica baseia-se na selecção de 5 espectros polínicos actuais (análogos<br />

actuais) que apresentam uma assinatura polínica semelhante àquela que<br />

caracteriza a amostra fóssil. A base de referência polínica moderna (modern<br />

pollen database) é constituída por cerca de 1328 espectros polínicos<br />

provenientes da Europa, Eurásia e norte de Africa (Peyron et al., 1998; Peyron<br />

et al., 2005), os quais incluem cerca de 103 taxa polínicos.<br />

Cada um destes espectros polínicos actuais é representado por uma<br />

série de parâmetros climáticos, tais como: precipitação média anual (PANN-<br />

annual mean precipitation), temperatura média anual (TANN-annual mean<br />

temperatures), temperatura média do mês mais frio (MTCO-temperature<br />

mean of the coldest month), e a temperatura média do mês mais quente<br />

(MTCO-temperature mean of the warmest month), os quais foram<br />

previamente interpolados utilizando uma técnica de ANN (Artificial Neural<br />

Network) (Peyron et al., 1998).<br />

1. 5. 4. 2 Indicadores paleoclimáticos marinhos<br />

A associação dos vários indicadores climáticos marinhos (IRD,<br />

associações de foraminíferos planctónicos, δ 18 O de de foraminíferos<br />

planctónicos, alcanonas) efectuada nas sondagens marinhas profundas<br />

(MD99-2331 e MD03-2697), permitiu detectar a variabilidade climática que<br />

ocorreu no Atlântico Norte. Este estudo foi efectuado na fracção grosseira<br />

(>150 μm) recuperada durante a lavagem das amostras utilizadas para<br />

efectuar a análise polínica.<br />

a) Identificação da dinâmica dos Icebergs no Atlântico Norte<br />

73<br />

F. Naughton, 2007


Após o colapso das grandes calotes glaciárias situadas no Hemisfério<br />

Norte, foi libertada uma grande quantidade de icebergues os quais foram<br />

dispersos através das correntes oceânicas. Os icebergues contêm, na sua<br />

base, detritos de dimensão grosseira (geralmente superior a 150 μm)<br />

designados por IRD (Ice-rafeted detritus) e, à medida que estes são<br />

transportados para sul, fundem e libertam os seus sedimentos nos fundos<br />

oceânicos. A contagem de detritos foi efectuada com a ajuda de uma lupa<br />

binocular em pelo menos 10 gramas de sedimento.<br />

b) Evolução da temperatura da massa de água superficial<br />

A reconstrução das condições de temperatura oceânica pode ser<br />

efectuada utilizando uma série de ferramentas distintas:<br />

b. 1) associações de foraminíferos planctónicos<br />

Os foraminíferos planctónicos pertencem ao Reino Protista e são<br />

constituídos por uma carapaça mineralizada carbonatada. A sua<br />

distribuição específica segue um padrão latitudinal de variação da<br />

temperatura da massa de água superficial, e depende das condições de<br />

salinidade local. A presença de foraminíferos planctónicos ao longo da<br />

coluna de água pode ser detectada até cerca de 4 000 m de<br />

profundidade. No entanto, estes organismos são mais abundantes nos<br />

primeiros 200 m de profundidade (próximo da zona fótica).<br />

Foram contados e identificados pelo menos cerca de 400 indivíduos<br />

em cada nível estudado, com base em Kennet & Srinivasan (1983). Os<br />

foraminíferos planctónicos foram agrupados em três associações<br />

bioclimáticas principais nomeadamente: polares (que inclui<br />

Neogloboquadrina pachyderma sinistral), subpolares (que inclui<br />

Neogloboquadrina pachyderma dextra, Globigerina bulloides e Turborotalia<br />

quinqueloba) e tropicais a subtropicais (que inclui ias espécies<br />

temperadas/frias, subtropicais, subtropicais quentes e tropicais, Globorotalia<br />

scitula, G.inflata, G.hirsuta, G.truncatulinoides, G.crassaformis, Globigerinita<br />

glutinata, Globigerina falconensis, G.calida, G.rubescens, G.digitata,<br />

Hastigerina aequilateralis, Orbulina universa, Globigerinoides ruber) (e.g. Bé,<br />

1977; Ottens, 1991; Duprat, 1983).<br />

74<br />

F. Naughton, 2007


O conhecimento sobre a distribuição actual dos foraminíferos<br />

planctónicos no oceano mundial, assim como das condições ecológicas<br />

exigidas pelas diferentes espécies, permite-nos reconstituir as variações da<br />

temperatura da massa de água superficial, durante o inverno (Fevereiro) e o<br />

verão (Agosto), no passado, através da utilização de uma função de<br />

transferência. Esta função de transferência é baseada na técnica dos<br />

melhores análogos modernos e cuja base de dados foi estabelecida por<br />

Pflaumann et al. (1996), e posteriormente melhorada por Elsa Cortijo<br />

(Laboratoire des Sciences du Climat et de l’Environnement-LSCE, Gif-sur-<br />

Yvette, França) e por Josette Duprat (UMR CNRS 5805 EPOC, Universidade de<br />

Bordéus 1, França).<br />

b.2) Estimativa da SST anual obtida a partir das alcanonas<br />

O cocolitóforo Emiliania huxleyi biosintetiza alcanonas sob a forma de<br />

uma série de componentes contendo átomos de carbono (C37-C39) os quais<br />

são compostos por duas ou três ligações (grau de insaturação) (Volkman et<br />

al., 1980; Volkman et al., 1995). O Índice U k 37, obtido através da equação de<br />

Prahl (Prahl et al., 1988), permite quantificar o grau de insaturação de uma<br />

dada série de alcanonas. Experiencias laboratoriais mostraram uma boa<br />

correlação entre o U k 37΄ e a temperatura do crescimento da espécie E.<br />

huxleyi permitindo a sua utilização como um indicador marinho na estimativa<br />

de SST (Prahl et al., 1988; Rostek et al., 1993; Rosell-Melé et al.,1995). Apesar<br />

de existirem ainda dúvidas relativamente ao impacto da sazonalidade no<br />

sinal da SST estimado a partir deste indicador paleoclimático (Sachs et al.,<br />

2000), iremos assumir que os resultados estimados ao longo deste trabalho<br />

representam os valores de SST anual.<br />

A reconstrução da SST anual foi efectuada apenas na sondagem<br />

marinha profunda MD99-2331.<br />

A extracção das alcanonas foi efectuada utilizando um “Dionex<br />

Accelerated Solvent Extractor (ASE-200)” automático. Previamente aos<br />

processos de extracção, a introdução de n-C36 na célula de ASE permite<br />

determinar a concentração de C37. A estimativa da SST foi calculada<br />

posteriormente utilizando a equação de Prahl et al. (1988).<br />

75<br />

F. Naughton, 2007


.3) Determinação dos valores de δ 18 O obtidos a partir das carapaças<br />

de foraminíferos planctónicos do tipo Globigerina bulloides<br />

Os isótopos de oxigénio encontram-se incorporados nas moléculas de<br />

água do mar. As moléculas de água enriquecidas em 16 O (isótopo leve) são<br />

preferencialmente sujeitas a processos de evaporação na zona equatorial e,<br />

posteriormente, transportadas para os pólos, deixando a água do mar<br />

enriquecida em 18 O (isótopo pesado). Durante o seu percurso (da zona<br />

equatorial para os pólos) ocorrem vários processos de condensação e de<br />

precipitação durante os quais o 18 O é removido da atmosfera, chegando<br />

aos pólos empobrecido neste isótopo pesado e enriquecido no isótopo leve.<br />

Ao chegar aos pólos, a precipitação de neve vai enriquecer o gelo em 16 O.<br />

Sabe-se ainda que a composição isotópica do oxigénio (δ 18 O)<br />

contido nas carapaças de foraminíferos planctónicos depende da<br />

temperatura, salinidade (do local) e da composição isotópica do oxigénio<br />

da água do mar. Por esta razão, estas carapaças vão estar enriquecidas em<br />

18 O quando as condições são favoráveis à acumulação de gelo nos pólos.<br />

A determinação de δ 18 O contido nas carapaças de foraminíferos<br />

planctónicos permite-nos determinar aproximadamente a temperatura da<br />

massa de água superficial. Contudo, deve ter-se em consideração o facto<br />

de que por vezes a salinidade local pode afectar ligeiramente o sinal<br />

isotópico. Por essas razões, ao longo deste trabalho foram utilizados 2 a 3<br />

indicadores distintos de SST tal como foi recentemente foi sugerido pelo<br />

grupo MARGO (Multiproxy Approach for the Reconstruction of the Glacial<br />

Ocean surface) (Kucera et al., 2005).<br />

A análise isotópica foi efectuada em foraminíferos planctónicos do<br />

tipo Globigerina bulloides. Os foraminíferos foram retirados da fracção<br />

granulométrica compreendida entre 250 e 315 µm e lavados com água<br />

destilada. A preparação de cada “aliquot” (4 a 10 indivíduos representando<br />

uma média de 80 μg) foi efectuada com a ajuda de um ataque químico<br />

individual no amostrador “Micromass Multiprep autosampler”. As análises<br />

isotópicas foram efectuadas no espectrómetro “Optima Micromass mass<br />

spectrometer” (UMR CNRS 5805 EPOC) e no “delta plus Finnigan” (LSCE). Os<br />

resultados obtidos são expressos versus PDB.<br />

76<br />

F. Naughton, 2007


1. 5. 4. 3 Variações do volume de gelo acumulado nos pólos<br />

Para além dos indicadores paleoclimáticos continentais e marinhos,<br />

foi ainda utilizado o registo de δ 18 O contido nas carapaças de foraminíferos<br />

bentónicos, o qual funciona como indicador de variações no volume de<br />

gelo acumulado nos pólos (Shackleton, 1987).<br />

Contrariamente ao que se passa nas massas de água de superfície,<br />

onde as condições de temperatura variam bastante, as águas de fundo são<br />

menos afectadas por variações de temperatura (Shackleton, 1987).<br />

Contudo, alguns registos mostram uma grande amplitude na variabilidade<br />

do δ 18 O que parece estar relacionada com variações importantes na<br />

temperatura das massas de água profunda (Labeyrie et al., 1987; McManus<br />

et al., 1999; Shackleton et al., 2000b). Uma forma de minimizar estes efeitos<br />

seria estudar zonas onde as variações de temperatura de fundo aparentem<br />

ser mínimas (Labeyrie et al., 1987).<br />

Apesar das dificuldades inerentes às variações da temperatura de<br />

fundo, utilizámos o δ 18 O contido nas carapaças de foraminíferos bentónicos<br />

como indicador de variações do volume de gelo acumulado nos pólos.<br />

A análise isotópica foi efectuada em foraminíferos bentónicos do tipo<br />

Cibicides wuellerstorfi e, o processo laboratorial seguiu a metodologia<br />

descrita para a análise isotópica dos foraminíferos planctónicos.<br />

Referências:<br />

Abrantes, F., Baas, J., Haflidason, H., Rasmussen, T., Klitgaard, D., Loncaric, N. and<br />

Gaspar, L., 1998. Sediment fluxes along the northeastern European Margin: inferring<br />

hydrological changes between 20 and 8 kyr. Marine Geology, 152: 7-23.<br />

Alcara Ariza, F., Asensi Marfil, A., de Bolos y Capdevilla, O., Costa Tales, M., Arco<br />

Aguilar, M., Diaz Gonzales, T.E., Diez Garretas, B., Fernandez Prieto, J.A., Fernandez<br />

Gonzales, F., Izco Sevillando, J., Loidi Arregui, J., Martinez Parras, J.M., Navarro Andres,<br />

F., Ninot I Sugranes, J.M., Peinado Lorca, M., Rivas Martinez, S., Sanchez Mata, D.,<br />

Valle Guitierrez, C., Vigo I Bonada, J. and Wildpret de la Torre, W., 1987. La<br />

vegetación de España. Collection Aula Abierta. Universidad de Alcala de Henares,<br />

544 pp.<br />

Allen, G.P. and Castaing, P., 1977. Carte de répartition des sédiments superficiels sur<br />

le plateau continental du Golfe de Gascogne. Bulletin de l'Institut de Géologie du<br />

Bassin d'Aquitaine. Talence France, 255-260.<br />

Allen, J.R.M., Huntley, B. and Watts, W.A., 1996. The vegetation and climate of<br />

northwest Iberia over the last 14 000 yr. Journal of Quaternary Science, 11: 125-147.<br />

77<br />

F. Naughton, 2007


Alley, R.B. and Ágústsdóttir, A.M., 2005. The 8 k event: cause and consequences of a<br />

major Holocene abrupt climate change. Quaternary Science Reviews, 24: 1123–1149.<br />

Alley, R.B., Meese, D.A., Shuman, C.A., Gow, A.J., Taylor, K.C., Grootes, P.M., White,<br />

J.W.C., Ram, M., Waddington, E.D., Mayewski, P.A. and Zielinski, G.A., 1993. Abrupt<br />

increase in Greenland snow accumulation at the end of the Younger Dryas event.<br />

Nature, 362: 527-529.<br />

Alley, R.B., Mayewski, P.A., Sowers, T., Stuiver, M., Taylor, K.C. and Clark, P.U., 1997.<br />

Holocene climatic instability: a prominent, widespread event 8200 yr ago. Geology,<br />

25: 483-486.<br />

Alley, R.B., Dupont, T.K., Parizek, B.R., Anandakrishnan, S., Lawson, D.E., Larson, G.J.<br />

and Evenson, E.B., 2006. Outburst flooding and the initiation of ice-stream surges in<br />

response to climatic cooling: A hypothesis. Geomorphology, 75: 76– 89.<br />

Álvarez–Salgado, X.A., Figueiras, F.G., Pérez, F.F., Groom, S., Nogueira, E., Borges, A.V.,<br />

Chou, L., Castro, C.G., Moncoiffé, G., Ríos, A.F., Miller, A.E.J., Frankignoulle, M.,<br />

Savidge, G. and Wollast, R., 2003. The Portugal coastal counter current off NW Spain:<br />

new insights on its biogeochemical variability. Progress in Oceanography, 56: 281-321.<br />

Ambar, I. and Fiúza, A., 1994. Some features of the Portugal Current System: A<br />

poleward slope undercurrent, an upwelling related southward flow and an autumnwinter<br />

poleward coastal surface current. 2 nd International Conference on Air-Sea<br />

Interaction and on Meterology and Oceanography of the Coastal Zone. American<br />

Meteorological Society, 286-287.<br />

Ambar, I., Fiúza, A.F.G., Boyd, T. and Frouin, R., 1986. Observations of a warm oceanic<br />

current flowing northward along the coasts of Portugal and Spain during Nov-Dec<br />

1983. EOS Transaction, American Geophysical Union, 67: 1054 pp.<br />

Andersen, A., Koç, N., Jennings, A., and Andrews, J.T., 2004a. Nonuniform response of<br />

the major surface currents in the Nordic Seas to Insolation forcing: implications for the<br />

Holocene climate variability. Paleoceanography, 19, 2003, DOI:<br />

10.1029/2002PA000873.<br />

Andersen, A., Koç, N. and Moros, M., 2004b. A highly unstable Holocene climate in<br />

the subpolar North Atlantic: Evidence from diatoms. Quaternary Science Reviews, 23:<br />

2155-2166.<br />

Andrews, J.T. and Giraudeau, J., 2002. Multi-proxy records showing Holocene<br />

oceanographic variability: the inner N. Iceland Shelf (Húnaflói). Quaternary Science<br />

Reviews, 22: 175-193.<br />

Andrews, J.T. and Tedesco, K., 1992. Detrital carbonate-rich sediments, northwestern<br />

Labrador Sea: implications for ice-sheet dynamics and iceberg rafting (Heinrich)<br />

events in the North Atlantic. Geology, 20: 1087-1090.<br />

Andrews, J.T., Cooper, T.A., Jennings, A.E., Stein, A.B. and Erlenkeuser, H., 1998. Late<br />

Quaternary iceberg-rafted detritus events on the Denmark Strait-Southeast<br />

Greenland continental slope (~65°N): related to North Atlantic Heinrich events?<br />

Marine Geology, 149: 211-228.<br />

Araújo, M. F., Jouanneau, J.M. and Dias, J.M.A., 1994. Chemical characterisation of<br />

the main fine sedimentary deposit at the northwester Portuguese shelf. Gaia, 9: 59-65.<br />

78<br />

F. Naughton, 2007


Araújo, M. F., Jouanneau, J.M., Valerio, P., Barbosa, T., Gouveia, A., Weber, O.,<br />

Oliveira, A., Rodrigues, A. and Dias, J.M.A., 2002. Geochemical tracers of northern<br />

Portuguese estuarine sediments on the Iberian Shelf. Progress in Oceanography, 52:<br />

277-297.<br />

Arhan, M., Colin de Verdière, A. and Memery, L., 1994. The eastern boundary of the<br />

subtropical North Atlantic. Journal of Physical Oceanography. 24: 1295–1316.<br />

Baas, J.H., Mienert, J., Abrantes, F. and Prins, M.A., 1997. Late Quaternary<br />

sedimentation on the Portuguese continental margin: climate-related processes and<br />

products. Palaeogeography, Palaeoclimatology, Palaeoecology, 130: 1-23.<br />

Baas, J.H., Schönfeld, J. and Zahn, R., 1998. Mid-depth oxygen drawdown during<br />

Heinrich events: evidence from benthic foraminiferal community structure, trace-fossil<br />

tiering, and benthic δ 13 C at the Portuguese Margin. Marine Geology, 152: 25-55.<br />

Baldini, J.U.L., McDermott, F. and Fairchild, I.J., 2002. Structure of the 8200-year cold<br />

event revealed by a speleothem trace element record. Science, 296: 2203-2206.<br />

Balsam, W.L. and Heusser, L.E., 1976. Direct correlation of sea surface<br />

paleotemperatures, deep circulation, and terrestrial <strong>paleoclimat</strong>es: foraminiferal and<br />

palynological evidences from two cores off Chesapeake Bay. Marine Geology, 21:<br />

121–147.<br />

Barber, D.C., Dyke, A., Hillaire-Marcel, C., Jennings, A.E., Andrews, J.T., Kerwin, M.W.,<br />

Bilodeau, G., McNeely, R., Southon, J., Morehead, M.D. and Gagnon, J.-M., 1999.<br />

Forcing of the cold event of 8,200 years ago by catastrophic drainage of Laurentide<br />

lakes. Nature, 400: 344-348.<br />

Barboni, D., Harrison, S.P., Bartlein, P.J., Jalut, G., New, M., Prentice, I.C., Sanchez-<br />

Goñi, M.-F., Spessa, A., Davis, B. and Stevenson, A.C., 2004. Relationships between<br />

plant traits and climate in the Mediterranean region: A pollen data analysis. Journal<br />

of Vegetation Science. 15: 635- 646.<br />

Bard, E., Hamelin, B., Fairbanks, R.G. and Zindler, A., 1990a. Calibration of the 14 C<br />

timescale over the past 30,000 years using mass spectrometric U-Th ages from<br />

Barbados corals. Nature, 345: 405-410.<br />

Bard, E., Hamelin, B. and Fairbanks, R.G., 1990b. U-Th ages obtained by mass<br />

spectrometry in corals from Barbados. Nature, 346: 456-458.<br />

Bard, E., Hamelin, B., Arnold, M., Montaggioni, L.F., Cabioch, G., Faure, G. and<br />

Rougerie, F., 1996. Deglacial sea-level record from Tahiti corals and the timing of<br />

global meltwater discharge. Nature, 382: 241-244.<br />

Bard, E., Rostek, F., Turon, J.L. and Gendreau, S., 2000. Hydrological impact of Heinrich<br />

events in the subtropical northeast Atlantic. Science, 289: 1321-1324.<br />

Bard, E., Rostek, F. and Menot-Combes, G., 2004. Radiocarbon calibration beyond<br />

20,000 14C yr B.P. by means of planktonic foraminifera of the Iberian Margin.<br />

Quaternary Research, 61: 204-214.<br />

Barton, E.D., 1998. Eastern boundary of the North Atlantic - northwest Africa and<br />

Iberia. Robinson, A.R., Brink, K. (Eds.), The Sea, 11: 633-657.<br />

79<br />

F. Naughton, 2007


Bauch, H.A., Erlenkeuser, H., Spielhagen, R.F., Struck, U., Matthiessen, J., Thiede, J. and<br />

Heinemeier, J., 2001. A multi-proxy reconstruction of the evolution of deep and<br />

surface waters in the subartic Nordic Seas over the last 30,000 yr. Quaternary Science<br />

Reviews, 20: 659-678.<br />

Bé, A.W.H., 1977. An ecological, zoological and taxonomic review of recent<br />

planctonic Foraminifera. Oeanic Micropaleontology, Volume 1: London, Academic<br />

Press Inc., 1-100 pp.<br />

Belknap, D.F. and Kraft, J.C., 1985. Influence of antecedent geology on stratigraphic<br />

preservation potential and evolution of Delaware’s barrier systems. Marine geology,<br />

63: 235-262.<br />

Bender, M., Sowers, T., Dickson, M.L., Orchardo, J., Grootes, P., Mayewski, P., and<br />

Meese, D.A., 1994. Climate correlations between Greenland and Antarctica during<br />

the past 100,000 years. Nature, 372: 663-666.<br />

Bennett, K.D., 1992. PSIMPOLL – a QuickBasic program that generates Post-script<br />

page description files of pollen diagrams. Newsletter (INQUA commission for the study<br />

of the Holocene: Working group on data-handling methods, 8: 11-12.<br />

Berger, A., 1978. Long-term variations of daily insolation and Quaternary climatic<br />

changes. Journal of Atmospheric Sciences, 35: 2362-2367.<br />

Berger A. and Loutre M.F., 2004. Théorie astronomique des paléoclimats. Comptes<br />

Rendus Geoscience, 336: 701-709.<br />

Bianchi, C. and Gersonde, R., 2004. Climate evolution at the last deglaciation: The<br />

role of the Southern Ocean. Earth and Planetary Science Letters, 228: 407-424.<br />

Bianchi, G.G. and McCave, I.N., 1999. Holocene periodicity in North Atlantic climate<br />

and deep-ocean flow south of Iceland. Nature, 397: 515-517.<br />

Birks, H.J.B. and Birks, H.H., 1980. Paleoecology, Quaternary Paleoecologie. London,<br />

Edward Arnold Limeted, 157-230.<br />

Blanco Castro, E., Casado González, M.A., Costa Tenorio, M., Escribano Bombín, R.,<br />

García Antón, M., Génova Fuster, M., Gómez Manzaneque, F., Moreno Sáis, J.C.,<br />

Morla Juaristi, C., Regato Pajares, P. and Sáiz Ollero, H., 1997. Los Bosques Ibéricos,<br />

Barcelona, 572 pp.<br />

Blunier, T. and Brook, E.J., 2001. Timing of Millennial-Scale Climate Change in<br />

Antarctica and Greenland During the Last Glacial Period. Science, 291: 109-112.<br />

Blunier, T., Chappelaz, J., Schwander, J., Daellenbach, A., Stauffer, B., Stocker, T.F.,<br />

Raynaud, D., Jouzel, J., Clausen, H.B., Hammer, C.U. and Johnsen, S.J., 1998.<br />

Asynchrony of Antarctic and Greenland climate change during the last glacial<br />

period. Nature, 394: 739-743.<br />

Boessenkool, K.P., Brinkhuis, H., Schonfeld, J. and Targarona, J., 2001. North Atlantic<br />

sea-surface temperature changes and the climate of western Iberia during the last<br />

deglaciation; a marine palynological approach. Global and Planetary Change, 30:<br />

33-39.<br />

Bond, G. and Lotti, R., 1995. Iceberg discharges into the North Atlantic on millennial<br />

time scales during the last glaciation. Science, 267: 1005-1010.<br />

80<br />

F. Naughton, 2007


Bond, G., Heinrich, H., Broecker, W., Labeyrie, L., McManus, L., Andrews, J., Huon, S.,<br />

Jantschik, R., Clasen, S., Simet, C., Tedesco, K., Klas, M., Bonani, G. and Ivy, S., 1992.<br />

Evidence for massive discharges of icebergs into the North Atlantic Ocean during the<br />

last glacial period. Nature, 360: 245-249.<br />

Bond, G., Broecker, W., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani,<br />

G., 1993. Correlations between climate records from North Atlantic sediments and<br />

Greenland ice. Nature, 365: 143-147.<br />

Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P.,<br />

Cullen, H., Hajdas, I. and Bonani, G.l., 1997. A pervasive millenial-scale cycle in North<br />

Atlantic Holocene and Glacial Climates. Science, 278: 1257-1266.<br />

Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M., Showers, W., Hoffmann, S.,<br />

Lotti-Bond, R., Hajdas, I. and Bonani, G., 2001. Persistant solar influence on North<br />

Atlantic climate during the Holocene. Science, 294: 2130-2136.<br />

Bottema, S. and van Straaten, L.M.J.U., 1966. Malacology and palynology of two<br />

cores from the Adriatic sea floor. Marine Geolology, 4: 553-564.<br />

Bourillet, J-F., Dubrulle, C., Goubert E., Jouanneau, J.-M., Cortijo, E., Weber, O.,<br />

Lesueur, P. and Vicaire, O., 2005. La Grande Vasière : architecture, mise en place et<br />

estimation des facteurs de son évolution. Ifremer Ed, Colloque Défi Golfe de<br />

Gascogne. Ifremer, Brest, 22-24.<br />

Boyle, E.A., 2000. Is ocean thermohaline circulation linked to abrupt<br />

stadial/interstadial transitions? Quaternary science reviews, 19: 255-272.<br />

Braun-Blanquet, J., Pinto da Silva, A.R., and Rozeira, A., 1956. Résultats de deux<br />

excursions géobotaniques à travers le Portugal septentrional et moyen. II. Chenaies à<br />

feuilles caduques (Quercion occidentale) et chenaies à feuilles persistentes<br />

(Quercion faginea) au Portugal. Agronomia Lusititana, 18: 167–234.<br />

Broecker, W.S., Bond, G. and Klas, M., 1990. A salt oscillator in the glacial Atlantic? 1.<br />

The concept. Paleoceanography, 5: 469-477.<br />

Broecker, W.S., 1991. The great ocean conveyor. Oceanography, 4: 79-89.<br />

Broecker, W.S., 1994. Massive iceberg discharges as triggers for global climate<br />

change. Nature, 372: 421-424.<br />

Broecker, W.S., 2000. Abrupt climate change: causal constraints provided by the<br />

<strong>paleoclimat</strong>e record. Earth-Science Reviews, 51: 137-154.<br />

Calvo, E., Grimalt, J. and Jansen, E., 2002. High resolution U37 K sea temperature<br />

reconstruction in the Norwegian Sea during the Holocene. Quaternary Science<br />

Reviews 21: 1385-1394.<br />

Campbell, I.D., Campbell, C., Apps, M.J., Rutter, N.W. and Bush, A.B.G., 1998. Late<br />

Holocene - 1500 yr climatic periodicities and their implications. Geology, 26: 471-473.<br />

Cane, M. and Clement, A.C., 1999. A role for the tropical Pacific coupled oceanatmosphere<br />

system on Milankovitch and millenial timescales. PartII: global impacts.<br />

Mechanisms of global climate change at millenial time scales. Clark P.U., Webb R.S.,<br />

Keigwin L.D. (edits), Geophys. Monographs, 112, AGU, Washington D.C., 373-383.<br />

81<br />

F. Naughton, 2007


Castaing, P. and Jouanneau, J.-M., 1987. Les apports sédimentaires actuels d’origine<br />

continentale aux océans. Bulletin de l’Institut Géologique du Bassin d’Aquitaine, 41:<br />

53–65.<br />

Cayre, O., Lancelot, Y. and Vincent, E., 1999. Paleoceanographic reconstructions<br />

from planktonic foraminifera off the Iberian Margin: Temperature, salinity and Heinrich<br />

events. Paleoceanography, 14: 384-396.<br />

Chapman, M.R. and Shackleton, N.J., 1998. Millennial-scale fluctuations in North<br />

Atlantic heat flux during the last 150,000 years. Earth and Planetary Science Letters,<br />

159: 57-70.<br />

Chapman, M.R. and Maslin, M.A., 1999. Low-latitude forcing of meridional<br />

temperature and salinity gradient in the subpolar North Atlantic and the growth of<br />

glacial ice sheets. Geology, 27: 875-878.<br />

Chapman, M.R., Shackleton, N.J. and Duplessy, J.C., 2000. Sea surface temperature<br />

variability during the last glacial-interglacial cycle: assessing the magnitude and<br />

pattern of climate change in the North Atlantic. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology, 157: 1 -25.<br />

Chappell, J. and Polach, H., 1991. Postglacial sea-level rise from a coral record at<br />

Huon Península, Papua New Guinea. Nature (London, U. K.), 349:147.<br />

Charles, C.D., Lynch-Stieglitz, J., Ninnemann, U.S. and Fairbanks, R., 1996. Climate<br />

connections between the hemispheres revealed by deep sea sediment core / ice<br />

core correlations. Earth and Planetary Science Letters, 142: 19-27.<br />

Chmura, G.L. and Eisma, D., 1995. A palynological study of surface and suspended<br />

sediment on a tidal flat: Implications for pollen transport and deposition in coastal<br />

waters. Marine Geology, 128:183-200.<br />

Chmura, G.L., Smirnov, A. and Campbell, I.D., 1999. Pollen transport through<br />

distributaries and depositional patterns in coastal waters. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology, 149: 257-270.<br />

Clark, P.U., Licciardi, J.M., MacAyeal, D.R. and Jenson, J.W., 1996. Numerical<br />

reconstruction of a soft-bedded Laurentide Ice Sheet during the last glacial<br />

maximum. Geology, 24: 679-682.<br />

Clark, P.U., Marshall, S.J., Clarke, G.K.C., Hostetler, S.W., Licciardi, J.M. and Teller, J.T.,<br />

2001. Freshwater forcing of abrupt climate change during the last glaciation.<br />

Science, 293: 283-287.<br />

CLIMAP Project Members, 1981. The surface of the ice-age earth. Science, 191:1131-<br />

1137.<br />

Combourieu Nebout, N., Londeix, L., Baudin, F., Turon, J.-L., von Grafnestein, R. and<br />

Zahn, R., 1999. Quaternary marine and continental paleoenvironments in the western<br />

Mediterranean (Site 976, Alboran Sea): Palynological evidence. R. Zahn, Comas, M.C.<br />

and Klaus, A. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 457-<br />

468.<br />

Combourieu-Nebout, N., Turon, J.-L., Zahn, R., Capotondi, L., Londeix, L. and Pahncke,<br />

K., 2002. Enhanced aridity and atmospheric high-pressure stability over the western<br />

82<br />

F. Naughton, 2007


Mediterranean during the North Atlantic cold events of the past 50 ky. Geology, 30:<br />

863-866.<br />

Cortijo, E., Labeyrie, L., Vidal, L., Vautravers, M., Chapman, M., Duplessy, J.C., Elliot,<br />

M., Arnold, M., Turon, J.L. and Auffret, G., 1997. Changes in sea surface hydrology<br />

associated with Heinrich event 4 in the North Atlantic Ocean between 40°N and<br />

60°N. Earth and Planetary Science Letters, 146: 29-45.<br />

Cortijo, E., Yiou, P., Labeyrie, L. and Cremer, M., 1995. Sedimentary record of rapid<br />

climatic variability in the North Atlantic Ocean during the last glacial cycle.<br />

Paleoceanography, 10: 911-926.<br />

Cour, P., Zheng, Z., Duzer, D., Calleja, M. and Yao, Z., 1999. Vegetational and climatic<br />

significance of modern pollen rain in northwestern Tibet. Review of Palaeobotany<br />

and Palynology, 104: 183-204.<br />

Cross, A.T., Thompson, G.G. and Zaitzeff, J.B., 1966. Source and distribution of<br />

palynomorphs in bottom sediments, southern part of Gulf of California. Marine<br />

Geology, 4: 467-524.<br />

Crucifix, M. and Loutre, M.F., 2002. Transient simulations over the last interglacial<br />

period (126-115 kyr BP): feedback and forcing analysis. Climate Dynamics, 19: 417-<br />

433.<br />

Dahl-Jensen, D., Monsegaard, K., Gundestrup, N., Clow, G.D., Johnsen, S.J., Hansen,<br />

A.W. and Balling, N., 1998. Past temperatures directly from the Greenland ice sheet.<br />

Science, 282: 268–271.<br />

Dansgaard, W. et al., 1993. Evidence for general instability of past climate from a 250kyr<br />

ice-core record. Nature, 364: 218-220.<br />

Dalrymple, R.A., Kirby, J.T., Martin, P.A., 1994. Spectral methods for forwardpropagating<br />

water waves in conformally-mapped channels. Applied Ocean<br />

Research, 16: 249-266.<br />

Davis, B.A.S., Brewer, S., Stevenson, A.C., Guiot J. and Data Contributors, 2003. The<br />

temperature of Europe during the Holocene reconstructed from pollen data.<br />

Quaternary Science Reviews, 22: 1701-1716.<br />

de Abreu, L., Shackleton, N.J., Schonfeld, J., Hall, M. and Chapman, M., 2003.<br />

Millennial-scale oceanic climate variability off the Western Iberian margin during the<br />

last two glacial periods. Marine Geology, 196: 1-20.<br />

de Beaulieu, J.L., Andrieu, V., Ponel, P., Reille, M. and Lowe, J.J., 1994. The Weichselian<br />

Late-glacial in southwestern Europe (Iberian Peninsula, Pyrenees, Massif Central,<br />

northern Apennines). Journal of Quaternary Science, 9: 101-107.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3: 16-27.<br />

de Vernal, A., Eynaud, F., Henry, M., Hillaire-Marcel, C., Londeix, L., Mangin, S.,<br />

Matthiessen, J., Marret, F., Radi, T., Rochon, A., Solignac, S. and Turon, J.-L., 2005a.<br />

Reconstruction of sea –surface conditions at middle to high latitudes of the Northern<br />

Hemisphere during the Last Glacial Maximum (LGM) based on dinoflagellate cyst<br />

assemblages. Quaternary Science Reviews, 24: 897-924.<br />

83<br />

F. Naughton, 2007


de Vernal, A., Hillaire-Marcel, C. and Darby, D.A., 2005b. Variability of sea ice cover in<br />

the Chukchi Sea (western Arctic Ocean) during the Holocene.<br />

Paleoceanography,20: 4018, DOI: 10.1029/2005PA001157.<br />

deMenocal, P.B, Ortiz, J., Guilderson, T., Adkins, J., Sarnthein, M., Baker, L. and<br />

Yarusinski, M., 2000. Abrupt onset and termination of the African Humid Period: rapid<br />

climate responses to gradual insolation forcing. Quaternary Science Reviews, 19: 347-<br />

361.<br />

Denton, G.H. and Karlén, W., 1973. Holocene climatic variations: their pattern and<br />

possible cause. Quaternary Research, 3: 155– 205.<br />

Denton, G.H., Alley, R.B., Comer, G.C. and Broecker, W.S., 2005. The role of seasonality<br />

in abrupt climate change. Quaternary Science Reviews, 24: 1159-1182.<br />

Desprat, S., 2005. Réponses climatiques marines et continentales du SW de l'Europe<br />

lors des derniers interglaciaires et des entrées en glaciation. PhD Thesis, Bordeaux 1<br />

University, Bordeaux (France), 282 pp.<br />

Desprat, S., Sánchez Goñi, M.F. and Loutre, M.F., 2003. Revealing climatic variability of<br />

the last three millennia in northwestern Iberia using pollen influx data. Earth and<br />

Planetary Science Letters, 213: 63-78.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., McManus, J.F., Loutre, M.F. Duprat, J.,<br />

Malaizé, B., Peyron, O. and Peypouquet, J.-P., 2005. Is vegetation responsible for<br />

glacial inception during periods of muted insolation changes? Quaternary Science<br />

Reviews, 24: 1361-1374.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., Duprat, J., Malaizé, B. and Peypouquet,<br />

J.-P., 2006. Climatic variability of Marine Isotope Stage 7: direct land-sea-ice<br />

correlation from a multiproxy analysis of a northwestern Iberian margin deep-sea<br />

core. Quaternary Science Reviews, 25: 1010-1026.<br />

Desprat S., Sánchez Goñi M.F., Naughton, F., Turon J.-L., Duprat J., Malaizé B., Cortijo,<br />

E. and Peypouquet J.-P., in press. Climate variability of the last five isotopic<br />

interglacials from direct land-sea-ice correlation. Sirocko, F., Litt, T., Claussen, M.,<br />

Sánchez Goñi, M.F. (eds.), Climate of past interglacials.<br />

Dias, J.M.A., 1987. Dinâmica sedimentar e evolução recente de plataforma<br />

continental Portuguesa setentrional. PhD Thesis, Lisbon University, Lisbon (Portugal),<br />

384 pp.<br />

Dias, J.M.A., Gonzalez, R., Garcia, C. and Diaz del Rio, V., 2002. Sediment distribution<br />

patterns on the Galicia-Minho continental shelf. Progress in Oceanography, 52: 215-<br />

231.<br />

Dietrich, G., Kalle, K., Krauss, W. and Siedler, G., 1980. General Oceanography. An<br />

Introduction, New York, John Wiley and Sons, 626 pp.<br />

Drago, T., Oliveira, A., Magalhães, F., Cascalho, J., Jouanneau, J.-M. and Vitorino, J.<br />

1998. Some evidences of northward fine sediment transport in the Portuguese<br />

continental shelf. Oceanologica Acta, 21: 223-231.<br />

Drago, T., Freitas, C., Rocha, F., Moreno, J., Cachão, M., Naughton, F., Fradique, C.,<br />

Araujo, F., Silveira, T., Oliveira, A., Cascalho, J., and Fatela, F., (in press):<br />

84<br />

F. Naughton, 2007


Palaeoenvironmental evolution of estuarine systems during the last 14000 years- the<br />

case of Douro estuary (NW Portugal). Journal of Coastal Research SI, 39: 7 p.<br />

Duplessy, J.C., Ivanova, E., Murdmaa, I., Paterne, M. and Labeyrie, L., 2001. Holocene<br />

paleoceanography of the northern Barents Sea and variations of the northward heat<br />

transport by the Atlantic Ocean. Boreas, 30: 2– 16.<br />

Dupont, L.M. and Wyputta, U., 2003. Reconstructing pathways of aeolian pollen<br />

transport to the marine sediments along the coastline of SW Africa. Quaternary<br />

Science Reviews, 22: 157-174.<br />

Dupont, L., Jahns, S., Marret, F. and Ning, S., 2000. Vegetation change in equatorial<br />

West Africa: time-slices for the last 150 ka Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 155: 95-122.<br />

Duprat, J., 1983. Les foraminifères planctoniques du Quaternaire Terminal d'un<br />

domaine pericontinental (Golfe de Gascogne, côtes Ouest-Ibériques, Mer<br />

d'Alboran): Ecologie-Biostratigraphie. Bulletin de l'Institut Géologie du Bassin<br />

d’Aquitaine, 33: 71-150.<br />

Edwards, R.L., Warren Beck, J., Burr, G.S., Donahue, D.J., Chappell, J.M.A., Bloom, A.L.,<br />

Druffel, E.R.M. and Taylor, F.W., 1993. A large drop in atmospheric 14 C/ 12 C and<br />

reduced melting in the Younger Dryas, documented with 230 Th ages of corals.<br />

Science, 260: 962-968.<br />

Elliot, M. Labeyrie, L., Bond, G., Cortijo, E., Turon, J.-L., Tisnerat, N. and Duplessy, J.-C.,<br />

1998. Millenial-scale iceberg discharges in the Irminger Basin during the last glacial<br />

period: Relationship with the Heinrich events and environmental settings.<br />

Paleoceanography, 13: 433-446.<br />

Elliot, M., Labeyrie, L. and Duplessy, J.-C., 2002. Changes in North Atlantic deep-water<br />

formation associated with the Dansgaard-Oeschger temperature oscillations (60-10<br />

ka). Quaternary Science Reviews, 21: 1153-1165.<br />

Ellison, C.R.W., Chapman, M.R. and Hall, I.R., 2006. Surface and deep ocean<br />

interactions during the cold climate event 8200 years ago. Science, 312: 1929-1932.<br />

Fairbanks, R., 1989. A 17,000-year glacio-eustatic sea-level record: influence of glacial<br />

melting rates on the Younger Dryas event and deep-ocean circulation. Nature, 342:<br />

637-642.<br />

Fawcett, P.J., Agustsdottir, A.M., Alley, R.B. and Shuman, C.A., 1997. The Younger<br />

Dryas termination and North Atlantic deepwater formation: insights from climate<br />

model simulations and Greenland ice core data. Paleoceanography, 12: 23-28.<br />

Fisk, H.N. and McFarlan, E., Jr., 1955. Late Quaternary Deltaic Deposits of the<br />

Mississippi River (Local Sedimentation and Basin Tectonics). The Crust of the Earth,<br />

Geological Society of America Special Paper 62.<br />

Fiúza, A.F.G., 1984. Hidrologia e Dinâmica das Águas Costeiras Portuguesas. PhD<br />

Thesis, Lisbon University, lisbon (Portugal), 294 pp.<br />

Fiúza, A.F.G., Macedo, M.E. and Guerreiro, M.R., 1982. Climatological space and time<br />

variation of the Portuguese coastal upwelling. Oceanologica Acta, 5: 31-40.<br />

85<br />

F. Naughton, 2007


Flückiger, J., Knutti, R. and White, J.W.C., 2006. Oceanic processes as potential trigger<br />

and amplifying mechanisms for Heinrich events. Paleoceanography, 21: PA2014,<br />

doi:10.1029/2005PA001204<br />

Fradique, C., Cascalho, J., Drago, T., Rocha, F. and Silveira, T., (in press). The Meaning<br />

of Heavy Minerals in the Recent Sedimentary Record of the Douro Estuary (Portugal).<br />

Journal of Coastal Research, ICS2004<br />

Freitas, M.C., and Andrade, C., 1997. Sistemas Fluvio-lagunares a sul da Tejo:<br />

evolução a escalas temporais diferentes. Colectânea de ideias sobre a zona<br />

costeira de Portugal. Associação Eurocoast-Portugal, 525-545.<br />

Freitas, M.C., Andrade, C., and Cruces, A., 2002. The geological record of<br />

environmental changes in southwestern Portuguese coastal lagoons since the<br />

Lateglacial. Quaternary International, 93-94: 161-170.<br />

Freitas, M.C. and Andrade, C., 2001. Marcadores ambientais na evolução dos<br />

espaços lagunares do Sw português durante o holocénico: a formação das barreiras<br />

arenosas e a influência antrópica. Comunicação apresentada na V Reunião do<br />

Quaternário Ibérico, 113-116.<br />

Freitas, M.C., Andrade, C., Rocha, F., Tassinari, C., Munhá, J.M., Cruces, A., Vidinha, J.,<br />

and Marques da Silva, C., 2003. Late glacial and Holocene environmental changes in<br />

Portuguese coastal lagoons 1: the sedimentological and geochemical records of the<br />

Santo André coastal area. The Holocene, 13: 433-446.<br />

Fronval, T., Jansen, E., Bloemendal, J. and Johnsen, S., 1995. Oceanic evidence for<br />

coherent fluctuations in Fennoscandian and Laurentide ice sheets on millenium<br />

timescales. Nature, 374: 443-446.<br />

Frouin, R., Fiúza, A.F.G., Ambar, I. and Boyd, T.J., 1990. Observations of a poleward<br />

surface current off the coasts of Portugal and Spain during winter. Journal of<br />

Geophysical Research, 95: 679-691.<br />

Ganopolski, A. and Rahmstorf, S., 2001. Rapid Changes of glacial climate simulated in<br />

a coupled climate model. Nature, 409: 153-158.<br />

Gil García, M.J., Dorado Valiño, M., Valdeolmillos Rodríguez, A. and Ruiz Zapata, M.B.,<br />

2002. Late-glacial and Holocene palaeoclimatic record from Sierra de Cebollera<br />

(northern Iberian Range, Spain). Quaternary International, 93-94: 13-18.<br />

Gomez-Orellana, L., Ramil-Rego, P. and Munoz Sobrino, C., 1998. Una nueva<br />

secuencia polinica y cronologica par el deposito pleistoceno de Mougas (NW de la<br />

Península Iberia). Revue de Paléobiologie, 17: 35-47.<br />

Goslar, T., Kuc, T., Ralska-Jasiewiczowa, M., Rozanski, K., Arnold, M., Bard, E., van Geel,<br />

B., Pazdur, M., Szeroczynska, K. and Wicik, B., 1993. Hight resolution lacustrine record<br />

of the late glacial/holocene transition in central Europe. Quaternary Science Reviews,<br />

12: 287-294.<br />

Goslar, T., Arnold, M., Tisnerat, L.N., Czenik, J. and Wieckowski, K., 2000. Variations of<br />

Younger Dryas atmospheric radiocarbon explicable without ocean circulation<br />

changes. Nature, 403, 877-880.<br />

Groot, J.J., Groot, C.R., 1966. Pollen spectra from deep-sea sediments as indicators of<br />

climatic changes in southern South America. Marine Geology, 4: 525-537.<br />

86<br />

F. Naughton, 2007


Grousset, F.E. Labeyrie, L., Sinko, J.A., Cremer, M., Bond, G., Duprat, J., Cortijo, E. and<br />

Huon, S., 1993. Patterns of ice-rafted detritus in the glacial North Atlantic (40-55°N).<br />

Paleoceanography, 8: 175-192.<br />

Grousset, F.E., Pujol, C., Labeyrie, L., Auffret, G. and Boelaert, A., 2000. Were the North<br />

Atlantic Heinrich events triggered by the behavior of the European ice sheets?<br />

Geology, 28: 123-126.<br />

Guerreiro, C., Cachão, M. and Drago, T., 2005. Calcareous nannoplankton as a<br />

tracer of marine influence on the NW coast of Portugal over the last 14 000 years.<br />

Journal of Nannoplankton Research, 27: 159-172.<br />

Guiot, J., 1990. Methodology of the last climatic cycle reconstruction from pollen<br />

data. Palaeogeography, Palaeoclimatology, Palaeoecology, 80: 49-69.<br />

Guiot, J., Pons, A., de Beaulieu, J.L. and Reille, M., 1989. A 140.000 years continental<br />

climate reconstruction from two European pollen records. Nature, 338: 309-313.<br />

Guiot, J. and Goeury, C., 1996. PPPBASE, a software for statistical analysis of<br />

paleoecological and <strong>paleoclimat</strong>ological data. Dendrochronologia, 14: 295-300.<br />

Haynes, R.D. and Barton. E.D., 1990. A poleward flow along the Atlantic coast of the<br />

Iberian Peninsula. Journal of Geophysical Research, 95: 11425-11442.<br />

Heinrich, H., 1988. Origin and consequences of cyclic ice rafting in the northeast<br />

Atlantic Ocean during the past 130,000 years. Quaternary Research, 29: 142-152.<br />

Heiri, O., Lotter, A.F., Hausmann, S. and Kienast, F., 2003. A chironomid-based<br />

Holocene summer air temperature reconstruction from the Swiss Alps. The Holocene,<br />

13: 477–484.<br />

Hemming, S., 2004. Heinrich events: massive late Pleistocene detritus layers of the<br />

North Atlantic and their global climate imprint. Review of Geophysics, 42: 1-43.<br />

Hemming, S.R., Broecker, W.S., Sharp, W.D., Bond, G.C., Gwiazda, R.H., McManus, J.F.,<br />

Klas, M. and Hajdas, I., 1998. Provenance of Heinrich layers in core V28-82,<br />

northeastern Atlantic: 40Ar/39Ar ages of ice-rafted hornblende, Pb isotopes in<br />

feldspar grains, and Nd-Sr-Pb isotopes in the fine sediment fraction. Earth and<br />

Planetary Science Letters, 164: 317-333.<br />

Heusser, L.E. and Balsam, W.L., 1977. Pollen distribution in the N.E. Pacific Ocean.<br />

Quaternary Research, 7: 45-62.<br />

Heusser, C.J. and Florer, L.E., 1973. Correlation of marine and continental quaternary<br />

pollen records from the Northeast Pacific and Western Washington. Quaternary<br />

Research, 3: 661-670.<br />

Heusser, L.E. and Shackleton, N.J., 1979. Direct marine-continental correlation:<br />

150,000-year oxygen isotope-pollen record from the North Pacific. Science, 204: 837-<br />

839.<br />

Heusser, L.E. and Stock, C.E., 1984. Preparation techniques for concentrating pollen<br />

from marine sediments and other sediments with low pollen density. Palynology, 8:<br />

225-227.<br />

87<br />

F. Naughton, 2007


Hooghiemstra, H., Lézine, A-M., Leroy, S., Dupont, L. and Marret, F., 2006. Late<br />

Quaternary palynology in marine sediments: A synthesis of the understanding of<br />

pollen distribution patterns in the NW African setting. Quaternary International, 148:<br />

29-44.<br />

Hooghiemstra, H., Stalling, H., Agwu, C.O.C. and Dupont, L.M., 1992. Vegetational<br />

and climatic changes at the northern fringe of the Sahara 250,000-5000 years BP:<br />

evidence from 4 marine pollen records located between Portugal and the Canary<br />

Islands. Review of Palaeobotany and Palynology, 74: 1-53.<br />

Hughen, K.A., Overpeck, J.T., Peterson, L.C. and Trumbore, S., 1996. Rapid climate<br />

changes in the tropical Atlantic region during the last deglaciation. Nature, 380: 51-<br />

54.<br />

Hughen, K.A., Southon, J.R., Lehman, S.J. and Overpeck, J.T., 2000. Synchronous<br />

radiocarbon and climate shifts during the last deglaciation, Science 290: 1951-1954.<br />

Hughen, K.A., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell,<br />

P.G., Buck, C.E., Burr, G., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G.,<br />

Friedrich, M., Guilderson, T.P., Kromer, B., McCormac, F.G., Manning, S., Bronk Ramsey,<br />

C., Reimer, P.J., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J. and Weyhenmeyer, C.E., 2004. Marine04 Marine<br />

Radiocarbon Age Calibration, 0-26 Cal Kyr BP. Radiocarbon, 46: 1059-1086.<br />

Hulbe, C.L., 1997. An ice shelf mechanism for Heinrich Layer production.<br />

Paleoceanography, 12: 711-717.<br />

Hulbe, C.L., MacAyeal, D.R., Denton, G.H., Kleman, J. and Lowell, T.V., 2004.<br />

Catastrophic ice shelf breakup as the source of Heinrich event icebergs.<br />

Paleoceanography, 19: PA1004.<br />

Hurrell, J.W. and Dickson, R.R., 2004. Climate variability over the North Atlantic.<br />

Stenseth N.C., Ottersen G., Hurrell, J.W. and Belgrano A. (eds.), Marine Ecosystems<br />

and Climate Variation - the North Atlantic. Oxford University Press, 15-31.<br />

Huthnance, J.M., Van Aken, H.M., White, M., Barton, E.D., Le Cann, B., Coelho, E.F.,<br />

Fanjul, E.A., Miller, P. and Vitorino, J., 2002. Ocean margin exchange-water flux<br />

estimates. Journal of Marine Systems, 32: 107-137.<br />

Imbrie, J., Boyle E.A., Clemens, S.C., Duffy, A., Howard, W.R., Kukla, G., Kutzbach, J.,<br />

Martinson, D.G., McIntyre, A., Mix, A.C., Molfino, B., Morley, J.J., Peterson, L.C., Pisias,<br />

N.G., Prell, W.L., Raymo, M.E., Shackleton, N.J. and Toggweiler, J.R., 1992. On the<br />

structure and origin of major glaciation cycles, 1. Linear response to Milankovitch<br />

forcing. Paleoceanography, 7: 701-738.<br />

Iversen, J., 1954. The Late-Glacial flora of Denmark and its relation to climate and soil.<br />

Danmarks Geologiske Undersogelse, 80: 87-119.<br />

Johnsen, S.J., Clausen, H.B., Dansgaard, W., Fuhrer, K.,Gundestrup, N., Hammer, C.U.,<br />

Iversen, P., Jouzel, J., Stauffer, B. and Steffensen, J.P.,1992. Irregular glacial<br />

interstadials in a new Greenland ice core. Nature, 359: 311-313.<br />

Johnsen, S., Dahl-Jensen, D., Gundestrup, N., Steffensen, J.P., Clausen, H.B., Miller, H.,<br />

Masson-Delmotte, V., Sveinbjornsdottir, A.E. and White, J., 2001. Oxygen isotope and<br />

palaeotemperature records from six Greenland ice-core stations: Camp Century,<br />

88<br />

F. Naughton, 2007


Dye-3, GRIP, GISP2, Renland and NorthGRIP. Journal of Quaternary Science, 16: 299-<br />

307.<br />

Johnson, R. G. and Lauritzen, S. E., 1995. Hudson Bay - Hudson strait jökulhlaups and<br />

Heinrich events: a hypothesis. Palaeogeography, palaeoclimatology,<br />

palaeoecology, 117: 123-137.<br />

Jouanneau, J.M., Garcia, C., Oliveira, A., Rodrigues, A., Dias, J.A. and Weber, O.,<br />

1998. Dispersal and deposition of suspended sediment on the shelf off the Tagus and<br />

Sado estuaries, S.W. Portugal. Progress in Oceanography, 42: 233-257.<br />

Jouanneau, J.M., Weber, O., Cremer, M. and Castaing, P., 1999. Fine-grained<br />

sediment budget on the continental margin of the Bay of Biscay. Deep Sea Research<br />

Part II: Topical Studies in Oceanography, 46: 2205-2220.<br />

Jouanneau, J.M., Weber, O., Drago, T., Rodrigues, A., Oliveira, A., Dias, J.M.A., Garcia,<br />

C., Schimdt, S., Reyss, J.L., 2002. Recent sedimentation and sedimentary budgets on<br />

the western Iberian shelf. Progress In Oceanography, 52: 261-275.<br />

Kanfoush, S.L., Hodell, D.A., Charles, C.D., Guilderson, T.P., Mortyn, P.G. and<br />

Ninnemann, U.S., 2000. Millenial-scale instability of the Antarctic ice sheet during the<br />

last glaciation. Science, 288: 1815-1818.<br />

Kaplan, M., Wolfe, A. and Miller, G., 2002. Holocene environmental variability in<br />

southern Greenland inferred from lake sediments. Quaternary Research, 58: 149–159.<br />

Kaufman, D.S., Ager, T.A., Anderson, N.J., Anderson, P.M., Andrews, J.T., Bartlein, P.J.,<br />

Brubaker, L.B., Coats, L.L., Cwynar, L.C., Duvall., M.L., Dyke, A.S., Edwards, M.E., Eisner,<br />

W.R., Gejewski, K., Geirsdóttir, A., Hu, F.S., Jennings, A.E., Kaplan, M.R., Kerwin, M.W.,<br />

Lozhkin, A.V., MacDonald, G.M., Miller, G.H., Mock, C.J., Oswald, W.W., Otto-Bliesner,<br />

B.L., Porinchu, D.F., Rühland, K., Smol, J.P., Steig, E.J. and Wolfe, B.B., 2004. Holocene<br />

thermal maximum in the Western Arctic (0 to 180W). Quaternary Science Reviews, 23:<br />

529-560.<br />

Keigwin, L. and Lehman, B., 1994. Deep circulation change linked to Heinrich event 1<br />

and Younger Dryas in a middepth North Atlantic core. Paleoceanography, 9: 185-194.<br />

Keigwin, L.D., Sachs, J.P., Rosenthal,Y. and Boyle, E.A., 2005. The 8200 year B.P. event<br />

in the slope water system, western subpolar North Atlantic. Palaeoceanography, 20:<br />

2003, DOI: 10.1029/2004PA00174.<br />

Kennet, J.P. and Srinivasan, S., 1983. Neogene planktonic foraminifera: a<br />

phylogenetic atlas. Hutchinson Ross, New York, 265 pp.<br />

Kim, J.-H., Schneider, R.R., Müller, P.J. and Wefer, G., 2002. Interhemispheric<br />

comparison of deglacial sea-surface temperature patterns in Atlantic eastern<br />

boundary currents. Earth and Planetary Science Letters, 194: 383-393.<br />

Klitgaard-Kristensen, D., Sejrup, H.P., Haflidason, H., Johnsen, S. and Spurk, M., 1998. A<br />

regional 8200 cal yr BP cooling event in northwest Europe, induced by final stages of<br />

the Laurentide ice sheet deglaciation? Journal of Quaternary Science, 13: 165–169.<br />

Knudsen, K.L., Jiang, H., Jansen, E., Eiriksson, J., Heinemeier, J. and Seidenkrantz, M.S.,<br />

2004. Environmental changes off North Iceland during the deglaciation and the<br />

Holocene: foraminifera, diatoms and stable isotopes. Marine Micropaleontology, 50:<br />

273–305.<br />

89<br />

F. Naughton, 2007


Knutti, R., Flückiger, J., Stocker, T.F. and Timmermann, A., 2004. Strong hemispheric<br />

coupling of glacial climate through freshwater discharge and ocean circulation.<br />

Nature, 430: 851-856.<br />

Koç, N. and Jansen, E., 1994. Response of the high latitude Northern Hemisphere to<br />

orbital climate forcing: evidence from the Nordic Seas. Geology, 22: 523–526.<br />

Koç, N., Jansen, E. and Haflidason, H., 1993. Paleooceanographic reconstruction of<br />

surface ocean conditions in the Greenland, Iceland and Norwegian Seas through the<br />

last 14 ka based on diatoms. Quaternary Science Reviews, 12: 115–140.<br />

Koreneva, E.V., 1966. Marine palynological researches in the U.S.S.R. Marine Geology,<br />

4: 565-574.<br />

Koreneva, E.V., 1971. Spores and pollen in Mediterranean bottom sediments. Funnel,<br />

B.M. and Reidel, W.R. (Eds.), The micropaleontology of Oceans, Cambridge, 361-371.<br />

Kotilainen, A.T. and Shackleton, N.J., 1995. Rapid climate variability in the North<br />

Pacific Ocean during the past 95.000 years. Nature, 377: 323-326.<br />

Kucera, M., Rosell-Melé, A., Schneider, R., Walbroeck, C. and Weilnet, M., 2005.<br />

Mutliproxy approach for the reconstruction of the glacial ocean surface (MARGO).<br />

Quaternary Science Reviews, 24: 813-819.<br />

Kudrass, H. R., Hofmann, A., Doose, H., Emeis, K. and Erlenkeuser, H., 2001. Modulation<br />

and amplification of climatic changes in the Northern Hemisphere by the Indian<br />

summer monsoon during the past 80 k.y. Geology, 29: 63-66.<br />

Kutzbach, J. E. and Gallimore, R. G., 1988. Sensitivity of a coupled atmosphere/mixed<br />

layer ocean model to changes in orbital forcing at 9000 years BP. Journal of<br />

Geophysical Research, 93: 803–821.<br />

Labeyrie, L., Duplessy, J.C. and Blanc, P.L., 1987. Variations in mode of formation and<br />

temperature of oceanic deep water over the past 125,000 yrs. Nature, 327: 477-482.<br />

Lambeck, K., Yokoyama, Y. and Purcell, T., 2002. Into and out of the Last Glacial<br />

Maximum: sea-level change during Oxygen Isotope Stages 3 and 2. Quaternary<br />

Science Reviews, 21: 343-360.<br />

Lambeck, K. and Bard, E., 2000. Sea-level change along the French Mediterranean<br />

coast for the past 30 000 years. Earth and Planetary Science Letters, 175: 203-222.<br />

Lebreiro, S.M., Moreno, J.C., Abrantes, F.F. and Pflaumann, U., 1997. Productivity and<br />

paleoceanographic implications on the Tore Seamount (Iberian Margin) during the<br />

last 225 kyr: Foraminiferal evidence. Paleoceanography, 12: 718-727.<br />

Lebreiro, S.M., Moreno, J.C., McCave, I.N. and Weaver, P.P.E., 1996. Evidence for<br />

Heinrich layers off Portugal (Tore Seamount: 39°N, 12°W). Marine Geology, 131: 47-56.<br />

Lehman, S.J. and Keigwin, L.D., 1992. Sudden changes in North Atlantic circulation<br />

during the last deglaciation. Nature, 356: 757-762.<br />

Leroy, S.A.G. and Dupont, L., 1994. Development of vegetation and continental<br />

aridity in northwestern Africa during the Late Pliocene: the pollen record of ODP Site<br />

658. Palaeogeography, Palaeoclimatology, Palaeoecology, 109: 295-316.<br />

90<br />

F. Naughton, 2007


Lesueur, P., Jouanneau, J-M., Boust, D., Tastet, J-P. and Weber, O., 2001.<br />

Sedimentation rates and fluxes in the continental shelf mud fields in the Bay of Biscay<br />

(France). Continental Shelf Research, 21: 1383-1401.<br />

Leuschner, D.C. and Sirocko, F., 2000. The low-latitude monsoon climate during<br />

Dansgaard-Oeschger cycles and Heinrich Events. Quaternary Science Reviews, 19:<br />

243-254.<br />

Levac, E., de Vernal, A. and Blake W.Jr., 2001. Sea-surface conditions in northernmost<br />

Baffin Bay during the Holocene: palynological evidence. Journal of Quaternary<br />

Science, 16: 353-363.<br />

Levesque, A.J., Mayle, F.E., Walker, I.R. and Cwynar, L.C., 1993. The Amphi-Atlantic<br />

Oscillation: A proposed Late-Glacial climatic event. Quaternary Science Reviews, 12:<br />

629-643.<br />

Lézine, A.M. and Vergnaud-Grazzini, C., 1993. Evidence of forest extension in West<br />

Africa since 22,000 BP: A pollen record from the eastern tropical Atlantic. Quaternary<br />

Science Reviews, 12: 203-210.<br />

Li, C., Wang, P., Sun, H., Zhang, J., Fan, D. and Deng, B., 2002. Late Quaternary<br />

incised-valley fill of the Yangtze delta (China): its stratigraphic framework and<br />

evolution. Sedimentary Geology, 152: 133-158.<br />

Libby, W.F., 1952. Radiocarbon dating. The University of Chicago Press, Chicago,<br />

Illinois, USA 161 pp.<br />

Linick, T.W., Long, A., Damon, P.E. and Ferguson, C.W., 1986. High-Precision<br />

Radiocarbon Dating of Bristlecone Pine from 6554 to 5350 BC. Radiocarbon, 28: 943-<br />

953.<br />

Linsley, B.K. and Thunell, R.C., 1990. The record of deglaciation in the Sulu Sea:<br />

Evidence for Younger Dryas event in the tropical western Pacific. Paleoceanography,<br />

5: 1025-1039.<br />

Lopez-Jamar, E., Cal, R.M., Gonzalez, G., Hanson, R.B., Rey, J., Santiago, G. and<br />

Tenore, K.R., 1992. Upwelling and outwelling effects on the benthic regime of the<br />

continental shelf off Galicia, NW Spain. Journal of Marine Research, 50: 465-488.<br />

Lorenz, S.J., Kim, J.-H., Rimbu, N., Schneider, R.R. and Lohmann, G., 2006. Orbitally<br />

driven insolation forcing on Holocene climate trends: evidence from alkenone data<br />

and climate modelling. Paleoceanography, 21: 1002, DOI: 10.1029/2005PA001152.<br />

Loureiro, J.J., Machado, M.L., Macedo, M.E., Nunes, M.N., Botelho, O.F., Sousa, M.L.,<br />

Almeida, M.C., and Martins, J.C., 1986. Monografias hidrológicas dos principias cursos<br />

de água de Portugal continental. Direcção Geral dos Serviços Hidráulicos, Lisbon,<br />

569 pp.<br />

MacAyeal, D.R., 1993. Binge/purge oscillations of the Laurentide Ice Sheet as a cause<br />

of the North Atlantic's Heinrich events. Paleoceanography, 8: 775-784.<br />

Magny, M., 2004. Holocene climate variability as reflected by mid-European lakelevel<br />

fluctuations, and its probable impact on prehistoric human settlements.<br />

Quaternary International, 113: 65-79.<br />

91<br />

F. Naughton, 2007


Magny, M., Bégeot, C., Guiot, J. and Peyron, O., 2003. Contrasting patterns of<br />

hydrological changes in Europe in response to Holocene climate cooling phases.<br />

Quaternary Science Reviews, 22: 1589–1596.<br />

Magri, D., 1995. Some questions on the late-Holocene vegetation of Europe. The<br />

Holocene, 5: 354-360.<br />

Maher, L.J. Jr., 1981. Statistics for microfossil concentration measurements employing<br />

samples spiked with marker grains. Review of Palaeobotany and Palynology, 32: 153-<br />

191.<br />

Mangerud, J., Andersen, S.T., Berglund, B.E. and Donner, J.J., 1974. Quaternary<br />

stratigraphy of Norden, a proposal for terminology and classification. Boreas, 3: 109-<br />

128.<br />

Mann, D.H. and Peteet, D.M., 1994. Extent and Timing of the Last Glacial Maximum in<br />

Southwestern Alaska. Quaternary Research, 42: 136-148.<br />

Manten, A.A., 1966. Marine palynology in progress. Marine Geology, 4: 385-386.<br />

Marchal, O., Cacho, I., Stocker, T.F., Grimalt, J.O., Calvo, E., Martrat, B., Shackleton,<br />

N., Vautravers, M., Cortijo, E., van Kreveld, S., Andersson, C., Koç, N., Chapman, M.,<br />

Sbaffi, L., Duplessy, J.-C., Sarnthein, M., Turon, J.-L., Duprat, J., and Jansen, E., 2002.<br />

Apparent long-term cooling of the sea surface in the northeast Atlantic and<br />

Mediterranean during the Holocene. Quaternary Science Reviews, 21: 455-483.<br />

Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, Jr. T.C., and Shackleton, N.J.,<br />

1987. Age dating and orbital theory of the Ice Ages: Development of a highresolution<br />

0 to 300,000-year chronostratigraphy. Quaternary Research, 27: 1-29.<br />

Maslin, M.A., Shackleton, N.J. and Pfaumann, U., 1995. Surface water temperature,<br />

salinity, and density changes in the northeast Atlantic during the last 45,000 years:<br />

Heinrich events, deep water formation, and climatic rebounds. Paleoceanography,<br />

10: 527-544.<br />

Mayewski, P.A., Meeker, L.D., Twickler, M.S., Whitlow, S., Yang, Q., Lyons, W.B. and<br />

Prentice, M., 1997. Major features and forcing of high-latitude northern hemisphere<br />

atmospheric circulation using a 110,000-year-long glaciochemical series. Journal of<br />

Geophysical Research, 102: 26345-26366.<br />

Mayewski, P.A., Rohling, E., Stager, C., Karlén, W., Maasch, K., Meeker, L.D., Meyerson,<br />

E., Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F.,<br />

Staubwasser, M. and Schneider, R., 2004. Holocene climate variability. Quaternary<br />

Research, 62: 243-255.<br />

Mazé, J.P., Arhan, M. and Mercier, H., 1997. Volume budget of the eastern boundary<br />

layer off the Iberian Peninsula. Deep-Sea Research, 44: 1543-1574.<br />

McAndrew, J.H. and King, J.E., 1976. Pollen of the North American Quaternary: The<br />

top twenty. Geosciences and Man, 15: 41-49.<br />

McCave, I.N. and Hall, I.R., 2002. Turbidity of waters over the Northwest Iberian<br />

continental margin. Progress in Oceanography, 52: 299-313.<br />

McManus, J.F., Oppo, D.W., and Cullen, J.L., 1999. A 0.5 million year record of<br />

millennial-scale climate variability in the North Atlantic. Science, 283: 971–975.<br />

92<br />

F. Naughton, 2007


Meland, M.Y., Jansen, E. and Elderfield, H., 2005. Constraints on SST estimates for the<br />

northern North Atlantic/Nordic Seas during the LGM. Quaternary Science Reviews, 24:<br />

835-852.<br />

Mix, A.C., Bard, E. and Schneider, R., 2001. Environmental processes of the ice age:<br />

land, oceans, glaciers (EPILOG). Quaternary Science Reviews, 20: 627-657.<br />

Moore, P.D., Webb, J.A. and Collinson, M.E., 1991. Pollen analysis. Blackwell scientific<br />

publication, Oxford, 216 pp.<br />

Moreno, A., Cacho, I., Canals, M., Prins, M.A., Sánchez-Goñi, M-F., Grimalt, J. O. and<br />

Weltje, G. J., 2002. Saharan dust transport and high-latitude glacial climate variability:<br />

the Alboran sea record. Quaternary Research, 58: 318-328.<br />

Moreno, J., Fatela, F., Andrade, C., Cascalho, J., Moreno, F. and Drago, T., 2005.<br />

Living Foraminiferal assemblages from Minho/Coura estuary (Northern Portugal): a<br />

stressfull environment. Thalassas, 21: 17-28.<br />

Moreno, J., Fatela, F., Andrade, C. and Drago, T., 2006. Distribution of “living”<br />

Pseudothurammina limnetis (Scott and Medioli): an occurence on the brackish tidal<br />

marsh of Minho/Coura estuary– Northern Portugal. Révue de Micropaléontologie, 49:<br />

45-53.<br />

Morey, A.E., Mix, A.C. and Pisias, N.G., 2005. Planktonic foraminiferal assemblages<br />

preserved in surface sediments correspond to multiple environmental variables.<br />

Quaternary Science Reviews, 24: 925-950.<br />

Moros, M., Emeis, K., Risebrobakken, B., Snowball, I., Kuipers, A., McManus, J. and<br />

Jansen, E., 2004. Sea surface temperatures and ice rafting in the Holocene North<br />

Atlantic: climate influences on northern Europe and Greenland. Quaternary Science<br />

Reviews, 23: 2113-2126.<br />

Mott, R.J., Grant, D.R., Stea, R. and Occhietti, S., 1986. Lake-glacial climatic oscillation<br />

in Atlantic Canada equivalent to the Allerød/Younger Dryas event. Nature, 323: 247-<br />

250.<br />

Mudie, P. J., McCarthy, M. G., 1994. Late Quaternary pollen transport processes,<br />

western North Atlantic: Data from box models, cross-margin and N-S transects. Marine<br />

Geology, 118: 79-105.<br />

Mudie, P. J., 1982. Pollen distribution in recent marine sediments, eastern Canada.<br />

Canadian Journal of Earth Sciences, 19: 729-747.<br />

Muller, J., 1959. Palynology of recent Orinoco delta and shelf sediments.<br />

Micropaleontology, 5: 1-32.<br />

Muñoz Sobrino, C., Ramil-Rego, P. and Rodríguez Guitián, M.A., 1997. Upland<br />

vegetation in the north-west peninsula after the last glaciation: forest history and<br />

deforestation dynamics. Vegetation History and Archeobotany, 6: 215-233.<br />

Muñoz Sobrino, C., Ramil-Rego, P. and Rodríguez Guitián, M.A., 2001. Vegetation in<br />

the mountains of northwest Iberia during the last glacial-interglacial transition.<br />

Vegetation History and Archeobotany, 10: 7-21.<br />

93<br />

F. Naughton, 2007


Muñoz Sobrino, C., Ramil-Rego, P. and Gomez-Orellana, L., 2004. Vegetation of the<br />

Lago Sanabria area (NW Iberia) since the end of the Pleistocene: a<br />

palaeoecological reconstruction on the basis of two new pollen sequences.<br />

Vegetation History and Archeobotany, 13: 1-22.<br />

Muscheler, R., Beer, J. and Vonmoos, M., 2004. Causes and timing of the 8200 yr BP<br />

event inferred from the comparison of the GRIP 10 Be and the tree ring Δ 14 C record.<br />

Quaternary Science Reviews, 23: 2101-2111.<br />

Narciso, A., Cachao, M. and de Abreu, L., 2006. Coccolithus pelagicus subsp.<br />

pelagicus versus Coccolithus pelagicus subsp. braarudii (Coccolithophore,<br />

Haptophyta): A proxy for surface subarctic Atlantic waters off Iberia during the last<br />

200 kyr. Marine Micropaleontology, 59: 15-34.<br />

Naughton, F., 2002. Evolução Morfodinâmica do Estuário do Douro com base na<br />

análise polínica e sedimentar. MSc thesis Lisbon University, Lisbon (Portugal),. 131 p.<br />

(not published).<br />

Nesje, A. and Dahl, S.O., 2001. The Greenland 8200 cal yr BP event detected in loss-on<br />

ignition profiles in Norwegian lacustrine sediment sequences. Journal of Quaternary<br />

Science, 16: 155–166.<br />

Nesje, A., Dahl, S.O. and Lie, Ø., 2004. Holocene millennial-scale summer temperature<br />

variability inferred from sediment parameters in a non-glacial mountain lake:<br />

Danntjørn, Jotunheimen, central southern Norway. Quaternary Science Reviews, 23:<br />

2183-2205.<br />

Nichol, S. L., Boyd, R., and Penland, S., 1996. Sequence stratigraphy of a coastal-plain<br />

incised valley estuary: Lake Calcasieu, Louisiana. Journal of Sedimentary Research,<br />

66: 847-857.<br />

O’Brien, S.R., Mayewski, P.A., Meeker, L.D., Meese, D.A., Twickler, M.S. and Whitlow,<br />

S.I., 1995. Complexity of Holocene climate as reconstructed from a Greenland ice<br />

core. Science, 270: 1962– 1964.<br />

Oliveira, A., Rocha, F., Rodrigues, A., Jouanneau, J.M., Dias, J.M.A., Weber, O. and<br />

Gomes, C., 2002. Clay mineral from the sedimentary cover from the Northwestern<br />

Iberian shelf. Progress in Oceanography, 52: 233-247.<br />

Oliveira, A., Rodrigues, A., Jouanneau, J.M., Weber, O., Dias, J.A. and Vitorino, J.,<br />

1999. The suspended matter distribution and composition in the northern portuguese<br />

margin. Boletin del Instituto Espanol de Oceanografia, 15: 101-109.<br />

Oppo, D.W. and Lehman, S.J., 1995. Suborbital timescale variability of North Atlantic<br />

Deep Water during the past 200,000 years. Paleoceanography, 10: 901-910.<br />

Ottens, J.J., 1991. Planktonic foraminifera as North Atlantic water mass indicators.<br />

Oceanologia Acta, 14: 123-140.<br />

Pailler, D. and Bard, E., 2002. High frequency palaeoceanographic changes during<br />

the past 140,000 yr recorded by the organic matter in sediments of the Iberian<br />

margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 181: 431-452.<br />

Papa, B., Mysak, L. and Wang, Z., 2006. Intermittent ice sheet discharge events in<br />

northeastern North America during the last glacial period. Climate Dynamics, 26: 201-<br />

216.<br />

94<br />

F. Naughton, 2007


Peck, R.M., 1973. Pollen budget studies in a small Yorkshire catchment. In Birks, H.J.B.<br />

and West, R.G. Eds. Quaternary Plant Ecology, The 14 th symposium of the British<br />

Ecological Society, University of Cambridge, Blackwell Scientific Publications, 43-60.<br />

Peinado Lorca, M. and Martínez-Parras, J.M., 1987. Castilla-La Mancha. In: M.R.M.<br />

Peinado Lorca, S. (Editor), La vegetación de España. Universidad de Alcalá de<br />

Henares, Alcala de Henares, 163-196.<br />

Peñalba, C., 1994. The history of the Holocene vegetation in northern Spain from<br />

pollen analysis. Journal of Ecology, 82: 815-832.<br />

Peñalba, M.C., Arnold, M., Guiot, J., Duplessy, J.-C. and de Beaulieu, J.-L., 1997.<br />

Termination of the Last Glaciation in the Iberian Peninsula inferred from the pollen<br />

sequence of Quintanar de la Sierra. Quaternary Research, 48: 205-214.<br />

Pérez-Obiol, R. and Julià, R., 1994. Climatic change on the Iberian Peninsula recorded<br />

in a 30,000-yr pollen record from Lake Banyoles. Quaternary Research, 41: 91-98.<br />

Peteet, D.M., Vogel, J.S., Nelson, D.E., Southon, J.R., Nickmann, R.J. and Heusser, L.E.,<br />

1990. Younger Dryas climatic reversal in northeastern USA? AMS ages for an old<br />

problem. Quaternary Research, 33: 219-230.<br />

Peteet, D.M., Daniels, R.A., Heusser, L. E., Vogel, J.S., Southon J.R. and Nelson, D.E.,<br />

1993. Late-glacial pollen, macrofossils and fish remains in northeastern U.S.A.: The<br />

Younger Dryas oscillation. Quaternary Science Reviews, 12: 597-612.<br />

Peterson, L.C., Haug, G.H., Hughen, K.A. and Röhl, U., 2000. Rapid changes in the<br />

hydrologic cycle of the tropical Atlantic during the last Glacial. Science, 290: 1947-<br />

1951.<br />

Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, J.-M., Basile, I., Bender, M.,<br />

Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, M., Legrand, M.,<br />

Lipenkov, Y., Lorius, C., Pepin, L., Ritz, C., Saltzman, E., and Stievenard, M., 1999.<br />

Climate and atmospheric history of the past 420,000 years from the Vostok ice core,<br />

Antarctica. Nature, 399: 429-436.<br />

Peyron, O., Bégeot, C., Brewer, S., Heiri, O., Magny, M., Millet, L., Ruffaldi, P., Van<br />

Campo E. and Yu, G., 2005. Lateglacial climatic changes in Eastern France (Lake<br />

Lautrey) from pollen, lake-levels, and chironomids. Quaternary Research, 64: 197-211.<br />

Peyron, O., Guiot, J., Cheddadi, R., Tarasov, P., Reille, M., de Beaulieu, J.-L., Bottema,<br />

S. and Andrieu, V., 1998. Climatic Reconstrustion in Europe for 18,000 yr BP from Pollen<br />

Data. Quaternary Research, 49: 183-196.<br />

Pflaumann, U., Duprat, J., Pujol, C. and Labeyrie, L.D., 1996. SIMMAX: A modern<br />

analog technique to deduce Atlantic sea surface temperatures from planktonic<br />

foraminifera in deep-sea sediments. Paleoceanography, 11: 15-35.<br />

Pingree, R.D. and Le Cann, B., 1990. Structure, strength and seasonality of the slope<br />

currents in the Bay of Biscay region. Journal of the Marine Biological Association of the<br />

United Kingdom, 70: 857-885.<br />

Pingree, R.D. and Le Cann, B., 1992. Three anti-cyclonic slope water oceanic eddies<br />

(SWODDIES) in the southern Bay of Biscay. Deep-Sea Research, 39: 1147-1175.<br />

95<br />

F. Naughton, 2007


Polunin, O. and Walters, M., 1985. A guide to the vegetation of Britain and Europe.<br />

Oxford University Press, New York, 238 pp.<br />

Pons, A. and Reille, M., 1986. Nouvelles recherches pollenanalytiques à Padul<br />

(Granada): La fin du dernier glaciaire et l’Holocene. E. López-Vera (Ed.). Quaternary<br />

climate in Western Mediterranean, University Autónoma de Madrid, 405-420.<br />

Pons, A. and Reille, M., 1988: The Holocene-and Upper Pleistocene pollen record from<br />

Padul (Granada, Spain): a new study. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 66: 243-263.<br />

Prahl, F.G., Muehlhausen, L.A. and Zahnle, D.L., 1988. Further evaluation of long-chain<br />

alkenones as indicators of paleoceanographic conditions. Geochimica et<br />

Cosmochimica Acta, 52: 2303-2310.<br />

Rahmstorf, S., 1995. Bifurcations of the Atlantic thermohaline circulation in response to<br />

changes in the hydrological cycle. Nature, 378: 145–149.<br />

Rahmstorf, S., 2002. Ocean circulation and climate during the past 120,000 years.<br />

Nature, 419: 207-214.<br />

Rasmussen, T.L., Thomsen, E., van Weering, T.C.E. and Labeyrie, L., 1996. Rapid<br />

changes in surface and deep water conditions at the Faeroe Islands Margin during<br />

the last 58 ka. Paleoceanography, 11: 757-771.<br />

Rasmussen, S.O., van Weering, T.C.E. and Labeyrie, L., 1997. Climatic instability, ice<br />

sheets and ocean dynamics at high northern latitudes during the last glacial period<br />

(58-10 ka BP). Quaternary Science Reviews, 16: 71-80.<br />

Reille, M., 1990. Leçons de palynologie et d'analyse pollinique. Centre National de la<br />

Recherche Scientifique, Paris, 206 pp.<br />

Reille, M., 1992. Pollen et spores d'Europe et d'Afrique du Nord. Laboratoire de<br />

botanique historique et palynologie, Marseille, 520 pp.<br />

Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell, P.G.,<br />

Buck, C.E., Burr, G.S., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G., Friedrich,<br />

M., Guilderson, T.P., Hog, A.G., Hughen, K.A., Kromer, B., McCormac, G., Manning, S.,<br />

Bronk Ramsey, C., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J., Weyhenmeyer, C.E., 2004. INTCAL04 terrestrial<br />

radiocarbon age calibration. Radiocarbon, 46: 1029-1058.<br />

Renssen, H., Goosse, H., Fichefet, T. and Campin, J.M., 2001. The 8.2 kyr BP event<br />

simulated by a global atmosphere–sea–ice–ocean model. Geophysical Research<br />

Letters, 28: 1567-1570.<br />

Renssen, H., Goosse, H., Fichefet, T., Brovkin, V., Driesschaert, E., and Wolk, F., 2005.<br />

Simulating the Holocene climate evolution at northern high latitudes using a coupled<br />

atmosphere-sea ice-ocean-vegetation mode. Climate Dynamics, 24: 23-43.<br />

Revel, M., Cremer, M., Grousset, F.E. and Labeyrie, L., 1996. Grain-size and Sr-Nd<br />

isotopes as tracer of paleo-bottom current strength, Northeast Atlantic Ocean.<br />

Marine Geology, 131: 233-249.<br />

Richter, T.O., Lassen, S., van Weering, T.C.E. and de Haas, H., 2001. Magnetic<br />

susceptibility patterns and provenance of ice-rafted material at Feni Drift, Rockall<br />

96<br />

F. Naughton, 2007


Trough: implications for the history of the British-Irish ice sheet. Marine Geology, 173:<br />

37-54.<br />

Rios, A.F., Pérez, F.F. and Fraga, F., 1992. Water masses in the upper and middle North<br />

Atlantic east of the Açores. Deep sea Research, 39: 645-658.<br />

Risebrobakken, B., Jansen, E., Andersson, C., Mjelde, E. and Hevroy, K., 2003. A highresolution<br />

study of Holocene <strong>paleoclimat</strong>ic and paleoceanographic changes in the<br />

Nordic seas. Paleoceanography, 18, 1017, doi: 10.1029/2002PA000764.<br />

Rohling, E.J. and Pälike, H., 2005. Centennial-scale climate cooling with a sudden<br />

cold event around 8,200 years ago. Nature, 434: 975-979.<br />

Rosell-Melé, A., Carter, J.F., Parry, A.T., and Eglinton, G., 1995. Determination of the<br />

UK37 Index in Geological Samples. Analytical Chemistry, 67: 1283-1289.<br />

Rosell-Mele, A., Maslin, M.A., Maxwell, J.R. and Schaeffer, P., 1997. Biomarker<br />

evidence for "Heinrich" events. Geochimica et Cosmochimica Acta, 61: 1671-1678.<br />

Rosenthal, J.S., Helvey, R.A., Battalino, T.E., Fisk, C. and Greiman, P.W., 2003. Ozone<br />

transport by mesoscale and diurnal wind circulations across southern California.<br />

Atmospheric Environment, 37: 51-71.<br />

Rossignol, M., 1969. Sédimentation palynologique dans le domaine marin quaternaire<br />

de Palestine. M.N.H.N. (Eds.). Notes et Mémoires du Moyen-Orient,10: Paris, 272 pp.<br />

Rossignol-Strick, M. and Duzer, D., 1979. West African vegetation and climate since 22<br />

500 BP from deep-sea cores palynology. Pollen Spores, 21: 105-134.<br />

Rostek, F., Ruhland, G., Bassinot, F.C., Müller, P.J., Labeyrie, L.D., Lancelot, Y. and Bard,<br />

E., 1993. Reconstructing sea surface temperature and salinity using δ 18 O and<br />

alkenone records, Nature, 364: 319-321.<br />

Roucoux, K.H., de Abreu, L., Shackleton, N.J. and Tzedakis, P.C., 2005. The response of<br />

NW Iberian vegetation to North Atlantic climate oscillations during the last 65 kyr.<br />

Quaternary Science Reviews, 24: 1637-1653.<br />

Roucoux, K.H., Shackleton, N.J., de Abreu, L., Schönfeld, J. and Tzedakis, P.C., 2001.<br />

Combined marine proxy and pollen analyses reveal rapid Iberian vegetation<br />

response to North Atlantic millennial-scale climate oscillations. Quaternary Science<br />

Reviews, 56: 128-132.<br />

Roy, P.S, Cowell, P.J. and Thom, B.G., 1995. Wave-dominated coasts. In: Cater,<br />

R.W.G., Woodroffe, C.D. (Eds.), Coastal Evolution. Cambridge University Press: 121-186.<br />

Ruddiman, W.F. and McIntyre, A., 1981. Oceanic mechanisms for amplification of the<br />

23000-year ice-volume cycle. Science, 212: 617-627.<br />

Ruddiman, W.F., 2001. Earth's Climate Past and Future. W. H. Freeman, New York, 465<br />

pp.<br />

Ruiz Zapata, M.B., Gil Garcia, M.J., Dorado Valiño, M., Valdeolmillos Rodrigues, A.,<br />

Vegas, J. and Pérez-Gonzalez, A., 2002. Clima y vegetation durante el tardiglaciar y<br />

el Holoceno en la Sierra de Neilla (Sistema Ibérico Noroccidental). Cuaternario y<br />

Geomorfologia, 16: 9-20.<br />

97<br />

F. Naughton, 2007


Rull, V., 1987. A note on pollen counting in paleoecology. Pollen Spores, 29: 471-480.<br />

Sachs, J.P., Schneider, R.R., Eglinton, T.I., Freeman, K.H., Ganssen, G., McManus, J.F.<br />

and Oppo, D.W., 2000. Alkenones as paleoceanographic proxies. Geochemistry,<br />

Geophysics, Geosystems, 1: 2000GC000059.<br />

Sánchez Goñi, M.F. and Hannon, G., 1999. High altitude vegetational patterns on the<br />

Iberian Mountain chain (north-central Spain) during the Holocene. The Holocene, 9:<br />

39-57.<br />

Sánchez Goñi, M.F., 2006. Interactions végétation-climat au cours des derniers 425<br />

000 ans en Europe Occidentale. Le message du pollen des archives marines.<br />

Quaternaire, 17: 1-23.<br />

Sánchez Goñi, M.F., Cacho, I., Turon, J.-L., Guiot, J., Sierro, F.J., Peypouquet, J.-P.,<br />

Grimalt, J.O. and Shackleton, N.J., 2002. Synchroneity between marine and terrestrial<br />

responses to millennial scale climatic variability during the last glacial period in the<br />

Mediterranean region. Climate Dynamics, 19: 95-105.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L. and Gendreau, S., 2000. European climatic<br />

response to millenial-scale climatic changes in the atmosphere-ocean system during<br />

the Last Glacial period. Quaternary Research, 54: 394-403.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L. and Shackleton, N.J., 1999. High resolution<br />

palynological record off the Iberian margin: direct land-sea correlation for the Last<br />

Interglacial complex. Earth and Planetary Science Letters, 171: 123-137.<br />

Sánchez Goñi, M.F., Loutre, M.F., Crucifix, M., Peyron, O., Santos, L., Duprat, J.,<br />

Malaizé, B., Turon, J.-L., Peypouquet, J.-P., 2005. Increasing vegetation and climate<br />

gradient in Western Europe over the Last Glacial Inception (122–110 ka): data-model<br />

comparison. Earth and Planetary Science Letters, 231: 111-130.<br />

Santos, L., Vidal Romani, J.R. and Jalut, G., 2000. History of vegetation during the<br />

Holocene in the Courel and Queixa Sierras, Galicia, northwest Iberian Peninsula.<br />

Journal of Quaternary Science, 15: 621-632.<br />

Schönfeld, J., Zahn, R. and de Abreu, L., 2003. Surface and deep water response to<br />

rapid climate changes at the Western Iberian Margin. Global and Planetary Change,<br />

36: 237-264.<br />

Schulz, H., von Rad, U. and Erlenkeuser, H., 1998. Correlation between Arabian Sea<br />

and Greenland climate oscillations of the past 110,000 years. Nature, 393: 54-57.<br />

Schulz, M., Paul, A. and Timmermann, A., 2004. Glacial-interglacial contrast in climate<br />

variability at centennial-to-millennial timescales: observations and conceptual model.<br />

Quaternary Science Reviews, 23: 2219-2230.<br />

Scourse, J.D., Hall, I.R., McCave, I.N., Young, J.R. and Sugdon, C., 2000. The origin of<br />

Heinrich layers: evidence from H2 for European precursor events. Earth and Planetary<br />

Science Letters, 182: 187-195.<br />

Seidov, D., and Maslin, M., 1999. North Atlantic Deep Water Circulation collapse<br />

during Heinrich events. Geology, 27: 23-26.<br />

98<br />

F. Naughton, 2007


Seppä, H. and Poska, A., 2004. Holocene annual mean temperature changes in<br />

Estonia and their relationship to solar insolation and atmospheric circulation patterns.<br />

Quaternary Research, 61: 22-31.<br />

Seppä, H., Hammarlund, D. and Antonsson, K., 2005. Low-frequency and highfrequency<br />

changes of temperature and effective humidity during the Holocene in<br />

South central Sweden: Implications for atmospheric and oceanic forcing of climate.<br />

Climate Dynamics, 25: 285-297.<br />

Severinghaus, J.P. and Brook, E.J., 1999. Abrupt Climate Change at the End of the<br />

Last Glacial Period Inferred from Trapped Air in Polar Ice. Science, 286: 930-934.<br />

Shackleton, N.J. and Opdyke, N.D., 1973. Oxygen isotope and paleomagnetic<br />

stratigraphy of Equatorial Pacific core V28-238: oxygen isotope temperatures and ice<br />

volumes a 105 year and 106 year scale. Quaternary Research, 3: 39-55.<br />

Shackleton, N.J., 1987. Oxygen isotopes, ice volume and sea level. Quaternary<br />

Science Reviews, 6: 183-190.<br />

Shackleton, N.J., Hall, M.A. and Vincent, E., 2000a. Phase relationships between<br />

millennial scale events 64,000-24,000 years ago. Paleoceanography, 15: 565-569.<br />

Shackleton, N.J., 2000b. The 100,000-year ice age cycle identified and found to lag<br />

temperature, carbon dioxide, and orbital eccentricity. Science, 289: 1897-1902.<br />

Sikes, E.L., Farrington, J.W., and Keigwin, L.D., 1991. Use of the alkenone unsaturation<br />

ratio Uk 37 to determine past sea surface temperatures: core-top SST calibrations and<br />

methodology considerations. Earth and Planetary Science Letters, 104: 36-47.<br />

Solignac, S., de Vernal, A. and Hillaire-Marcel, C., 2004. Holocene sea-surface<br />

conditions in the North Atlantic—contrasted trends and regimes in the western and<br />

eastern sectors (Labrador Sea vs. Iceland Basin). Quaternary Science Reviews, 23:<br />

319–334.<br />

Sowers, T. and Bender, M., 1995. Climate records covering the last deglaciation.<br />

Science, 269: 210-214.<br />

Sprangers, M., Dammers, N., Brinkhuis, H., van Weering, T.C.E. and Lotter, A.F., 2004.<br />

Modern organic-walled dinoflagellate cyst distribution offshore NW Iberia; tracing the<br />

Upwelling System. Review of palaeobotany and Palynolology, 128: 97-106.<br />

Spurk, M., Leuschner, H.H., Baillie, M.G.L., Briffa, K.R. and Friedrich, M., 2002.<br />

Depositional frequency of German subfossil oaks: climatically and non-climatically<br />

induced fluctuations in the Holocene. The Holocene, 12: 707–715.<br />

Stanley, E.A., 1966. The application of palynology to oceanology with reference to<br />

the northwestern Atlantic. Deep-Sea Research, 13: 921-939.<br />

Stuiver, M., Braziunas, T.F., Becker, B. and Kromer, B., 1991. Climatic, solar, oceanic,<br />

and geomagnetic influences on late-glacial and Holocene atmospheric 14 C/ 12 C<br />

change. Quaternary Research, 35: 1-24<br />

Stuiver, M. and Reimer, P., 1993. Extended 14 C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

99<br />

F. Naughton, 2007


Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A., Kromer, B.,<br />

McCormac, G., and Van der Plicht, J., 1998. INTCAL98 Radiocarbon age calibration,<br />

24.000-0 cal BP. Radiocarbon, 40: 1041-1083.<br />

Stuiver, M., Reimer, P.J. and Reimer, R.W., 2005. CALIB 5.0. (WWW program and<br />

documentation).<br />

Telford, R.J., Heegaard, E. and Birks, H.J.B., 2004. The intercept is a poor estimate of a<br />

calibrated radiocarbon age. The Holocene, 14: 296-298.<br />

Teller, J.T., Leverington, D.W. and Mann, J.D., 2002. Freshwater outbursts to the oceans<br />

from glacial Lake Agassiz and their role in climate change during the last<br />

deglaciation. Quaternary Science Reviews, 21: 879–887.<br />

Thouveny, N., Moreno, E., Delanghe, D., Candon, L., Lancelot, Y. and Shackleton,<br />

N.J., 2000. Rock magnetic detection of distal ice-rafted debries: clue for the<br />

identification of Heinrich layers on the Portuguese margin. Earth and Planetary<br />

Science Letters, 180: 61-75.<br />

Tinner, W. and Lotter, A.F., 2001. Central European vegetation response to abrupt<br />

climate change at 8.2 ka. Geology, 29: 551–554.<br />

Tinner, W. and Lotter, A.F., 2006. Holocene expansions of Fagus silvatica and Abies<br />

alba in Central Europe: where are we after eight decades of debate? Quaternary<br />

Science Reviews, 25: 526-549.<br />

Torres, R., Barton, E. D., Miller, P. and Fanjul, E., 2003. Spatial pattern of wind and sea<br />

surface temperature in the Galician upwelling region. Journal of Geophysical<br />

Research, 108: 3130.<br />

Traverse, A. and Ginsburg, R. N., 1966. Palynology of the surface sediments of Great<br />

Bahama Bank, as related to water movement and sedimentation. Marine Geology, 4:<br />

417-459.<br />

Turon, J.-L., 1984. Le palynoplancton dans l'environnement actuel de l'Atlantique<br />

nord-oriental. Evolution climatique et hydrologique depuis le dernier maximum<br />

glaciaire. Mémoires de l'Institut de Géologie du bassin d'Aquitaine, 17. Université de<br />

Bordeaux I, Bordeaux, 313 pp.<br />

Turon, J.-L., Lézine, A.-M. and Denèfle, M., 2003. Land-sea correlations for the last<br />

glaciation inferred from a pollen and dinocyst record from the Portuguese margin.<br />

Quaternary Research, 59: 88-96.<br />

Tzedakis, P.C., Frogley, M.R. and Heaton, T.H.E., 2003. Last interglacial conditions in<br />

southern Europe: evidence from Ioannina, northwest Greece. Global and Planetary<br />

Change, 36: 157-170.<br />

Tzedakis, P.C., Roucoux, K.H., de Abreu, L. and Shackleton, N.J., 2004. The duration of<br />

forest stages in southern Europe and interglacial climate variability. Science, 306:<br />

2231-2235.<br />

Valdès, C.M. and Gil Sanchez, L., 2001. La transformación histórica del paisaje forestal<br />

en Galicia, Ministerio de Medio Ambiente, 159 pp.<br />

100<br />

F. Naughton, 2007


Van Campo, M.M., 1984. Relations entre la végétation de l'Europe et les<br />

températures de surface océaniques après le dernier maximum glaciaire. Pollen et<br />

Spores, XXVI(3-4): 497-518.<br />

van der Hammen, T and Hooghiemstra, H., 1995. The EL abra stadial, a younger dryas<br />

equivalent in Colombia. Quaternary Science Reviews, 14: 841-851.<br />

Van der Hammen, T., Wijmstra, T.A. and Zagwijn, W.H., 1971. The floral record of the<br />

late Cenozoic of Europe. The late Cenozoic glacial ages (K.K. Turekian, ed.), Yale<br />

University Press, New Haven: 391-423.<br />

Van der Kaars, S., 2001. Pollen distribution in marine sediments from the south-eastern<br />

Indonesian waters. Palaeogeography, Palaeoclimatology, Palaeoecology, 171: 341-<br />

361.<br />

Van der Kaars, S. and de Deckker, P., 2003. Pollen distribution in marine surface<br />

sediments offshore Western Australia. Review of Palaeobotany and Palynology, 124:<br />

113-129.<br />

Van der Knaap, W.O. and Van Leeuwen, J.F.N., 1995. Holocene vegetation<br />

succession and degradation as responses to climatic change and human activity in<br />

the Serra de Estrela, Portugal. Review of Palaeobotany and Palynology, 89: 153-211.<br />

Van Geel, B., van der Plicht, J. and Renssen, H., 2003. Major Δ 14 C excursions during<br />

the late glacial and early Holocene: changes in ocean ventilation or solar forcing of<br />

climate change? Quaternary International, 105: 71-76.<br />

Van Kreveld, S., Sarnthein, M., Erienkeuser, H., Grootes, P., Jung, S., Nadeau, M.J.,<br />

Pflaumann, U., and Voelker, A., 2000. Potential links between surging ice sheets,<br />

circulation changes, and the Dansgaard-Oeschger cycles in the Irminger Sea, 60-18<br />

kyr. Paleoceanography, 15: 425-442.<br />

Van Weering, T.C.E., McCave, I.N., 2002. Dedication to Roland Wollast. Progress In<br />

Oceanography, 52: 121-122.<br />

Van Weering, T.C.E., de Stigter, H.C., Boer, W. and Haas, H., 2002. Recent sediment<br />

transport and accumulation on the NW Iberian margin. Progress In Oceanography,<br />

52: 349-371.<br />

Vanney, J.R. and Mougenot, D., 1981. La plate-forme continentale du Portugal et les<br />

provinces adjacentes: Analyse Géomorphologique, Memórias dos serviços<br />

geológicos de Portugal, 145 pp.<br />

Verbitsky, M. and Saltzman, B., 1995. A diagnostic analysis of Heinrich glacial surge<br />

events. Paleoceanography, 10: 59-65.<br />

Vidal, L., Labeyrie, L., Cortijo, E., Arnold, M., Duplessy, J. C., Michel, E., Becqué, S. and<br />

van Weering, T. C. E., 1997. Evidence for changes in the North Atlantic Deep Water<br />

linked to meltwater surges during the Heinrich events. Earth and Planetary Science<br />

Letters, 146: 13-27.<br />

Vidal, L., Schneider, R.R., Marchal, O., Bickert, T., Stocker, T.F. and Wefer, G., 1999. Link<br />

between the North and South Atlantic during the Heinrich events of the last glacial<br />

period. Climate Dynamics, 15: 909-919.<br />

101<br />

F. Naughton, 2007


Vitorino, J., Oliveira, A., Jouanneau, J.M. and Drago, T., 2002. Winter dynamics on the<br />

northern Portuguese shelf. Part1: physical processes. Progress in Oceanography, 52:<br />

129-153.<br />

Voelker, A.H.L., Sarnthein, M., Grootes, P.M., Erlenkeuser, H., Laj, C., Mazaud, A.,<br />

Nadeau, M.-J., and Schleicher, M., 1998. Correlation of marine 14 C ages from the<br />

Nordic seas with the GISP2 isotope record: Implications for 14 C calibration beyond 25<br />

ky BP. Radiocarbon, 40: 517-534.<br />

Voelker, A.H.L. and workshop participants, 2002. Global distribution of centennialscale<br />

records of centennial-scale records for Marine Isotope Stage (MIS) 3: a<br />

database. Quaternary Science Reviews, 21: 1185-1212.<br />

Voelker, A.H.L., Lebreiro, S. M., Schönfeld, J., Cacho, I., Erlenkeuser, H. and Abrantes,<br />

F., 2006. Mediterranean outflow strengthening during northern hemisphere coolings: A<br />

salt source for the glacial Atlantic? Earth and Planetary Science Letters, 245: 39-55.<br />

Volkman, J.K., Barrett, S.M., Blackburn, S.I. and Sikes, E.L., 1995. Alkenones in<br />

Gephyrocapsa oceanica: implications for studies of <strong>paleoclimat</strong>e. Geochimica et<br />

Cosmochimica Acta, 59: 513-520.<br />

Volkman, J.K., Eglinton, G., Comer, E.D.S. and Forsberg, T.E.V., 1980. Long-chain<br />

alkenes and alkenones in the marine coccolithophorid Emiliania huxleyi.<br />

Phytochemistry, 19: 2619-2622.<br />

Von Engelbrechten, S., 1998. Late-glacial and Holocene vegetation and<br />

environmental history of the sierra de Urbión, North-Central Spain. PhD , Dublin<br />

University, Trinity College, (Ireland), 212 pp.<br />

von Grafenstein, U., Erlenkeuser, H., Müller, J., Jouzel, J. and Johnsen, S., 1998. The<br />

cold event 8200 years ago documented in oxygen isotope records of precipitation in<br />

Europe and Greenland. Climate Dynamics, 14: 73-81.<br />

Wang, Z. and Mysak, L.A., 2006. Glacial abrupt climate changes and Dansgaard-<br />

Oeschger oscillations in a coupled climate model. Paleoceanography (in press).<br />

Weber, S.L. and Oerlemans, J., 2003. Holocene glacier variability: three case studies<br />

using an intermediate-complexity climate model. The Holocene, 13: 353-363.<br />

Wunsch, C., 2006. Abrupt climate change: An alternative view. Quaternary Research,<br />

65: 191-203.<br />

Yokoyama, Y., Lambeck, K., DE Deckker, P., Johnston, P., Fifield, K., 2000. Timing of the<br />

Last Glacial Maximum from observed sea-level minima. Nature, 406: 713-716.<br />

Yokoyama, Y., Esat, T.M. and Lambeck, K., 2001. Coupled climate and sea-level<br />

changes deduced from Huon Peninsula coral terraces of the last ice age. Earth and<br />

Planetary Science Letters, 193: 579-587.<br />

Zahn, R., Schönfeld, J., Kudrass, H.R., Park, M.H. and Erlenkeuser, H., 1997.<br />

Thermohaline instability in the North Atlantic during meltwater events: stable isotopes<br />

and detritus records from core SO75-26KL, Portuguese margin. Paleoceanography,<br />

12: 696-710.<br />

Zaitlin, B.A., Dalrymple, R.W. and Boyd, R., 1994. The stratigraphic organisation of<br />

incised valley systems associated with relative sea-level change. R.W. Dalrymple, R.J.<br />

102<br />

F. Naughton, 2007


Boyd, B.A. Zaitlin, (eds)., Incised valley systems: Origin and sedimentary sequences:<br />

SEPM Special Publication. 51: 45-60.<br />

103<br />

F. Naughton, 2007


104<br />

F. Naughton, 2007


Capítulo 2| Present-day and past (last 25 000 years)<br />

marine pollen signal off western Iberia<br />

Sinal polínico marinho presente e passado (últimos 25 000<br />

anos) ao longo da margem oeste Ibérica<br />

Répresentation pollinique actuelle et passée (des derniers 25<br />

000 ans) dans les sédiments marins de la marge ibérique<br />

occidentale<br />

F. Naughton<br />

a, e<br />

Marine Micropaleontology<br />

in press. 2006<br />

, M.F. Sánchez Goñi a, b , S. Desprat a, b , J-L. Turon a , J. Duprat a , B.<br />

Malaizé a , C. Joly a , E. Cortijo c , T. Drago d , M.C. Freitas e<br />

a Environnements et Paléoenvironnements Océaniques (UMR CNRS 5805 EPOC),<br />

Université Bordeaux 1, Av. des Facultés, 33405 Talence, France<br />

b Ecole Pratique des Hautes Etudes, Environnements et Paléoenvironnements<br />

Océaniques (UMR CNRS 5805 EPOC), Université Bordeaux 1, Av. des Facultés, 33405<br />

Talence, France<br />

c Laboratoire des Sciences du Climat et de l’Environnement (LSCE-Vallée), Bât. 12,<br />

avenue de la Terrasse, F-91198 Gif-sur-Yvette cedex, France<br />

d Centro Regional de Investigação Pesqueira do Sul , Instituto Nacional de<br />

Investigação Agrária e Pescas (INIAP) (IPIMAR-CRIPSUL), Av. 5 de Outubro, 8700-305<br />

Olhão, Portugal<br />

e Departamento e Centro de Geologia -Universidade de Lisboa, Bloco C6, 3º piso,<br />

Campo Grande, 1749-016 Lisboa, Portugal<br />

105<br />

F. Naughton, 2007


Resumo<br />

A comparação entre assinaturas polínicas actuais terrestres e<br />

marinhas, ao longo da margem e Península Ibérica, mostra que, as<br />

associações polínicas marinhas fornecem uma imagem integral da<br />

vegetação regional que coloniza o continente adjacente. As comunidades<br />

florestais mediterrânicas e Atlânticas são facilmente discriminadas pelos<br />

espectros polínicos marinhos do sul e norte, respectivamente. Os resultados<br />

obtidos a partir das concentrações polínicas totais, juntamente com, o<br />

conhecimento sobre a dinâmica actual das partículas sedimentares finas ao<br />

longo da margem Ibérica, permitiu-nos estabelecer o padrão actual de<br />

dispersão polínica, nesta região.<br />

O registo paleoclimático, relativo aos últimos 25 000 anos,<br />

representado por indicadores climáticos continentais (pólen) e marinhos<br />

(δ 18 O de foraminíferos planctónicos do tipo G. bulloides, Ice-rafted detritus-<br />

IRD e percentages de N. pachyderma sinistrógira), ao longo de duas<br />

sondagens (MD99-2331 e MD03-2697) colhidas na margem noroeste Ibérica,<br />

mostra que a vegetação do noroeste da Península Ibérica respondeu<br />

contemporaneamente à variabilidade climática detectada no Atlântico<br />

Norte. A resposta da vegetação aos eventos de Heinrich 2 e 1 é contudo<br />

complexa, sendo caracterizada por duas fases distintas nas médias e baixas<br />

altitudes do noroeste Ibérico. O início de cada evento de Heinrich é<br />

marcado, no continente, por uma forte regressão da floresta de pinheiros,<br />

assim como pela expansão de Calluna. Este sinal é síncrono dos valores mais<br />

pesados em δ 18 O e ainda da máxima expansão de N. pachyderma<br />

sinistrógira. Isto sugere que as primeiras fases de cada evento de Heinrich<br />

terão sido frias e húmidas, nesta região. A segunda fase de cada evento de<br />

Heinrich é marcada pela expansão da floresta de pinheiros a qual indica a<br />

presença de um episódio ligeiramente menos frio. Esta segunda fase é ainda<br />

marcada pelo aumento da aridez a qual é testemunhada pelo<br />

desenvolvimento de plantas semi-desérticas, no continente. A comparação<br />

do nosso registo multi-proxy com uma série de sequências polínicas<br />

continentais e marinhas Ibéricas permitiu-nos demonstrar que o evento H1 é<br />

o equivalente marinho do evento de Drias Antigo no continente.<br />

106<br />

F. Naughton, 2007


A ocorrência de árvores de clima temperado durante o último<br />

máximo glaciar (LGM- Last Glacial Maximum) e a rápida expansão de<br />

Quercus caducifolio durante o evento de Bölling-Allerød, ao longo da nossa<br />

sequência composta Galega, mostra que não só o sul mas também o norte<br />

da Península Ibérica reagiram como uma zona refugio para as árvores<br />

caducifolias durante o último período glaciário, especialmente nas médias e<br />

baixas altitudes do Noroeste Ibérico.<br />

Finalmente, a comparação entre as sequências terrestres e marinhas<br />

do sul e norte da Península e margem Ibérica permitiu-nos verificar que a<br />

vegetação respondeu ao aquecimento que caracteriza o Bölling-Allerød, ao<br />

evento frio conhecido por Drias recente e a melhoria climática que<br />

caracteriza o Holocénico de forma mais rápida no sul e nas médias e baixas<br />

altitudes do noroeste da Península ibérica do que nas altas altitudes da<br />

região nortenha como resultado da grande densidade de zonas refugio<br />

nessas zonas durante o LGM.<br />

Résumé<br />

La comparaison entre le signal pollinique actuel des sédiments de la<br />

marge Ibérique occidentale et celui de la Péninsule Ibérique montre que les<br />

assemblages polliniques marins représentent une image intégrée de la<br />

végétation régionale qui colonise le continent adjacent. Les communautés<br />

forestières présentes sur les zones biogéographiques Ibériques<br />

Méditerranéenne et Atlantique sont bien discriminées par les spectres<br />

polliniques marins du Sud et du Nord de la marge Ibérique, respectivement.<br />

L’étude de la concentration pollinique totale, ainsi que la connaissance des<br />

modèles actuels sur la dynamique des particules fines de la marge ibérique,<br />

nous a permis d’établir les scénarii de dispersion pollinique actuelle sur cette<br />

région.<br />

L’enregistrement climatique des derniers 25 000 ans, obtenu à partir<br />

des indicateurs continentaux (pollen) et marins (δ 18 O des foraminifères<br />

planctoniques du type G. bulloides, Ice-rafted detritus-IRD et les<br />

pourcentages de N. pachyderma senestre), d’une séquence marine<br />

composite de la marge Galicienne (MD99-2331 et MD03-2697) montre que la<br />

végétation du Nord-ouest Ibérique à répondu de façon synchrone aux<br />

107<br />

F. Naughton, 2007


variations climatiques détectées dans l’Atlantique Nord. La réponse de la<br />

végétation aux événements d’Heinrich 2 et 1 (H2 et H1) est cependant<br />

complexe et, essentiellement caractérisée par deux phases distinctes dans<br />

les basses et moyennes altitudes du Nord-ouest Ibérique. Le début de<br />

chaque événement d’Heinrich marqué sur la marge Ibérique par le<br />

refroidissement des eaux de surface et par l’alourdissement du δ 18 O de<br />

foraminifères planctoniques, est représenté sur le continent par une forte<br />

régression de la forêt de pin et l’expansion des bruyères (heathers). Cela<br />

suggère que la première phase de l’Heinrich a été froide et humide dans le<br />

Nord-ouest de la Péninsule Ibérique. L’expansion de la forêt de pin<br />

caractérise la deuxième phase de chacun des événements d’Heinrich,<br />

indiquant des conditions légèrement moins froides. Pendant la deuxième<br />

phase de l’événement H1 une augmentation de la sécheresse est indiquée<br />

montrée par le développement des plantes semi-désertiques. La<br />

comparaison entre notre enregistrement multi-proxy de la marge Galicienne<br />

avec des séquences polliniques Ibériques et d’autres marines de la même<br />

marge a permis de démontrer que l’événement H1 est l’équivalent marin de<br />

l’événement du Dryas ancien premièrement défini sur le continent.<br />

La présence d’arbres tempérés au cours du dernier maximum glaciaire<br />

(LGM) ainsi que l’expansion rapide du chêne caducifolié pendant<br />

l’événement du Bolling-Allerod, déduit de notre enregistrement marin, montre<br />

que non seulement le Sud mais aussi le Nord de la Péninsule Ibérique a été<br />

une zone refuge d’arbres caducifoliés durant la dernière période glaciaire et,<br />

en particulier, dans les zones de basses et moyennes altitudes.<br />

De plus, la comparaison des enregistrements marins et continentaux<br />

entre le Sud et le Nord permet de confirmer que la végétation a répondu au<br />

réchauffement du Bolling-Allerod, au refroidissement correspondant au Dryas<br />

récent et à l’amélioration climatique qui caractérise l’Holocène plus<br />

rapidement dans le sud et dans les basses et moyennes altitudes du Nord-<br />

ouest Ibérique que dans les hautes altitudes de la région Nord Ibérique. Cela<br />

indique que la densité de zones refuges pour des arbres tempérés au LGM<br />

étais plus importante dans ces zones que dans les hautes altitudes de la<br />

Péninsule Ibérique.<br />

108<br />

F. Naughton, 2007


Abstract<br />

The comparison between modern terrestrial and marine pollen signals<br />

in and off western Iberia shows that marine pollen assemblages give an<br />

integrated image of the regional vegetation colonising the adjacent<br />

continent. Present-day Mediterranean and Atlantic forest communities of<br />

Iberia are well discriminated by south and north marine pollen spectra,<br />

respectively. Results from Total Pollen Concentration together with recognized<br />

conceptual models of fine particle dynamics in the Iberian margin have<br />

allowed us to establish the present-day pattern of pollen dispersion in this<br />

region.<br />

The 25 000 year-long record of continental (pollen) and marine (δ 18 O of<br />

G. bulloides, Ice-rafted detritus-IRD and N. pachyderma s.) proxies, from the<br />

Galician margin composite core (MD99-2331 and MD03-2697), show that<br />

vegetation cover in north-western Iberia has responded contemporaneously<br />

to the climate variability of the North-Atlantic. The vegetation response to the<br />

well known North Atlantic Heinrich events 2 and 1 (H2 and H1) is however<br />

complex and characterised by two vegetation phases at low and mid-<br />

atitudes of north-western Iberia. The beginning of each Heinrich event is<br />

marked on land by an important pine forest reduction and the expansion of<br />

heathers which are synchronous with the heaviest planktonic δ 18 O values and<br />

the maxima of N. pachyderma (s.) suggesting that these first phases were<br />

cold and wet. Pinus forest expansion characterising the second phase of<br />

each Heinrich event indicates a less cold episode associated, during H1, with<br />

an increase of dryness as suggested by the development of semi-desert<br />

associations. The comparison of our Galician margin multi-proxy record with<br />

several pollen sequences from in and off Iberia allows us to demonstrate that<br />

H1 event is the marine equivalent of the Oldest Dryas on the continent.<br />

The occurrence of temperate trees during the last glacial maximum<br />

(LGM) and the rapid expansion of deciduous Quercus during the Bölling-<br />

Allerød period in our Galician margin composite sequence show that not only<br />

the southern but also north-western Iberia was a refugium zone for deciduous<br />

trees during the last glacial period, especially at low and mid-altitude zones.<br />

Furthermore, the comparison between southern and northern marine<br />

and terrestrial sequences allows us to confirm that vegetation responded to<br />

109<br />

F. Naughton, 2007


the Bölling-Allerød warming, the Younger Dryas cold event and the Holocene<br />

more quickly in low and mid-altitudes of north-western Iberia and in the south<br />

than in the high altitude northern region most likely as the result of the higher<br />

density of refugia for temperate trees in these zones during the LGM.<br />

110<br />

F. Naughton, 2007


2. 1 Introduction<br />

During the last decade several studies had been carried out in marine<br />

deep-sea cores off Iberia (Hooghiemstra et al., 1992; Sánchez Goñi et al.,<br />

1999, 2000, 2002, 2005; Boessenkool et al., 2001; Roucoux et al., 2001, 2005;<br />

Turon et al., 2003; Tzedakis et al., 2004; Desprat, 2005; Desprat et al., 2005,<br />

2006, in press) to understand vegetation responses to the climate variability<br />

detected in the North Atlantic. Among these sequences, those covering the<br />

last 25 000 years show similar vegetation changes to those recorded by the<br />

available 25 000 year-long terrestrial records. However, no experimental<br />

studies have been conducted in order to demonstrate that pollen grains<br />

preserved in those marine sequences represent the regional vegetation of<br />

the nearby continent or to understand the mechanisms involved in the<br />

transport and dispersion of these grains from the continent to the sea. To fill<br />

these gaps, we have compared present-day continental (including coastal<br />

systems) pollen signatures with modern marine (including shelf and slope)<br />

pollen assemblages. We have also determined total pollen concentration<br />

(TPC) of those surface samples to recognize present-day patterns of pollen<br />

dispersion in the Iberian margin.<br />

Having assessed the reliability of the present-day pollen signal in the<br />

upper layer sediments of MD99-2331 deep-sea core, we will compare their 25<br />

000-year high-resolution pollen record with other marine and terrestrial pollen<br />

sequences (Pons and Reille, 1988; Hooghiemstra et al., 1992; Peñalba, 1994;<br />

Pérez-Obiol and Julià, 1994; Allen et al., 1996; Muñoz Sobrino et al., 1997; 2001;<br />

2004; Peñalba et al., 1997; Von Engelbrechten, 1998; Combourieu Nebout, et<br />

al., 1999; 2002; Sánchez Goñi and Hannon, 1999; Santos et al., 2000;<br />

Boessenkool et al., 2001; Roucoux et al., 2001; 2005; Gil-Garcia et al., 2002;<br />

Ruiz Zapata et al., 2002; Turon et al., 2003) to document accurately western<br />

Iberian vegetation changes over this period. Furthermore, the direct<br />

correlation between sea surface temperature and vegetation changes in<br />

and off Iberia from the multiproxy study of MD99-2331 and MD03-2697 deep-<br />

sea cores will allow us to link several well known terrestrial climate events with<br />

those detected elsewhere in the North Atlantic and over Greenland.<br />

111<br />

F. Naughton, 2007


2. 2 Environmental Setting<br />

2. 2. 1 Study area and present-day vegetation and climate<br />

Western Iberia including Portugal and the north-western part of Spain<br />

extends from 37°N to 43°N and comprises essentially the Minho and Sado<br />

basins and the western part of the Douro and Tagus basins (Fig. II.1). North-<br />

western Spain, including the Minho basin, is influenced by the wet, relatively<br />

cool and weakly seasonal Atlantic climate (annual precipitation mean: 900-<br />

1400 mm and temperature range: -7 to 10° C) and is dominated by<br />

deciduous Quercus forest (Q. robur, Q. pyrenaica and Q. petraea), heath<br />

communities (Ericaceae and Calluna) and Ulex. There are also locally birch<br />

(Betula pubescens subsp. celtiberica) and hazel (Corylus avellana) groves,<br />

and brooms (Genista) (Alcara Ariza et al., 1987).<br />

In the south, the Tagus and Sado basins, influenced by Mediterranean<br />

climate (mean annual precipitation: 200-600 mm and temperature range: 4<br />

to 14° C), are dominated by evergreen sclerophyllous forests. Q. rotundifolia<br />

and Q. suber forests with Phillyrea angustifolia and Pistacia terebinthus<br />

colonise the western basins while Q. rotundifolia and Q. coccifera woodlands<br />

associated with Juniperus communis and Pinus halepensis occupies the<br />

eastern part. In the warmest zones, thermophilous elements such as Pistacia<br />

lentiscus and Olea sylvestris form the forests. Middle altitudes (700-1000 m<br />

a.s.l.) are dominated by deciduous Quercus forest (Q. pyrenaica and Q.<br />

faginea) associated with northern European species such as Taxus baccata.<br />

The degradation of this forest produces two types of brush communities:<br />

rockrose shrublands (Cistaceae) in zones with precipitation between 600 and<br />

1000 mm and heath communities (Ericaceae) in wetter zones.<br />

Between both regions, there is a transitional zone which includes the<br />

hydrographic basin of the Douro. This zone is characterised by high<br />

precipitation values (700 to 1000 mm/year) and winter temperatures between<br />

4 and -4°C. At high altitudes, the wettest and coldest zones reach 1600<br />

mm/year and -8°C, respectively (Polunin and Walters, 1985). The oceanic<br />

influence is particularly important in the northwest of the basin, where the Q.<br />

robur and Q. suber association predominates (Braun-Blanquet et al., 1956).<br />

The spread of both Pinus pinaster and Eucalyptus globulus has been favoured<br />

112<br />

F. Naughton, 2007


y anthropic impact. The understory vegetation is largely dominated by Ulex,<br />

in association with heaths. The river margins are colonized by Alnus glutinosa,<br />

Fraxinus angustifolia, Ulmus spp., Salix spp. and Populus spp..<br />

Fig. II.1 | Fig. 1- Study area. Dashed line divides the Atlantic and Mediterranean biogeographical zones<br />

(Blanco Castro et al., 1997). White circles with a dark point represent the top samples analysed in this study;<br />

white circles represent the modern samples from the European Pollen Database; white circles with a cross<br />

represent the studied cores sites (MD03-2697 and MD99-2331); dark circles represent marine and terrestrial<br />

core sites used for comparison with our study. Continental sequences: a) Square A locates sequences 1 to<br />

5: 1- Laguna de la Roya (Allen et al., 1996), 2- Sanabria March (Allen et al., 1996), 3-Laguna de las<br />

Sanguijuelas (Muñoz Sobrino et al., 2004), 4-Lleguna (Muñoz Sobrino et al., 2004), 5- Pozo do Carballal<br />

(Muñoz Sobrino et al., 1997); b) Sites 6 to 13 correspond to: 6- Laguna Lucenza (Santos et al., 2000); 7-<br />

Lagoa Lucenza (Muñoz Sobrino et al., 2001); 8-Lago de Ajo (Allen et al., 1996); 9- Los Tornos (Peñalba,<br />

1994); 10-Saldropo (Peñalba, 1994); 11-Belate (Peñalba, 1994); 12- Atxuri (Peñalba, 1994); 13- Banyoles<br />

(Pérez-Obiol and Julià, 1994); c) Square B includes sequences 14 to 19: 14- Quintanar de la Sierra (Peñalba<br />

et al., 1997); 15- Sierra de Neila - Quintanar de la Sierra (Ruiz Zapata et al., 2002); 16- Hoyos de Iregua (Gil-<br />

Garcia et al., 2002); 17- Laguna Masegosa (Von Engelbrechten, 1998); 18- Laguna Negra (Von<br />

Engelbrechten, 1998); 19- Las Pardillas lake (Sánchez Goñi and Hannon, 1999); 20- Padul (Pons and Reille,<br />

1988); 21- Mougás (Gómez-Orellana et al., 1998); 22- Charco da Candieira (Van der Knaap and Van<br />

Leeuwen, 1995). The marine cores represented on the map are: 8057 B (Hooghiemstra et al., 1992), SO75-<br />

6KL (Boessenkool et al., 2001), SU81-18 (Turon et al., 2003) and ODP 976 (Combourieu Nebout, et al., 1999;<br />

2002) and MD95-2039 (Roucoux et al., 2001; 2005).<br />

113<br />

F. Naughton, 2007


2. 2. 2 Oceanography<br />

The western Iberian margin is dominated by the surface Portugal<br />

Current system (PCS) which is composed of the slow equatorward current in<br />

the open ocean (Arhan et al., 1994) and the fast, seasonally reversing coastal<br />

current (Ambar and Fiúza, 1994; Barton, 1998) (Fig. II.2). During the summer,<br />

the Azores high pressure cell is located in the central North Atlantic and the<br />

Greenland low is weak. This situation generates northerly and northwesterly<br />

prevailing winds (Fig. II.1) which favour the occurrence of upwelling events<br />

and a southward surface circulation (Fiúza et al., 1982; Haynes and Barton,<br />

1990) near the shelf break in the upper 50-100 m (Álvarez-Salgado et al.,<br />

2003). The resultant upwelled cold and nutrient-rich Eastern North Atlantic<br />

Central Water of subpolar sources (<strong>EN</strong>ACWsp) is transported northward of 45°<br />

N. Warm, salty and nutrient-poor Eastern North Atlantic Central Water of<br />

subtropical origin (<strong>EN</strong>ACWst) is transported to the south of 40° N (Fiúza, 1984;<br />

Rios et al., 1992) (Fig. II.2).<br />

Fig. II.2 | West to east scheme of the different water masses from the western Iberian margin (adapted from<br />

Sprangers et al., 2004). White circles with a dark point represent southward water flow and white circle with<br />

a cross represent northward water flow. PCS-Portugal Current System; <strong>EN</strong>ACW st-Eastern North Atlantic<br />

Central Water of subtropical origin; <strong>EN</strong>ACW sp-Eastern North Atlantic Central Water of subpolar origin; MSW-<br />

Mediterranean Sea Water; LSW-Labrador Sea Water; NADW-North Atlantic Deep Water.<br />

During the winter the Azores high pressure cell is located off the<br />

northwest African coast and the Greenland low is deep and situated off<br />

south-eastern Greenland. The pressure gradient between the two systems<br />

results in an onshore and slightly northward wind off Iberia (Fig. II.1) triggering<br />

114<br />

F. Naughton, 2007


downwelling processes and a northward surface circulation (Frouin et al.,<br />

1990; Haynes and Barton, 1990). This reversion of hydrological paths starts in<br />

the end of summer in September-October and it persists until March-April<br />

representing the well known Portugal Coastal Counter Current (PCCC)<br />

(Ambar et al., 1986). This poleward flow is narrow (30 km wide) and it<br />

transports warm and salty waters (<strong>EN</strong>ACWst) in the upper 200-300 m to the<br />

North (Pingree and Le Cann, 1990).<br />

Below the Central Waters system, between 550 m and 1500 m depth, the<br />

Mediterranean Sea Water (MSW) consisting of high salinity and relatively<br />

warm water mass is transported northward (Mazé et al., 1997) (Fig. II.2).<br />

However, the salinity of the MSW decreases highly at latitudes higher than 41°<br />

N by mixing with the underlying low-salininity Labrador Sea water (LSW)<br />

(McCave and Hall, 2002). This LSW is one of the three water masses included<br />

in the North Atlantic Deep Water (NADW) over the western Iberian margin<br />

(Huthnance et al., 2002).<br />

2. 2. 3 Morphology and recent sedimentation<br />

The Iberian margin is characterised by a relatively narrow shelf (30-50<br />

km wide) with a steep irregular slope plunging to the oceanic abyssal plain<br />

(Fig. II.3a). This margin is cut off by deep canyons like Mugia, Porto, Aveiro,<br />

Nazaré, Cascais, Lisbon, Setúbal and S.Vicente. The largest canyons (Nazaré,<br />

Setúbal) dissect the entire continental shelf, capturing sediments carried over<br />

the shelf and upper slope by alongshore currents, providing a direct conduit<br />

of particles from the upper shelf to the deep-sea (Vanney and Mougenot,<br />

1981). Some canyons, e.g. Setúbal, start close by the present-day coastline<br />

and have a direct connection to the river mouth, while others, such as the<br />

Porto Canyon, begin only at the shelf edge and play a minor role in the<br />

interception of shelf material at the present-day sea level. All Iberian canyons<br />

were probably more active during the period of low sea-level (Van Weering<br />

and McCave, 2002). The lower and upper slopes are also intersected by<br />

several seamounts as Vigo (VS), Vasco da Gama (VDGS), Porto (PS), Tore (TS),<br />

by the Galicia Bank and several tectonic depressions (Vanney and<br />

Mougenot, 1981).<br />

115<br />

F. Naughton, 2007


Fig. II.3 | a) Morphology of the Iberian margin. Location of the surface samples from b) north-western<br />

Iberian margin and c) south-western Iberian margin. White arrows indicate the present-day pattern of<br />

pollen dispersion in the western Iberian margin.<br />

116<br />

F. Naughton, 2007


2. 2. 3. 1 North-western Iberian margin<br />

In north-western Iberia, five rivers (Douro, Ave, Cávado, Lima and<br />

Minho) release large amounts of sediments to the adjacent continental<br />

margin. The Douro is the main sediment supplier to the adjacent shelf (~8.2 x<br />

10 9 m 3 annual mean discharge) followed by the Minho river (Dias et al., 2002;<br />

Jouanneau et al., 2002; Oliveira et al., 2002) (Fig. II.3b). They are 927 km and<br />

300 km long, draining a catchment area of 97 700 km 2 and 17 100 km 2 ,<br />

respectively (Loureiro et al., 1986). Above 42° N, rivers are replaced by rias<br />

(Vigo, Pontevedra, Arousa and Muros), which act essentially as sediment<br />

traps, preventing particle input to the adjacent margin (Dias et al., 2002;<br />

Jouanneau et al., 2002).<br />

The northern Portuguese continental shelf is composed of a) an inner<br />

shelf zone (


BNL (Vitorino et al., 2002) nourishing the INL (Oliveira et al. 2002). These<br />

extreme events contribute to an important export of fine sediments (Vitorino<br />

et al., 2002) and occasionally of coarse fraction (Dias, 1987) to the upper<br />

slope. Current reversals, probably caused by the presence of local slope<br />

eddies, can also allow some down-slope transport of particles (Pingree and<br />

LeCann, 1992). During upwelling conditions, fine sediment export is restricted<br />

to the shelf edge (McCave and Hall 2002; Van Weering et al., 2002).<br />

However, lateral sediment exchange can be favoured by offshore filaments<br />

stretching westward (Huthnance et al., 2002).<br />

MD99-2331 and MD03-2697 twin deep-sea cores, located north-<br />

western of the Mesozoic and Cenozoic outcrops, mostly receive sediments<br />

coming from the Douro and Minho rivers, especially during downwelling<br />

conditions.<br />

2. 2. 3. 2 South-western Iberian margin<br />

In the south-western Iberian margin, the Tagus river is the primary<br />

sediment supplier followed by Sado river to the shore (Dias, 1987; Jouanneau<br />

et al., 1998) (Fig. II.3a and Fig. II.3c). The Tagus river is 1110 km long draining a<br />

catchment area of 80 600 km 2 with 400 m 3 . s -1 of annual mean flow (Vale,<br />

1990). The Sado river is 175 km long, drains a catchment area of 7 640 km 2<br />

and yields less than 10 m 3 . s -1 of annual mean discharge (Loureiro et al., 1986).<br />

Differences between both river discharge and littoral currents influence the<br />

sediment distribution along the shelf (Jouanneau et al., 1998). The mud patch<br />

is located offshore of Tagus river basin and covers the entire continental shelf<br />

(Araújo et al., 2002). During summer, suspended particulate matter (SPM)<br />

concentration in the mouth of the Tagus estuary is four times higher than that<br />

of the Sado, and the n<strong>ephe</strong>loid layer can extend 30 km westward<br />

(Jouanneau et al., 1998). Fine sediments are essentially exported to the slope<br />

and adjacent abyssal plains through the canyons of Cascais, Lisbon and<br />

Setúbal (Jouanneau et al., 1998) and by offshore filaments (Huthnance et al.,<br />

2002).<br />

118<br />

F. Naughton, 2007


2. 3 Material and methods<br />

2. 3. 1 Deep-sea cores: MD99-2331 and MD03-2697<br />

MD99-2331 (42° 09’ 00 N, 09° 40’ 90 W; 2110 m depth) and MD03-2697<br />

(42° 09’ 59 N, 59° 42’ 10 W; 2164 m depth) deep-sea cores were retrieved in<br />

the Galician margin (north-west of Iberia) using a CALYPSO corer during the<br />

GINNA (IMAGES V) and PICABIA oceanographic cruises on board the R/V<br />

Marion Dufresne (Fig. II.1). MD99-2331 and MD03-2697 are 37.2 m and 41.23 m<br />

long, respectively, covering Marine Isotopic Stages (MIS) 1 to 11. X-ray analysis<br />

using SCOPIX image-processing (Migeon et al., 1999) has shown a well<br />

preserved sedimentary sequence in core MD03-2697 while core MD99-2331<br />

sees a sediment mixing zone between 1.10 m and 1.90 m of core depth. In<br />

order to obtain a detailed palaeoclimatic sequence for the last 25 000 years<br />

in and off NW Iberia, we have built a composite record assembling the MIS 1<br />

interval of core MD03-2697 with the MIS 2 interval of core MD99-2331.<br />

2. 3. 1. 1 Radiometric dating<br />

Seven levels of MD03-2697 and twenty levels from MD99-2331 were<br />

dated by AMS 14 C on Globigerina bulloides and Neogloboquadrina<br />

pachyderma (s.) at Beta Analytic Inc (Beta), at Gif-sur-Yvette (Gif) and at<br />

Laboratoire de Mesure du Carbone 14-Saclay (LMC), indicating that this<br />

sequence covers the last 25 000 years (Tab. II.1). All radiocarbon dates were<br />

corrected for marine age reservoir difference (400 years) (Bard et al., 2004).<br />

The samples presenting conventional AMS 14 C younger than 21 786 BP<br />

were calibrated by using CALIB Rev 5.0 program and "global" marine<br />

calibration dataset (marine 04.14c) (Stuiver and Reimer, 1993; Hughen et al.,<br />

2004; Stuiver et al., 2005). We use 95.4% (2 sigma) confidence intervals and<br />

their relative areas under the probability curve as well as the median<br />

probability of the probability distribution (Telford et al., 2004). 14 C radiometric<br />

ages older than 21 786 yr BP were calibrated by matching the obtained<br />

conventional AMS 14 C with the calendar ages estimated for MD95-2042 deep-<br />

sea core by Bard et al. (2004).<br />

119<br />

F. Naughton, 2007


In this paper, we will use 14 C ages (yr BP) corrected for the marine<br />

reservoir effect (of 400 years) instead of calibrated ages (cal yr BP) because<br />

in most of the Iberian terrestrial sequences, calendar ages are not available.<br />

Lab<br />

code<br />

Beta<br />

2131134<br />

Beta<br />

2131135<br />

Beta<br />

003257<br />

Beta<br />

2131136<br />

Beta<br />

2131137<br />

Beta<br />

003258<br />

Beta<br />

003259<br />

LMC14<br />

001231<br />

LMC14<br />

001232<br />

GIF<br />

102377<br />

LMC14<br />

001233<br />

LMC14<br />

001235<br />

LMC14<br />

001236<br />

LMC14<br />

001237<br />

LMC14<br />

002445<br />

GIF<br />

101109<br />

GIF<br />

102373<br />

LMC14<br />

002446<br />

LMC14<br />

001845<br />

LMC14<br />

001846<br />

LMC14<br />

001847<br />

LMC14<br />

001849<br />

LMC14<br />

001850<br />

GIF<br />

102378<br />

LMC14<br />

001851<br />

LMC14<br />

001852<br />

LMC14<br />

001853<br />

Core<br />

depth<br />

(cm)<br />

MD03 2697<br />

20<br />

MD03 2697<br />

40<br />

MD03 2697<br />

70<br />

MD03 2697<br />

80<br />

MD03 2697<br />

110<br />

MD03 2697<br />

150<br />

MD03 2697<br />

200<br />

MD99 2331<br />

200<br />

MD99 2331<br />

205<br />

MD99 2331<br />

220<br />

MD99 2331<br />

222<br />

MD99 2331<br />

228<br />

MD99 2331<br />

235<br />

MD99 2331<br />

242<br />

MD99 2331<br />

260<br />

MD99 2331<br />

290<br />

MD99 2331<br />

570<br />

MD99 2331<br />

590<br />

MD99 2331<br />

595<br />

MD99 2331<br />

600<br />

MD99 2331<br />

607<br />

MD99 2331<br />

620<br />

MD99 2331<br />

623<br />

MD99 2331<br />

630<br />

MD99 2331<br />

637<br />

MD99 2331<br />

650<br />

MD99 2331<br />

655<br />

Material<br />

Conv.<br />

AMS 14C<br />

age BP<br />

120<br />

Conv.<br />

AMS 14C<br />

age BP<br />

(-400 yr)<br />

error<br />

95.4 % (2σ)<br />

Cal BP<br />

age ranges<br />

Cal BP age<br />

median<br />

probability<br />

G. bulloides 2880 2480 40 2501:2739 2656<br />

G. bulloides 4760 4360 40 4866:5198 5008<br />

G. bulloides 7435 7035 50 7783:7998 7895<br />

G. bulloides 7470 7070 40 7835:8014 7930<br />

G. bulloides 9940 9540 40 10705:11084 10896<br />

G. bulloides 11920 11520 60 13233:13486 13353<br />

G. bulloides 12520 12120 60 13816:14111 13965<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

N.<br />

pachyderma<br />

13640 13240 80 15303:16099 15679<br />

13810 13410 80 15524:16359 15922<br />

14130 13730 120 15898:16828 16342<br />

13920 13520 90<br />

13930 13530 80<br />

15658:16520<br />

a<br />

15686:16521<br />

a<br />

16067 a<br />

16081 a<br />

15130 14730 90 17250:18182 17848<br />

15060 14660 90 17170:18038 17722<br />

G. bulloides 15540 15140 90 18405:18723 18520<br />

G. bulloides 16170 15770 130 18787:19265 18983<br />

G. bulloides 19770 19370 170 22534:23622 23038<br />

G. bulloides 22290 21890 170 ~ 25950 b,c ~ 25950 b,c<br />

G. bulloides 20860 20460 250 23931:25369 24542<br />

G. bulloides 20550 20150 240 23450:24803 24119<br />

G. bulloides 20460 20060 140 23652:24405 24016<br />

G. bulloides 21620 21220 160 25301:26000 25626<br />

G. bulloides 21740 21340 160 25439:26000 25730<br />

N.<br />

pachyderma<br />

22690 22290 180 ~ 26350 b,c ~ 26350 b,c<br />

G. bulloides 22150 21750 170 ~ 25800 c ~ 25800 c<br />

G. bulloides 22440 22040 170 ~ 26000 c ~ 26000 c<br />

G. bulloides 22430 22030 180 ~ 26000 c ~ 26000 c<br />

Tab. II.1| Radiocarbon ages of MD99-2331 and MD03-2697 deep-sea cores. a Not acceptable dating<br />

(bioturbated layers); b Radiocarbon dates too old (not used); c dates calibrated by matching conventional<br />

AMS 14C with calendar ages estimated for MD95-2042 deep-sea core by Bard et al. (2004).<br />

F. Naughton, 2007


2. 3. 1. 2 Marine proxy analyses<br />

The planktonic isotopic record of MD99-2331 covering MIS 6 to MIS 1<br />

has been published in Gouzy et al. (2004). However, additional stable isotope<br />

measurements of planktonic foraminifera have been done to refine the<br />

planktonic isotopic record of the MIS 2 interval.<br />

In total, 56 measurements have been made at 2 to 10 cm sample<br />

resolution. For MIS 1, 29 levels with a sample spacing of 5 to 10 cm have been<br />

analysed in the MD03-2697 sequence. These measurements have been<br />

carried out on the 250–315 µm fraction of Globigerina bulloides previously<br />

cleaned with distilled water. Each aliquot, including 8–10 specimens and<br />

representing a mean weight of 80 μg, was prepared in the Micromass<br />

Multiprep autosampler, using an individual acid attack for each sample.<br />

The CO2 gas extracted has been analyzed against NBS 19 standard,<br />

taken as an international reference standard.<br />

Isotopic analysis of MD99-2331 has been carried out using an Optima<br />

Micromass mass spectrometer in the UMR CNRS 5805 EPOC (Environnements<br />

et Paléoenvironnements Océaniques) at Bordeaux 1 University and, those of<br />

MD03-2697 were performed using a delta plus Finnigan at the LSCE<br />

(Laboratoire des Sciences du Climat et de l’Environnement). The mean<br />

external reproducibility of powdered carbonate standards is ±0.05‰ for<br />

oxygen. Results are presented versus PDB.<br />

Polar foraminifera, N. pachyderma (s.), counting include 79 levels (2 to<br />

10 cm of sample spacing) and 40 levels (5 to 10 cm of sample spacing) from<br />

MD99-2331 and MD03-2697, respectively.<br />

IRD semiquantitative analysis as been carried out in 78 levels (2 to 10<br />

cm sample spacing) and 30 levels (5 to 10 cm sample spacing) from MD99-<br />

2331 and MD03-2697, respectively. In this study, only the total concentrations<br />

of the lithic grains were considered.<br />

Both analyses were performed on the >150 μm sand-size fraction which<br />

was obtained according to classic sedimentological procedure.<br />

121<br />

F. Naughton, 2007


2. 3. 1. 3 Pollen analysis<br />

110 and 22 samples with a sample spacing of 2 to 10 cm and 5 to 10<br />

cm were analysed from MD99-2331 and MD03-2697, respectively. In each 1<br />

cm-thickness sample, 3 to 5 cm 3 of sediment were treated for pollen analysis.<br />

The treatment of the samples from both deep-sea cores (MD99-2331<br />

and MD03-2697) followed the procedure described by de Vernal et al. (1996),<br />

slightly modified at the UMR CNRS 5805 EPOC (Desprat, 2005).<br />

Palynological treatment consists of pollen concentration by chemical<br />

digestion using cold HCl (at 10%, 25% and 50%) and cold HF (at 40% and 70%)<br />

to eliminate carbonates and silicates, respectively. A Lycopodium spike of<br />

known concentration has been added to each sample to calculate total<br />

pollen (including spores) concentrations. The residue was sieved through a 10<br />

µm nylon mesh screen (Heusser and Stock, 1984) and mounted in bidistillate<br />

glycerine. A Zeiss microscope with x550 and x1250 (oil immersion)<br />

magnifications was used for pollen observation and counting.<br />

Pollen identifications were achieved via comparison with specialised<br />

atlases (Moore et al., 1991; Reille, 1992) together with the pollen reference<br />

collection of the UMR CNRS 5805 EPOC. At least 100 pollen grains (excluding<br />

Pinus, aquatic plants and spores) and 20 pollen types were counted in each<br />

of the 142 samples (deep-sea cores and modern samples, cf. section 2. 3. 2)<br />

analysed to obtain statistically reliable pollen spectra (McAndrew and King,<br />

1976).<br />

Pollen percentages were calculated based on the main pollen sum<br />

which excludes aquatic plants, spores, indeterminate and unknown pollen<br />

grains. Because Pinus grains are usually over-represented in marine sediments<br />

(Heusser and Balsam, 1977), they are also excluded from the main sum and<br />

their percentages are determined by using the total sum (pollen + spores +<br />

indeterminable + unknowns).<br />

2. 3. 2 Modern pollen samples<br />

We have analysed the pollen grains of 10 top samples from several<br />

estuarine, shelf and marine sedimentary sequences retrieved in and off<br />

western Iberia (Fig. II.1, Tab. II.2).<br />

122<br />

F. Naughton, 2007


The high percentages of Pinus detected in these top samples confirm<br />

that they represent the last 0-350 yr, since it is well known that Pinus<br />

reforestation in western Iberia started in the seventeenth century (Valdès and<br />

Gil Sanchez, 2001). Because major vegetation changes are not detected in<br />

percentage pollen diagrams for the last centuries in this region (Desprat et al.,<br />

2003), we assume that our modern samples represent present-day pollen<br />

signatures.<br />

The resulting marine and coastal modern pollen assemblages have<br />

been compared with 12 terrestrial pollen samples including moss samples,<br />

surface sediments and top of peat bog and lake sequences, of both the<br />

Mediterranean and Atlantic parts of western Iberia stored in the European<br />

Pollen Database, http:/www.imep-cnrs.com/pages/EPD.htm, (Peyron et al.,<br />

1998; Barboni et al., 2004) (Fig. II.1).<br />

Sample name<br />

MD95-2042<br />

1FP8-1<br />

MD99-2332<br />

Barreiro<br />

MD04-2814 CQ<br />

Laquasup<br />

Po 287-13-2G<br />

CG11<br />

Depth<br />

(cm)<br />

Top<br />

(0-1)<br />

Top<br />

(0-1)<br />

Top<br />

(0-1)<br />

Top<br />

(0-1)<br />

Top<br />

(0-1)<br />

Top<br />

(0-5)<br />

Top<br />

(0-1)<br />

Top<br />

(0-1)<br />

Latitude Longitude<br />

123<br />

Water depth<br />

(m)<br />

Year of<br />

sampling<br />

37°48’N 10°10’W 3148 1995<br />

38°01’N 09°20’W 980 2003<br />

38°33’N 09°22’W 97 1999<br />

38°40’N 09°07’W 0 1999<br />

40°37’N 09°52’W 2449 2004<br />

41°09’N 08°38’W 0 2001<br />

41°09’N 09°01’W 81 2002<br />

41°48’N 09°04’W 107 1992<br />

MD99-2331 3-4 42°09’N 09°42’W 2120 1999<br />

Vir-18<br />

Top<br />

(0-1)<br />

42°14’N 08°47’W 45 1990<br />

Tab. II.2| Location, water depth and year of sample sampling from coastal, shelf and slope sequences of<br />

the Iberian margin.<br />

F. Naughton, 2007


2. 4 Results and Discussion<br />

2. 4. 1 Present day pollen signature<br />

2. 4. 1. 1 Western Iberian terrestrial sites<br />

Figure II.4 shows pollen spectra from several modern samples collected<br />

in western Iberian Península. Pollen assemblages from surface samples<br />

located above 42°N (COV1, MUN1, MUN3, ES09, E258 and ES62), record the<br />

Atlantic deciduous forest (Figs. II.1 and II.4). However, the dominant tree<br />

species differs from place to place reflecting the heterogeneity of the<br />

vegetation cover of this region (Fig. II.4). For example, deciduous Quercus is<br />

the most important tree pollen in samples ES09, ES258, MUN3 and ES62 while<br />

Corylus dominates COV1 pollen spectrum and Betula that of MUN1. The<br />

pollen signal from the southern samples (FRA1, FRA4, GAT1, EXT1 and EXT2)<br />

represents Mediterranean plant communities, essentially composed of<br />

evergreen Quercus (Quercus ilex-type) and Olea (Figs. II.1 and II.4). However,<br />

FRA4 sample also includes relatively high percentages of pollen of deciduous<br />

trees, similar to those found in pollen spectra from north-western Iberia. This<br />

sample, though located in the Mediterranean region, comes from a high<br />

altitude deciduous oak forest zone. Within this southern region, sample ESO6<br />

collected in the coastal area reflects open vegetation resulting of saline<br />

conditions and sandy soils, preventing the development of deciduous and<br />

perennial forests. These southern samples also reflect the mosaic of the<br />

vegetation colonising present day southern Iberia.<br />

We notice that, the southern pollen samples show higher percentages<br />

of Mediterranean plants than the north-western Iberian samples (Fig. II.4),<br />

clearly discriminating between Mediterranean and Atlantic plant community<br />

sources, respectively.<br />

124<br />

F. Naughton, 2007


Fig. II.4 | Pollen spectra from western Iberian modern samples. Total temperate and humid (Tot.<br />

Temp./Hum.) trees includes: Alnus, Betula, Corylus, deciduous Quercus and other temperate and humid<br />

species (Acer, Fagus, Fraxinus, Salix, Tilia, Ulmus, Hedera helix, Myrica and Vitis). Total mediterranean (Tot.<br />

Mediter.) plants includes: evergreen Quercus, Olea and Cistus. Taraxacum-type, Asteraceae, Poaceae,<br />

Ericaceae and Calluna represent the ubiquist group. Semi-desert plants include Ephedra,<br />

Chenopodiaceae and Artemisia. Climate parameters: Alt: Altitude; PP: Precipitation; MTCO: Mean<br />

temperature of the coldest month; MTWA: Mean temperature of the warmest month; TANN: Annual<br />

temperature.<br />

125<br />

F. Naughton, 2007


2. 4. 1. 2 Western Iberian estuarine and margin sites<br />

Estuarine pollen samples VIR-18 (Ría de Vigo) and Laquasup (Douro<br />

estuary), are marked by relatively high percentages of deciduous forest<br />

reflecting the present-day vegetation of north-western Iberia (Figs. II.5 and<br />

II.1). Shelf and slope samples (MD99-2331, CG11, Po 287-13-2G and MD04-<br />

2814 CQ), located in the adjacent margin, reproduce the same pollen signal<br />

as that of these northern estuarine samples. Southern samples from estuarine<br />

(Barreiro) and margin (MD99-2332, FP8-1 and MD95-2042) sites present in turn<br />

higher percentages of Mediterranean plants than northern sites (Fig. II.5). As<br />

in terrestrial samples, Mediterranean and Atlantic plant communities are well<br />

discriminated in the pollen signal from estuarine and margin sites (Figs. II.5<br />

and II.1). It is important to note that the estuarine pollen assemblages are<br />

more similar to the marine ones than to the terrestrial pollen spectra. Indeed,<br />

estuarine sediments contain pollen from the regional vegetation which<br />

colonises the hydrographic basin while terrestrial samples mainly reflect local<br />

vegetation (Figs. II.4 and II.5). This indicates, as previous studies have already<br />

shown for the south-western French margin (Turon, 1984), that pollen spectra<br />

off north-western Iberia reflect an integrated image of the regional<br />

vegetation of the adjacent continent. Pinus percentages from western<br />

Iberian samples are relatively low when compared with estuarine, shelf and<br />

slope samples off this region (Fig. II.5). This is in agreement with the observed<br />

overrepresentation of Pinus pollen in marine sediments in general and, in<br />

particular, off south-western Europe (Turon, 1984). Other works have further<br />

shown that Pinus pollen percentages increase seawards (Heusser and Balsam,<br />

1977; Heusser and Shackleton, 1979). Our study confirms this pattern in the<br />

north and south-western margins. However, MD95-2042 site, representing the<br />

farthest sample from the coast line, presents weaker percentages of Pinus<br />

pollen than estuarine and the other marine samples.<br />

Despite the general Pinus overrepresentation in marine sediments, the<br />

good correlation between both terrestrial and marine present-day pollen<br />

signatures in and off western Iberia confirm the reliability of past vegetation<br />

and climate change reconstructions of this region proposed by previous<br />

works on western Iberian margin cores (Hooghiemstra et al., 1992; Sánchez<br />

Goñi et al., 1999, 2000, 2002, 2005; Boessenkool et al., 2001; Roucoux et al.,<br />

126<br />

F. Naughton, 2007


2001, 2005; Turon et al., 2003; Tzedakis et al., 2004; Desprat et al., 2005, 2006, in<br />

press).<br />

Fig. II.5 | Pollen assemblages of top samples from coastal and marine western Iberian sites (see also<br />

caption of Fig. II.4). TPC: total pollen concentration.<br />

127<br />

F. Naughton, 2007


2. 4. 2 Present-day pollen transport patterns<br />

Previous works on coastal zones with complex fluvial systems have<br />

shown that pollen is mainly transported to the sea by rivers and streams<br />

(Muller 1959; Bottema and Van Straaten, 1966; Peck, 1973; Heusser and<br />

Balsam, 1977). The western Iberian margin, close to several important<br />

hydrographic basins such as Tagus and Sado in the south and Douro and<br />

Minho in the north, mainly receives pollen through fluvial transport (Fig. II.3).<br />

Furthermore, north-western prevailing winds in both north and southern<br />

regions probably impede substantial direct airborne transport of pollen<br />

seaward. This pattern of fluvial transport contrasts with others, e.g. north-<br />

western Africa, associated with an arid environment, where pollen grains are<br />

mainly seaward transported by the wind (Dupont et al., 2000; Hooghiemstra<br />

et al., 2006). Indeed, the distribution of the TPC shows (Fig. II.5) that the<br />

highest TPC values are found in samples from coastal areas such as the Douro<br />

estuary (Laquasup: 44,399x10 3 grains/cm 3 ) and the Ría de Vigo (VIR-18:<br />

65,443x10 3 grains/cm 3 ). Barreiro sample is an exceptional case with low TPC<br />

(18 x10 3 grains/cm 3 ) probably because it was collected far away from the<br />

Tagus main channel and likely receiving pollen only from the local<br />

vegetation. Shelf surface samples present intermediate concentration values<br />

(CG11: 30,468x10 3 grains/cm 3 , Po 287-13-2G: 42, 600x10 3 grains/cm 3 and<br />

MD99-32b: 56,396x10 3 grains/cm 3 ) and finally slope samples attain the lowest<br />

TPC values (MD99-2331: 1,924x10 3 grains/cm 3 , MD04-2814 CQ: 1,886x10 3<br />

grains/cm 3 , MD95-2042: 2,153x10 3 grains/cm 3 and IFP8: 3,489x10 3 grains/cm 3 ).<br />

Our work shows that a seaward decrease of total pollen<br />

concentrations occurs on the Iberian margin following the estuary-shelf-slope<br />

transect (Fig. II.3). This pattern coincides with that observed in other margin<br />

zones around the world showing a seaward decrease in total pollen<br />

concentration with maximum values close to the mouth of the river systems<br />

(Muller, 1959; Bottema and Van Straaten, 1966; Cross et al., 1966; Groot and<br />

Groot, 1966, 1971; Koreneva, 1966; Stanley, 1966; Mudie, 1982; Turon, 1984;<br />

Van der Kaars and de Deckker, 2003).<br />

Based on several studies of sedimentary dynamics on the north-western<br />

Iberian margin (Araújo et al., 1994; Drago et al., 1998; Dias et al., 2002;<br />

Huthnance et al., 2002; Jouanneau et al., 2002; Oliveira et al., 2002; Van<br />

128<br />

F. Naughton, 2007


Weering et al., 2002; Vitorino et al., 2002), we propose a pattern of pollen<br />

dispersion for this region (Fig. II.3b). This pattern is similar to the distribution<br />

model of fine terrigenous particles proposed by Dias et al. (2002). Pollen and<br />

spores, once immersed behave in a similar manner to fine sedimentary<br />

particles (Chmura and Eisma, 1995). After being released by rivers (mainly<br />

Douro followed by Minho), pollen grains, are enclosed in n<strong>ephe</strong>loid layers<br />

and transported to the shelf until getting blocked by the rocky outcrops. In<br />

winter, during downwelling conditions pollen grains are then transported<br />

polewards, firstly deposited in the Douro mud patch (S-N direction) then in the<br />

Galicia mud patch, and finally they flow westward to the deep-sea. Only<br />

small quantities of pollen grains can be transported directly to the outer shelf<br />

and upper slope under extreme stormy events. In summer, under upwelling<br />

conditions, pollen transfer to the slope must be restricted to offshore filaments<br />

as suggested by Huthnance et al. (2002) for the fine sediments.<br />

In the southern Iberian margin, TPC values also decrease seawards as<br />

in the northern region (Fig. II.5). Our study suggests that pollen grains released<br />

by the Tagus and to a lesser extent by the Sado river, are partially deposited<br />

in the shelf and transported to the south and seaward by littoral and oceanic<br />

currents probably during upwelling conditions (Fig. II.3c). Pollen grains are<br />

probably transported by the southern canyons from the shelf to the slope and<br />

abyssal plain following the fine particle pathway suggested by several works<br />

on sedimentary dynamics in this region (Dias, 1987; Jouanneau et al., 1998;<br />

Araújo et al., 2002).<br />

2. 4. 3 Climatic and vegetational response in western Iberia to North<br />

Atlantic climatic events over the last 25 000 years<br />

The comparison of the high resolution pollen composite record from<br />

the Galician margin (Fig. II.6; Tab. II.3), with other marine and terrestrial pollen<br />

sequences (Figs. II.1 and II.7; Tabs. II.4 and II.5) document the vegetation<br />

changes that occurred in the Iberian Península over the last 25 000 years.<br />

Moreover, the direct correlation between marine proxies and vegetation<br />

changes from this record will allow us to accurately evaluate the vegetation<br />

response to the climatic events detected elsewhere in the North Atlantic<br />

Ocean and over Greenland.<br />

129<br />

F. Naughton, 2007


Fig. II.6 | Galician margin composite record (MD99-2331 and MD03-2697 deep-sea cores). From the left to<br />

the right: corrected radiocarbon ages; marine proxies: δ 18O of G. bulloides, % N. pachyderma (s.), icerafted<br />

detritus (IRD), Marine and Greenland climatic events; % pollen taxa; pollen zones and<br />

chronostratigraphy. Pollen zones were established using qualitative and quantitative fluctuations of a<br />

minimum of 2 curves of ecologically important taxa (Pons and Reille, 1986). They are defined by the<br />

abbreviated name of the core (MD31 or MD97) followed by the number of the marine isotopic stage (1 or 2)<br />

and numbered from the bottom to the top (MD31-2-1 to MD31-2-5 and MD97-1-1 to MD97-1-6).<br />

130<br />

F. Naughton, 2007


Pollen<br />

zones<br />

MD97-1-6<br />

MD97-1-5<br />

MD97-1-4<br />

MD97-1-3<br />

MD97-1-2<br />

MD97-1-1<br />

MD31-2-5<br />

MD31-2-4<br />

MD31-2-3<br />

MD31-2-2<br />

MD31-2-1<br />

Pollen signature Chronostratigraphy<br />

Strong increase of Pinus (15-70%).<br />

Continuous decrease of deciduous Quercus.<br />

Ericaceae (55%), Poaceae (10%) and Taraxacum-type (10%).<br />

Continuous decline of Pinus (40-15%),<br />

deciduous Quercus (40-15%), Corylus and evergreen Quercus,<br />

presence of Alnus (8-12%). Ericaceae increase (30-55%).<br />

Gradual decline of Pinus (60-40%) and deciduous Quercus (60-<br />

40%), maximum expansion of Corylus (6-10%),<br />

beginning of Alnus continuous presence.<br />

Gradual increase of Ericaceae (10-30%) and decrease of<br />

herbaceous pollen percentages: Poaceae (


2. 4. 3. 1 Marine Isotopic Stage 2<br />

2. 4. 3. 1. 1 Heinrich events (H2 and H1)<br />

Our Galician margin composite record reveals two periods marked by<br />

the dominance of herbaceous communities (Poaceae, Ericaceae, Calluna,<br />

Cyperaceae, Aster-type, Taraxacum-type) along with a Pinus forest reduction<br />

indicating two major cold events in north-western Iberia. These events, pollen<br />

zones MD31-2-2 and MD31-2-4, are centred at around 21 700 yr BP and 14 700<br />

yr BP, respectively. In the ocean, our record identifies H2 and H1 events on the<br />

basis of, as usual in other North Atlantic cores, peaks in ice rafted detritus<br />

(IRD), high polar foraminifera (N. pachyderma s.) percentages and heavy<br />

planktonic δ 18 O values (e.g., Heinrich, 1988; Bond et al., 1993; Duplessy et al.,<br />

1993; Grousset el al., 1993; Bond and Lotti, 1995; Lebreiro et al., 1996; Baas et<br />

al., 1997; Abrantes et al., 1998; Cayre et al., 1999; Bard et al., 2000; Shackleton<br />

et al., 2000; Thouveny et al., 2000; Broecker and Hemming, 2001; de Abreu et<br />

al., 2003; Hemming, 2004). Radiocarbon ages obtained for H2 (~22 000 to ~20<br />

000 yr BP) and H1 (~15 350 to ~13 000 yr BP) intervals in our Galician margin<br />

record are in agreement with the age limits of these events, proposed by Elliot<br />

et al. (1998) for the North Atlantic.<br />

Direct correlation between pollen and marine proxies performed in this<br />

record (Fig. II.6; Tab. II.3) shows that these major cold events in north-western<br />

Iberia are only associated with the first part of H2 and H1. Indeed, H2 and H1<br />

encompass two vegetational phases. Besides the Pinus forest contraction, the<br />

first part of H2 (~22 000 to 21 500 yr BP; MD31-2-2 pollen zone) and H1 (~15 350<br />

to 14 500 yr BP; MD31-2-4 pollen zone) is characterised by the expansion of<br />

Calluna. Calluna vulgaris is a light demanding species (Calvo et al., 2002)<br />

favoured by forest regression and moist conditions. This indicates that the first<br />

part of both Heinrich events was cold and humid. Furthermore, the first part of<br />

H1 is marked by the continuous presence of the Isoetes fern suggesting also<br />

moist conditions.<br />

The second part of H2 (21 500 to 20 000 yr BP; first 1 500 yr of the MD31-<br />

2-3 pollen zone) and that of H1 (14 500 to 13 000 yr BP, MD31-2-5) are marked<br />

by a Pinus expansion, indicating less cold conditions than the previous<br />

phases. Furthermore, during the second part of H1, a gradual increase of<br />

semi-desert plants (Artemisia, Chenopodiaceae and Ephedra) reflects a<br />

132<br />

F. Naughton, 2007


gradual dryness on land. Our multiproxy palaeoclimatic record (Fig. II.6; Tab.<br />

II.3) indicates therefore that H2 and H1 events display a complex pattern on<br />

the adjacent continent.<br />

This two-phase climatic succession on land within H2 and H1 agrees<br />

with the changes detected in the marine proxy data from the same record<br />

(Fig. II.6). The first phase is represented by the heaviest δ 18 O values of G.<br />

bulloides and the increase of N. pachyderma (s.) percentages, suggesting a<br />

strong decrease in sea surface temperatures (SST), and the absence of IRD in<br />

this region. In contrast, the second phase records the lightening of the<br />

planktonic isotopic signal and the decrease in the polar foraminifera<br />

population although the presence of IRD testifies to iceberg melting off<br />

Galicia. This complex marine pattern within H2 and H1 events has already<br />

been detected further south in the SU81-18 deep-sea record (Fig. II.1) (Bard et<br />

al., 2000).<br />

Comparison between our multiproxy palaeoclimatic record (Fig. II.6)<br />

and the available terrestrial and marine pollen sequences in and off Iberia<br />

(Fig. II.1) indicate that the impact of H2 and H1 events in Iberia is spatially<br />

variable. The Pinus reduction associated with H2, dated in the Galician<br />

margin record between 22 040±180 and 21 340±160 years BP, has been<br />

already detected by terrestrial and marine pollen sequences in and off<br />

northern Iberia (Fig. II.1) (Pérez-Obiol and Julià, 1994; Roucoux et al., 2005). In<br />

north-eastern Spain, the decrease in Pinus percentages before 19,900 yrs BP<br />

detected in the Banyoles sequence and explained as the result of local<br />

factors (Pérez-Obiol and Julià, 1994) can now be interpreted as the<br />

consequence of the climatic change associated to H2. The slight expansion<br />

of Calluna recorded in the Galician margin sequence during the first phase of<br />

H2 has not been detected, however, at low altitude in north-eastern Spain<br />

where Poaceae was the dominant taxa. This suggests that heathers grow<br />

preferentially in the north-western Iberia favoured probably by Atlantic wet<br />

conditions.<br />

Southern Iberian margin cores (Fig. II.1) reveal the expansion of semi-<br />

desert associations (Artemisia, Chenopodiaceae, Ephedra), suggesting an<br />

increase of dryness during the entire H2, 22 000-20 000 years BP, although no<br />

decrease of Pinus forest was detected (SU81-18, Turon et al., 2003; ODP 976,<br />

133<br />

F. Naughton, 2007


Combourieu-Nebout et al., 2002 and SO75-6KL, Boessenkool et al., 2001). In<br />

contrast, the Padul record (Pons and Reille, 1988; Fig. II.1; Tab. II.4) shows<br />

between 23 600 and 19 800 yrs BP an alternation between periods of high<br />

Pinus pollen values and periods of high percentages of semi-desert plants. This<br />

suggests dryness variability in Sierra Nevada at that time or changes in pollen-<br />

input related with local factors.<br />

The two-phase climatic succession of H1 event characterised in our<br />

Galician margin record (MD31-2-4 and MD31-2-5 pollen zones, Fig. II.6; Tab.<br />

II.3) by a first cold and humid episode followed by a dry and cool phase is<br />

contemporaneous, as H2, with a unique aridity interval and no Pinus forest<br />

reduction in south-western Iberia (Fig. II.1) (Boessenkool et al., 2001; Turon et<br />

al., 2003). Seemingly, high altitude sites of northern Iberia and eastern Iberian<br />

sites (Padul and Banyoles) detect one event of dryness between 15 000 and<br />

13 000 (Tab. II.4, Figs. II.1 and II.7; Laguna de la Roya, Allen et al., 1996;<br />

Quintanar de la Sierra, Peñalba et al., 1997; Laguna Masegosa, Von<br />

Engelbrechten, 1998; Lagoa Lucenza, Muñoz Sobrino et al., 2001; Laguna<br />

Lleguna and Laguna de las Sanguijuelas; Muñoz Sobrino et al., 2004). The<br />

westernmost sequences record the expansion of Calluna and Isoetes, as we<br />

observed in the first part of H1 of our record, showing that these sites are also<br />

affected by wet Atlantic influence (Lagoa Lucenza, Muñoz Sobrino et al.,<br />

2001; Laguna Lleguna and Laguna de las Sanguijuelas, Muñoz Sobrino et al.,<br />

2004; Mougás, Gómez-Orellana et al., 1998). Pinus forest cover around all high<br />

altitude sites remains weak over this time-interval while our Galician margin<br />

record (Figs. II.6 and II.7; Tabs. II.3 and II.4) representing also the vegetation<br />

of low and mid altitudes sees a Pinus expansion in the second part of H1. This<br />

suggests that the temperature increase was not enough to trigger Pinus<br />

expansion in high altitude areas (Fig. II.1). These vegetational changes<br />

related with cold conditions in Iberia coincide with the Oldest Dryas originally<br />

identified in Danish deposits one century ago and dated older than 13 000<br />

years BP (Mangerud et al., 1974) and not with the Older Dryas as erroneously<br />

correlated by Turon et al. (2003). Therefore, our work demonstrates that the<br />

Oldest Dryas is the terrestrial counterpart of the H1 event.<br />

134<br />

F. Naughton, 2007


Years 14 C BP/<br />

continental<br />

sequences<br />

1200 -present<br />

day<br />

3000-1200<br />

6000-3000<br />

10000-8200<br />

Younger Dryas<br />

Late Glacial<br />

interstadial<br />

Late Glacial<br />

(Q S and LR)/<br />

Oldest Dryas<br />

(Padul)<br />

yr BP<br />

Laguna de la Roya (1608 m a.s.l)<br />

(Allen et al., 1996)<br />

Local increase of Betula (30-50%).<br />

Poaceae (20-30%), Ericaceae (5-10%),<br />

Rumex (~2%).<br />

Culture presence: Olea, Castanea and<br />

Cerealia.<br />

Absence of Pinus and Quercus decrease<br />

Poaceae (30-40%) well represented,<br />

spread of Ericaceae (10%).<br />

Decrease of Pinus, Betula and slight<br />

decrease of Quercus<br />

Succession of Juniperus, Betula, Quercus,<br />

Corylus and Alnus<br />

10290 Pinus presence (10-40%)<br />

Poaceae (20-30%), Artemisia (10%),<br />

Chenopodiaceae, Plantago,<br />

Caryophyllaceae, Anthemis-type and<br />

Calluna)<br />

Succession of Juniperus, Betula, Quercus.<br />

Pinus well represented (40-60%)<br />

12940 Pinus presence (10-30%)<br />

Poaceae (20-40%) , Artemisia (20-40%),<br />

Chenopodiaceae (2-5%) , Plantago (~2%),<br />

Caryophyllaceae (2%), Aster-type (2%) and<br />

Calluna (2%)<br />

- - -<br />

yr BP<br />

Quintanar de la Sierra (1470 m a.s.l.)<br />

(Penalba, 1994; Penalba et al., 1997)<br />

Pinus, Fagus<br />

and herbs (Ericaceae, Cerealia)<br />

135<br />

Spread of Fagus<br />

3060 Spread of Corylus<br />

Maximum percentages of trees<br />

(Betula, Pinus, deciduous Quercus, Quercus<br />

ilex and Corylus)<br />

8200 Succession of Juniperus, Betula and<br />

deciduous Quercus and Q. ilex<br />

10120 Pinus presence (20-40%)<br />

Poaceae (20-30%), Artemisia (10%),<br />

Apiaceae (10-15%), Plantago (5%), Aster<br />

(2%), Cyperaceae (2-10%),<br />

Chenopodiaceae and Calluna<br />

11050 Succession of Juniperus, Salix, Betula<br />

Pinus well represented (60-80%)<br />

13350<br />

Pinus presence (10-15%)<br />

Poaceae (20-40%), Artemisia (15-30%),<br />

Chenopodiaceae (~5%), Plantago (0-7%),<br />

Cyperaceae (5-10%), Aster (2-5%)<br />

- - -<br />

yr BP<br />

4450<br />

8200<br />

Padul (785 m a.s.l.)<br />

(Pons and Reille, 1988)<br />

- - -<br />

- - -<br />

Deciduous Quercus, Quercus ilex, Quercus<br />

suber, Pistacia.<br />

Pinus is almost absent (10%), Ephedra (5%),<br />

Cyperaceae (20-30%)<br />

Slight increase of Pinus<br />

13200<br />

Juniperus, Betula, deciduous Quercus,<br />

Quercus ilex and Pistacia<br />

Pinus decrease (10%), Cyperaceae (60-<br />

100%)<br />

15200 Pinus increase (50-75%)<br />

Artemisia (10-20%), Poaceae (10%),<br />

Chenopodiaceae (


Fig. II.7 | - Comparison between continental (Quintanar de la Sierra; Peñalba et al., 1997) and marine<br />

(MD99-2331 and MD03-2697) pollen sequences.<br />

136<br />

F. Naughton, 2007


2. 4. 3. 1. 2 The LGM<br />

In our record, the Last Glacial Maximum (LGM), bracketed by H2 and<br />

H1 events as established by the EPILOG program (Environmental processes of<br />

Ice age: Land, Oceans, Glaciers) (Mix et al., 2001), is characterised by the<br />

expansion of Pinus in an herbaceous-dominant environment along with<br />

scattered pockets of deciduous trees (MD31-2-3 pollen zone) (Fig. II.6, Tab.<br />

II.3). Planktonic δ 18 O values are slightly lower than during H2 and H1 events<br />

and N. pachyderma s. decrease to very low percentages. Pinus percentages<br />

stay constant over this interval and there is an almost continuous slight<br />

presence of deciduous tree pollen (deciduous Quercus, Betula, Corylus and<br />

Alnus) over this period. Nevertheless herbaceous communities remain the<br />

dominant group. The presence of deciduous trees has also been detected in<br />

and off southern Iberia between 20 000 and 15 000 yr BP (Tab. II.4; Pons and<br />

Reille, 1988; Boessenkool et al., 2001; Combourieu-Nebout et al., 2002)<br />

associated with the LGM (Turon et al., 2003). Our record clearly shows that not<br />

only southern but also north-western Iberia acted as a refugium zone for<br />

certain temperate trees (deciduous Quercus, Corylus, Alnus and Betula)<br />

during the last glacial maximum corroborating what has been suggested by<br />

previous studies (Roucoux et al., 2005). However, it must be noted that<br />

deciduous trees presence is weak and that they attain their maximum<br />

expression in southern Iberia. Another interesting feature within the LGM<br />

concerns the sustaining of Ericaceae communities in north-western Iberia<br />

indicated by our Galician margin core and their slight expansion in southern<br />

Iberia detected by marine cores SU81-18 (Turon et al., 2003) and ODP 976<br />

(Combourieu-Nebout et al., 2002). This is contemporaneous with the slight<br />

decrease of semi-desert associations in the middle altitudes of Sierra Nevada<br />

(Padul), indicating an increase of humidity in Iberia at that time.<br />

2. 4. 3. 2 Marine Isotopic Stage 1<br />

2. 4. 3. 2. 1 The Bölling-Allerød<br />

Following the H1 event, a drastic change in the pollen assemblage<br />

and planktonic stable oxygen isotopic values identifies the Bölling-Allerød (B-<br />

A) temperate period (Greenland Interstadial 1- GIS 1, Lateglacial interstadial).<br />

137<br />

F. Naughton, 2007


Our Galician margin pollen record (Fig. II.6, Tab. II.3) (MD97-1-1 pollen<br />

zone) detects a fast deciduous Quercus expansion and the slight<br />

development of pioneer species (Betula, Cupressaceae and Hippophae), a<br />

decrease of herbaceous associations and Pinus percentages reach<br />

maximum values. In the ocean, surface waters show an important lowering of<br />

the δ 18 O values suggesting, in absence of freshwater input, an oceanic<br />

warming at these North Atlantic mid-latitudes.<br />

In the northern Iberian Península, the Lateglacial interstadial (B-A) is<br />

characterised by the succession of pioneer associations (Juniperus-Betula-<br />

Pinus) and the more or less important expansion of deciduous trees (Fig. II.1<br />

and II.7; Tab. II.4; Allen et al., 1996; Peñalba et al., 1997; Von Engelbrechten,<br />

1998). High altitude sites of northern Iberia such as Laguna Masegosa (Von<br />

Engelbrechten, 1998), Laguna de la Roya (Allen et al., 1996), Hojos de Iregua<br />

(Gil-Garcia et al., 2002) and Laguna de las Sanguijuelas (Muñoz Sobrino et al.,<br />

2004) record a deciduous Quercus expansion between 13 000 and 11 000 yrs<br />

BP above 1 000 m a.s.l.. Our pollen analysis records higher pollen percentages<br />

of deciduous Quercus than the high altitude sequences suggesting that<br />

deciduous Quercus woodlands expanded preferentially in the lowlands and<br />

mid-altitudes of northern Iberia. Brewer et al. (2002), based on a small number<br />

of high altitude northern and low altitude southern sequences, suggested that<br />

only southern Iberia acted as a refugium zone for deciduous oak during the<br />

last glacial period. However, as previously shown by our Galician record, the<br />

mid and low-altitudes of north-western Iberia were a refugium zone for<br />

deciduous Quercus species allowing the fast spread of these taxa during B-A<br />

climate improvement.<br />

In southern Iberia, a rapid expansion of deciduous and evergreen<br />

Quercus and other Mediterranean elements is recorded in the Lateglacial<br />

interstadial of the Padul peat-bog sequence and marine records SU81-18<br />

(Turon et al., 2003), 8057 B (Hooghiemstra et al., 1992), SO75-6KL (Boessenkool<br />

et al., 2001) and ODP 976 (Combourieu Nebout et al., 1999, 2002). Indeed, a<br />

distinct phase of pioneer trees is not reflected at the beginning of this<br />

interstadial neither in southern Iberia nor in low and mid altitudinal sites of the<br />

north-western Iberia as shown by our Galician margin pollen record.<br />

138<br />

F. Naughton, 2007


2. 4. 3. 2. 2 The Younger Dryas cold event<br />

Following the B-A warm phase, our Galician margin pollen record<br />

(MD97-1-2 pollen zone) sees the increase of pioneer species (Betula), grasses<br />

and semi-desert associations (Artemisia and Ephedra) at the expense of the<br />

temperate forest. These vegetational features characterise the Younger Dryas<br />

cold event (Fig. II.6, Tab. II.3).<br />

This cold episode has been detected in and off Iberia (Pons and Reille, 1988;<br />

Pérez-Obiol and Julià, 1994; Allen et al., 1996; Peñalba et al., 1997; von<br />

Engelbrechten, 1998; Gil Garcia et al., 2002; Turon et al., 2003) (Fig. II.1, Tab.<br />

II.4) and is associated with a slight increase of the planktonic δ 18 O values (Hall<br />

and McCave, 2000; Schönfeld and Zahn, 2000; Löwemark et al., 2004; Turon<br />

et al., 2003).<br />

The slight decrease of deciduous Quercus, the increase of pioneer<br />

species (Betula) and the expansion of semi-desert and herbaceous plants are<br />

also represented in north-eastern Iberia as well as in continental and offshore<br />

southern Iberian sequences (Pons and Reille, 1988; Pérez-Obiol and Julià,<br />

1994; Boessenkool et al., 2001; Turon et al., 2003). Vegetation changes related<br />

to this short event appear more drastic at high altitudinal sites of northern<br />

Iberia than at low and mid altitude sites of north-western Iberia (this study)<br />

and those from the south as indicated by the slight decrease of temperate<br />

trees in terrestrial and marine pollen sequences in and off southern Iberia (Fig.<br />

II.7).<br />

2. 4. 3. 2. 3 The Holocene<br />

After the YD, the tree succession of deciduous Quercus, Corylus and<br />

Alnus defines the Holocene in our Galician margin record (Fig. II.6, Tab. II.3).<br />

Remarkably, during the onset of the Holocene (MD97-1-3 pollen zone), the<br />

increase of deciduous Quercus pollen percentages tightly parallels the<br />

lightening of δ 18 O values (Fig. II.6). However, deciduous forest attains its<br />

maximum expression slightly before sea surface water experiences its lightest<br />

δ 18 O values. Pinus pollen values decline steadily in the pollen zones MD97-1-3<br />

through MD97-1-5. The end of the maximum expansion of deciduous Quercus<br />

trees and the beginning of the Corylus increase (MD97-1-4) occur<br />

139<br />

F. Naughton, 2007


contemporaneously with the beginning of the lightest δ 18 O isotopic values in<br />

the ocean. Deciduous Quercus gradually decreases until the end of the<br />

Holocene (from MD97-1-4 to MD97-1-6).<br />

The expansion of heaths (Ericaceae and Calluna) and ferns (as<br />

indicated by psilate Trilete spores) along with the reduction of trees<br />

(deciduous Quercus, Corylus and Betula) mark the late Holocene phases<br />

(MD97-1-5 and MD97-1-6). The minimum pollen values of Pinus are recorded in<br />

zone MD97-1-5 while maxima pollen percentages of this tree, in the<br />

successive zone MD97-1-6, probably reflect the reforestation of the last 350<br />

years.<br />

In Iberia, vegetation response to climate amelioration that<br />

characterises the Holocene period looks quite similar to that of the B-A event.<br />

The settlement of the Mediterranean forest occurred very fast in southern sites<br />

(Fig. II.1) as illustrated by the pollen sequences of Padul (Tab. II.4, Pons and<br />

Reille, 1988), Charco da Candieira (Van der Knaap and Van Leeuwen, 1995),<br />

SU81-18 (Turon et al., 2003), ODP 976 (Combourieu-Nebout et al. 2002), 8057 B<br />

(Hooghiemstra et al., 1992), and SO75-6KL (Boessenkool et al., 2001). In the<br />

north, vegetation response to the Holocene climate appears slower than in<br />

the south. The expansion of pioneer trees (Juniperus, Betula and Pinus) marks<br />

the beginning of this period in the high altitude sites of northern Iberia<br />

followed by the development of deciduous Quercus, Corylus and Alnus (Tab.<br />

II.5; Fig. II.1). This succession is also clearly detected by our Galician marine<br />

record synchronously with the decrease of planktonic δ 18 O values which are<br />

contemporaneous with the sea surface gradual warming detected by de<br />

Abreu et al. (2003) and Schönfeld et al. (2003) in the Iberian margin. However,<br />

this vegetation succession, recorded in our Galician sequence, begins earlier,<br />

during the Younger Dryas event, in the low and mid altitude sites than in the<br />

high altitudinal sites of north-western Iberia. The Galician margin record further<br />

suggests that the maximum development of deciduous Quercus forest leads<br />

the lightest values of planktonic δ 18 O during the Holocene. Finally, the late<br />

Holocene interval of all Iberian marine and terrestrial sequences indicates the<br />

decline of the temperate forest during the last 5000 years.<br />

140<br />

F. Naughton, 2007


Quintanar de<br />

la Sierra<br />

(42°02’N,3°W)<br />

1608 m asl<br />

Laguna<br />

de La<br />

Roya<br />

(42°6’N,6°44’W)<br />

1085 m asl<br />

Lago<br />

de Ajo<br />

(43°3’N,6°9’W)<br />

1570 m asl<br />

Laguna<br />

Masegosa<br />

(42°57’N,2°49’W)<br />

1600 m asl<br />

Laguna Negra<br />

(42°0’N,2°52’W)<br />

1760 m asl<br />

Banyoles<br />

Lake<br />

(42°08’N,2°45’E)<br />

173 m asl<br />

Las Pardillas<br />

Lake<br />

(42°2’N,3°2’W)<br />

1850 m asl<br />

141<br />

Laguna<br />

Lucenza<br />

(Sierra de<br />

Queixa)<br />

(Galicia)<br />

1420 m asl<br />

Hoyos de<br />

Iregua<br />

(42°01’N,2°45’W)<br />

1780 m asl<br />

Pinus Betula Betula Pinus Pinus Pinus Pinus Betula Pinus Betula<br />

Fagus<br />

Taxus<br />

Fagus<br />

Taxus<br />

Corylus Alnus<br />

Ulmus<br />

Fraxinus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Pinus<br />

Salix<br />

Betula<br />

Juniperus<br />

Fagus Fagus Fagus<br />

Taxus<br />

Ulmus<br />

Alnus<br />

Fraxinus<br />

Corylus<br />

Alnus<br />

Ulmus<br />

Taxus<br />

Fraxinus<br />

Corylus<br />

Dec.Quercus Quercus Eve.Quercus<br />

Dec.Quercus<br />

Betula<br />

Juniperus<br />

Betula<br />

Juniperus<br />

Pinus<br />

Salix<br />

Betula<br />

Juniperus<br />

Alnus<br />

Ulmus<br />

Fraxinus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Pinus<br />

Salix<br />

Betula<br />

Juniperus<br />

Abies<br />

Alnus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Acer<br />

Betula<br />

Pinus<br />

Juniperus<br />

Fagus Salix Fagus Fagus<br />

Alnus<br />

Ulmus<br />

Fraxinus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Salix<br />

Betula<br />

Pinus<br />

Juniperus<br />

Alnus<br />

Ulmus<br />

Corylus<br />

Alnus<br />

Salix Corylus<br />

Dec.Quercus Eve.Quercus<br />

Dec.Quercus<br />

Salix<br />

Betula<br />

Pinus<br />

Pinus<br />

Betula<br />

Tab. II.5| Holocene tree succession in north-western Iberia.<br />

Pozo do<br />

Carballal<br />

(42°42’N,7°07’W)<br />

1330 m asl<br />

Alnus<br />

Ulmus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Salix<br />

Betula<br />

Juniperus<br />

Pinus<br />

Lagoa Lucenza<br />

(42°35’N,7°07’W)<br />

1375 m asl<br />

Betula<br />

Fraxinus<br />

Alnus<br />

Ulmus Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Betula<br />

Pinus<br />

Juniperus<br />

Laguna de las<br />

Sanguijuelas<br />

(42°08’N,6°42’W)<br />

1080 m asl<br />

Fagus<br />

Ulmus<br />

Alnus<br />

Fraxinus<br />

Corylus<br />

Eve.Quercus<br />

Dec.Quercus<br />

Salix<br />

Juniperus<br />

Pinus<br />

Betula<br />

F. Naughton, 2007<br />

Holocene subphases<br />

Late-Holocene<br />

5000-4000-3000<br />

yr BP<br />

until present<br />

day<br />

Mid-Holocene<br />

9000-8000 yr BP<br />

to<br />

5000-4000 yr BP<br />

Early Holocene<br />

10500 yr BP to<br />

9000-8000 yr BP


2. 5 Conclusions<br />

The comparison of present-day terrestrial and marine pollen samples in<br />

and off western Iberia shows that the pollen signature from the Iberian margin<br />

is similar to that of the Iberian terrestrial deposits, and, in particular, to that of<br />

the estuarine samples which recruit pollen from the vegetation colonising the<br />

adjacent hydrographic basins. Therefore, western Iberian margin pollen<br />

spectra reflect an integrated image of the regional vegetation of the<br />

adjacent continent. Furthermore, our study shows that marine pollen spectra<br />

clearly discriminate both the Mediterranean and the Atlantic plant<br />

communities colonising southern and northern Iberian Península, respectively.<br />

It also identifies the present-day pattern of pollen transport in northern and<br />

southern Iberian margin during downwelling and upwelling conditions.<br />

High resolution pollen and marine proxies analysis from the Galician<br />

margin composite core (MD99-2331 and MD03-2697) shows a synchronicity of<br />

the vegetation response to the North Atlantic climatic variability during H2,<br />

LGM, H1, B-A, YD events. Comparison of this palaeoclimatic record with other<br />

marine and terrestrial pollen records shows that the beginning of both H2 and<br />

H1 cold events are associated with Pinus forest reduction in northern Iberia. It<br />

also shows the presence of two vegetation phases within H1 and H2 events,<br />

associated with an initial cold and wet episode followed by a cool and,<br />

particularly, dry episode during H1. Furthermore this comparison allows us to<br />

demonstrate that the Oldest Dryas event on the continent corresponds to the<br />

H1 event in the ocean.<br />

The slight presence of deciduous Quercus, Corylus and Alnus during<br />

the Last Glacial Maximum shows that not only southern Iberia but also<br />

northern Iberia acted as a refugium zone for these trees, though at a smaller<br />

scale. Bölling-Allerød interstadial in our sequence, which mainly represents low<br />

and mid-altitude zones, show a more rapid and great expansion of<br />

deciduous Quercus than the high altitude sites of north-western Iberia,<br />

indicating that the vegetation of low and mid-altitudes responded more<br />

rapidly to the climate variability of the North Atlantic during this interstadial.<br />

Because deciduous forest attained it maximum expression during the B-A<br />

interstadial in low and mid-altitudes of the north-western Iberia, the climate


eversal characterizing the Younger Dryas event is less marked in these zones<br />

than in the high altitude ones. The response of deciduous forest to the climate<br />

improvement that characterises the onset of the Holocene at low and mid<br />

altitudes of north-western Iberia seems to lead those observed in the high<br />

altitude sites, although the same succession of trees is observed in all these<br />

northern regions.<br />

This study confirms that marine pollen sequences from western Iberian<br />

margin are a powerful tool for accurate reconstruction of vegetation<br />

response to oceanic and atmospheric climate changes within a reliable<br />

chronological framework. Furthermore, it demonstrates the importance of<br />

including vegetation reconstructions from marine pollen sequences in future<br />

efforts to refine and model vegetation and climate dynamics in the Iberian<br />

Península.<br />

References:<br />

Abrantes, F., Baas, J., Haflidason, H., Rasmussen, T., Klitgaard, D., Loncaric, N. and<br />

Gaspar, L., 1998. Sediment fluxes along the northeastern European Margin: inferring<br />

hydrological changes between 20 and 8 kyr. Marine Geology, 152: 7-23.<br />

Alcara Ariza, F. Asensi Marfil, A., de Bolos y Capdevilla, O., Costa Tales, M., Arco<br />

Aguilar, M., Diaz Gonzales, T.E., Diez Garretas, B., Fernandez Prieto, J.A., Fernandez<br />

Gonzales, F., Izco Sevillando, J., Loidi Arregui, J., Martinez Parras, J.M., Navarro Andres,<br />

F., Ninot I Sugranes, J.M., Peinado Lorca, M., Rivas Martinez, S., Sanchez Mata, D.,<br />

Valle Guitierrez, C., Vigo I Bonada, J. and Wildpret de la Torre, W., 1987. La<br />

vegetación de España. Collection Aula Abierta. Universidad de Alcala de Henares,<br />

544 pp.<br />

Allen, J.R.M., Huntley, B. and Watts, W.A., 1996. The vegetation and climate of<br />

northwest Iberia over the last 14 000 yr. Journal of Quaternary Science, 11: 125-147.<br />

Álvarez–Salgado, X.A., Figueiras, F.G., Pérez, F.F., Groom, S., Nogueira, E., Borges, A.V.,<br />

Chou, L., Castro, C.G., Moncoiffé, G., Ríos, A.F., Miller, A.E.J., Frankignoulle, M.,<br />

Savidge, G. and Wollast, R., 2003. The Portugal coastal counter current off NW Spain:<br />

new insights on its biogeochemical variability. Progress in Oceanography, 56: 281-321.<br />

Ambar, I. and Fiúza, A., 1994. Some features of the Portugal Current System: A<br />

poleward slope undercurrent, an upwelling related southward flow and an autumnwinter<br />

poleward coastal surface current. 2 nd International Conference on Air-Sea<br />

Interaction and on Meterology and Oceanography of the Coastal Zone. American<br />

Meteorological Society: 286-287.<br />

Ambar, I., Fiúza, A.F.G., Boyd, T. and Frouin, R., 1986. Observations of a warm oceanic<br />

current flowing northward along the coasts of Portugal and Spain during Nov-Dec<br />

1983. EOS Transaction, American Geophysical Union, 67, 1054 pp.<br />

Araújo, M. F., Jouanneau, J.M. and Dias, J.M.A., 1994. Chemical characterisation of<br />

the main fine sedimentary deposit at the northwester Portuguese shelf. Gaia, 9: 59-65.<br />

143<br />

F. Naughton, 2007


Araújo, M. F., Jouanneau, J.M., Valerio, P., Barbosa, T., Gouveia, A., Weber, O.,<br />

Oliveira, A., Rodrigues, A. and Dias, J.M.A., 2002. Geochemical tracers of northern<br />

Portuguese estuarine sediments on the Iberian Shelf. Progress in Oceanography, 52:<br />

277-297.<br />

Arhan, M., Colin de Verdière, A. and Memery, L., 1994. The eastern boundary of the<br />

subtropical North Atlantic. Journal of Physical Oceanography, 24: 1295–1316.<br />

Baas, J.H., Mienert, J., Abrantes, F. and Prins, M.A., 1997. Late Quaternary<br />

sedimentation on the Portuguese continental margin: climate-related processes and<br />

products. Palaeogeography, Palaeoclimatology, Palaeoecology, 130: 1-23.<br />

Barboni, D., Harrison, S.P., Bartlein, P.J., Jalut, G., New, M., Prentice, I.C., Sanchez-<br />

Goñi, M.-F., Spessa, A., Davis, B. and Stevenson, A.C., 2004. Relationships between<br />

plant traits and climate in the Mediterranean region: A pollen data analysis. Journal<br />

of Vegetation Science, 15: 635- 646.<br />

Bard, E., Rostek, F., Turon, J.L. and Gendreau, S., 2000. Hydrological impact of Heinrich<br />

events in the subtropical northeast Atlantic. Science, 289: 1321-1324.<br />

Bard, E., Rostek, F. and Menot-Combes, G., 2004. Radiocarbon calibration beyond<br />

20,000 14C yr B.P. by means of planktonic foraminifera of the Iberian Margin.<br />

Quaternary Research, 61: 204-214.<br />

Barton, E.D., 1998. Eastern boundary of the North Atlantic - northwest Africa and<br />

Iberia. Robinson, A.R., Brink, K. (Eds.), The Sea, 11, 633-657 pp.<br />

Blanco Castro, E., Casado González, M.A., Costa Tenorio, M., Escribano Bombín R.,<br />

García Antón, M., Génova Fuster, M., Gómez Manzaneque, F., Moreno Sáis, J.C.,<br />

Morla Juaristi, C., Regato Pajares, P. and Sáiz Ollero, H., 1997. Los Bosques Ibéricos,<br />

Barcelona, 572 pp.<br />

Boessenkool, K.P., Brinkhuis, H., Schonfeld, J. and Targarona, J., 2001. North Atlantic<br />

sea-surface temperature changes and the climate of western Iberia during the last<br />

deglaciation; a marine palynological approach. Global and Planetary Change, 30:<br />

33-39.<br />

Bond, G. and Lotti, R., 1995. Iceberg discharges into the North Atlantic on millennial<br />

time scales during the last glaciation. Science, 267: 1005-1010.<br />

Bond, G., Broecker, W., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani,<br />

G., 1993. Correlations between climate records from North Atlantic sediments and<br />

Greenland ice. Nature, 365: 143-147.<br />

Bottema, S. and van Straaten, L.M.J.U., 1966. Malacology and palynology of two<br />

cores from the Adriatic sea floor. Marine Geolology, 4: 553-564.<br />

Braun-Blanquet, J., Pinto da Silva, A.R., and Rozeira, A., 1956. Résultats de deux<br />

excursions géobotaniques à travers le Portugal septentrional et moyen. II. Chenaies à<br />

feuilles caduques (Quercion occidentale) et chenaies à feuilles persistentes<br />

(Quercion faginea) au Portugal. Agronomia Lusititana, 18 : 167–234.<br />

Brewer, S., Cheddadi, R., de Beaulieu, J.-L., Reille, M. and contributors, D., 2002. The<br />

spread of deciduous Quercus throuout Europe since the last glacial period. Forest<br />

Ecology and Management, 156: 27-48.<br />

144<br />

F. Naughton, 2007


Broecker, W.S. and Hemming, S., 2001. Climate swings come into focus. Science, 294:<br />

2308-2309.<br />

Calvo, L., Tarrega, R., Luis, E., 2002. Regeneration patterns in a Calluna vulgaris<br />

heathland in the Cantabrian mountains (NW Spain): effects of burning, cutting and<br />

ploughing. Acta Oecol. 23, 81-90.<br />

Cayre, O., Lancelot, Y. and Vincent, E., 1999. Paleoceanographic reconstructions<br />

from plaktonic foraminifera off the Iberian Margin: Temperature, salinity, and Heinrich<br />

events. Paleoceanography, 14: 384-396.<br />

Chmura, G.L. and Eisma, D., 1995. A palynological study of surface and suspended<br />

sediment on a tidal flat: Implications for pollen transport and deposition in coastal<br />

waters. Marine Geology 128:183-200.<br />

Combourieu Nebout, N., Londeix, L., Baudin, F., Turon, J.-L., von Grafnestein, R. and<br />

Zahn, R., 1999. Quaternary marine and continental paleoenvironments in the western<br />

Mediterranean (Site 976, Alboran Sea): Palynological evidence. R. Zahn, Comas, M.C.<br />

and Klaus, A. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results., pp<br />

457-468.<br />

Combourieu-Nebout, N., Turon, J.-L., Zahn, R., Capotondi, L., Londeix, L. and Pahncke,<br />

K., 2002. Enhanced aridity and atmospheric high-pressure stability over the western<br />

Mediterranean during the North Atlantic cold events of the past 50 ky. Geology, 30:<br />

863-866.<br />

Cross, A.T., Thompson, G.G. and Zaitzeff, J.B., 1966. Source and distribution of<br />

palynomorphs in bottom sediments, southern part of Gulf of California. Marine<br />

Geology, 4: 467-524.<br />

de Abreu, L., Shackleton, N.J., Schonfeld, J., Hall, M. and Chapman, M., 2003.<br />

Millennial-scale oceanic climate variability off the Western Iberian margin during the<br />

last two glacial periods. Marine Geology, 196: 1-20.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3: 16-27.<br />

Desprat, S., 2005. Réponses climatiques marines et continentales du SW de l'Europe<br />

lors des derniers interglaciaires et des entrées en glaciation. PhD Thesis, Bordeaux 1<br />

University, Bordeaux (France), 282 pp.<br />

Desprat, S., Sánchez Goñi, M.F. and Loutre, M.F., 2003. Revealing climatic variability of<br />

the last three millennia in northwestern Iberia using pollen influx data. Earth and<br />

Planetary Science Letters, 213: 63-78.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., McManus, J.F., Loutre, M.F. Duprat, J.,<br />

Malaizé, B., Peyron, O. and Peypouquet, J.-P., 2005. Is vegetation responsible for<br />

glacial inception during periods of muted insolation changes? Quaternary Science<br />

Reviews, 24: 1361-1374.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., Duprat, J., Malaizé, B. and Peypouquet,<br />

J.-P., 2006. Climatic variability of Marine Isotope Stage 7: direct land-sea-ice<br />

correlation from a multiproxy analysis of a northwestern Iberian margin deep-sea<br />

core. Quaternary Science Reviews, 25: 1010-1026.<br />

145<br />

F. Naughton, 2007


Desprat, S., Sánchez Goñi, M.F., Naughton, F., Turon, J.-L., Duprat, J., Malaizé, B.,<br />

Cortijo, E. and Peypouquet, J.-P., in press. Climate variability of the last five isotopic<br />

interglacials from direct land-sea-ice correlation. Sirocko, F., Litt, T., Claussen, M.,<br />

Sánchez Goñi, M.F. (eds.), Climate of past interglacials.<br />

Dias, J.M.A., 1987. Dinâmica sedimentar e evolução recente de plataforma<br />

continental Portuguesa setentrional. PhD Thesis, Lisbon University, Lisbon (portugal),<br />

384 pp.<br />

Dias, J.M.A., Gonzalez, R., Garcia, C. and Diaz del Rio, V., 2002. Sediment distribution<br />

patterns on the Galicia-Minho continental shelf. Progress in Oceanography, 52: 215-<br />

231.<br />

Drago, T., Oliveira, A., Magalhães, F., Cascalho, J., Jouanneau, J.-M. and Vitorino, J.<br />

1998. Some evidences of northward fine sediment transport in the Portuguese<br />

continental shelf. Oceanologica Acta, 21: 223-231.<br />

Drago, T., Araújo, F., Valério, P., Weber, O. and Jouanneau, J.M., 1999.<br />

Geomorphological control of fine sedimentation on the northern Portuguese shelf.<br />

Boletin del Instituto Espanol de Oceanografia, 15: 111-122.<br />

Duplessy, J.C., Bard, E., Labeyrie, L., Duprat, J. and Moyes, J., 1993. Oxygen isotope<br />

records and salinity changes in the Northeastern Atlantic Ocean during the last<br />

18,000 years. Paleoceanography, 8: 341-350.<br />

Dupont, L., Jahns, S., Marret, F. and Ning, S., 2000. Vegetation change in equatorial<br />

West Africa: time-slices for the last 150 ka Palaeogeography Palaeoclimatology<br />

Palaeoecology, 155: 95-122.<br />

Elliot, M. Labeyrie, L., Bond, G., Cortijo, E., Turon, J.-L., Tisnerat, N. and Duplessy, J.-C.,<br />

1998. Millenial-scale iceberg discharges in the Irminger Basin during the last glacial<br />

period: Relationship with the Heinrich events and environmental settings.<br />

Paleoceanography, 13: 433-446.<br />

Fiúza, A.F.G., 1984. Hidrologia e Dinâmica das Águas Costeiras Portuguesas. PhD<br />

Thesis, Lisbon University, Lisbon (Portugal), 294 pp.<br />

Fiúza, A.F.G., Macedo, M.E. and Guerreiro, M.R., 1982. Climatological space and time<br />

variation of the Portuguese coastal upwelling. Oceanologica Acta, 5: 31-40.<br />

Frouin, R., Fiúza, A.F.G., Ambar, I. and Boyd, T.J., 1990. Observations of a poleward<br />

surface current off the coasts of Portugal and Spain during winter. Journal of<br />

Geophysical Research, 95: 679-691.<br />

Gil García, M.J., Dorado Valiño, M., Valdeolmillos Rodríguez, A. and Ruiz Zapata, M.B.,<br />

2002. Late-glacial and Holocene palaeoclimatic record from Sierra de Cebollera<br />

(northern Iberian Range, Spain). Quaternary International, 93-94: 13-18.<br />

Gomez-Orellana, L., Ramil-Rego, P. and Munoz Sobrino, C., 1998. Una nueva<br />

secuencia polinica y cronologica par el deposito pleistoceno de Mougas (NW de la<br />

Peninsula Iberia). Revue de Paléobiologie, 17, 35-47.<br />

Gouzy, A., Malaizé, B., Pujol C. and Charlier K., 2004. Climatic "pause" during<br />

Termination II identified in shallow and intermediate waters off the Iberian margin.<br />

Quaternary Science Reviews, 23: 1523-1528.<br />

146<br />

F. Naughton, 2007


Groot, J.J., Groot, C.R., 1966. Pollen spectra from deep-sea sediments as indicators of<br />

climatic changes in southern South America. Marine Geology, 4: 525-537.<br />

Groot, J.J. and Groot, C.R., 1971. Horizontal and vertical distribution of pollen and<br />

spores in Quaternary sequences. Funnel, B.M., Reidel, W.R. (Eds.), The<br />

micropaleontology of Oceans. Cambridge University Press, Cambridge: 493-504.<br />

Grousset, F.E. Labeyrie, L., Sinko, J.A., Cremer, M., Bond, G., Duprat, J., Cortijo, E. and<br />

Huon, S.,1993. Patterns of ice-rafted detritus in the glacial North Atlantic (40-55°N).<br />

Paleoceanography, 8: 175-192.<br />

Hall, I.R. and McCave, I.N., 2000. Palaeocurrent reconstruction, sediment and thorium<br />

focussing on the Iberian margin over the last 140 ka. Earth and Planetary Science<br />

Letters, 178: 151-164.<br />

Haynes, R.D. and Barton. E.D., 1990. A poleward flow along the Atlantic coast of the<br />

Iberian Peninsula. Journal of Geophysical Research, 95: 11425-11442.<br />

Heinrich, H., 1988. Origin and consequences of cyclic ice rafting in the northeast<br />

Atlantic ocean during the past 130,000 years. Quaternary Research, 29: 142-152.<br />

Hemming, S., 2004. Heinrich events: massive late Pleistocene detritus layers of the<br />

North Atlantic and their global climate imprint. Review of Geophysics, 42: 1-43.<br />

Heusser, L.E. and Balsam, W.L., 1977. Pollen distribution in the N.E. Pacific Ocean.<br />

Quaternary Research, 7: 45-62.<br />

Heusser, L.E. and Shackleton, N.J., 1979. Direct marine-continental correlation:<br />

150,000-year oxygen isotope-pollen record from the North Pacific. Science, 204: 837-<br />

839.<br />

Heusser, L.E. and Stock, C.E., 1984. Preparation techniques for concentrating pollen<br />

from marine sediments and other sediments with low pollen density. Palynology, 8:<br />

225-227.<br />

Hooghiemstra, H., Stalling, H., Agwu, C.O.C. and Dupont, L.M., 1992. Vegetational<br />

and climatic changes at the northern fringe of the Sahara 250,000-5000 years BP:<br />

evidence from 4 marine pollen records located between Portugal and the Canary<br />

Islands. Review of Palaeobotany and Palynology, 74: 1-53.<br />

Hooghiemstra, H., Lézine, A-M., Leroy, S., Dupont, L. and Marret, F., 2006. Late<br />

Quaternary palynology in marine sediments: A synthesis of the understanding of<br />

pollen distribution patterns in the NW African setting. Quaternary International, 148:<br />

29-44.<br />

Hughen, K.A., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell,<br />

P.G., Buck, C.E., Burr, G., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G.,<br />

Friedrich, M., Guilderson, T.P., Kromer, B., McCormac, F.G., Manning, S., Bronk Ramsey,<br />

C., Reimer, P.J., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J. and Weyhenmeyer, C.E., 2004. Marine04 Marine<br />

Radiocarbon Age Calibration, 0-26 Cal Kyr BP. Radiocarbon, 46: 1059-1086.<br />

Huthnance, J.M., Van Aken, H.M., White, M., Barton, E.D., Le Cann, B., Coelho, E.F.,<br />

Fanjul, E.A., Miller, P. and Vitorino, J., 2002. Ocean margin exchange-water flux<br />

estimates. Journal of Marine Systems, 32: 107-137.<br />

147<br />

F. Naughton, 2007


Jouanneau, J.M. Garcia, C., Oliveira, A., Rodrigues, A., Dias, J.A., Weber, O., 1998.<br />

Dispersal and deposition of suspended sediment on the shelf off the Tagus and Sado<br />

estuaries, S.W. Portugal. Progress in Oceanography, 42: 233-257.<br />

Jouanneau, J.M. Weber, O., Drago, T., Rodrigues, A., Oliveira, A., Dias, J.M.A., Garcia,<br />

C., Schimdt, S., Reyss, J.L., 2002. Recent sedimentation and sedimentary budgets on<br />

the western Iberian shelf. Progress In Oceanography, 52: 261-275.<br />

Koreneva, E.V., 1966. Marine palynological researches in the U.S.S.R. Marine Geology,<br />

4: 565-574.<br />

Lebreiro, S.M., Moreno, J.C., McCave, I.N. and Weaver, P.P.E., 1996. Evidence for<br />

Heinrich layers off Portugal (Tore Seamount: 39°N, 12°W). Marine Geology, 131: 47-56.<br />

Lopez-Jamar, E., Cal, R.M., Gonzalez, G., Hanson, R.B., Rey, J., Santiago, G. and<br />

Tenore, K.R., 1992. Upwelling and outwelling effects on the benthic regime of the<br />

continental shelf off Galicia, NW Spain. Journal of Marine Research, 50: 465-488.<br />

Loureiro, J.J., Machado, M.L., Macedo, M.E., Nunes, M.N., Botelho, O.F., Sousa, M.L.,<br />

Almeida, M.C., and Martins, J.C., 1986. Monografias hidrológicas dos principias cursos<br />

de água de Portugal continental. Direcção Geral dos Serviços Hidráulicos, Lisbon<br />

(portugal), 569pp.<br />

Löwemark, L., Schönfeld, J., Werner, F. and Schäfer, P., 2004. Trace fossils as a<br />

paleoceanographic tool: evidence from Late Quaternary sediments of the<br />

southwestern Iberian margin. Marine Geology, 204: 27-41.<br />

Mangerud, J., Andersen, S.T., Berglund, B.E. and Donner, J.J., 1974. Quaternary<br />

stratigraphy of Norden, a proposal for terminology and classification. Boreas, 3: 109-<br />

128.<br />

Mazé, J.P., Arhan, M. and Mercier, H., 1997. Volume budget of the eastern boundary<br />

layer off the Iberian Peninsula. Deep-Sea Research, 44: 1543-1574.<br />

McAndrew, J.H. and King, J.E., 1976. Pollen of the North American Quaternary: The<br />

top twenty. Geoscience and Man, 15: 41-49.<br />

McCave, I.N., Hall, I.R., 2002. Turbidity of waters over the Northwest Iberian continental<br />

margin. Progress In Oceanography, 52: 299-313.<br />

Migeon, S., Weber, O., Faugeres, J.C. and Saint-Paul, J., 1999. SCOPIX: A new imaging<br />

system for core analysis. Geo-Marine Letters, 18: 251-255.<br />

Mix, A.C., Bard, E. and Schneider, R., 2001. Environmental processes of the ice age:<br />

land, oceans, glaciers (EPILOG). Quaternary Science Reviews, 20: 627-657.<br />

Moore, P.D., Webb, J.A. and Collinson, M.E., 1991. Pollen analysis. Blackwell scientific<br />

publication, Oxford, 216 pp.<br />

Mudie, P. J., 1982. Pollen distribution in recent marine sediments, eastern Canada.<br />

Canadian Journal of Earth Sciences, 19: 729-747.<br />

Muller, J., 1959. Palynology of recent Orinoco delta and shelf sediments.<br />

Micropaleontology, 5: 1-32.<br />

148<br />

F. Naughton, 2007


Muñoz Sobrino, C., Ramil-Rego, P. and Rodríguez Guitián, M.A., 1997. Upland<br />

vegetation in the north-west peninsula after the last glaciation: forest history and<br />

deforestation dynamics. Vegetation History and Archeobotany, 6: 215-233.<br />

Muñoz Sobrino, C., Ramil-Rego, P. and Rodríguez Guitián, M.A., 2001. Vegetation in<br />

the mountains of northwest Iberia during the last glacial-interglacial transition.<br />

Vegetation History and Archeobotany, 10: 7-21.<br />

Muñoz Sobrino, C., Ramil-Rego, P. and Gomez-Orellana, L., 2004. Vegetation of the<br />

Lago Sanabria area (NW Iberia) since the end of the Pleistocene: a<br />

palaeoecological reconstruction on the basis of two new pollen sequences.<br />

Vegetation History and Archeobotany, 13: 1-22.<br />

Oliveira, A., Rodrigues, A., Jouanneau, J.M., Weber, O., Dias, J.A. and Vitorino, J.,<br />

1999. The suspended matter distribution and composition in the northern portuguese<br />

margin. Boletin del Instituto Espanol de Oceanografia, 15: 101-109.<br />

Oliveira, A., Rocha, F., Rodrigues, A., Jouanneau, J.M., Dias, J.M.A., Weber, O. and<br />

Gomes, C., 2002. Clay mineral from the sedimentary cover from the Northwestern<br />

Iberian shelf. Progress in Oceanography, 52: 233-247.<br />

Peck, R.M., 1973. Pollen budget studies in a small Yorkshire catchment. In Birks, H.J.B.<br />

and West, R.G. Eds. Quaternary Plant Ecology, The 14 th symposium of the British<br />

Ecological Society, University of Cambridge, Blackwell Scientific Publications, 43-60.<br />

Peñalba, C., 1994. The history of the Holocene vegetation in northern Spain from<br />

pollen analysis. Journal of Ecology, 82: 815-832.<br />

Peñalba, M.C., Arnold, M., Guiot, J., Duplessy, J.-C. and de Beaulieu, J.-L., 1997.<br />

Termination of the Last Glaciation in the Iberian Peninsula inferred from the pollen<br />

sequence of Quintanar de la Sierra. Quaternary Research, 48: 205-214.<br />

Pérez-Obiol, R. and Julià, R., 1994. Climatic change on the Iberian Peninsula recorded<br />

in a 30,000-yr pollen record from Lake Banyoles. Quaternary Research, 41: 91-98.<br />

Peyron, O., Guiot, J., Cheddadi, R., Tarasov, P., Reille, M., de Beaulieu, J.-L., Bottema,<br />

S. and Andrieu, V., 1998. Climatic Reconstrustion in Europe for 18,000 yr BP from Pollen<br />

Data. Quaternary Research, 49: 183-196.<br />

Pingree, R.D. and Le Cann, B., 1990. Structure, strength and seasonality of the slope<br />

currents in the Bay of Biscay region. Journal of the Marine Biological Association of the<br />

United Kingdom, 70: 857-885.<br />

Pingree, R.D. and Le Cann, B., 1992. Three anti-cyclonic slope water oceanic eddies<br />

(SWODDIES) in the southern Bay of Biscay. Deep-Sea Research, 39: 1147-1175.<br />

Polunin, O. and Walters, M., 1985. A guide to the vegetation of Britain and Europe.<br />

Oxford University Press, New York, 238 pp.<br />

Pons A. and Reille, M., 1986. Nouvelles recherches pollenanalytiques à Padul<br />

(Granada) : La fin du dernier glaciaire et l’Holocene. E. López-Vera (Ed.). Quaternary<br />

climate in Western Mediterranean, University Autónoma de Madrid, 405-420 pp.<br />

Pons, A. and Reille, M., 1988: The Holocene-and Upper Pleistocene pollen record from<br />

Padul (Granada, Spain): a new study. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 66: 243-263.<br />

149<br />

F. Naughton, 2007


Reille, M., 1992. Pollen et spores d'Europe et d'Afrique du Nord. Laboratoire de<br />

botanique historique et palynologie, Marseille, 520 pp.<br />

Rios, A.F., Pérez, F.F. and Fraga, F., 1992. Water masses in the upper and middle North<br />

Atlantic east of the Açores. Deep sea Research, 39: 645-658.<br />

Roucoux, K.H., de Abreu, L., Shackleton, N.J. and Tzedakis, P.C., 2005. The response of<br />

NW Iberian vegetation to North Atlantic climate oscillations during the last 65 kyr.<br />

Quaternary Science Reviews, 24: 1637-1653.<br />

Roucoux, K.H., Shackleton, N.J., de Abreu, L., Schönfeld, J. and Tzedakis, P.C., 2001.<br />

Combined marine proxy and pollen analyses reveal rapid Iberian vegetation<br />

response to North Atlantic millennial-scale climate oscillations. Quaternary Science<br />

Reviews, 56: 128-132.<br />

Ruiz Zapata, M.B., Gil Garcia, M.J., Dorado Valiño, M., Valdeolmillos Rodrigues, A.,<br />

Vegas, J. and Pérez-Gonzalez, A., 2002. Clima y vegetation durante el tardiglaciar y<br />

el Holoceno en la Sierra de Neilla (Sistema Ibérico Noroccidental). Cuaternario y<br />

Geomorfologia, 16: 9-20.<br />

Sánchez Goñi, M.F. and Hannon, G., 1999. High altitude vegetational patterns on the<br />

Iberian Mountain chain (north-central Spain) during the Holocene. The Holocene, 9:<br />

39-57.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L. and Shackleton, N.J., 1999. High resolution<br />

palynological record off the Iberian margin: direct land-sea correlation for the Last<br />

Interglacial complex. Earth and Planetary Science Letters, 171: 123-137.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L. and Gendreau, S., 2000. European climatic<br />

response to millenial-scale climatic changes in the atmosphere-ocean system during<br />

the Last Glacial period. Quaternary Research, 54: 394-403.<br />

Sánchez Goñi, M.F., Cacho, I., Turon, J.-L., Guiot, J., Sierro, F.J., Peypouquet, J.-P.,<br />

Grimalt, J.O. and Shackleton, N.J., 2002. Synchroneity between marine and terrestrial<br />

responses to millennial scale climatic variability during the last glacial period in the<br />

Mediterranean region. Climate Dynamics, 19: 95-105.<br />

Sánchez Goñi, M.F., Loutre, M.F., Crucifix, M., Peyron, O., Santos, L., Duprat, J.,<br />

Malaizé, B., Turon, J.-L., Peypouquet, J.-P., 2005. Increasing vegetation and climate<br />

gradient in Western Europe over the Last Glacial Inception (122–110 ka): data-model<br />

comparison. Earth and Planetary Science Letters, 231: 111-130.<br />

Santos, L., Vidal Romani, J.R. and Jalut, G., 2000. History of vegetation during the<br />

Holocene in the Courel and Queixa Sierras, Galicia, northwest Iberian Peninsula.<br />

Journal of Quaternary Science, 15: 621-632.<br />

Schönfeld, J. and Zahn, R., 2000. Late glacial to Holocene history of the<br />

Mediterranean outflow: Evidence from benthic foraminiferal assemblages and stable<br />

isotopes at the Portuguese margin. Palaeogeography, Palaeoclimatology,<br />

Palaeoecolecology, 159: 85-111.<br />

Schönfeld, J., Zahn, R. and de Abreu, L., 2003. Surface and deep water response to<br />

rapid climate changes at the Western Iberian Margin. Global and Planetary Change,<br />

36: 237-264.<br />

150<br />

F. Naughton, 2007


Shackleton, N.J., Hall, M.A. and Vincent, E., 2000. Phase relationships between<br />

millennial scale events 64,000-24,000 years ago. Paleoceanography, 15: 565-569.<br />

Sprangers, M., Dammers, N., Brinkhuis, H., van Weering, T.C.E. and Lotter, A.F., 2004.<br />

Modern organic-walled dinoflagellate cyst distribution offshore NW Iberia; tracing the<br />

Upwelling System. Review of palaeobotany and Palynolology, 128: 97-106.<br />

Stanley, E.A., 1966. The application of palynology to oceanology with reference to<br />

the northwestern Atlantic. Deep-Sea Research, 13: 921-939.<br />

Stuiver, M. and Reimer, P., 1993. Extended 14C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

Stuiver, M., Reimer, P.J. and Reimer, R.W., 2005. CALIB 5.0. (WWW program and<br />

documentation).<br />

Telford, R.J., Heegaard, E. and Birks, H.J.B., 2004. The intercept is a poor estimate of a<br />

calibrated radiocarbon age. The Holocene, 14: 296-298.<br />

Thouveny, N., Moreno, E., Delanghe, D., Candon, L., Lancelot, Y. and Shackleton N.J.,<br />

2000. Rock magnetic detection of distal ice-rafted debries: clue for the identification<br />

of Heinrich layers on the Portuguese margin. Earth and Planetary Science Letters, 180:<br />

61-75.<br />

Turon, J.-L., 1984. Le palynoplancton dans l'environnement actuel de l'Atlantique<br />

nord-oriental. Evolution climatique et hydrologique depuis le dernier maximum<br />

glaciaire. Mémoires de l'Institut de Géologie du bassin d'Aquitaine, 17. Université de<br />

Bordeaux I, Bordeaux, 313 pp.<br />

Turon, J.-L., Lézine, A.-M. and Denèfle, M., 2003. Land-sea correlations for the last<br />

glaciation inferred from a pollen and dinocyst record from the Portuguese margin.<br />

Quaternary Research, 59: 88-96.<br />

Tzedakis, P.C., Roucoux, K.H., de Abreu, L. and Shackleton, N.J., 2004. The duration of<br />

forest stages in southern Europe and interglacial climate variability. Science, 306:<br />

2231-2235.<br />

Valdès, C.M., Gil Sanchez, L., 2001. La transformación histórica del paisaje forestal en<br />

Galicia, Ministerio de Medio Ambiente, 159 pp.<br />

Vale, C., 1990. Temporal variations of particulate metals in the Tagus river estuary.<br />

Science of the Total Environment, 97-98: 137-154.<br />

Van der Kaars, S. and de Deckker, P., 2003. Pollen distribution in marine surface<br />

sediments offshore Western Australia. Review of Palaeobotany and Palynology, 124:<br />

113-129.<br />

Van der Knaap, W.O. and Van Leeuwen, J.F.N., 1995. Holocene vegetation<br />

succession and degradation as responses to climatic change and human activity in<br />

the Serra de Estrela, Portugal. Review of Palaeobotany and Palynology, 89: 153-211.<br />

Van Weering, T.C.E., McCave, I.N., 2002. Dedication to Roland Wollast. Progress In<br />

Oceanography, 52: 121-122.<br />

151<br />

F. Naughton, 2007


Van Weering, T.C.E., de Stigter, H.C., Boer, W. and Haas, H., 2002. Recent sediment<br />

transport and accumulation on the NW Iberian margin. Progress In Oceanography,<br />

52: 349-371.<br />

Vanney, J.R. and Mougenot, D., 1981. La plate-forme continentale du Portugal et les<br />

provinces adjacentes: Analyse Géomorphologique, Memórias dos serviços<br />

geológicos de Portugal, 145 pp.<br />

Vitorino, J., Oliveira, A., Jouanneau, J.M. and Drago, T., 2002. Winter dynamics on the<br />

northern Portuguese shelf. Part1: physical processes. Progress in Oceanography, 52:<br />

129-153.<br />

Von Engelbrechten, S., 1998. Late-glacial and Holocene vegetation and<br />

environmental history of the sierra de Urbión, North-Central Spain. PhD Thesis, Dublin<br />

University, Trinity College, Ireland, 212 pp.<br />

152<br />

F. Naughton, 2007


Capítulo 3| New insights on the impact of Heinrich events<br />

and LGM in the mid-latitudes of the eastern North Atlantic<br />

and in the adjacent continent<br />

Novos conhecimentos sobre o impacto dos eventos de<br />

Heinrich e do último máximo glaciar nas latitudes médias do<br />

Atlântico Nordeste e no continente adjacente<br />

Nouvelles approches sur l’impact des évènements d’Heinrich<br />

et du Dernier Maximum Glaciaire dans les moyennes<br />

latitudes dans l’est de l’Atlantique Nord et sur le continent<br />

adjacent<br />

A reduced version of this chapter will be submitted in December 2006 to:<br />

Earth and Planetary Science Letters<br />

F. Naughton a, b , M.F. Sánchez Goñi a , J. Duprat a , E. Cortijo c , B. Malaizé a , C. Joly a , S.<br />

E. Bard d , F. Rostek d and J-L. Turon a<br />

a Environnements et Paléoenvironnements Océaniques (UMR CNRS 5805 EPOC), EPHE,<br />

Université Bordeaux 1, Av. des Facultés, 33405 Talence, France<br />

b Departamento e Centro de Geologia -Universidade de Lisboa, Bloco C6, 3º piso,<br />

Campo Grande, 1749-016 Lisboa, Portugal<br />

c Laboratoire des Sciences du Climat et de l’Environnement (LSCE-Vallée), Bât. 12,<br />

avenue de la Terrasse, F-91198 Gif-sur-Yvette cedex, France<br />

d CEREGE, UMR-CNRS 6635, Aix-en-Provence, France<br />

153<br />

F. Naughton, 2007


Resumo<br />

O estudo multidisciplinar de alta resolução da sondagem marinha,<br />

MD99-2331, recolhida no noroeste da margem Ibérica, mostra que a resposta<br />

da vegetação temperada e do oceano das médias latitudes do Atlântico<br />

Norte ao último máximo glaciar (LGM) é contraditório ao que foi observado<br />

durante os episódios interestadiais do MIS3 tardio (MIS3-Marine isotopic Stage<br />

3). O aumento do albedo, o forte contraste sazonal das altas latitudes do<br />

Atlântico Norte e a diminuição da concentarção de CO2 na atmosfera,<br />

durante o LGM, podem ter sido os principais responsáveis pelo fraco<br />

desenvolvimento da floresta temperada nesta região. Contudo, o aumento<br />

da intensidade da circulação termohalina do Atlântico Norte (MOC)<br />

favoreceu a transferência de humidade para o noroeste da Península<br />

Ibérica provocando o desenvolvimento de Ericaceae e Calluna nesta<br />

região.<br />

Este estudo mostra ainda duas fases principais na vegetação Ibérica<br />

as quais parecem estar intimamente ligadas ao complexo sinal deixado<br />

pelos típicos eventos de Heinrich (H4, H2 and H1) ao longo desta margem. A<br />

primeira fase, anterior à chegada máxima de icebergues a esta margem, é<br />

marcada por condições oceânicas de superfície muito frias (evidenciada<br />

pelas associações de foraminíferos, δ 18 O e estimativa da temperatura<br />

baseada nas alcanonas) assim como por um arrefecimento extremo no<br />

continente revelado pelo forte declínio da floresta de Pinus. O aumento de<br />

Calluna e elevadas concentrações polínicas indicam o aumento das<br />

condições húmidas durante esta primeira fase. A segunda fase, associada à<br />

chegada máxima dos icebergs nesta margem, é caracterizada por<br />

condições oceânicas de superfície e continentais menos frias e pelo<br />

aumento da aridez a qual é representada pelo desenvolvimento de plantas<br />

semi-desérticas.<br />

O drástico arrefecimento detectado durante os eventos H4, H3, H2 e<br />

H1 foi provavelmente provocado pela interrupção da MOC seguida de<br />

rápidas reorganizações entre o oceano e a atmosfera as quais favoreceram<br />

a transmissão das condições frias para o noroeste Ibérico. Para além deste<br />

mecanismo oceanográfico, variações semelhantes ao actual modo<br />

negativo e positivo do índice da Oscilação Norte Atlântica (NAO) parecem<br />

154<br />

F. Naughton, 2007


ter tido um papel crucial no padrão climático deixado pelos eventos de<br />

Heinrich no noroeste Ibérico. Durante a primeira fase, a predominância do<br />

modo negativo da NAO-like gerou um aumento da precipitação no inverno<br />

e importantes descargas fluviais nesta região. Estas condições favoreceram<br />

ainda a fusão dos icebergues na cintura de IRD (IRD belt), onde a<br />

temperatura da água superficial seria relativamente quente, impedindo o<br />

seu transporte para as latitudes médias do Atlântico Norte. Durante a<br />

segunda fase, a predominância do modo positivo da NAO-like provocou a<br />

intensificação e a migração para norte dos ventos de oeste provocando um<br />

aumento da aridez ao longo da Península Ibérica. Estas condições<br />

favoreceram ainda a migração dos icebergues para sul e a sua fusão nas<br />

latitudes médias do Atlântico Norte.<br />

O evento atípico H3 reflecte condições húmidas durante quase todo<br />

o episódio sugerindo uma não intensificação ventos de oeste nesta região.<br />

Résumé<br />

L’étude multiproxy et à haute résolution de la carotte marine profonde<br />

MD99-2331 prélevée au nord-ouest de la marge Ibérique montre un<br />

découplage entre la réponse de la forêt tempérée et les températures<br />

relativement élevées des eaux océaniques de surface (SST) pendant le<br />

Dernier Maximum Glaciaire (LGM) contrairement à ce qui est détecté<br />

pendant les plus récents interstadiaires de Dansgaard-Oeschger du Stade<br />

Isotopique Marin 3 (MIS3). L’augmentation de l’albédo, le fort contraste<br />

saisonnier ainsi que la diminution de la concentration de CO2 pourraient<br />

expliquer le faible développement des arbres tempérés dans cette région.<br />

Cependant, l’augmentation de l’intensité de la circulation méridienne<br />

Atlantique de renversement (MOC), transportant l’humidité vers la marge<br />

nord-ouest Ibérique, a sûrement joué un rôle dans l’augmentation de SST et le<br />

développement observé des bruyères.<br />

Cette étude met également en évidence deux phases principales de<br />

développement de la végétation dans le nord-ouest de la Péninsule Ibérique<br />

intimement liées à la complexité de l’impact des évènements typiques<br />

d’Heinrich (H4, H2 et H1) sur la marge Ibérique. La première phase, précédant<br />

155<br />

F. Naughton, 2007


l’arrivée maximale d’icebergs sur cette marge, est marquée par des<br />

températures des eaux de surface particulièrement froides, détectées par les<br />

assemblages de foraminifères planctoniques, les analyses de δ 18 O et les<br />

mesures de Uk 37 , associées au refroidissement important sur le continent<br />

adjacent détecté par la réduction de la forêt de Pin. L’expansion de Calluna,<br />

conjointement à l’augmentation de concentration pollinique, indique une<br />

augmentation de l’humidité pendant cette phase. La deuxième phase,<br />

associée à l’arrivée massive d’icebergs dans la marge Ibérique, est<br />

caractérisée par des eaux de surface et des conditions atmosphériques<br />

moins froides et par une augmentation de l’aridité identifiée par le<br />

développement des plantes semi-désertiques.<br />

Les conditions froides au cours des événements d’Heinrich 4, 3, 2 et 1<br />

ont été probablement engendrées par l’arrêt de la MOC, suivi de<br />

réorganisations rapides entre l’océan et l’atmosphère favorisant le transfert<br />

instantané de conditions froides dans le nord-ouest de la Péninsule Ibérique.<br />

Derrière ce mécanisme océanographique, des changements d’index<br />

(négatif ou positif) de l’Oscillation Nord Atlantique (NAO-like) semblent avoir<br />

joué un rôle crucial dans le scénario climatique complexe laissé par les<br />

évènements d’Heinrich dans cette région. En effet, pendant la première<br />

phase, un mode dominant d’index négatif de la NAO-like aurait généré,<br />

comme c’est le cas à présent, une augmentation des précipitations<br />

hivernales et donc des décharges des rivières dans la Péninsule Ibérique. Ces<br />

conditions favorisent la fonte des icebergs au niveau de la ceinture de dépôt<br />

préférentielle d’IRD où les eaux de surface sont relativement chaudes,<br />

empêchant leur migration vers le sud jusqu’aux moyennes latitudes. Pendant<br />

la seconde phase, une plus forte fréquence des situations de NAO-like en<br />

mode positif aurait conduit à une intensification des vents d’ouest et leur<br />

déplacement vers le nord, produisant une augmentation de la sécheresse<br />

dans la Péninsule Ibérique. Ces conditions auraient favorisé la migration vers<br />

le sud des icebergs vers les moyennes latitudes.<br />

Par ailleurs, l’évènement atypique H3 est caractérisé dans cette région<br />

par des conditions humides pendant presque toute sa totalité, certainement<br />

liées au maintien sur cette région de vent d’ouest affaiblis.<br />

156<br />

F. Naughton, 2007


Abstract<br />

High resolution multi-proxy study of MD99-2331 deep sea core retrieved<br />

in north-western Iberian margin shows that the response of temperate forest<br />

and mid-latitude sea surface temperatures to the Last Glacial Maximum<br />

(LGM) period was decoupled contrarily to what has been detected during<br />

the late MIS 3 interstadials. Albedo increase, high seasonality and CO2<br />

concentration decrease could explain the weak development of temperate<br />

trees in this region. However, the more vigorous Meridional overturning<br />

circulation (MOC) transferring moisture to north-western Iberia could be<br />

responsible for the observed heathland development.<br />

Furthermore, this study evidences two main vegetation phases in north-<br />

western Iberia linked to the complex imprint left by the typical Heinrich events<br />

(H4, H2 and H1) in the Iberian margin. The first phase, before the maximal<br />

arrival of icebergs into this margin, is marked by extremely cold sea surface<br />

temperatures indicated by the planktonic foraminifera assemblages and δ 18 O<br />

analyses together with Uk 37 measurements and by the strong cooling of the<br />

adjacent continent revealed by the Pinus forest decline. Calluna expansion in<br />

concert with the highest total pollen concentration indicates moisture<br />

increase during this phase. The second phase, associated with the maximum<br />

arrival of icebergs into the Iberian margin, is characterised by less cold sea<br />

surface and atmospheric conditions and by an increase of dryness identified<br />

by the development of semi-desert plants.<br />

The cold conditions during Heinrich events 4, 3, 2 and 1 were probably<br />

triggered by the Atlantic MOC shutdown followed by ocean-atmosphere<br />

rapid reorganizations favouring the transfer of cold conditions into north-<br />

western Iberia. Besides this oceanographic mechanism, changes similar to<br />

that of prevailing (negative and positive) North Atlantic Oscillation (NAO)<br />

index seems to have played a crucial role on the complex climatic pattern<br />

left by Heinrich events in north-western Iberia. Indeed during the first phase,<br />

prevailing negative mode of NAO-like index generates winter precipitations<br />

and river flow increase in Iberia. These prevailing conditions, favoured iceberg<br />

melting in the IRD belt (where sea surface temperature was relatively warm)<br />

preventing their southern migration to the mid-latitudes. During the second<br />

157<br />

F. Naughton, 2007


phase, prevailing positive mode of NAO-like index leads to westerlies<br />

intensification and northward displacement triggering an increase of dryness<br />

in Iberia. These prevailing conditions favoured the southward migration of the<br />

icebergs to the mid-latitudes sites.<br />

The atypical H3 is characterised in this region by wet conditions over<br />

almost the entire event probably due to maintaining reduced westerlies in this<br />

region.<br />

158<br />

F. Naughton, 2007


3. 1 Introduction<br />

In the last years, many studies have been carried out to understand<br />

the sources, trigger mechanisms and the global impact of the well known<br />

massive episodes of iceberg discharges into the North Atlantic that occurred<br />

during the last glacial period (see Hemming, 2004). These extreme episodes,<br />

named Heinrich events, have been firstly documented in the Ruddiman belt<br />

from several North Atlantic deep-sea cores between 45 and 50° N (Heinrich,<br />

1988, Bond and Lotti, 1995). They were identified by the anomalous presence<br />

of ice-rafted detritus (IRD) that were transported to the ocean by drifting<br />

icebergs from Laurentide and northern European ice sheets (Heinrich, 1988)<br />

as well as by peaks of N. pachyderma (s) (e.g. Bond and Lotti, 1995;<br />

Hemming, 2004) and magnetic susceptibility (Grousset et al., 1993). These<br />

coarse fraction intervals, representing the well known IRD layers, were also<br />

detected out of the Ruddiman belt i.e. north of 50°N (e.g. Fronval et al., 1995;<br />

Rasmussen et al., 1996; Elliot et al., 1998; Voelker et al., 1998; Van Kreveld et<br />

al., 2000) as well as below 40°N (e.g. Lebreiro et al., 1996; Baas et al., 1997;<br />

Zahn et al., 1997; Chapman et al., 2000; Bard et al., 2000; de Abreu et al.,<br />

2003). The thickness of the IRD layers and the magnetic signal is, however,<br />

smaller in the mid-latitude sites than in the northern ones (Thouveny et al.,<br />

2000; Dowdeswell et al., 1995). Also, the duration of the impact of these<br />

extreme events on the sea surface temperatures (SST) is in this region longer<br />

than that of the IRD layers (e.g. Bard et al., 2000; Chapman et al., 2000;<br />

Sánchez Goñi et al., 2000). Indeed, the Heinrich events have left a complex<br />

pattern imprint along the Iberian margin (e.g Abrantes et al., 1998; Bard et al.,<br />

2000; Thouveny et al., 2000; Schönfeld et al., 2003; Narciso et al., 2006).<br />

Western Iberian vegetation further shows a complex pattern<br />

associated with Heinrich events (H) 2 and 1 (Naughton et al., 2006) and with<br />

H3, H4 and H5 (Sánchez Goñi et al., 2000). The first hypothesis based on land-<br />

sea direct correlation has linked the relatively wet conditions in south western<br />

Iberia at the beginning of H5 to H3 events to “European iceberg discharges”,<br />

and the semi-desert plant development to the successive massive iceberg<br />

discharges from the Laurentide ice sheet (Sánchez Goñi et al., 2000). This<br />

extreme dryness has been attributed to a prevailing positive North Atlantic<br />

159<br />

F. Naughton, 2007


Oscillation (NAO) index (Sánchez Goñi et al., 2002). However, the<br />

mechanisms proposed for explaining this complex pattern within Heinrich<br />

events in the Iberian margin are not conclusive so far.<br />

Most of the direct correlations between terrestrial and marine proxies<br />

from Iberian margin deep-sea cores does not deeply discuss about the Last<br />

Glacial maximum (LGM) period. However, the LGM is considered as a key<br />

interval for understanding the sensitivity of global environmental systems to<br />

change (Mix et al., 2001), because it represents the extreme opposite<br />

situation to an interglacial and a period of relatively stable glacial maximum<br />

conditions. Recently, several sea surface temperature (SST) reconstructions for<br />

the LGM period have been carried out around the world by MARGO project<br />

(Multiproxy Approach for the Reconstruction of the Glacial Ocean surface)<br />

suggesting the presence of seasonal ice cover in the North Atlantic and in the<br />

Nordic Seas (de Vernal et al., 2005). This sea-ice free season would allow the<br />

Meridional Overturning Circulation to supply moisture to the northern<br />

hemisphere high latitudes (Meland et al., 2005). Previous SST reconstructions<br />

carried out in the Iberian margin have clearly demonstrated that during the<br />

LGM sea surface conditions were warmer than during Heinrich events (e.g.<br />

Bard et al., 1987; Lebreiro et al., 1997; Cayre et al., 1999; Bard et al., 2000;<br />

Pailler and Bard, 2002; de Abreu et al., 2003). Because LGM SST values are<br />

similar or even higher than those characterising Late Marine Isotopic Stage 3<br />

(MIS3) Dansgaard-Oeschger (D-O) interstadials, we should expect similar<br />

amplitude of temperate trees expansion in north-western Iberia what it is not<br />

the case. Indeed, north-western Iberian margin MD95-2039 record (Roucoux<br />

et al., 2005) does not reveal the same amplitude of temperate tree expansion<br />

during both the MIS 3 D-O interstadials and the LGM. Therefore, other<br />

mechanisms than the weakening of the MOC has to be seek to explain the<br />

weak development of temperate trees in this region during the LGM.<br />

The aim of this work is firstly to propose the possible trigger mechanisms<br />

for the complex pattern of vegetation observed in north western Iberia during<br />

Heinrich events. Secondly, we will discuss the paradox observed during the<br />

LGM when relatively high SST in the mid-latitudes of the eastern North Atlantic<br />

were associated with cool environments on land.<br />

160<br />

F. Naughton, 2007


3. 2 Environmental Setting<br />

MD99-2331 deep sea core was recovered on the Galician margin (42°<br />

09’ 00 N, 09° 40’ 90 W) at ~100 km from the coast and at 2110 m water depth<br />

(Fig. III.1). Morphology, recent sedimentation and hydrology of the Iberian<br />

margin have been thoroughly described in Naughton et al. (2006).<br />

The northernmost part of north-western Iberia is influenced by the wet,<br />

cool and weakly seasonal Atlantic climate (annual precipitation = 900-1400<br />

mm and annual temperature ranges = -7 and 10° C). This region is dominated<br />

by deciduous Quercus forest (Q. robur, Q. pyrenaica and Q. petraea), heaths<br />

communities (Ericaceae and Calluna) and Ulex (Alcara Ariza et al., 1987;<br />

Polunin and Walters, 1985).<br />

Fig. III.1 | Map showing MD99-2331 location and sites of the cores referred in the text: 1: MD95-2040 (Pailler<br />

and Bard, 2002; de Abreu et al., 2003; Schönfeld et al., 2003; Narciso et al., 2006), 2: MD95-2039 (Thouveny<br />

et al., 2000; Roucoux et al., 2001; 2005; Schönfeld et al., 2003), 3: PO 28-1 (Abrantes et al., 1998), 4: D11957P<br />

(Lebreiro et al., 1996; 1997), 5: SO75-26KL (Zahn et al., 1997; Boessenkool et al., 2001), 6: PO 8-2 (Abrantes et<br />

al., 1998), 7: MD95-2042 (Cayre et al., 1999; Sánchez Goñi et al., 2000; 2002; Thouveny et al., 2000; Pailler<br />

and Bard, 2002), 8: SU81-18 (Bard et al., 2000; Turon et al., 2003); 9: ODP 976 (Combourieu et al., 2002), 10:<br />

SU90-03 (Chapman et al., 2000), 11: ESSCAMP-KS02 (Loncaric et al., 1998; Zaragosi et al., 2001), 12: MD95-<br />

2002 (Grousset et al., 2000; Zaragosi et al., 2001), 13: AKS01 (Zaragosi et al., 2001), 14: VM 23-81 (Bond and<br />

Lotti, 1995), 15: MD04-2845 (work in progress), 16: SU90-11 (Jullien et al., in press.), 17: MD03-2705 (Jullien et<br />

al., submitted), 18: OCE326-GGC5 (McManus et al., 2004).<br />

161<br />

F. Naughton, 2007


3. 3 Material and methods<br />

Core MD99-2331 was retrieved using a giant CALYPSO corer during the<br />

Ginna (IMAGES V) oceanographic cruise on board the R/V Marion Dufresne<br />

(Fig. III.1). This sedimentary record, mainly composed of hemi-pelagic clay, is<br />

37.2 m long and covers the last Marine Isotopic Stages (MIS) 7 to 1. In this<br />

study, we will focus on the last 40 000 years where sedimentary rates vary from<br />

43 cm kyr -1 to 30 cm kyr -1 , providing a high-resolution palaeoclimatic record<br />

for this time period off north-western Iberia and in the adjacent continent.<br />

X-ray analysis on this core using SCOPIX image-processing (Migeon et<br />

al., 1999) shows a rather well preserved sedimentary sequence between 2<br />

and 11 m core depth.<br />

3. 3. 1 Chronostratigraphy<br />

The age model of MD99-2331 deep sea core is based on 55<br />

accelerator mass spectrometer (AMS) 14 C dates obtained at “Laboratoire de<br />

Mesure du Carbone 14” (LMC) in Saclay and at AMS laboratory (GifA) in Gif-<br />

sur-Yvette on monospecific samples with maxima of Globigerina bulloides or<br />

Neogloboquadrina pachyderma (s.) abundances (Fig. III.2 and Tab. III.1).<br />

AMS 14 C younger than 21 786 BP were calibrated using CALIB Rev 5.0<br />

program and the "global" marine calibration dataset (marine 04.14c) (Stuiver<br />

and Reimer, 1993; Hughen et al., 2004; Stuiver et al., 2005). We used the 95.4%<br />

(2 sigma) confidence intervals and their relative areas under the probability<br />

curve as well as the median probability of the probability distribution (Telford<br />

et al., 2004) as suggested by Stuiver et al. (2005).<br />

14 C radiometric ages older than 21 786 yr BP and younger than 40 000<br />

yr BP were firstly corrected for the regional marine age reservoir of -400 yr and<br />

then, calibrated using a simple second-order polynomial (Age cal yr BP = -<br />

6.2724 x 10 -6 x [Age 14 C yr BP] 2 + 1.3818 x [Age 14 C yr BP] – 1388) which has<br />

been constructed by means of Iberian margin data tuned with GISP2 (Bard et<br />

al., 2004).<br />

All conventional radiocarbon dates as well as their calibration results<br />

are plotted against core depth in Fig. III.2. We use both the <strong>paleoclimat</strong>e<br />

162<br />

F. Naughton, 2007


multi-proxy record of MD99-2331 as well as the age limits of Heinrich events<br />

proposed by Elliot et al., 2002 to delimit H4, H3, H2 and H1 events.<br />

Some levels associated with Heinrich events reflect age inversions<br />

(older and younger) when compared with adjacent host sediments (Fig. III.2<br />

and Tab. III.1). They were rejected because they likely reveal reworking<br />

processes due to substantial changes in sedimentary rate. Others,<br />

representing bioturbated levels (presence of zoophycos burrows) with<br />

younger material as well as those which are too old compared with Elliot et<br />

al. (2002) age limits were also rejected. Finally, we have also rejected some<br />

anomalous dates associated with isolated peaks of IRD that occurred after H4<br />

and H3.<br />

Therefore we only use 29 AMS 14 C levels for establishing the chronology<br />

of MD99-2331 during MIS2 and late MIS3 (Fig. III.2 and Tab. III.1).<br />

Fig. III.2 | Chronostratigraphy of the MD99-2331 record. AMS radiocarbon dates are represented by<br />

triangles while the calibrated ones are represented by squares. North Atlantic Heinrich events are delimited<br />

by both the age limits (not calibrated) proposed by Elliot et al. (2002) and by the results obtained from the<br />

multi-proxy study of the MD99-2331 record (see below). White triangles and squares reflect the rejected<br />

levels for the model age while the dark ones represent the accepted ages.<br />

163<br />

F. Naughton, 2007


Lab code Core<br />

depth<br />

(cm)<br />

Material Conv.<br />

AMS<br />

14C<br />

age BP<br />

164<br />

Conv.<br />

AMS 14C<br />

age BP<br />

(-400 yr)<br />

error 95.4 % (2σ)<br />

Cal BP<br />

age ranges<br />

Cal BP<br />

age<br />

Median<br />

probability<br />

F. Naughton, 2007<br />

Bard<br />

et al.,<br />

2004<br />

LMC14-001231 200 N. pachyderma 13640 13240 80 15303:16099 15679<br />

LMC14-001232 205 N. pachyderma 13810 13410 80 15524:16359 15922<br />

GIF-102377 220 N. pachyderma 14130 13730 120 15898:16828 16342<br />

LMC14-001233 222 N. pachyderma 13920 13520 90 15658:16520 16067<br />

LMC14-001235 228 N. pachyderma 13930 13530 80 15686:16521 16081<br />

LMC14-001236 235 N. pachyderma 15130 14730 90 17250:18182 17848<br />

LMC14-001237 242 N. pachyderma 15060 14660 90 17170:18038 17722<br />

LMC14-002445 260 G. bulloides 15540 15140 90 18405:18723 18520<br />

GIF-101109 290 G. bulloides 16170 15770 130 18787:19265 18983<br />

GIF-102373 570 G. bulloides 19770 19370 170 22534:23622 23038<br />

LMC14-002446 590 G. bulloides 22290 21890 170 25854<br />

LMC14-001845 595 G. bulloides 20860 20460 250 23931:25369 24542<br />

LMC14-001846 600 G. bulloides 20550 20150 240 23450:24803 24119<br />

LMC14-001847 607 G. bulloides 20460 20060 140 23652:24405 24016<br />

LMC14-001849 620 G. bulloides 21620 21220 160 25301:26000 25626<br />

LMC14-001850 623 G. bulloides 21740 21340 160 25439:26000 25730<br />

GIF-102378 630 N. pachyderma 22690 22290 180 26296<br />

LMC14-001851 637 G. bulloides 22150 21750 170 25699<br />

LMC14-001852 650 G. bulloides 22440 22040 170 26020<br />

LMC14-001853 655 G. bulloides 22430 22030 180 26009<br />

LMC14-001854 670 G. bulloides 23080 22680 190 26725<br />

LMC14-001855 700 G. bulloides 24140 23740 210 27881<br />

GifA-102374 740 G. bulloides 25880 25480 250 29748<br />

LMC14-001856 740 G. bulloides 25140 24740 230 28959<br />

LMC14-002447 760 N. pachyderma 26310 25910 260 30204<br />

LMC14-001857 770 G. bulloides 23960 23560 200 27686<br />

LMC14-001858 780 G. bulloides 24390 23990 210 28151<br />

LMC14-002448 800 N. pachyderma 25870 25470 250 29737<br />

LMC14-001859 810 N. pachyderma 26050 25650 280 29928<br />

LMC14-001860 820 N. pachyderma 26370 25970 260 30267<br />

LMC14-001861 830 N. pachyderma 27010 26610 280 30940<br />

GifA-102375 840 G. bulloides 27790 27390 320 31754<br />

LMC14-001862 850 G. bulloides 28530 28130 350 32519<br />

LMC14-001863 860 G. bulloides 28510 28110 350 32498<br />

LMC14-001864 870 G. bulloides 28860 28460 370 32858<br />

LMC14-001865 880 N. pachyderma 29030 28630 380 33032<br />

LMC14-001866 890 N. pachyderma 29470 29070 390 33480<br />

LMC14-001867 895 G. bulloides 29540 29140 390 33552<br />

LMC14-001868 920 G. bulloides 30260 29860 430 34280<br />

LMC14-001869 925 G. bulloides 30020 29620 420 34038<br />

LMC14-001870 945 N. pachyderma 31960 31560 560 35974<br />

LMC14-001871 960 G. bulloides 32720 32320 570 36720<br />

LMC14-001872 970 G. bulloides 32530 32130 570 36534<br />

GifA-102376 980 G. bulloides 33410 33010 500 37390<br />

LMC14-001873 985 N. pachyderma 34440 34040 700 38381<br />

LMC14-002449 990 N. pachyderma 34810 34410 700 38733<br />

LMC14-001874 995 N. pachyderma 34490 34090 700 38428<br />

LMC14-001875 1000 N. pachyderma 35350 34950 790 39244<br />

LMC14-001876 1005 N. pachyderma 34600 34200 710 38533<br />

LMC14-001877 1010 N. pachyderma 31490 31090 510 35509<br />

LMC14-001878 1015 N. pachyderma 32950 32550 620 36944<br />

GifA-102379 1020 N. pachyderma 33850 33450 440 37815<br />

LMC14-001879 1025 N. pachyderma 34920 34520 760 38837<br />

LMC14-001880 1030 N. pachyderma 36940 36540 920 40728<br />

LMC14-002450 1040 G. bulloides 37100 36700 940 40876<br />

Tab. III.1| Radiocarbon ages of MD99-2331 deep sea cores. Bold levels represent the accepted ages while<br />

not bold ones represent the rejected ages for the age model.


3. 3. 2 Pollen analysis<br />

MD99-2331 core was sub-sampled for pollen analysis between 2.00 and<br />

11.00 m. The sample spacing was between 2 and 5 cm for MIS2 and 5 to 30<br />

cm for the late MIS3, allowing a time resolution average of 110 years and 210<br />

years, respectively.<br />

The Galician margin is influenced by north-western prevailing winds<br />

which impede substantial direct airborne transport of pollen grains from the<br />

continent to the sea. Because, north-western Iberia is composed of a<br />

complex fluvial system, pollen grains are mainly released by the Douro river<br />

followed by the Minho river and transferred seaward following a estuary-shelf-<br />

slope transect (Naughton et al., 2006). Furthermore, the comparison between<br />

modern terrestrial and marine pollen signals in and off western Iberia<br />

demonstrate that marine pollen spectra reflect an integrated image of the<br />

regional vegetation of the adjacent continent (Naughton et al., 2006).<br />

Sample preparation technique follows de Vernal et al. (1996) modified<br />

at the UMR CNRS 5805 EPOC (Desprat, 2005). An exotic, Lycopodium, spike of<br />

known concentration has been added to each sample to calculate total<br />

pollen concentrations (including Pinus, aquatic plants, spores, indeterminate<br />

and unknown pollen grains).<br />

After chemical treatments (cold 10%, 25% and 50% HCl as well as cold<br />

40% and 70% HF) the samples were sieved through a 10 µm nylon mesh<br />

screen (Heusser and Stock, 1984) and mounted in bidistillate glycerine. Pollen<br />

and spores were counted using a Zeiss Axioscope light microscope at x550<br />

and x1250 (oil immersion) magnifications. A minimum of 100 pollen grains<br />

excluding the over-represented Pinus grains in the Iberian margin (Naughton<br />

et al., 2006) have been counted. At least 20 taxa and 100 Lycopodium were<br />

also counted in each of the 165 samples analysed. Pollen percentages were<br />

calculated based on the main pollen sum which excludes Pinus, aquatic<br />

plants, spores, indeterminate and unknown pollen grains.<br />

Detailed description of pollen data from MIS2 has been previously<br />

published in Naughton et al. (2006).<br />

165<br />

F. Naughton, 2007


3. 3. 3 Marine proxy analysis<br />

3. 3. 3. 1 Isotopic analyses<br />

A total of 101 oxygen isotopic measurements were carried out on<br />

Globigerina bulloides planktonic foraminifera (Gouzy et al., 2004; Naughton et<br />

al., 2006) at Bordeaux University with a sample spacing of 2 to 20 cm (time<br />

resolution average of 240 years). Sixty-one measurements were made on<br />

Cibicides wuellerstorfi benthic foraminifera every 2 to 90 cm (time resolution<br />

average of 400 years) at the Laboratoire des Sciences du Climat et de<br />

l’Environnement (LSCE), Gif-sur-Yvette, France.<br />

Each specimen has been picked up within the 250–315 µm fraction<br />

and cleaned with distilled water. The preparation of each aliquot (4–10<br />

specimens with 80 μg of weight) has been carried out in the Micromass<br />

Multiprep autosampler by using an individual acid attack. The CO2 gas<br />

extracted has been analysed against NBS 19 standard, taken as an<br />

international reference standard. Planktonic Isotopic analysis has been<br />

carried out using an Optima Micromass mass spectrometer at the UMR CNRS<br />

5805 EPOC and benthic isotopic analysis were performed using a delta plus<br />

Finnigan at the LSCE. The mean external reproducibility of powdered<br />

carbonate standards is ±0.05‰ for oxygen. Results from oxygen isotopic<br />

analysis are presented versus PDB.<br />

3. 3. 3. 2 Ice rafted detritus (IRD)<br />

188 levels (2 to 10 cm sample spacing with an age resolution average<br />

of 130 yr) were washed with distilled water and wet sieved trough a 150 μm<br />

mesh screen. After this classic sedimentological procedure, IRD<br />

semiquantitative analysis was performed on the >150 μm sand-size fraction. In<br />

this study, only the total concentrations of the lithic grains were considered.<br />

The presence of two anomalous small peaks of IRD after H4 and H3<br />

associated with age inversions likely shows reworked levels as the result of<br />

destabilizations of the slope through changes of sea-level (Figs. III.2 and III.4).<br />

3. 3. 3. 3 Planktonic foraminifer-derived SST<br />

154 levels (2 to 20 cm of sample spacing with a time resolution average<br />

of 150 years) were treated for planktonic foraminifera analysis. Sub samples<br />

166<br />

F. Naughton, 2007


were washed with distilled water and wet sieved trough a 150 mesh screen.<br />

At least 400 specimens per sample were counted for semiquantitative analysis<br />

and identified based on Kennet and Srinivasan (1983).<br />

Planktonic foraminifera were grouped into three main bioclimatic<br />

assemblages: polar (Neogloboquadrina pachyderma sinistral), subpolar<br />

(Neogloboquadrina pachyderma dextra, Globigerina bulloides and<br />

Turborotalia quinqueloba) and warm including temperate/cold, subtropical,<br />

warm subtropical and tropical, (Globorotalia scitula, G.inflata, G.hirsuta,<br />

G.truncatulinoides, G.crassaformis, Globigerinita glutinata, Globigerina<br />

falconensis, G.calida,G.rubescens, G.digitata, Hastigerina aequilateralis,<br />

Orbulina universa, Globigerinoides ruber) (e.g. Bé, 1977; Ottens, 1991; Duprat,<br />

1983).<br />

Winter (February) and summer (August) sea-surface temperatures SST<br />

were estimated by using the modern analogue technique transfer function<br />

from the database of Pflaumann et al. (1996) improved by E. Cortijo (LSCE,<br />

Gif-sur-Yvette, France) and J. Duprat (UMR CNRS 5805 EPOC) on planktonic<br />

foraminifera assemblages.<br />

3. 3. 3. 4 Alkenone-derived SST<br />

Long-chain C37-C39 ketones, or alkenones are biomarkers synthesised by<br />

algae (Prymnesiophyceae class) such as the coccolithophores E. huxleyi and<br />

Gephyrocapsa oceanica (Volkman et al., 1980; Volkman et al., 1995). The<br />

unsaturation ratio of C37 alkenone (U k 37΄) is directly correlated with the<br />

temperature of Emiliania huxleyi growth in laboratory cultures and therefore, it<br />

is used as proxy for sea surface temperature estimations (Prahl et al., 1988;<br />

Rostek et al., 1993; Rosell-Melé et al., 1995).<br />

Calibration on U k 37΄ had been applied to modern sediments (Sikes et<br />

al., 1991; Rosell-Melé et al., 1995) agreeing with previous calibration on the E.<br />

huxleyi cultures. We will assume that our data represent the annual SST even if<br />

the impact of seasonality on the SST signal remains unsolved (Sachs et al.,<br />

2000).<br />

136 levels (5 to 10 cm of sample spacing) were used for alkenone<br />

analysis providing a time resolution average of 180 yr. The extraction of<br />

alkenone from the sediment was performed using an automated Dionex<br />

167<br />

F. Naughton, 2007


Accelerated Solvent Extractor (ASE-200) in Cerege, Aix-Marseille III, CNRS<br />

UMR-6635. C37 concentration was determined using n-C36 added to the ASE-<br />

cell before the extraction processes and SST values were calculated based<br />

on Prahl et al. (1988) equation.<br />

The weak C37 concentration detected at around 10 m core depth<br />

prevented reliable SST estimates for H4 event.<br />

3. 4 Results and discussion<br />

Direct correlation of both continental and marine high-resolution<br />

records from the MD99-2331 deep-sea core allows us to recognise the<br />

vegetation changes in north-western Iberia linked with Heinrich events 4 to 1<br />

and the LGM period, and to discriminate the main mechanisms underlying<br />

these changes.<br />

3. 4. 1 Long-term climate variability and the LGM period<br />

Both the late MIS3 (starting at around 40 000 cal yr BP) and almost the<br />

entire MIS2, representing the final stages of the last glacial period, are<br />

characterised by the gradual increase of the global ice extent which is<br />

indirectly testified by the steady increase of heavy benthic δ 18 O values in the<br />

MD99-2331 record (Fig. III.3) as previously suggested by Shackleton (1987).<br />

This gradual increase of the global ice is synchronous with the long-term<br />

Greenland atmosphere temperature decrease (Sánchez Goñi et al., in prep.)<br />

during a period characterised by weak high-latitude summer insolation<br />

(Berger, 1978). The gradual long-term cooling is synchronous with the long-<br />

term pattern of Pinus and temperate forest contraction as well as Poaceae<br />

expansion in north-western Iberia (Fig. III.3). The general trend of gradual<br />

forest tree decline has been previously detected in this region (Roucoux et al.,<br />

2001; Roucoux et al., 2005) and elsewhere in Europe further south (Sánchez<br />

Goñi et al., 2000) and east (Tzedakis et al., 2004).<br />

168<br />

F. Naughton, 2007


Fig. III.3 | Comparison between long term trends of MD99-2331 record and Greenland temperatures<br />

(Sánchez Goñi et al., in prep.) during the Late MIS 3 and MIS 2 against age (cal yr BP). From the bottom to<br />

the top: MD99-2331 benthic δ 13C; percentages of Pinus, temperate and humid trees and Poaceae and<br />

Greenland temperatures. Dashed line represents the long term trend of each signature.<br />

The maximum ice sheet extension took place between 30 000 yr and<br />

19 000 yr contemporaneously with a minimum of the sea-level (Lambeck et<br />

169<br />

F. Naughton, 2007


al., 2002). This period is however punctuated by a number of D-O and H<br />

events. Its steady state is only observed in the interval bracketed between H2<br />

and H1 events (Mix et al., 2001). This period of relatively stable climate which<br />

was dramatically different from that of today, representing maximum glacial<br />

conditions, has been defined by EPILOG as the LGM.<br />

Paradoxically, the LGM interval in the MD99-2331 record is<br />

characterised by a huge decrease of polar foraminifera and an increase of<br />

sub-polar and warm planktonic assemblages suggesting an increase of SST<br />

values in the north-western Iberian margin at this time (Fig. III.4). Planktonic<br />

foraminifera climate reconstruction estimate 14° to 17°C and 9° to 13°C for<br />

summer and winter SST, respectively, while alkenone-based estimates show<br />

annual SST means of about 12°-13°C (Fig. III.4).<br />

Although LGM SST values are similar or even higher than those<br />

characterising the late MIS3 D-O interstadials (GIS8 to GIS3) in the MD99-2331<br />

record, north-western Iberia vegetation did not reply in the same way (Figs.<br />

III.3 and III.4). During GIS8, 7, 6 and 5, temperate and humid trees slightly<br />

developed in this region but more than during GIS 4 and 3 (Fig. III.4). They<br />

were slightly present over the LGM (Fig. III.4). It has been proposed (Sánchez<br />

Goñi et al., 2000; Sánchez Goñi, 2006) that during D-O interstadials of MIS3<br />

temperate forest expansion started when summer SST reached 12°C. At<br />

present, the distribution of temperate forest in both sides of the North Atlantic<br />

coincides with summer SST between 12 and 18°C (Van Campo, 1984).<br />

Because throughout the LGM period summer SST off north-western<br />

Iberian was higher than 12°C we should expect the expansion of temperate<br />

forest in this region. However, the vegetation cover of north-western Iberia<br />

during the LGM period was dominated by pine and herbaceous communities<br />

(including heaths and central-European steppe species) and by the slight<br />

presence of temperate and humid trees, suggesting that only scattered<br />

pockets of deciduous trees colonised this region (Naugthon et al., 2006).<br />

Furthermore, here and in south-western Iberia (Turon et al., 2003) wet<br />

conditions prevailed during the LGM likely as a response to a more vigorous<br />

Meridional Overturning Circulation (MOC) testified by 231 Pa/ 230 Th<br />

measurements which estimate only less than 30-40% of MOC slowdown<br />

170<br />

F. Naughton, 2007


through this period (McManus et al., 2004) and predicted by numerical<br />

climate models (Ganopolsky et al., 1998).<br />

In the framework of the MARGO program, a compilation of several<br />

climate reconstructions, including the Iberian margin region, further shows a<br />

gradual latitudinal SST decrease from a southern to northern transect (e.g<br />

Morey et al., 2005) during the LGM period. This latitudinal climatic gradient is<br />

right expressed in pollen diagrams from this region (Boessenkool et al., 2001;<br />

Turon et al., 2003; Combourieu-Nebout et al., 2002; Pons and Reille, 1988;<br />

Roucoux et al., 2005; Naughton et al., 2006) suggesting that deciduous<br />

Quercus forest expanded more easily in southern than in northern Iberia. Even<br />

if this latitudinal SST gradient has contributed for the different amplitude<br />

response between southern and northern vegetation during the LGM, the<br />

magnitude of the temperate tree expansion is not linearly related to the SST<br />

increase as it was during the D-O interstadials (Fig. III.3). Therefore other<br />

mechanisms seem to have played a crucial role for precluding the large<br />

expansion of temperate trees during the LGM in both regions.<br />

One of the mechanisms that could explain the reduction of the<br />

temperate tree forest, during this period, would be the increase of albedo<br />

due to the maximum expansion of the northern ice-sheets and North Atlantic<br />

sea-ice cover (Broccoli, 2000) which affected the steady-state global-mean<br />

temperature (Rahmstorf, 2002). Indeed, the long term decline of temperate<br />

and humid trees in western Iberia and the gradual amplitude decrease of the<br />

D-O interstadials tightly parallels benthic foraminifera δ 18 O curve which<br />

despite its low resolution clearly shows the general trend of ice volume<br />

increase (Fig. III.3). This suggests that the ice accumulation via the albedo<br />

feedback probably masked the impact of the mid-latitudes high SST values in<br />

western Iberia during the LGM period. Moreover, this period is marked by the<br />

increase of seasonal contrast as the result of sea-ice expansion during winter<br />

and its reduction during summer in the North Atlantic and Nordic Seas (de<br />

Vernal et al., 2005) which would also contribute to prevent deciduous trees<br />

expansion in western Iberia. Also, the CO2 concentration average of 200<br />

ppm during the LGM, representing 35% less than the present-day values<br />

(Cowling and Sykes, 1999), could had play an important role for precluding<br />

the expansion of temperate forest in western Iberia.<br />

171<br />

F. Naughton, 2007


Fig. III.4 | Multi-proxy results of MD99-2331 record. From bottom to top: ice-rafted detritus (IRD)<br />

concentrations, percentages of planktonic foraminifera associations (polar, sub-polar and warm),<br />

planktonic foraminifera-based winter and summer SST estimates, alkenone-based annual SST<br />

reconstruction, δ 18O of G. bulloides, percentages of temperate and humid trees, Pinus percentages and<br />

Greenland temperatures (Sánchez Goñi et al., in prep.). Grey lines represent the Heinrich events. Note: the<br />

age limits of H4 that are based on GISP2 chronology are slightly different from those estimated from the<br />

calibration using NGRIP.<br />

172<br />

F. Naughton, 2007


3. 4. 2 Heinrich events<br />

Superimposed to the late MIS3 and MIS2 long term-cooling, Heinrich<br />

events, in particular H4, H3, H2 and H1, display a complex pattern in the north-<br />

western Iberian margin (Fig. III.4), similar to that previously revealed by other<br />

mid-latitudes North Atlantic deep sea cores (e.g. Zahn, 1997; Abrantes et al.,<br />

1998; Loncaric et al., 1998; Bard et al., 2000; Chapman et al., 2000; Grousset<br />

et al., 2000; Zaragosi et al., 2001; Thouveny et al., 2000; Schönfeld et al., 2003;<br />

Narciso et al., 2006) (Fig. III.1).<br />

Although IRD are not detected in the MD99-2331 record at the<br />

beginning of the typical Heinrich events, a first drop of annual, summer and<br />

winter SST estimates, the increase of planktonic δ 18 O values and that of the<br />

planktonic polar foraminifera clearly show that these events have left an<br />

imprint on north-western Iberian margin before the maximal arrival of icebergs<br />

in this region (Fig. III.4). The evidence of this impact is reliable dated in north-<br />

western Iberian margin, and are in agreement with the age limits proposed<br />

by Elliot et al. (2002) elsewhere in the North Atlantic region.<br />

North-western Iberia vegetation has reacted contemporaneously with<br />

this complex pattern left by the Heinrich events in the Iberian margin. The<br />

beginning of H4, H3, H2 and H1, marked by a SST drop in the ocean, is<br />

synchronous with an important episode of Pinus forest contraction on land<br />

suggesting a strong decrease of atmospheric temperatures (Fig. III.4). North-<br />

western Iberia atmospheric cold conditions are also testified by a drop in<br />

temperate and humid trees even though they played a secondary role in the<br />

total tree cover during the glacial period (Fig. III.4).<br />

Pinus forest contraction occurred, therefore, slightly before the<br />

maximal presence of IRD in the north-western Iberian margin during every<br />

Heinrich events recorded in the MD99-2331 deep-sea core (Figs. III.4 and<br />

III.5). Within H4 interval, the second Pinus forest contraction is, however,<br />

synchronous with the maxima of IRD. Higher resolution analysis is required to<br />

better constrain this episode. The multiproxy diagram of MD95-2039 deep sea<br />

core (40° 34’ N, 10° 20’ W) (Roucoux et al., 2005) also illustrates that during<br />

Heinrich events 5 to 1 the maximum of Pinus forest contraction occurred<br />

before the maxima of IRD in this region. Pinus forest decline has been<br />

173<br />

F. Naughton, 2007


detected further north in MD04-2845 deep sea core retrieved in the Bay of<br />

Biscay (work in progress).<br />

Another important feature within the H4, H2 and H1 in north-western<br />

Iberia is that the first phase is marked by an increase of Calluna and of total<br />

pollen concentration, indicating wet conditions (Fig. III.5). Indeed, the sharp<br />

decline of Pinus forest favoured continental erosion while the increase of<br />

Calluna indicates an increase of moisture, intensifying river discharges and<br />

seaward pollen transfer (Fig. III.5). Following this, the second phase is<br />

characterised by the expansion of semi-desert plants reflecting an increase of<br />

dryness contemporaneously with the maxima of IRD arrival in the north-<br />

western Iberian margin and less cold sea surface conditions (Figs. III.4 and<br />

III.5). The impact of H3 in north-western Iberia is peculiar and associated with<br />

wet conditions over almost the entire event.<br />

Fig. III.5 | Response of the north-western Iberia vegetation to the complex pattern of Heinrich events in the<br />

Iberian margin. From bottom to top: Heinrich events, percentages of Pinus, percentages of Calluna<br />

representing wet conditions, percentages of semi-desert plants reflecting continental dryness and total of<br />

pollen concentration. Dashed line separated wet from dryness conditions during H events.<br />

174<br />

F. Naughton, 2007


Evidence for a gradual increase of dryness during each of the last five<br />

Heinrich events have been detected further south in Iberia (Sánchez Goñi et<br />

al., 2000) and in the Mauritanian margin MD03-2705 deep sea core (Jullien et<br />

al., submitted) (Fig. III.1). Peaks in dust from this tropical core occur at the end<br />

of the putative Heinrich events (Jullien et al., submitted). This suggests as our<br />

record does that the mid- and low- latitudes of the eastern North Atlantic<br />

region have experienced a first relatively wet period followed by increasing<br />

dryness conditions during at least the last five Heinrich events. However, dust<br />

content was lower during H3 than through H4, H2 and H1, suggesting that this<br />

episode was less dry than the others also in the north-eastern tropical region.<br />

3. 4. 3 Possible mechanisms triggering the complex pattern signal of<br />

Heinrich events in and off north-western Iberia<br />

Different hypotheses had been proposed to explain the complex<br />

pattern observed during Heinrich events in the mid-latitudes of the North<br />

Atlantic: a) multiple IRD sources (Bard et al., 2000; Thouveny et al., 2000;<br />

Sánchez Goñi et al., 2000); b) multiple pulses of the Laurentide Ice sheet<br />

(Abrantes et al., 1998), and c) shifts in the polar front position (Chapman et al,<br />

2000). However, none of these hypotheses can explain either the succession<br />

of wet and dry conditions during the typical H4, H2 and H1 or the drastic<br />

atmospheric and sea surface cooling characterising the first stage of these<br />

extreme events in north-western Iberia margin and adjacent landmasses.<br />

Furthermore, the increase of dryness in Iberia coinciding with the maximum<br />

dust input off Mauritania revealed by core MD03-2705 (Jullien et al.,<br />

submitted) suggests that mechanisms triggering both atmospheric signatures<br />

in temperate and tropical North Atlantic latitudes were probably connected.<br />

Indeed, both regions are influenced by changes of the North Atlantic<br />

Meridional Overturning Circulation (MOC) coupled with changes in prevailing<br />

(negative or positive) mode of the North Atlantic Oscillation (NAO) index in<br />

north-western Iberia and with shifts of the Intertropical Convergence Zone<br />

(ITCZ) in north of Africa (Marshall et al., 2001).<br />

Present-day relatively warm and moistness conditions of western<br />

Europe are sustained by the poleward surface branch of the Atlantic Ocean<br />

thermohaline circulation (THC) transferring heat from the tropics to the high<br />

175<br />

F. Naughton, 2007


latitudes of the northern hemisphere (Rahmstorf, 1995). However, this system is<br />

highly sensitive to freshwater input which disturbs the strength of the THC<br />

affecting regional and global climate (e.g., Rahmstorf, 1995, Broecker and<br />

Hemming, 2001; Clark et al., 2002; Vellinga and Wood, 2002; Timmermann et<br />

al., 2005).<br />

In order to understand the impact of a Heinrich event on climate,<br />

several climate models have introduced anomalous freshwater pulses into<br />

the North Atlantic to force the MOC to shutdown (e.g. Seidov and Maslin,<br />

1999, Ganopolski and Rahmstorf, 2001; Knuti et al., 2004; Rahmstorf et al.,<br />

2005) triggering a substantial SST drop in that region (e.g. Paillard and<br />

Labeyrie, 1994; Seidov and Maslin, 1999). Besides modelling simulations, the<br />

episode of drastic and complete MOC shutdown has been confirmed for H1<br />

by 231 Pa/ 230 Th measurements from OCE326-GGC5 deep sea core (McManus<br />

et al., 2004) (Fig. III.1). Following this, rapid oceanic and atmospheric<br />

reorganizations favoured the transfer of atmospheric cold conditions to the<br />

north-western Iberia triggering the Pinus and temperate forests decline in this<br />

region.<br />

MOC shutdown should also preclude moisture transfer to Europe during<br />

the episode of Pinus forest decline. On the contrary, an increase of wet<br />

conditions in north-western Iberia is observed. Therefore other mechanisms<br />

must be invoked to explain the increase of wet conditions during the first<br />

phase of Heinrich events in north-western Iberia. One of the mechanisms that<br />

could explain the increase of wet conditions during the first phase of typical<br />

Heinrich events is an atmospheric one related with weak pressure gradient<br />

between high and low latitudes of the North Atlantic (negative mode of the<br />

North Atlantic Oscillation-NAO). The north-western Iberian dryness during the<br />

second part of typical Heinrich events would be, in turn, interpreted as a<br />

prevailing positive mode of the NAO (Figs. III.6 and III.7).<br />

Despite the importance of the NAO conditions upon present-day<br />

wintertime climate it is still unknown how the NAO have changed in the past.<br />

During the last glacial period the great extension of ice in the Northern<br />

Hemisphere has probably affected the latitudinal and longitudinal position of<br />

the low- and high-pressure cells. Therefore, the use of NAO mode changes as<br />

an interpretive tool for explaining the complex climatic pattern of north-<br />

176<br />

F. Naughton, 2007


western Iberian Heinrich events must be made with caution. For this reason<br />

we will use the term NAO-like from now on.<br />

Indeed, during the wet first phase of a Heinrich event in this region both<br />

the Azores high and the Icelandic low were weak, giving rise to reduced<br />

westerlies over the eastern North Atlantic triggering prevailing negative NAO-<br />

like index conditions over Europe (Fig. III.6). This would favour, as at present-<br />

day (Trigo et al., 2004), the increase of winter precipitation and river flow in<br />

western Iberia as revealed by the development of Calluna and the increase<br />

of total pollen concentrations in our record. Furthermore, this atmospheric<br />

mechanism produces nowadays warm sea surface conditions in the north-<br />

western Atlantic above 45°N (Wanner et al., 2001) (Fig. III.6). During the first<br />

phase of a Heinrich event, this warming might have facilitated iceberg<br />

melting in the IRD belt as supported by the highest IRD concentrations (e.g.<br />

Heinrich, 1988; Andrews and Tedesco, 1992; Grousset el al., 1993; Bond and<br />

Lotti, 1995; Gwiazda et al., 1996; Hemming et al., 1998) and consequent<br />

cooling of SST in that region. This likely hindered iceberg southern migration to<br />

the mid-latitudes and explain the almost absence of IRD at the beginning of<br />

H4, H3, H2 and H1 in western Iberia margin (Fig. III.6). The proposed prevailing<br />

negative mode of the NAO-like index also favours the decrease of SST along<br />

north-western Iberian margin until the Greater North Sea (Fig. III.6).<br />

Fig. III.6 | Prevailing NAO negative conditions scheme (adapted from Wanner et al., 2001).<br />

177<br />

F. Naughton, 2007


The dry second phase of H4, H2 and H1 in Iberia, demonstrated by the<br />

expansion of semi-desert plants from our MD99-2331 record was probably the<br />

result of the strong pressure gradient between Açores and Iceland which<br />

favoured the northward displacement and intensification of the westerlies<br />

(Fig. III.7). At present-day, during prevailing positive mode of NAO, warm sea<br />

surface conditions move towards the western mid-latitudes of the North<br />

Atlantic (Wanner et al., 2001) (Fig. III.7). During the second phase of Heinrich<br />

event the observed increase of IRD in the Iberian margin deep sea cores<br />

could be explained by the warmth of mid-latitudes SST which favoured the<br />

iceberg maximum arrival and melting at these latitudes (Fig. III.7). At this<br />

moment most of deep sea cores from the IRD belt displays a gradual<br />

decrease of IRD content. The prevailing positive mode of NAO-like index can<br />

probably explain such a relatively warming of SST although iceberg melted at<br />

this time in western Iberian margin.<br />

Fig. III.7 | Prevailing NAO positive conditions scheme (adapted from Wanner et al., 2001).<br />

To confirm that changes in the NAO modes operated within Heinrich<br />

events we need to know whether there is an asymmetry between the east<br />

and west of the North Atlantic climate during these events. Grimm et al (2006)<br />

suggest based on Lake Tulane pollen record that Heinrich events in Florida<br />

178<br />

F. Naughton, 2007


were characterised by warm and wet conditions. However, the low resolution<br />

data of Lake Tulane record and the radiometric dating-based correlation<br />

between land and ocean climatic proxies precludes the identification of two<br />

possible vegetation phases in this pollen sequence during Heinrich events.<br />

Therefore, future work on pollen-rich deep-sea cores from the western North<br />

Atlantic is crucial to support our hypothesis.<br />

The impact of H3 in Iberia is different from H4, H2 and H1. Indeed H3<br />

has been considered as an “atypical” Heinrich event (Snoeckx et al., 1999)<br />

being characterised by a European ice sheet dominant signal (Bond et al.,<br />

1992; Grousset et al., 1993; Snoeckx et al., 1999). Furthermore, studies carried<br />

out in marine sediments from the high-latitudes of the North Atlantic and in<br />

the western North Atlantic suggest that during H3 Laurentide ice sheet (Elliot<br />

et al., 1998; Jullien et al., in press) as well as the others pan-Atlantic ice sheets<br />

(Jullien et al., in press) may have discharged icebergs on a much smaller<br />

scale than during the typical Heinrich events. The relatively wet conditions<br />

detected in north-western Iberia (this study) and off Mauritanian coast (Jullien<br />

et al., submitted) over the whole H3 would be the result of reduced wind-field<br />

intensification in both regions likely due to a reduced pressure gradient<br />

between Açores and Iceland in the North Atlantic region and to relatively<br />

small southern migration of the ITCZ in the tropical zone.<br />

3. 5 Conclusions<br />

Direct correlation between marine and terrestrial proxies from MD99-<br />

2331 deep sea core shows that during the LGM temperate tree expansion<br />

was largely reduced when compared with the previous late MIS 3 D-O<br />

interstadials although SST were similar. We propose three mechanisms to<br />

explain this decoupling between the ocean and atmosphere temperatures:<br />

a) albedo increase which enhanced the cooling produced by low summer<br />

insolation in the Northern Hemisphere, b) high seasonality, and c) weak CO2<br />

concentration. Nonetheless, the wet conditions in Iberia during the LGM were<br />

likely the result of the strengthening of the MOC which was more vigorous<br />

than during the bracketing Heinrich events.<br />

179<br />

F. Naughton, 2007


During Heinrich events 4, 3, 2 and 1 the introduction of large amounts<br />

of freshwater via Northern hemisphere icebergs drifting and consequent<br />

melting triggered a shutdown of the Atlantic Meridional Overturning<br />

Circulation (MOC) and a drop of SST in the North Atlantic region. This<br />

produced ocean-atmosphere rapid reorganizations which allow the fast cold<br />

conditions transfer into north-western Iberia triggering Pinus forest decline.<br />

Superimposed to this important cooling, changes of prevailing<br />

(negative and positive) North Atlantic Oscillation (NAO-like) index seems to<br />

have played a crucial role for explaining the complex pattern of Heinrich<br />

events in north-western Iberian margin and in the adjacent continent.<br />

Indeed, Heinrich events in this region is characterised by two main<br />

phases: a) the first is marked by a drop of SST and the virtual absence of<br />

icebergs in the Iberian margin, an important fall of atmospheric temperatures<br />

(strong Pinus forest contraction) and an increase of moisture conditions<br />

(Calluna expansion in concert with the increase of total pollen<br />

concentration), b) the second phase is characterised by less cold conditions,<br />

maximal arrival of icebergs and an increase of dryness (semi-desert plants<br />

expansion). During the first phase of H4, H2 and H1, prevailing negative NAO-<br />

like index likely triggered the increase of winter precipitation in Iberia and<br />

enhanced river flow favouring the seaward pollen transfer. Furthermore, these<br />

prevailing conditions allowed iceberg melting in the IRD belt preventing their<br />

southern migration to the mid-latitudes. This prevailing NAO-like mode also<br />

favoured the drop of SST in north-western Iberian margin. Through the second<br />

phase of H4, H2 and H1, prevailing positive NAO-like index conditions<br />

intensified and moved northward the westerlies triggering an increase of<br />

dryness in this region. The displacement of warm sea surface conditions to the<br />

mid-latitudes of the North Atlantic favoured the southward migration of the<br />

icebergs to the mid-latitudes sites. This mechanism also contributes to a<br />

relatively warming of sea surface conditions in north-western Iberia. H3 is an<br />

exceptional case of prevailing wet conditions during almost the entire event<br />

probably due to maintaining reduced westerlies in this region.<br />

References:<br />

Abrantes, F., Baas, J., Haflidason, H., Rasmussen, T., Klitgaard, D., Loncaric, N. and<br />

Gaspar, L., 1998. Sediment fluxes along the northeastern European Margin: inferring<br />

hydrological changes between 20 and 8 kyr. Marine Geology, 152: 7-23.<br />

180<br />

F. Naughton, 2007


Alcara Ariza, F. Asensi Marfil, A., de Bolos y Capdevilla, O., Costa Tales, M., Arco<br />

Aguilar, M., Diaz Gonzales, T.E., Diez Garretas, B., Fernandez Prieto, J.A., Fernandez<br />

Gonzales, F., Izco Sevillando, J., Loidi Arregui, J., Martinez Parras, J.M., Navarro Andres,<br />

F., Ninot I Sugranes, J.M., Peinado Lorca, M., Rivas Martinez, S., Sanchez Mata, D.,<br />

Valle Guitierrez, C., Vigo I Bonada, J. and Wildpret de la Torre, W., 1987. La<br />

vegetación de España. Collection Aula Abierta. Universidad de Alcala de Henares,<br />

544 pp.<br />

Andrews, J.T. and Tedesco, K., 1992. Detrital carbonate-rich sediments, northwestern<br />

Labrador Sea: implications for ice-sheet dynamics and iceberg rafting (Heinrich)<br />

events in the North Atlantic. Geology, 20: 1087-1090.<br />

Baas, J.H., Mienert, J., Abrantes, F. and Prins, M.A., 1997. Late Quaternary<br />

sedimentation on the Portuguese continental margin: climate-related processes and<br />

products. Palaeogeography, Palaeoclimatology, Palaeoecology, 130: 1-23.<br />

Bard, E., Arnold, M., Maurice, P., Duprat, J., Moyes, J. and Duplessy, J.C., 1987. Retreat<br />

velocity of the North-Atlantic polar front during the last deglaciation determined by<br />

14 C accelerator mass spectrometry. Nature, 328: 791-794.<br />

Bard, E., Rostek, F. and Menot-Combes, G., 2004. Radiocarbon calibration beyond<br />

20,000 14C yr B.P. by means of planktonic foraminifera of the Iberian Margin.<br />

Quaternary Research, 61: 204-214.<br />

Bard, E., Rostek, F., Turon, J.L. and Gendreau, S., 2000. Hydrological impact of Heinrich<br />

events in the subtropical northeast Atlantic. Science, 289: 1321-1324.<br />

Bé, A.W.H., 1977. An ecological, zoological and taxonomic review of recent<br />

planctonic Foraminifera. Oeanic Micropaleontology, Volume 1: London, Academic<br />

Press Inc., 1-100 pp.<br />

Berger, A., 1978. Long-term variations of daily insolation and Quaternary climatic<br />

changes. Journal of Atmospheric Sciences, 35: 2362-2367.<br />

Boessenkool, K.P., Brinkhuis, H., Schonfeld, J. and Targarona, J., 2001. North Atlantic<br />

sea-surface temperature changes and the climate of western Iberia during the last<br />

deglaciation; a marine palynological approach. Global and Planetary Change, 30:<br />

33-39.<br />

Bond, G. and Lotti, R., 1995. Iceberg discharges into the North Atlantic on millennial<br />

time scales during the last glaciation. Science, 267: 1005-1010.<br />

Bond, G., Heinrich, H., Broecker, W., Labeyrie, L., McManus, L., Andrews, J., Huon, S.,<br />

Jantschik, R., Clasen, S., Simet, C., Tedesco, K., Klas, M., Bonani, G. and Ivy, S., 1992.<br />

Evidence for massive discharges of icebergs into the North Atlantic Ocean during the<br />

last glacial period. Nature, 360: 245-249.<br />

Broccoli, Anthony J., 2000. Tropical cooling at the Last Glacial Maximum: An<br />

atmosphere-mixed layer ocean model simulation. Journal of Climate, 13: 951-976.<br />

Broecker, W.S. and Hemming, S., 2001. Climate swings come into focus. Science, 294:<br />

2308-2309.<br />

181<br />

F. Naughton, 2007


Cayre, O., Lancelot, Y. and Vincent, E., 1999. Paleoceanographic reconstructions<br />

from plaktonic foraminifera off the Iberian Margin: Temperature, salinity, and Heinrich<br />

events. Paleoceanography, 14: 384-396.<br />

Chapman, M. R., Shackleton, N. J. and Duplessy, J. C., 2000. Sea surface temperature<br />

variability during the last glacial-interglacial cycle: assessing the magnitude and<br />

pattern of climate change in the North Atlantic. Palaeogeography-<br />

Palaeoclimatology- Palaeoecology, 157: 1 -25.<br />

Clark, P.U., Pisias, N.G., Stocker, T.F. and Weaver, A.J., 2002. The role of the<br />

thermohaline circulation in abrupt climate change. Nature, 415: 863-869.<br />

Combourieu-Nebout, N., Turon, J.-L., Zahn, R., Capotondi, L., Londeix, L. and Pahncke,<br />

K., 2002. Enhanced aridity and atmospheric high-pressure stability over the western<br />

Mediterranean during the North Atlantic cold events of the past 50 ky. Geology, 30:<br />

863-866.<br />

Cowling, S.A., and Sykes, M.T., 1999. Physiological significance of low atmospheric<br />

CO2 for plant-climate interactions. Quaternary Research, 52: 237-242.<br />

de Abreu, L., Shackleton, N.J., Schonfeld, J., Hall, M. and Chapman, M., 2003.<br />

Millennial-scale oceanic climate variability off the Western Iberian margin during the<br />

last two glacial periods. Marine Geology, 196: 1-20.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3: 16-27.<br />

de Vernal, A., Eynaud, F., Henry, M., Hillaire-Marcel, C., Londeix, L., Mangin, S.,<br />

Matthiessen, J., Marret, F., Radi, T., Rochon, A., Solignac, S. and Turon, J.-L., 2005.<br />

Reconstruction of sea –surface conditions at middle to high latitudes of the Northern<br />

Hemisphere during the Last Glacial Maximum (LGM) based on dinoflagellate cyst<br />

assemblages. Quaternary Science Reviews, 24: 897-924.<br />

Desprat, S., 2005. Réponses climatiques marines et continentales du SW de l'Europe<br />

lors des derniers interglaciaires et des entrées en glaciation. PhD Thesis, Bordeaux 1<br />

University, Bordeaux (France), 282 pp.<br />

Dowdeswell, J.A., Maslin, M.A., J.T, A. and McCave, I.N., 1995. Iceberg production,<br />

debris rafting, and the extent and thickness of Heinrich layers (H-1, H-2) in North<br />

Atlantic sediments. Geology, 33: 301-304.<br />

Duprat, J., 1983. Les foraminifères planctoniques du Quaternaire Terminal d'un<br />

domaine pericontinental (Golfe de Gascogne, côtes Ouest-Ibériques, Mer<br />

d'Alboran): Ecologie-Biostratigraphie. Bulletin de l'Institut Géologie du Bassin d'<br />

Aquitaine, 33: pp. 71-150.<br />

Elliot, M. Labeyrie, L., Bond, G., Cortijo, E., Turon, J.-L., Tisnerat, N. and Duplessy, J.-C.,<br />

1998. Millenial-scale iceberg discharges in the Irminger Basin during the last glacial<br />

period: Relationship with the Heinrich events and environmental settings.<br />

Paleoceanography, 13: 433-446.<br />

Elliot, M., Labeyrie, L. and Duplessy, J.-C., 2002. Changes in North Atlantic deep-water<br />

formation associated with the Dansgaard-Oeschger temperature oscillations (60-10<br />

ka). Quaternary Science Reviews, 21: 1153-1165.<br />

182<br />

F. Naughton, 2007


Fronval, T., Jansen, E., Bloemendal, J. and Johnsen, S., 1995. Oceanic evidence for<br />

coherent fluctuations in Fennoscandian and Laurentide ice sheets on millenium<br />

timescales. Nature, 374: 443-446.<br />

Ganopolski, A. and Rahmstorf, S., 2001. Rapid Changes of glacial climate simulated in<br />

a coupled climate model. Nature, 409: 153-158.<br />

Ganopolsky, A., Rahmstorf, S., Petroukhov, V. and Claussen, M., 1998. Simulation of<br />

modern and glacial climates with a coupled global model of intermediate<br />

complexity. Nature, 391: 351–356.<br />

Gouzy, A., Malaizé, B., Pujol C. and Charlier K., 2004. Climatic "pause" during<br />

Termination II identified in shallow and intermediate waters off the Iberian margin.<br />

Quaternary Science Reviews, 23: 1523-1528.<br />

Grimm, E.C., Watts, W.A.Jr., Jacobson, G.L., Hansen, B.C.S., Almquist, H.R. and<br />

Dieffenbacher-Krall, A.C., 2006. Evidence for warm wet Heinrich events in Florida.<br />

Quaternary Science Reviews, 25: 2197-2211.<br />

Grousset, F.E. Labeyrie, L., Sinko, J.A., Cremer, M., Bond, G., Duprat, J., Cortijo, E. and<br />

Huon, S.,1993. Patterns of ice-rafted detritus in the glacial North Atlantic (40-55°N).<br />

Paleoceanography, 8: 175-192.<br />

Grousset, F.E., Pujol, C., Labeyrie, L., Auffret, G. and Boelaert, A., 2000. Were the North<br />

Atlantic Heinrich events triggered by the behavior of the European ice sheets?<br />

Geology, 28: 123-126.<br />

Gwiazda, R.H., Hemming, S. and Broecker, W.S., 1996. Tracking the sources of<br />

icebergs with lead isotopes: the provenance of ice-rafted debris in Heinrich layer 2.<br />

Paleoceanography, 11: 77-93.<br />

Heinrich, H., 1988. Origin and consequences of cyclic ice rafting in the northeast<br />

Atlantic ocean during the past 130,000 years. Quaternary Research, 29: 142-152.<br />

Hemming, S.R., Broecker, W.S., Sharp, W.D., Bond, G.C., Gwiazda, R.H., McManus, J.F.,<br />

Klas, M. and Hajdas, I., 1998. Provenance of Heinrich layers in core V28-82,<br />

northeastern Atlantic: 40Ar/39Ar ages of ice-rafted hornblende, Pb isotopes in<br />

feldspar grains, and Nd-Sr-Pb isotopes in the fine sediment fraction. Earth and<br />

Planetary Science Letters, 164: 317-333.<br />

Hemming, S., 2004. Heinrich events: massive late Pleistocene detritus layers of the<br />

North Atlantic and their global climate imprint. Review of Geophysics, 42: 1-43.<br />

Heusser, L.E. and Stock, C.E., 1984. Preparation techniques for concentrating pollen<br />

from marine sediments and other sediments with low pollen density. Palynology, 8:<br />

225-227.<br />

Hughen, K.A., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell,<br />

P.G., Buck, C.E., Burr, G., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G.,<br />

Friedrich, M., Guilderson, T.P., Kromer, B., McCormac, F.G., Manning, S., Bronk Ramsey,<br />

C., Reimer, P.J., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J. and Weyhenmeyer, C.E., 2004. Marine04 Marine<br />

Radiocarbon Age Calibration, 0-26 Cal Kyr BP. Radiocarbon, 46: 1059-1086.<br />

183<br />

F. Naughton, 2007


Jullien, E., Grousset, F.E., Hemming, S.R., Peck, V.L., Hall, I.R., Jeantet, C. and Billy, I.,<br />

2006. Constrating conditions preceding MIS3 and MIS2 Heinrich events. Global and<br />

Planetary Change (in press).<br />

Jullien, E., Grousset, F.E., Malaizé, B., Duprat, J., Sanchez-Goni, M.F., Eynaud, F.,<br />

Charlier, K., Schneider, R., Bory, A., Bout, V. and Flores J.A., submitted. Tropical "dusty<br />

Heinrich-like events" vs. high-latitude "icy Heinrich events". Quaternary Research.<br />

Kennet, J.P. and Srinivasan, S., 1983. Neogene planktonic foraminifera: a<br />

phylogenetic atlas. Hutchinson Ross, New York, 265p.<br />

Knutti, R., Flückiger, J., Stocker, T.F. and Timmermann, A., 2004. Strong hemispheric<br />

coupling of glacial climate through freshwater discharge and ocean circulation.<br />

Nature, 430: 851-856.<br />

Lambeck, K., Yokoyama, Y. and Purcell, T., 2002. Into and out of the Last Glacial<br />

Maximum: sea-level change during Oxygen Isotope Stages 3 and 2. Quaternary<br />

Science Reviews, 21: 343-360.<br />

Lebreiro, S.M., Moreno, J.C., McCave, I.N. and Weaver, P.P.E., 1996. Evidence for<br />

Heinrich layers off Portugal (Tore Seamount: 39°N, 12°W). Marine Geology, 131: 47-56.<br />

Lebreiro, S.M., Moreno, J.C., Abrantes, F.F. and Pflaumann, U., 1997. Productivity and<br />

paleoceanographic implications on the Tore Seamount (Iberian Margin) during the<br />

last 225 kyr: Foraminiferal evidence. Paleoceanography, 12: 718-727.<br />

Loncaric, N. et al., 1998. Late Quaternary sedimentation patterns on the Meriadzek<br />

Terrace, Bay of Biscay (ESSCAMP 02 core: 47°N 9°W). Marine Geology, 152: 57-73.<br />

Marshall, J., Kushnir, Y., Battisti, D., Chang, P., Czaja, A., Dickson, R., Hurrell, J.,<br />

McCartney, M., Saravanan, R., Visbeck, M., 2001. North Atlantic climate variability:<br />

Phenomena, impacts and mechanisms. International Journal of Climatology, 21:<br />

1863–1898.<br />

McManus, J. F., Francois, R., Gherardi, J.-M., Keigwin, L. D., Brown-Leger, S., 2004.<br />

Collapse and rapid resumption of Atlantic meridional circulation linked to deglacial<br />

climate changes. Nature 428 ,834-837.<br />

Meland, M.Y., Jansen, E. and Elderfield, H., 2005. Constraints on SST estimates for the<br />

northern North Atlantic/Nordic Seas during the LGM. Quaternary Science Reviews, 24:<br />

835-852.<br />

Migeon, S., Weber, O., Faugeres, J.C. and Saint-Paul, J., 1999. SCOPIX: A new imaging<br />

system for core analysis. Geo-Marine Letters, 18: 251-255.<br />

Mix, A.C., Bard, E. and Schneider, R., 2001. Environmental processes of the ice age:<br />

land, oceans, glaciers (EPILOG). Quaternary Science Reviews, 20: 627-657.<br />

Morey, A.E., Mix, A.C. and Pisias, N.G., 2005. Planktonic foraminiferal assemblages<br />

preserved in surface sediments correspond to multiple environmental variables.<br />

Quaternary Science Reviews, 24: 925-950.<br />

Narciso, A., Cachao, M. and de Abreu, L., 2006. Coccolithus pelagicus subsp.<br />

pelagicus versus Coccolithus pelagicus subsp. braarudii (Coccolithophore,<br />

Haptophyta): A proxy for surface subarctic Atlantic waters off Iberia during the last<br />

200 kyr. Marine Micropaleontology, 59: 15-34.<br />

184<br />

F. Naughton, 2007


Naughton, F., Sánchez Goñi, M.F., Desprat, S., Turon, J.-L., Duprat, J., Malaizé, B., Joli,<br />

C., Cortijo, E., Drago T., and Freitas, M.C., 2006. Present-day and past (last 25 000<br />

years) marine pollen signal off western Iberia. Marine Micropaleontology, in press.<br />

Ottens, J.J., 1991. Planktic foraminifera as North Atlantic water mass indicators.<br />

Oceanologia Acta, 14: 123-140.<br />

Paillard, D. and Labeyrie, L., 1994. Role of the thermohaline circulation in the abrupt<br />

warming after Heinrich events. Nature, 372: 162-164.<br />

Pailler, D. and Bard, E., 2002. High frequency palaeoceanographic changes during<br />

the past 140,000 yr recorded by the organic matter in sediments of the Iberian<br />

margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 181: 431-452.<br />

Pflaumann, U., Duprat, J., Pujol, C. and Labeyrie, L. D., 1996. SIMMAX: A modern<br />

analog technique to deduce Atlantic sea surface temperatures from planktonic<br />

foraminifera in deep-sea sediments. Paleoceanography, 11: 15-35.<br />

Polunin, O. and Walters, M., 1985. A guide to the vegetation of Britain and Europe.<br />

Oxford University Press, New York, 238 pp.<br />

Pons, A. and Reille, M., 1988: The Holocene-and Upper Pleistocene pollen record from<br />

Padul (Granada, Spain): a new study. Palaeogeography Palaeoclimatology<br />

Palaeoecology, 66: 243-263.<br />

Prahl, F.G., Muehlhausen, L.A. and Zahnle, D.L., 1988. Further evaluation of long-chain<br />

alkenones as indicators of paleoceanographic conditions. Geochimica et<br />

Cosmochimica Acta, 52: 2303-2310.<br />

Rahmstorf, S., Crucifix, M., Ganapolsky, A., Goose, H., Kamenkovich, I., Knutti, R.,<br />

Lohmann, G., Marsh, B., Mysak, L., Wang, Z. and Weaver, A., 2005. Thermohaline<br />

circulation hysteresis: a model Intercomparison. Geophysical Research Letters, 32:<br />

L23605, doi:10.1029/2005GL023655.<br />

Rahmstorf, S., 1995. Bifurcations of the Atlantic thermohaline circulation in response to<br />

changes in the hydrological cycle, Nature, 378: 145–149.<br />

Rahmstorf, S., 2002. Ocean circulation and climate during the past 120,000 years.<br />

Nature 419: 207-214.<br />

Rasmussen, T.L., Thomsen, E., van Weering, T.C.E. and Labeyrie, L., 1996. Rapid<br />

changes in surface and deep water conditions at the Faeroe Islands Margin during<br />

the last 58 ka. Paleoceanography, 11: 757-771.<br />

Rosell-Melé, A., Carter, J.F., Parry, A.T., and Eglinton, G., 1995. Determination of the<br />

UK37 Index in Geological Samples. Analytical Chemistry, 67: 1283-1289.<br />

Rostek F., Ruhland, G., Bassinot, F.C., Müller, P.J., Labeyrie, L.D., Lancelot, Y. and Bard,<br />

E., 1993. Reconstructing sea surface temperature and salinity using δ 18 O and<br />

alkenone records, Nature, 364: 319-321.<br />

Roucoux, K.H., de Abreu, L., Shackleton, N.J. and Tzedakis, P.C., 2005. The response of<br />

NW Iberian vegetation to North Atlantic climate oscillations during the last 65 kyr.<br />

Quaternary Science Reviews, 24: 1637-1653.<br />

185<br />

F. Naughton, 2007


Roucoux, K.H., Shackleton, N.J., de Abreu, L., Schönfeld, J. and Tzedakis, P.C., 2001.<br />

Combined marine proxy and pollen analyses reveal rapid Iberian vegetation<br />

response to North Atlantic millennial-scale climate oscillations. Quaternary Science<br />

Reviews, 56: 128-132.<br />

Sachs, J.P., Schneider, R.R., Eglinton, T.I., Freeman, K.H., Ganssen, G., McManus, J.F.<br />

and Oppo, D.W., 2000. Alkenones as paleoceanographic proxies. Geochemistry,<br />

Geophysics, Geosystems, 1: 2000GC000059.<br />

Sánchez Goñi, M.F., Cacho, I., Turon, J.-L., Guiot, J., Sierro, F.J., Peypouquet, J.-P.,<br />

Grimalt, J.O. and Shackleton, N.J., 2002. Synchroneity between marine and terrestrial<br />

responses to millennial scale climatic variability during the last glacial period in the<br />

Mediterranean region. Climate Dynamics, 19: 95-105.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L. and Gendreau, S., 2000. European climatic<br />

response to millenial-scale climatic changes in the atmosphere-ocean system during<br />

the Last Glacial period. Quaternary Research, 54: 394-403.<br />

Sánchez Goñi, M.F., 2006. Interactions végétation-climat au cours des derniers<br />

425 000 ans en Europe Occidentale. Le message du pollen des archives marines.<br />

Quaternaire, 17: 1-23.<br />

Sánchez Goñi, M.F., Landais, A. Flückiger, J., Desprat, S., Naughton, F. and Fletcher,<br />

W., in prep. Evidence for major sources of methane in the temperate latitudes (20°N-<br />

40°N) during the last climatic cycle.<br />

Schönfeld, J., Zahn, R. and de Abreu, L., 2003. Surface and deep water response to<br />

rapid climate changes at the Western Iberian Margin. Global and Planetary Change,<br />

36: 237-264.<br />

Seidov, D., and Maslin, M., 1999. North Atlantic Deep Water Circulation collapse<br />

during Heinrich events. Geology, 27: 23-26.<br />

Shackleton, N.J., 1987. Oxygen isotopes, ice volume and sea level. Quaternary<br />

Science Reviews, 6: 183-190.<br />

Sikes, E.L., Farrington, J.W., and Keigwin, L.D., 1991. Use of the alkenone unsaturation<br />

ratio Uk 37 to determine past sea surface temperatures: core-top SST calibrations and<br />

methodology considerations. Earth and Planetary Science Letters, 104: 36-47.<br />

Snoeckx, H., Grousset, F., Revel, M. and Boelaert, A., 1999. European contribution of<br />

ice-rafted sand to Heinrich layers H3 and H4. Marine Geology, 158: 197-208.<br />

Stuiver, M. and Reimer, P., 1993. Extended 14C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

Stuiver, M., Reimer, P.J. and Reimer, R.W., 2005. CALIB 5.0. (WWW program and<br />

documentation).<br />

Telford, R.J., Heegaard, E. and Birks, H.J.B., 2004. The intercept is a poor estimate of a<br />

calibrated radiocarbon age. The Holocene, 14: 296-298.<br />

Thouveny, N., Moreno, E., Delanghe, D., Candon, L., Lancelot, Y. and Shackleton N.J.,<br />

2000. Rock magnetic detection of distal ice-rafted debries: clue for the identification<br />

of Heinrich layers on the Portuguese margin. Earth and Planetary Science Letters, 180:<br />

61-75.<br />

186<br />

F. Naughton, 2007


Timmermann, A., An, S.-I., Krebs, U. and Goosse, H., 2005. <strong>EN</strong>SO suppression due to<br />

weakening of the North Atlantic Thermohaline Circulation. Journal of Climate, 18:<br />

3122-3139.<br />

Trigo, R.M., Pozo-Vazquez, D., Osborn, T.J, Castro-Diez, Y., Gámis-Fortis, S. and<br />

Esteban-Parra, M.J., 2004. North Atlantic oscillation influence on precipitation, river<br />

flow and water resources in the Iberian Peninsula. International Journal of<br />

Climatology, 24: 925-944.<br />

Turon, J.-L., Lézine, A.-M. and Denèfle, M., 2003. Land-sea correlations for the last<br />

glaciation inferred from a pollen and dinocyst record from the Portuguese margin.<br />

Quaternary Research, 59: 88-96.<br />

Tzedakis, P. C., Frogley, M. R., Heaton, T. H. E., 2003. Last interglacial conditions in<br />

southern Europe: evidence from Ioannina, northwest Greece. Global and Planetary<br />

Change, 36: 157-170.<br />

Van Campo, M.M., 1984. Relations entre la végétation de l'Europe et les<br />

températures de surface océaniques après le dernier maximum glaciaire. Pollen et<br />

Spores, XXVI: 497-518.<br />

Van Kreveld, S., Sarnthein, M., Erienkeuser, H., Grootes, P., Jung, S., Nadeau, M.J.,<br />

Pflaumann, U., and Voelker, A., 2000. Potential links between surging ice sheets,<br />

circulation changes, and the Dansgaard-Oeschger cycles in the Irminger Sea, 60-18<br />

kyr. Paleoceanography, 15: 425-442.<br />

Vellinga, M. and Wood, R.A., 2002. Global Climatic Impacts of a Collapse of the<br />

Atlantic Thermohaline Circulation. Climatic Change, 54: 251-267.<br />

Voelker, A.H.L., Sarnthein, M., Grootes, P.M., Erlenkeuser, H., Laj, C., Mazaud, A.,<br />

Nadeau, M.-J., and Schleicher, M., 1998. Correlation of marine 14C ages from the<br />

Nordic seas with the GISP2 isotope record: Implications for 14C calibration beyond 25<br />

ka BP. Radiocarbon, 40: 517-534.<br />

Volkman, J.K., Barrett, S.M., Blackburn, S.I., and Sikes, E.L., 1995. Alkenones in<br />

Gephyrocapsa oceanica: implications for studies of <strong>paleoclimat</strong>e. Geochimica et<br />

Cosmochimica Acta, 59: 513-520.<br />

Volkman, J.K., Eglinton, G., Comer, E.D.S. and Forsberg, T.E.V., 1980. Long-chain<br />

alkenes and alkenones in the marine coccolithophorid Emiliania huxleyi.<br />

Phytochemistry, 19: 2619-2622.<br />

Wanner, H., Brönnimann, S., Casty, C., Gyalistras, D., Luterbacher, J., Schmutz, C.,<br />

St<strong>ephe</strong>nson, D.B. and Xoplaki, E., 2001. North Atlantic Oscillation-Concepts and<br />

studies. Surveys in Geophysics, 22: 321– 382, 2001.<br />

Zahn, R., Schönfeld, J., Kudrass, H.R., Park, M.H. and Erlenkeuser, H., 1997.<br />

Thermohaline instability in the North Atlantic during meltwater events: stable isotopes<br />

and detritus records from core SO75-26KL, Portuguese margin. Paleoceanography,<br />

12: 696-710.<br />

Zaragosi, S. Eynaud, F., Pujol, C., Auffret, G.A., Turon, J.-L. and Garlan, T., 2001.<br />

Initiation of the European deglaciation as recorded in the northwestern Bay of Biscay<br />

slope environments (Meriadzek Terrace and Trevelyan Escarpment): a multi-proxy<br />

approach. Earth and Planetary Science Letters, 188: 493-507.<br />

187<br />

F. Naughton, 2007


188<br />

F. Naughton, 2007


Capítulo 4| Climate variability during the last deglaciation<br />

in north-western Iberian margin and adjacent continent<br />

Variabilidade climática durante a última deglaciação no<br />

noroeste da Península Ibérica e margem continental<br />

adjacente<br />

Variabilité climatique au cours de la dernière déglaciation<br />

dans le nord-ouest de la marge Ibérique et sur le continent<br />

adjacent<br />

A reduced version of this chapter will be soon submitted to:<br />

Paleogeography, Paleoclimatology, Paleoecology<br />

F. Naughton a, b , M.F. Sánchez Goñi a , J. Duprat a , E. Cortijo c , S. Zaragosi a<br />

a Environnements et Paléoenvironnements Océaniques (UMR CNRS 5805 EPOC), EPHE,<br />

Université Bordeaux 1, Av. des Facultés, 33405 Talence, France<br />

b Departamento e Centro de Geologia -Universidade de Lisboa, Bloco C6, 3º piso,<br />

Campo Grande, 1749-016 Lisboa, Portugal<br />

c Laboratoire des Sciences du Climat et de l’Environnement (LSCE-Vallée), Bât. 12,<br />

avenue de la Terrasse, F-91198 Gif-sur-Yvette cedex, France<br />

189<br />

F. Naughton, 2007


Resumo<br />

A correlação directa entre indicadores paleoclimáticos terrestres e<br />

marinhos, obtidos a partir do estudo de uma sondagem marinha profunda<br />

(MD03-2697), recolhida no noroeste da margem Ibérica, permitiu detectar<br />

uma variabilidade climática sub-orbital, nas médias latitudes do Atlântico<br />

Norte, durante a última deglaciação.<br />

Durante o último máximo glaciar (LGM), o noroeste da Península<br />

Ibérica foi afectado por um clima relativamente frio e húmido enquanto que<br />

o oceano adjacente apresentava temperaturas de superfície relativamente<br />

quentes. O aumento da intensidade da circulação termohalina do Atlântico<br />

Norte (MOC) favoreceu a transferência de humidade para o noroeste da<br />

Península Ibérica enquanto que o aumento do albedo, o forte contrate<br />

sazonal das altas latitudes do Atlântico Norte e a diminuição do conteúdo<br />

em CO2 atmosférico contribuíram para a manutenção de condições frias<br />

sobre o continente.<br />

Nas médias latitudes do Atlântico Norte, o evento de Heinrich 1 está<br />

marcado por um padrão complexo, constítuido por duas fases distintas. A<br />

primeira fase é marcada por condições oceânicas extremamente frias e<br />

uma fraca presença de icebergues, enquanto que a segunda fase é<br />

caracterizada por temperaturas da massa de água superficial menos frias e<br />

por uma grande quantidade de detritos provenientes da fusão de<br />

icebergues. No noroeste da Península Ibérica a vegetação respondeu<br />

contemporaneamente a estas duas fases que caracterizam o evento de<br />

Heinrich 1. O declínio da floresta de Pinus e a expansão de Ericaceae<br />

(incluindo Calluna), durante a primeira fase, marcam condições<br />

extremamente frias e húmidas no noroeste da Península Ibérica, enquanto<br />

que na segunda fase a expansão de plantas semi-desérticas e da floresta de<br />

Pinus reflectem condições relativamente frias e áridas no continente. O<br />

drástico arrefecimento que caracteriza a primeira fase foi provavelmente<br />

provocado pela interrupção da MOC seguido por reorganizações rápidas<br />

entre o oceano e a atmosfera, enquanto que as fases húmida e seca<br />

resultam do predomínio de condições semelhantes aos modos negativo e<br />

positivo do índice da oscilação Norte Atlântica (NAO-like), respectivamente.<br />

190<br />

F. Naughton, 2007


O aquecimento continental (expansão de Quercus deciduous) e<br />

oceânico que caracteriza o evento Bölling-Alleröd (B-A) foi favorecido pelo<br />

aumento da insolação de verão nas latitudes médias do Hemisfério Norte<br />

assim como pela intensificação da MOC. Há cerca de 14 000 anos a<br />

expansão máxima de Quercus deciduous a qual reflecte um aquecimento<br />

continental extremo é síncrona do pico máximo na temperatura da<br />

Gronelândia e do episódio de subida súbita do nível do mar (MWP 1A). Este<br />

aquecimento severo do Hemisfério Norte poderá ter sido o impulsionador<br />

deste drástico evento.<br />

Durante o Dryas recente (YD), a diminuição da floresta de Quercus e a<br />

expansão de plantas semi-desérticas reflectem um arrefecimento e aumento<br />

da aridez continental contemporâneo da diminuição da temperatura da<br />

massa de água superficial no oceano adjacente. A redução da intensidade<br />

da MOC, embora associada ao máximo de insolação de verão das latitudes<br />

médias do Hemisfério Norte, favoreceu uma diminuição em vez de um<br />

declínio total da floresta decídua de Quercus no noroeste Ibérico. Para além<br />

da redução da intensidade da MOC, a predominância de condições<br />

semelhantes ao actual índice positivo da NAO (NAO-like) parece justificar o<br />

aumento da aridez observado ao longo deste período.<br />

Nesta região, o máximo térmico do Holocénico (HTM) foi detectado<br />

entre 11 700 e 8 200 anos cal BP. Por volta dos 8 200 anos cal BP um rápido<br />

arrefecimento continental (diminuição da floresta decídua de Quercus e<br />

Corylus) e oceânico marca o evento frio “8.2 ky” no noroeste da Península<br />

Ibérica e margem adjacente. Este episódio resulta do culminar de uma<br />

sucessão de episódios relaccionados com o colapso da calote glaciar da<br />

“Laurentide” a qual provocou a amplificação da diminuição da<br />

temperatura na Europa e na Gronelândia.<br />

Após o evento “8.2 ky” a diminuição gradual da floresta temperada é<br />

contemporânea da diminuição da temperatura induzida pela diminuição<br />

da insolação de verão das latitudes médias do Hemisfério Norte. Isto sugere<br />

que a regressão da floresta temperada parece ter sido mais afectada pelas<br />

variações orbitais do que pelo impacto antrópico.<br />

191<br />

F. Naughton, 2007


Résumé<br />

La corrélation directe entre des indicateurs climatiques terrestres<br />

(pollen) et marins d’une carotte marine profonde, MD03-2697, prélevée dans<br />

le nord-ouest de la marge Ibérique a permis de détecter une variabilité<br />

climatique d’ordre millénaire au cours de la dernière déglaciation dans les<br />

moyennes latitudes de l’Atlantique Nord.<br />

Le dernier maximum glaciaire (LGM) a été relativement froid et<br />

humide dans le nord-ouest de la Péninsule Ibérique, alors que les<br />

températures des eaux de surface étaient chaudes. L’intensification de la<br />

circulation méridienne Atlantique de renversement (MOC) a favorisé le<br />

transfert d’humidité des moyennes latitudes de l’Atlantique Nord vers la<br />

marge ouest Ibérique tandis que l’augmentation de l’albedo, le fort<br />

contraste saisonnier et la chute de la concentration de CO2 atmosphérique<br />

ont maintenu des températures froides sur le continent.<br />

L’événement d’Heinrich 1 (H1) a été marqué par un scenario<br />

complexe dans les moyennes latitudes de l’Atlantique Nord. La première<br />

phase est caractérisée par des températures des eaux de surface<br />

extrêmement froides et par la quasi-absence d’icebergs alors que la<br />

deuxième phase est légèrement moins froide et caractérisée par une forte<br />

quantité de débris provenant de la fonte d’icebergs (IRD). La végétation du<br />

nord-ouest de la Péninsule Ibérique a répondu de façon synchrone au<br />

scenario complexe qui caractérise l’H1. Au cours de la première phase, la<br />

réduction drastique de la forêt de pins et l’expansion des bruyères révèlent<br />

des conditions extraordinairement froides et humides. La deuxième phase est<br />

caractérisée par l’expansion de la forêt de pins et des plantes semi-<br />

désertiques indiquant des conditions relativement froides et arides. Le<br />

refroidissement qui chartérise la première phase a été probablement<br />

engendrée par une coupure de la MOC, suivi par des réorganisations rapides<br />

entre l’océan et l’atmosphère tandis que les phases humides et arides<br />

résulteraient de situation dominante caractérisée par un index négatif et<br />

positif de l’oscillation Nord Atlantique (NAO-like), respectivement.<br />

Le réchauffement atmosphérique et océanique, marqué par<br />

l’expansion de la forêt de chêne caducifolié et l’augmentation des<br />

192<br />

F. Naughton, 2007


températures des eaux de surface, caractérisant l’événement du Bölling-<br />

Alleröd (B-A), est la conséquence de l’augmentation de l’insolation d’été des<br />

latitudes moyennes de l’Hémisphère Nord et de l’intensification de la MOC.<br />

L’expansion maximale de la forêt de chêne reflète un incident extrêmement<br />

chaud vers 14 000 cal ans BP lequel est synchrone à la fois du réchauffement<br />

maximal au Groenland et de l’épisode nommé Meltwater Pulse 1A (MWP 1A).<br />

Ce réchauffement sévère de l’Hémisphère Nord pourrait être le principal<br />

responsable de la fonte drastique et soudaine des calottes glaciaires.<br />

La réduction de la forêt de chêne et l’expansion des plantes semi-<br />

désertiques pendant l’événement du Dryas récent, indiquent un<br />

refroidissement et une augmentation de l’aridité sur le continent. Cet<br />

événement est contemporain de la diminution des températures des eaux<br />

de surface. La réduction de l’intensité de la MOC et l’augmentation de<br />

l’insolation d’été des moyennes latitudes de l’Hémisphère Nord ont favorisé la<br />

réduction plutôt que la disparition complète de la chênaie dans le nord-<br />

ouest de la Péninsule Ibérique. Au-delà de la diminution de l’intensité de la<br />

MOC, la prévalence de l’index positif de la NAO-like pourrait expliquer<br />

l’aridité détectée dans cette région.<br />

Le maximum thermique de l’Holocène (HTM) à été détecté dans la<br />

Péninsule Ibérique entre 11 700 et 8 200 cal ans BP. Un soudain refroidissement<br />

continental (diminution de la forêt de chêne et du noisetier) et océanique<br />

marque l’événement « 8.2 kyrs » dans cette région. Ce refroidissement serait<br />

le résultat d’épisodes successifs associés à la réduction de la calotte<br />

Laurentidienne produisant un fort refroidissement sur l’Europe et au<br />

Groenland.<br />

La diminution graduelle à long terme de la forêt tempérée après<br />

l’événement « 8.2 kyrs » serait la réponse à un refroidissement induit par des<br />

changements orbitaux plutôt qu’à un impact anthropique.<br />

Abstract<br />

Direct correlation between terrestrial (pollen) and marine climatic<br />

indicators from deep sea core MD03-2697 (north-western Iberian margin)<br />

193<br />

F. Naughton, 2007


allows the detection of millennial scale climate variability for the last<br />

deglaciation in the mid-latitudes of the North Atlantic realm.<br />

The Last Glacial Maximum (LGM) was relatively cold and humid in<br />

north-western Iberia while sea surface conditions were warm. More vigorous<br />

Meridional Overturning Circulation (MOC) favoured moisture transfer from the<br />

mid-latitudes of the North Atlantic to the western Iberia whereas increasing<br />

albedo, high seasonality and atmospheric CO2 drop maintain the continent<br />

cold.<br />

The mid-latitudes of the North Atlantic were marked by a complex<br />

pattern within Heinrich 1 (H1) event. In the first phase, sea surface conditions<br />

were extremely cold with almost no evidence for iceberg calving while the<br />

second one was less cold with high quantity of Ice-rafted detritus (IRD). In<br />

north-western Iberia vegetation has responded synchronously to this H1<br />

pattern. During the first phase a drastic Pinus forest decline and heaths<br />

expansion reflect extremely cold and moist conditions whereas the second<br />

phase reveals the expansion of Pinus forest and semi-desert plants<br />

representing relatively cold and dry conditions on the continent. The first<br />

coldest phase was probably triggered by the MOC shutdown followed by<br />

ocean-atmosphere rapid reorganizations while the wet and dry phases were<br />

the result of prevailing negative and positive North Atlantic Oscillation (NAO-<br />

like) indexes, respectively.<br />

The continental (deciduous Quercus expansion) and sea-surface<br />

warming characterizing the Bölling-Alleröd (B-A) event was produced by both<br />

the increase of mid-latitude summer insolation of the northern Hemisphere<br />

and the strengthening of the MOC. Maxima of deciduous Quercus expansion<br />

reflecting extremely warming at 14 000 cal yr BP is synchronous with both the<br />

peak of Greenland temperatures and the Meltwater Pulse 1A (MWP 1A). This<br />

severe warmth of the Northern Hemisphere could be the trigger of this drastic<br />

meting episode.<br />

During the Younger Dryas (YD) the decrease of deciduous Quercus<br />

forest and the expansion of semi-desert plants reflect continental cooling and<br />

dryness which is contemporaneous with sea surface cooling. MOC reduction<br />

but increasing northern mid-latitudes summer insolation favoured a decrease<br />

rather than a complete decline of deciduous Quercus forest in north-western<br />

194<br />

F. Naughton, 2007


Iberia. Beyond the MOC reduction, a prevailing positive NAO-like index could<br />

explain the observed dryness. Following this, the Holocene Thermal Maximum<br />

in this region is identified between 11 700 and 8 200 cal yr BP. At around 8 200<br />

cal yr BP a sudden land (decrease of deciduous Quercus forest and Corylus<br />

woodlands) and sea cooling marks the 8.2 Ky event in Iberia as the result of<br />

the culmination of the successive episodes of the Laurentide Ice sheet decay<br />

which enhanced the cooling over Greenland and Europe. After the 8.2 Ky<br />

event the long-term temperate forest decrease has responded to the<br />

orbitally-induced cooling rather than to human impact.<br />

195<br />

F. Naughton, 2007


196<br />

F. Naughton, 2007


4. 1 Introduction<br />

Abrupt widespread millennial-scale climate changes during the last<br />

deglaciation have been widely documented for the North Atlantic high and<br />

mid-latitudes (e.g. Bond et al., 1993; Keigwin and Lehman, 1994; Andrews et<br />

al., 1995; 1999; McManus et al., 2004) as well as from the tropics (Hughen et<br />

al., 1996; Arz et al., 1999; Rühlemann et al., 1999; Peterson et al., 2000) and<br />

Southern Atlantic Region (Kim et al., 2002; Shemesh et al., 2002). However, the<br />

mechanisms responsible for inter-hemispheric and tropical-pole<br />

teleconnections are far from being completely understood (Hughen et al.,<br />

1996; Blunier and Brook, 2001; Wunsch, 2003). Changes in heat transfer via<br />

termohaline circulation, resulting from iceberg melting, are one of the<br />

mechanisms proposed for explaining sub-orbital scale climate variability<br />

(Bond and Lotti, 1995; Knuti et al., 2004; McManus et al., 2004) while others,<br />

propose that these oscillations are a consequence of changes in the<br />

topography of continental ice sheets leading to windfield shifts (Wunsch,<br />

2006).<br />

Among the mechanisms that appear to be involved in this sub orbital<br />

climatic variability, shifts in the mode of prevailing North Atlantic Oscillation<br />

(NAO-like) index coupled with changes in the Meridional Overturning<br />

Circulation (MOC) could explain the complex climatic pattern of Heinrich<br />

events during the late Marine Isotopic Stage (MIS) 3 and MIS 2 in north-<br />

western Iberian margin and the adjacent continent (Naughton et al., in<br />

prep.). However, these events occurred in a period characterised by a<br />

relatively high and stable ice volume while the climatic variability of the last<br />

deglaciation is affected by ongoing increase in summer insolation and<br />

substantial reduction of ice volume in northern high latitudes. Previous works<br />

have shown, for example, that the decrease in the MOC strength during the<br />

Younger Dryas together with the insolation maximum lead to a mitigated cold<br />

event (McManus et al., 2004).<br />

To further explore the links between MOC strength and atmospheric<br />

variability in the North Atlantic region superimposed to the gradual evolution<br />

of insolation, we correlate north-western Iberia climatic and vegetation<br />

197<br />

F. Naughton, 2007


changes with North Atlantic mid-latitudes sea surface conditions and with<br />

temperature estimates of Greenland.<br />

4. 2 Environmental Setting<br />

Deep sea core MD03-2697 (42° 09’ 59 N, 59° 42’ 10 W) was retrieved at<br />

~100 km off the Galician margin (north-west of Iberia) and at 2164 m of water<br />

depth (Fig. IV.1). This site is at present day under the influence of the North<br />

Atlantic Deep water mass. Morphology, recent sedimentation, detailed<br />

hydrology and regional climate of this region have been previously described<br />

in Naughton et al. (2006).<br />

Briefly, north-western Iberia is influenced by a temperate and humid<br />

climate (mean annual temperature of about 12. 5 °C and precipitation<br />

varying from 1000 to 2000 mm) (Atlas Nacional de España, 1992); is incised by<br />

two important seaward sediment and pollen suppliers such as the Douro river<br />

followed by Minho, especially during downwelling conditions (Dias et al.,<br />

2002; Jouanneau et al., 2002; Oliveira et al., 2002; Naughton et al., 2006) and;<br />

is dominated by oak woodlands (Quercus robur, Q. pyrenaica and Q.<br />

petraea), heaths communities (Ericaceae including Calluna), brooms<br />

(Genista) and gorses (Ulex) (Alcara Ariza et al., 1987).<br />

Fig. IV.1 | Study area. Location of deep-sea cores referred in the text: OCE326-GGC5 (McManus et al.,<br />

2004); MD95-2002 (Zaragosi et al., 2001; Auffret et al., 2002; Ménot et al., 2006) and MD99-2331 (Naughton<br />

et al., 2006; in prep.).<br />

198<br />

F. Naughton, 2007


4. 3 Material and methods<br />

Core MD03-2697 was recovered using a CALYPSO corer during PICABIA<br />

oceanographic cruise on board the R/V Marion Dufresne (Fig. IV.1). MD03-<br />

2697, mainly composed of hemipelagic clays, is 41.23 m long covering the<br />

last 425 000 years (from Marine Isotopic Stages (MIS) 1 to 11). In this study, we<br />

will focus on the last deglaciation period.<br />

Mean sedimentary rate is high (30 cm kyr -1 ) between ~ 20 000 yr cal BP<br />

and 10 000 yr cal BP providing a high-resolution palaeoclimatic record for this<br />

period in north-western Iberian margin and the adjacent continent. In<br />

contrast, during the Holocene, mean sedimentary rate is relatively weak (10<br />

cm kyr -1 ) preventing a detailed <strong>paleoclimat</strong>e record for the present-day<br />

interglacial.<br />

X-ray analysis using SCOPIX image-processing (Migeon et al., 1999),<br />

shows a well preserved sedimentary sequence in core MD03-2697 for the first<br />

4. 10 m representing the last 20 000 years. Nonetheless, sediment levels of 1<br />

cm thickness at 249 and 275 cm show the presence of Zoophycos borrows<br />

and therefore they were not included in this work. This benthic burrowing<br />

organism can reach a vertical extension of more than 1 m (Löwemark and<br />

Werner, 2001; Leuschner et al., 2002) by doing a downward helicoidally<br />

movement along a centred vertical axial shaft producing several more or less<br />

horizontal burrows (Löwemark et al., 2004). Furthermore the successive up to<br />

downward displacement of young material can produce important changes<br />

in the original sediment when compared with the adjacent host sediment.<br />

This is particularly true for the Iberian margin where foraminifera tests from<br />

Zoophycos spreiten show younger ages of about 1 000 to 2 500 yr BP than the<br />

adjacent sediment (Löwemark and Werner, 2001).<br />

4. 3. 1 Stratigraphy and age model<br />

The age model of MD03-2697 has been established by using 10 AMS<br />

14 C dates from this core together with 1 AMS 14 C date from the twin core<br />

MD99-2331 (42° 09’ 00 N, 09° 40’ 90 W; 2110 m depth) (Fig. IV.1 and Tab. IV.1).<br />

These 11 dated levels were obtained at “Laboratoire de Mesure du Carbone<br />

14” (LMC, in Saclay; France) and at Beta Analytic Inc (Beta; USA) in<br />

199<br />

F. Naughton, 2007


monospecific samples with maxima of Globigerina bulloides or<br />

Neogloboquadrina pachyderma (s.) abundances.<br />

AMS 14 C dates were calibrated using CALIB Rev 5.0 program and the<br />

"global" marine calibration dataset (marine 04.14c) (Stuiver and Reimer, 1993;<br />

Hughen et al., 2004; Stuiver et al., 2005). We used the 95.4% (2 sigma)<br />

confidence intervals and their relative areas under the probability curve as<br />

well as the median probability of the probability distribution (Telford et al.,<br />

2004) as suggested by Stuiver et al. (2005).<br />

Some levels, such as 249 and 275 cm, were automatically rejected for<br />

the age model reconstruction because they reflect contaminating levels of<br />

young material introduced when the Zoophycos burrow was active<br />

(Leuschner et al., 2002). Level 200 cm has also been considered too young<br />

when compared with the age limits proposed by Elliot et al. (2002) for the<br />

Heinrich 1 (H1) event. Therefore we have introduced one AMS 14 C date from<br />

the twin core MD99-2331 for delimitate the end of H1 event in the MD03-2697.<br />

We have also rejected level 150 cm because it has been considered too old<br />

when compared with pollen results which place the beginning of Younger<br />

Dryas event at 160 cm of depth.<br />

Lab code<br />

Beta-2131134<br />

Beta-2131135<br />

LMC14-003257<br />

Beta-2131136<br />

Beta-2131137<br />

LMC14-003258<br />

LMC14-003259<br />

Beta-213138<br />

Beta-213139<br />

LMC14-003260<br />

LMC14-001231<br />

Core- depth<br />

(cm)<br />

MD03 2697-<br />

20<br />

MD03 2697-<br />

40<br />

MD03 2697-<br />

70<br />

MD03 2697-<br />

80<br />

MD03 2697-<br />

110<br />

MD03 2697-<br />

150<br />

MD03 2697-<br />

200<br />

MD03 2697-<br />

249<br />

MD03 2697-<br />

275<br />

MD03 2697-<br />

340<br />

MD99 2331-<br />

200<br />

Material<br />

Conv.<br />

AMS 14 C<br />

age BP<br />

200<br />

Conv.<br />

AMS 14 C<br />

age BP<br />

(-400 yr)<br />

error<br />

95.4 % (2σ)<br />

Cal BP age ranges<br />

Cal BP<br />

age<br />

median<br />

probability<br />

G. bulloides 2880 2480 40 2501 BP:2739 BP 2656<br />

G. bulloides 4760 4360 40 4866 BP:5198 BP 5008<br />

G. bulloides 7435 7035 50 7783 BP:7998 BP 7895<br />

G. bulloides 7470 7070 40 7835 BP:8014 BP 7930 a<br />

G. bulloides 9940 9540 40 10705 BP:11084 BP 10896<br />

G. bulloides 11920 11520 60 13233 BP:13486 BP 13353 b<br />

G. bulloides 12520 12120 70 13805 BP:14127 BP 13965 b<br />

N. pachy (s.) 13240 12840 50 14934 BP:15424 BP 15152 a<br />

N. pachy (s.) 13980 13580 60 15779 BP:16569 BP 16151 a<br />

G. bulloides 15410 15010 70 18044 BP:18596 BP 18313<br />

N. pachy (s.) 13640 13240 80 15303 BP:16099 BP 15679<br />

Tab. IV.1| Radiocarbon ages of MD03-2697 deep-sea core and one level from the twin core (MD99-2331). a<br />

Radiocarbon dates too young or too old and b Not acceptable dating (bioturbated layers).<br />

F. Naughton, 2007


4. 3. 2 Pollen analysis<br />

54 samples covering the first 4.10 m of the MD03-2697 core were used<br />

for pollen analysis, with a sample spacing of 1 to 10 cm. Pollen concentration<br />

of two 1cm-thick samples (at 40 and 60 cm) was weak and therefore they<br />

have been considered as sterile. Sample preparation technique follows de<br />

Vernal et al. (1996) modified at the UMR CNRS 5805 EPOC (Desprat, 2005).<br />

After chemical treatments (cold 10%, 25% and 50% HCl as well as cold<br />

40% and 70% HF) the samples were sieved through a 10 µm nylon mesh<br />

screen (Heusser and Stock, 1984) and mounted in bidistillate glycerine. Pollen<br />

and spores were counted using a Zeiss Axioscope light microscope at x550<br />

and x1250 (oil immersion) magnifications. A minimum of 100 pollen grains<br />

excluding the over-represented Pinus grains in the Iberian margin (Naughton<br />

et al., 2006), have been counted. At least 20 taxa and 100 grains of the<br />

added exotic Lycopodium fern were also counted. Pollen percentages were<br />

calculated based on the main pollen sum which excludes Pinus, aquatic<br />

plants, spores, indeterminate and unknown pollen grains.<br />

Pollen analysis from 0 to 2 m of the MD03-2697 deep sea core have<br />

been previously published in Naughton et al. (2006). This previous work also<br />

shows that pollen grains included in marine cores from western Iberia margin<br />

reflect an integrated image of the regional vegetation of the adjacent<br />

continent, being therefore an important tool for establishing a direct sea-land<br />

correlation in this region.<br />

4. 3. 3 Marine proxy analyses<br />

assemblages<br />

4. 3. 3. 1 Ice rafted detritus (IRD) and planktonic foraminiferal<br />

60 and 43 levels from MD03-2697 deep sea core, with 1 to 10 cm of<br />

sample spacing have been used for IRD and planktonic foraminifera<br />

semiquantitative analysis, respectively. IRD and polar foraminifera analyses<br />

from 0 to 2 m of core depth have been previously published in Naughton et<br />

al. (2006).<br />

Samples were washed with distilled water and wet sieved trough a 150<br />

μm mesh screen. Planktonic foraminifera were grouped into three main<br />

201<br />

F. Naughton, 2007


ioclimatic assemblages: polar (Neogloboquadrina pachyderma sinistral),<br />

subpolar (Neogloboquadrina pachyderma dextra, Globigerina bulloides and<br />

Turborotalia quinqueloba) and warm (which includes temperate/cold,<br />

subtropical, warm subtropical and tropical, i.e. Globorotalia scitula, G.inflata,<br />

G.hirsuta, G.truncatulinoides, G.crassaformis, Globigerinita glutinata,<br />

Globigerina falconensis, G.calida,G.rubescens, G.digitata, Hastigerina<br />

aequilateralis, Orbulina universa, Globigerinoides ruber) (e.g. Bé, 1977; Ottens,<br />

1991; Duprat, 1983).<br />

Winter (February) and summer (August) sea-surface temperatures (SST)<br />

were estimated by using the modern analogue technique transfer function<br />

from the database of Pflaumann et al. (1996) improved by E. Cortijo<br />

(Laboratoire des Sciences du Climat et de l’Environnement-LSCE, Gif-sur-<br />

Yvette, France) and J. Duprat (UMR CNRS 5805 EPOC) on planktonic<br />

foraminifera assemblages.<br />

4. 3. 3. 2 Isotopic analyses<br />

Several isotopic analyses were carried out on the first 4.10 m of MD03-<br />

2697 deep sea core. A total of 59 oxygen isotopic measurements were<br />

carried out on Globigerina bulloides planktonic foraminifera and 35 on<br />

Cibicides wuellerstorfi benthic foraminifera with a sample spacing of 1 to 10<br />

cm and 5 to 10 cm, respectively, at the Laboratoire des Sciences du Climat et<br />

de l’Environnement (LSCE), Gif-sur-Yvette, France. Planktonic oxygen isotopic<br />

data from the first 2 m of MD03-2697 record has been previously published in<br />

Naughton et al. (2006).<br />

Each specimen has been picked up within the 250–315 µm fraction<br />

and cleaned with distilled water. The preparation of each aliquot (4–10<br />

specimens with 80 μg of weight) has been carried out in the Micromass<br />

Multiprep autosampler by using an individual acid attack. The CO2 gas<br />

extracted has been analysed against NBS 19 standard, taken as an<br />

international reference standard. Planktonic and benthic isotopic analysis<br />

were performed using a delta plus Finnigan at the LSCE. The mean external<br />

reproducibility of powdered carbonate standards is ±0.05‰ for oxygen.<br />

Results from oxygen isotopic analysis are presented versus PDB.<br />

202<br />

F. Naughton, 2007


4. 4 Vegetation and climate changes in north-western Iberia and<br />

adjacent margin during the last deglaciation<br />

High resolution continental (pollen) (Fig. IV.2) and marine proxies<br />

(including Ice rafted detritus-IRD, planktonic foraminiferal assemblages, sea<br />

surface temperature-SST and planktonic oxygen isotopic) were analysed<br />

together with benthic oxygen isotopes (ice-volume indicator) from MD03-2697<br />

deep-sea core. This has allowed the detection of sub-orbital climate<br />

variability during the last deglaciation in north-western Iberia (Fig. IV.3):<br />

Fig. IV.2 | Pollen diagram of MD03-2697 deep-sea core against depth. From left to right: calibrated ages<br />

and percentages of selected pollen taxa. The stratigraphy is based on previous work by (Naughton et al.,<br />

2006) where the Oldest Dryas represents the continental counterpart of Heinrich 1 event in the ocean.<br />

203<br />

F. Naughton, 2007


Fig. IV.3 | Multi-proxy record of MD03-2697 against calibrated ages. From bottom to top: percentages of<br />

selected pollen taxa (trees: Betula, Corylus, deciduous Quercus, Pinus; Calluna and semi-desert plants:<br />

Artemisia, Chenopodiaceae and Ephedra); Ice-rafted detritus (IRD); planktonic foraminifera associations;<br />

Sea Surface Temperature (SST) estimates; δ 18O of planktonic and benthic foraminifera and Greenland<br />

temperatures (Sánchez Goñi et al., in prep.).<br />

204<br />

F. Naughton, 2007


4. 4. 1 The end of the Last Glacial maximum (LGM)<br />

North-western Iberia was dominated by herbaceous plants with Pinus<br />

at the end of the late pleniglacial period (Fig. IV.2). Previous works on the twin<br />

MD99-2331 deep sea core, clearly show that this period represents the end of<br />

the last glacial maximum (LGM) in the ocean (Naughton et al., 2006). The<br />

expansion of heath communities (Ericaceae more Calluna) recorded in the<br />

MD03-2697 deep sea core (Fig. IV.2) indicates moist conditions at that time as<br />

it has been previously detected by the twin core MD99-2331 (Naughton et al.,<br />

2006) and further south in the SU81-18 deep sea record (Turon et al., 2003).<br />

This increase of moisture conditions in western Iberia has probably been<br />

favoured by a more vigorous Meridional Overturning Circulation (MOC)<br />

(Naughton et al., in prep.) which is estimated to be reduced by less than 30-<br />

40% during the LGM period in comparison with the strongest reduction during<br />

the Heinrich event 1 (H1) (McManus et al., 2004).<br />

Sea surface conditions were warm, 11° C in winter and 16° during<br />

summer time, corroborating previous SST reconstructions from mid-latitudes<br />

North Atlantic deep-sea cores (Lebreiro et al., 1997; Cayre et al., 1999; Bard et<br />

al., 2000; Chapman et al., 2000, Pailler and Bard, 2002; de Abreu et al., 2003;<br />

de Vernal et al., 2005; Morey et al., 2005; Naughton et al., in prep.) (Fig. IV.3).<br />

Although sea surface conditions were warmer than the previous and<br />

subsequent Heinrich events (H2 and H1) and MOC was more vigorous,<br />

temperate trees did not expand during the LGM interval (Naughton et al., in<br />

prep.), though their continuous presence (~5%) suggests that scattered<br />

woodlands survived in north-western Iberia at this time (Naughton et al.,<br />

2006). Increasing albedo, high seasonality and low CO2 concentrations are<br />

pointed as the major forcing mechanisms for preventing temperate tree<br />

expansion in Iberia during this period (Naughton et al., in prep.).<br />

4. 4. 2 The Heinrich 1 (H1)<br />

MD03-2697 deep sea core shows a complex pattern within H1 event<br />

(Fig. IV.3). H1 is marked by a drop in sea surface temperatures (expansion of<br />

polar planktonic foraminifera, a decrease of summer and winter SST and<br />

increase of planktonic δ 18 O values) that started at around 18 300 cal yr BP<br />

lasting until 15 700 cal yr BP (Fig. IV.3). Although IRD presence is recorded in<br />

205<br />

F. Naughton, 2007


MD03-2697 deep sea core at around 18 000 cal yr BP, the maximal arrival of<br />

Icebergs in north-western Iberian margin occurred after 16 400 cal yr BP (Fig.<br />

IV.3).<br />

At the same time, north-western Iberia vegetation reflects two distinct<br />

phases. The first phase is represented by a drastic decline of Pinus forest,<br />

testifying an important atmospheric cooling in this region (Figs. IV.2 and IV.3).<br />

Paralleling this atmospheric cooling, a second expansion of Ericaceae and<br />

Calluna, suggests an increase of moisture (Fig. IV.2), showing that north-<br />

western Iberia has been influenced by different climatic conditions than those<br />

characterizing the most continental central and eastern high-altitude sites of<br />

Iberia, which were affected by dryer climate (see Naughton et al., 2006). In<br />

the second phase, Pinus forest expands and semi-desert plants increase<br />

gradually until getting a maximum synchronously with the IRD highest values<br />

(Figs. IV.2 and IV.3).<br />

These results are in agreement with those obtained for the late MIS 3<br />

and during MIS 2 on the twin core MD99-2331, which suggests that each of<br />

the typical Heinrich events (H4, H2 and H1) are marked by a first oceanic and<br />

atmospheric extreme cold and wet episode followed by a dry and cool one<br />

on land and in the ocean (Naughton et al., 2006; in prep.).<br />

Recently, a study about European rivers reactivation during the last<br />

deglaciation in the north-western French margin also shows a humid phase<br />

preceding IRD maximal arrival (Ménot et al., 2006) (Fig. IV.1). However, this<br />

phase has been considered as an episode preceding H1 event and, it has<br />

been erroneously correlated with a warm continental phase. Furthermore,<br />

previous studies on the same core, MD95-2002, have clearly showed that cold<br />

sea surface conditions preceded IRD maximal arrival into this region (Zaragosi<br />

et al., 2001; Auffret et al., 2002).<br />

This strong cooling has been probably triggered by the MOC shutdown<br />

followed by rapid oceanic and atmospheric reorganizations favouring the<br />

transfer of atmospheric cold conditions to this region (Naughton et al., in<br />

prep.). Because MOC shutdown prevents the moisture transfer from the North<br />

Atlantic to Iberia, a prevailing negative mode of the NAO-like has been<br />

proposed as the main responsible for the increase of wet conditions in north-<br />

western Iberia during the first phase of H1 (Naughton et al., in prep.). At<br />

206<br />

F. Naughton, 2007


present, these prevailing atmospheric conditions could also trigger an<br />

increase of moisture in north-western of France and generate SST increase in<br />

north-western Atlantic (at latitudes northern than 45°-50°N) (Wanner et al.,<br />

2001). This likely favours icebergs melting in the IRD belt and therefore, the<br />

substantial deposition of IRD and the consequent cooling of the surface<br />

waters in that region during the first phase of H1. So it prevents Laurentide<br />

iceberg displacement to regions located far away from that belt, including<br />

Iberian margin and Bay of Biscay. Prevailing negative NAO-like index also<br />

favours the decrease of SST along north-western Iberian margin until the<br />

Greater North Sea, passing by north-western French margin, and the slight<br />

warmth over Greenland as shown by temperature estimates (Fig. IV.3).<br />

On the contrary, during the second phase of H1 north-western Iberia<br />

has been probably influenced by prevailing positive NAO-like index<br />

(Naughton et al., in prep.). The strengthening and the northward<br />

displacement of the westerlies, caused by the increase of the pressure<br />

gradient between Açores and Iceland, favours nowadays the increase of<br />

dryness in south-western Europe up to 47 °N and warm sea surface conditions<br />

at mid-latitudes of the North Atlantic (20°- 40°N) (Wanner et al., 2001). This<br />

atmospheric situation could lead the southward migration of the icebergs<br />

until the Iberian margin during the second part of H1. During this episode of<br />

maximum arrival of icebergs into the mid-latitudes of the North Atlantic sea<br />

surface conditions off north-western Iberia were, paradoxically, slightly<br />

warmer than the precedent phase. This could be explained by the influence<br />

of prevailing positive NAO-like index along north-western Iberian margin and<br />

Greater North Sea branch which override the cooling effect of the maximal,<br />

although weak (Downswell et al., 1995), arrival of icebergs on this margin. The<br />

slight warming of SST is not observed in north-western French margin probably<br />

because the high quantity of icebergs from the British Ice Sheet arriving to this<br />

margin. The lightening of planktonic foraminifera δ 18 O observed in north-<br />

western Iberian margin within the second phase of H1 (starting at around 16<br />

500 cal yr BP) has also been detected in other mid-latitudes deep-sea cores<br />

such as OCE326-GGC5 (33° 42’N, 57°35’W) (McManus et al., 2004). This<br />

decrease of δ 18 O values is also corroborated by the slight warming at the<br />

North Atlantic mid-latitudes during the second phase of H1 (Fig. IV.3).<br />

207<br />

F. Naughton, 2007


4. 4. 3 The Bölling-Alleröd (B-A)<br />

Deciduous Quercus expansion and herbaceous communities decline<br />

marks the warm Bölling-Alleröd (B-A) event in north-western Iberia (Fig. IV.2).<br />

At the same time, Iberian margin warmed as testified by the decrease of<br />

polar planktonic foraminifera and the increase of warm and sub-polar<br />

species. The decrease of the planktonic foraminifera δ 18 O values of about 1‰<br />

PDB also suggest, beyond a decrease in salinity, an increase of sea surface<br />

temperature during this period (Fig. IV.3).<br />

The maximum expansion of deciduous Quercus woodlans, in north-<br />

western Iberia, occurred synchronously with the well known meltwater pulse<br />

1A (MWP 1A) (Fairbanks, 1989; Bard et al., 1996; 1990) and with a peak in the<br />

Greenland temperature (Sánchez Goñi et al., in prep.), at around 14 000 cal<br />

yr BP (Fig. IV.3).<br />

It has been proposed by McManus et al. (2004) that the abrupt<br />

resumption of the MOC at the beginning of B-A event together with the<br />

increase of high latitude summer insolation accelerate the Laurentide Ice<br />

sheet melting and trigger the MWP 1A. This ice sheet melting increased the<br />

input of freshwater in the ocean. However, this has not produced either a<br />

reduction or a shutdown of the conveyor belt as it would be expected.<br />

Few hypotheses have been pointed out for explaining this controversial<br />

situation revealed by the North Atlantic deep-sea cores. Clark et al. (2002)<br />

proposed that meltwater pulses during the B-A were primarily originated from<br />

Antarctica with only a small contribution of the Northern Hemisphere Ice<br />

sheet melting while McManus et al. (2004) suggest that meltwater from the<br />

Laurentide experienced substantial mixing with sea water before reaching<br />

the zones of deep convection or that the site of deep-water production have<br />

migrated further north away from the influence of the Laurentide meltwater.<br />

Our data has shown that the timing of maximum warming during the B-<br />

A period at mid-latitudes of north-western Iberia together with maxima<br />

temperatures over Greenland coincides with that of MWP 1A. This suggests<br />

that an atmospheric warming in the Northern Hemisphere triggered the MWP<br />

1A event rather than the Southern Hemisphere.<br />

208<br />

F. Naughton, 2007


Fig. IV.4 | Long-term and small-scalle pattern of vegetation changes in north-western Iberia. From bottom<br />

to top: benthic foraminifera δ 18O; percentages of temperate trees includes (Acer, Alnus, Betula, Corylus,<br />

Cupressaceae, deciduous Quercus, Fagus, Fraxinus excelsior-type, Pinus, Quercus ilex-type, Salix, Tilia and<br />

Ulmus); percentages of Pinus, percentages of deciduous Quercus and summer insolation at 45° N (after<br />

Berger, 1978).<br />

209<br />

F. Naughton, 2007


4. 4. 4 The Younger Dryas (YD)<br />

Deciduous Quercus forest reduction together with the re-expansion of<br />

semi-desert and pioneer species (Betula) mark returning cold conditions<br />

which characterises the Younger Dryas event in north-western Iberia (Figs. IV.2<br />

and IV.3). Synchronously with the atmospheric cooling at mid-latitudes and<br />

over Greenland, the ocean experienced a decrease of surface temperatures<br />

as suggested by the increase of planktonic foraminifera δ 18 O values (Fig.<br />

IV.3).<br />

These cold conditions are weaker than those characterizing H1 event (Fig.<br />

IV.3).<br />

Both the YD and the H1 are associated with substantial meltwater<br />

pulses from Laurentide ice-sheet (Andrews et al., 1995). However, the YD is<br />

only related with the reduction of the MOC while H1 with its shutdown<br />

(McManus et al., 2004). The reduction of the MOC together with the mid-<br />

latitude maximum increase of summer insolation (Fig. IV.4) could explain the<br />

decrease rather than the complete decline of deciduous Quercus forest in<br />

the mid- and low-altitudes of north-western Iberia. Because MOC was<br />

activated, though in a reduced way, a weak quantity of moisture was still<br />

transferred from the tropics to the high-latitudes of the Northern Hemisphere.<br />

Nevertheless, Iberia (this study and see Naughton et al., 2006) and other<br />

European regions (Watts, 1980; Lotter et al., 1992; Reille et al., 2000) were<br />

influenced by fairly dry conditions during this period.<br />

At the same time, the increase of trade wind strength in the Cariaco<br />

Basin (Hughen et al., 1996) and the decrease of riverine run-off from adjacent<br />

land masses (Peterson et al., 2000) testifies the southward migration of the<br />

Intertropical Convergence Zone (ITCZ) which is supported by climate<br />

modelling (Dahl et al., 2005).<br />

Because shifts on the ITCZ are intimately connected with the NAO by<br />

reorganizations of the Hadley cells (Marshall et al., 2001) one could think that<br />

the southward migration of the ITCZ is accompanied by the increase of<br />

pressure gradient between Açores and Iceland. This favoured the<br />

intensification and northward displacement of the westerlies over Europe and<br />

therefore an increase of dryness conditions in this region.<br />

210<br />

F. Naughton, 2007


4. 4. 5 The Holocene<br />

The beginning of the Holocene is marked by the expansion of<br />

deciduous Quercus forest and herbaceous communities decline reflecting an<br />

increase of temperatures in north-western Iberia (Fig. IV.2). This is<br />

contemporaneous with a substantial warming over Greenland (Fig. IV.3) and<br />

with mid-latitude high summer insolation (Fig. IV.4).<br />

In the ocean, SST increases as showed by the decrease of planktonic<br />

foraminifera δ 18 O values and the decrease of polar planktonic foraminifera<br />

percentages (Fig. IV.3). MOC became more active as reflected by 231 Pa/ 230 Th<br />

data from mid-latitude OCE326-GGC5 deep-sea core (McManus et al., 2004)<br />

(Fig. IV.1).<br />

The maximum spread of deciduous Quercus in north-western Iberia<br />

representing the Holocene Thermal Maximum (HTM) in this region occurred in<br />

the early Holocene before the 8.2 Ky event (Figs. IV.3 and IV.4). This episode<br />

of rapid increase of atmospheric temperatures documented by the<br />

expansion of deciduous Quercus forest in the Iberia Peninsula has also been<br />

detected in other marine and continental pollen sequences (see Naughton<br />

et al., 2006). Nonetheless, the HTM has been identified later, between 8000<br />

and 4500 cal yr BP, in northern Europe (Seppä and Poska, 2004; Seppä et al.,<br />

2005). This suggests that the Holocene maximum warming on land occurred<br />

at different times depending of places as previously suggested for the ocean<br />

realm by Kaufman et al. (2004).<br />

Following the HTM in north-western Iberia, temperate forest, including<br />

Pinus, gradually decreases and parallels the long-term trend of northern mid-<br />

latitudes summer insolation (Figs. IV.3 and IV.4), suggesting that vegetation<br />

has responded to the orbitally-induced long-term cooling that characterises<br />

this interglacial period. This pattern has also been observed further north in<br />

north-western France after the 8.2 Ky event (Naughton et al. submitted). This<br />

suggests that long-term vegetation change during the mid- and late-<br />

Holocene in both regions has been triggered by orbital forcing rather than by<br />

human impact, agreeing with what has been previously proposed by Magri<br />

(1995).<br />

211<br />

F. Naughton, 2007


4. 4. 5. 1 The 8.2 k yr event<br />

Superimposed to the long-term cooling trend of the Holocene, a<br />

sudden decline of temperate forest, mainly involving deciduous Quercus and<br />

Corylus trees, together with a slight increase of planktonic foraminifera δ 18 O<br />

values and a decrease of SST, marks the 8.2 Ky event in north-western Iberian<br />

margin and adjacent landmasses (Fig. IV.3). The low resolution multi-proxy<br />

analysis for the Holocene interval in our core MD03-2697 precludes, however,<br />

the identification of the multi-centennial cooling event between ~8600 and<br />

8000 cal yr BP noticed by Rohling and Pälike, 2005 based in several archives<br />

around the world and detected in the eastern North Atlantic ocean and<br />

borderlands by Ellison et al. 2006 and Naughton et al. (submitted),<br />

respectively.<br />

Corylus woodlands decline at the time of the well known 8.2 ky cooling<br />

event has been detected in several pollen sequences from central (Tinner<br />

and Lotter, 2001; 2006) and northern Europe (Seppä and Poska, 2004; Veski et<br />

al., 2004; Seppä et al., 2005) and in a marine pollen record from the French<br />

margin (Naughton et al. submitted). This event was triggered by the<br />

culminating of successive episodes that started 400 to 500 yr before this<br />

drastic event. Laurentide Ice sheet decay and following catastrophic<br />

outburst episodes from the lakes Agassiz and Ojibway into the Hudson Bay<br />

(Barber et al., 1999; Teller et al., 2002; Clarke et al., 2004) allow the<br />

introduction of freshwater into the North Atlantic (Alley et al., 1997; Clark et<br />

al., 2001) leading to a substantial decrease of SST (Knudsen et al., 2004;<br />

Keigwin et al., 2005, Ellison et al., 2006) and a gradual reduction of the flow<br />

speed of the Iceland-Scotland Overflow water (ISOW), a component of the<br />

NADW which peaked at around 8 290 yr cal BP (Ellison et al., 2006). The<br />

impact of this drastic last event, associated with the maximum reduction of<br />

the flow speed of the ISOW is synchronous with a temperature drop over<br />

Greenland and Europe (Fig. IV.3).<br />

212<br />

F. Naughton, 2007


4. 5 Conclusion<br />

The last deglacial period is marked in north-western Iberian margin and<br />

adjacent landmasses by millennial scale climate variability. The end of the<br />

Last Glacial Maximum (LGM) is marked in north-western Iberia by relatively<br />

cold and humid conditions while mid-latitudes of the North Atlantic were<br />

warm. We explain this apparent contradiction as the interplay of a more<br />

vigorous Meridional Overturning Circulation (MOC) with the increasing<br />

albedo, high seasonality and/or atmospheric CO2 drop.<br />

Heinrich 1 is marked by a complex pattern off Iberia and in the<br />

adjacent landmasses and represented by two main phases. The first one is<br />

characterised by extremely cold sea surface and atmospheric conditions, a<br />

low quantity of IRD and continental moisture. The second phase shows a high<br />

quantity of IRD associated with continental dryness and with a slight warming<br />

in SST and in the atmosphere although these temperatures were relatively less<br />

cold. The extremely cold phase in north-western Iberia was probably<br />

triggered by the MOC shutdown followed by ocean-atmosphere rapid<br />

reorganization. We explain the shift between wet and dry conditions as the<br />

result of changes in prevailing negative and positive NAO-like indexes,<br />

respectively. This last mechanism can also explain the fact that the maximal<br />

arrival of IRD is asynchronous in the different regions of the North-Atlantic<br />

ocean. The Bölling-Alleröd (B-A) in this region is marked by a substantial<br />

atmospheric and oceanic warming favoured by both the increase of mid-<br />

latitude summer insolation and the strengthening of the MOC. The maximum<br />

of deciduous Quercus forest expansion reflects the highest warming<br />

conditions of north-western Iberia at 14 000 cal yr BP. This expansion is<br />

synchronous with the Greenland temperature peak of GIS (Greenland<br />

Interstadial)1 and with the Meltwater Pulse 1A (MWP 1A) suggesting that this<br />

drastic melting episode must have been initiated in the Northern Hemisphere<br />

rather that in the Southern Hemisphere.<br />

A returning to glacial conditions characterises the Younger Dryas (YD)<br />

event in north-western Iberian margin and in the adjacent continent. MOC<br />

reduction, instead of shutdown, and the increase of northern mid-latitude<br />

213<br />

F. Naughton, 2007


summer insolation favoured the decrease rather than the complete decline<br />

of deciduous Quercus forest in north-western Iberia. Besides the cooling,<br />

Iberian Peninsula has been affected by substantial dryness which was<br />

probably the result of prevailing positive NAO-like index. Oceanic and<br />

continental warming in the early Holocene parallels Greenland temperature<br />

curve and define the Holocene Thermal Maximum between 11 700 and 8 200<br />

cal yr BP. Following this, the decrease of deciduous Quercus and Corylus<br />

woodlands together with an oceanic cooling marks the 8.2 Ky event in this<br />

region in response to the culmination of the successive episodes of the<br />

Laurentide Ice sheet decay which enhanced the cooling over Greenland<br />

and Europe.<br />

The long-term trend decline of the temperate forest in north-western<br />

Iberia during the mid- and late-Holocene parallels the gradual decrease of<br />

northern mid-latitude summer insolation revealing that this vegetation pattern<br />

is orbitally-induced rather than due to the human activity.<br />

References:<br />

Alcara Ariza, F. Asensi Marfil, A., de Bolos y Capdevilla, O., Costa Tales, M., Arco<br />

Aguilar, M., Diaz Gonzales, T.E., Diez Garretas, B., Fernandez Prieto, J.A., Fernandez<br />

Gonzales, F., Izco Sevillando, J., Loidi Arregui, J., Martinez Parras, J.M., Navarro Andres,<br />

F., Ninot I Sugranes, J.M., Peinado Lorca, M., Rivas Martinez, S., Sanchez Mata, D.,<br />

Valle Guitierrez, C., Vigo I Bonada, J. and Wildpret de la Torre, W., 1987. La<br />

vegetación de España. Collection Aula Abierta. Universidad de Alcala de Henares,<br />

544 pp.<br />

Alley, R.B., Mayewski, P.A., Sowers, T., Stuiver, M., Taylor, K.C. and Clark, P.U., 1997.<br />

Holocene climatic instability: a prominent, widespread event 8200 yr ago. Geology,<br />

25: 483-486.<br />

Andrews, J.T., Bond, G., Jennings, A.E., Kerwin, M., Kirby, M., MacLean, B., Manley, W.<br />

and Miller, G.H., 1995. A Heinrich-like event, H-0 (DC-0); Source(s) for detrital<br />

carbonate in the North Atlantic during the Younger Dryas chronozone.<br />

Paleoceanography, 10: 943-952.<br />

Andrews, J.T., Keigwin, L., Hall, F. and Jennings, A.E., 1999. Abrupt deglaciation events<br />

and Holocene palaeoceanography from high-resolution cores, Cartwright Saddle,<br />

Labrador Shelf, Canada. Journal of Quaternary Science, 14: 383-397.<br />

Arz, H.W., Patzold, J., Wefer, G., 1999. The deglacial history of the western tropical<br />

Atlantic as inferred from high resolution stable isotope records off northeastern Brazil.<br />

Earth and Planetary Science Letters, 167: 105–117.<br />

Atlas Nacional de España, 1992. Climatologia. Ministerio de Obras Publicas y<br />

Transportes. Direccion General del Instituto Geografico Nacional, Madrid.<br />

214<br />

F. Naughton, 2007


Auffret, G., Zaragosi, S., Dennielou, B., Cortijo, E., Van Rooij, D., Grousset, F., Pujol, C.,<br />

Eynaud F. and Siegert, M., 2002. Terrigenous fluxes at the Celtic margin during the last<br />

glacial cycle. Marine Geology, 188: 79-108.<br />

Barber, D.C., Dyke, A., Hillaire-Marcel, C., Jennings, A.E., Andrews, J.T., Kerwin, M.W.,<br />

Bilodeau, G., McNeely, R., Southon, J., Morehead, M.D. and Gagnon, J.-M., 1999.<br />

Forcing of the cold event of 8,200 years ago by catastrophic drainage of Laurentide<br />

lakes. Nature, 400: 344-348.<br />

Bard, E., Hamelin, B., Arnold, M., Montaggioni, L.F., Cabioch, G., Faure, G. and<br />

Rougerie, F., 1996. Deglacial sea-level record from Tahiti corals and the timing of<br />

global meltwater discharge. Nature, 382: 241-244.<br />

Bard, E., Hamelin, B., Fairbanks, R.G. and Zindler, A., 1990. Calibration of the 14 C<br />

timescale over the past 30,000 years using mass spectrometric U-Th ages from<br />

Barbados corals. Nature, 345: 405-410.<br />

Bard, E., Rostek, F., Turon, J.L. and Gendreau, S., 2000. Hydrological impact of Heinrich<br />

events in the subtropical northeast Atlantic. Science, 289: 1321-1324.<br />

Bé, A.W.H., 1977. An ecological, zoological and taxonomic review of recent<br />

planktonic Foraminifera. Oceanic Micropaleontology, Volume 1: London, Academic<br />

Press Inc., 1-100 pp.<br />

Berger, A., 1978. Long-term variations of daily insolation and Quaternary climatic<br />

changes. Journal of Atmospheric Sciences, 35: 2362-2367.<br />

Blunier, T. and Brook, E.J., 2001. Timing of Millennial-Scale Climate Change in<br />

Antarctica and Greenland During the Last Glacial Period. Science, 291: 109-112.<br />

Bond, G. and Lotti, R., 1995. Iceberg discharges into the North Atlantic on millennial<br />

time scales during the last glaciation. Science, 267: 1005-1010.<br />

Bond, G., Broecker, W., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani,<br />

G., 1993. Correlations between climate records from North Atlantic sediments and<br />

Greenland ice. Nature, 365: 143-147.<br />

Cayre, O., Lancelot, Y. and Vincent, E., 1999. Paleoceanographic reconstructions<br />

from plaktonic foraminifera off the Iberian Margin: Temperature, salinity, and Heinrich<br />

events. Paleoceanography, 14: 384-396.<br />

Chapman, M. R., Shackleton, N. J. and Duplessy, J. C., 2000. Sea surface temperature<br />

variability during the last glacial-interglacial cycle: assessing the magnitude and<br />

pattern of climate change in the North Atlantic. Palaeogeography-<br />

Palaeoclimatology- Palaeoecology, 157: 1 -25.<br />

Clark, P. U., Mitrovica, J.X., Milne, G.A. and Tamisiea, M.E., 2002. Sea-Level<br />

Fingerprinting as a Direct Test for the Source of Global Meltwater Pulse IA. Science,<br />

295: 2438-2441.<br />

Clark, P.U., Marshall, S.J., Clarke, G.K.C., Hostetler, S.W., Licciardi, J.M. and Teller, J.T.,<br />

2001. Freshwater forcing of abrupt climate change during the last glaciation.<br />

Science, 293: 283-287.<br />

215<br />

F. Naughton, 2007


Clarke, G.K.C., Leverington, D.W., Teller, J.T. and Dyke, A.S., 2004. Paleohydraulics of<br />

the last outburst flood from glacial Lake Agassiz and the 8200 BP cold event.<br />

Quaternary Science Reviews, 23: 389–407.<br />

Dahl, K.A., Broccoli, A.J. and Stouffer, R.J., 2005. Assessing the role of North Atlantic<br />

freshwater forcing in millennial scale climate variability: a tropical Atlantic<br />

perspective. Climate Dynamics, 24: 325-346.<br />

de Abreu, L., Shackleton, N.J., Schonfeld, J., Hall, M. and Chapman, M., 2003.<br />

Millennial-scale oceanic climate variability off the Western Iberian margin during the<br />

last two glacial periods. Marine Geology, 196: 1-20.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3: 16-27.<br />

de Vernal, A., Eynaud, F., Henry, M., Hillaire-Marcel, C., Londeix, L., Mangin, S.,<br />

Matthiessen, J., Marret, F., Radi, T., Rochon, A., Solignac, S. and Turon, J.-L., 2005.<br />

Reconstruction of sea –surface conditions at middle to high latitudes of the Northern<br />

Hemisphere during the Last Glacial Maximum (LGM) based on dinoflagellate cyst<br />

assemblages. Quaternary Science Reviews, 24: 897-924.<br />

Desprat, S., 2005. Réponses climatiques marines et continentales du SW de l'Europe<br />

lors des derniers interglaciaires et des entrées en glaciation. PhD Thesis, Bordeaux 1<br />

University, Bordeaux (France), 282 pp.<br />

Dias, J.M.A., Gonzalez, R., Garcia, C. and Diaz del Rio, V., 2002. Sediment distribution<br />

patterns on the Galicia-Minho continental shelf. Progress in Oceanography, 52: 215-<br />

231.<br />

Dowdeswell, J.A., Maslin, M.A., J.T, A. and McCave, I.N., 1995. Iceberg production,<br />

debris rafting, and the extent and thickness of Heinrich layers (H-1, H-2) in North<br />

Atlantic sediments. Geology, 33: 301-304.<br />

Duprat, J., 1983. Les foraminifères planctoniques du Quaternaire Terminal d'un<br />

domaine pericontinental (Golfe de Gascogne, côtes Ouest-Ibériques, Mer<br />

d'Alboran): Ecologie-Biostratigraphie. Bulletin de l'Institut Géologie du Bassin d'<br />

Aquitaine, 33: pp. 71-150.<br />

Elliot, M., Labeyrie, L. and Duplessy, J.-C., 2002. Changes in North Atlantic deep-water<br />

formation associated with the Dansgaard-Oeschger temperature oscillations (60-10<br />

ka). Quaternary Science Reviews, 21: 1153-1165.<br />

Ellison, C.R.W., Chapman, M.R. and Hall, I.R., 2006. Surface and deep ocean<br />

interactions during the cold climate event 8200 years ago. Science, 312: 1929-1932.<br />

Fairbanks, R., 1989. A 17,000-year glacio-eustatic sea-level record: influence of glacial<br />

melting rates on the Younger Dryas event and deep-ocean circulation. Nature, 342:<br />

637-642.<br />

Heusser, L.E. and Stock, C.E., 1984. Preparation techniques for concentrating pollen<br />

from marine sediments and other sediments with low pollen density. Palynology, 8:<br />

225-227.<br />

216<br />

F. Naughton, 2007


Hughen, K.A., Overpeck, J.T., Peterson, L.C. and Trumbore, S., 1996. Rapid climate<br />

changes in the tropical Atlantic region during the last deglaciation. Nature, 380: 51-<br />

54.<br />

Hughen, K.A., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell,<br />

P.G., Buck, C.E., Burr, G., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G.,<br />

Friedrich, M., Guilderson, T.P., Kromer, B., McCormac, F.G., Manning, S., Bronk Ramsey,<br />

C., Reimer, P.J., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J. and Weyhenmeyer, C.E., 2004. Marine04 Marine<br />

Radiocarbon Age Calibration, 0-26 Cal Kyr BP. Radiocarbon, 46: 1059-1086.<br />

Jouanneau, J.M. Weber, O., Drago, T., Rodrigues, A., Oliveira, A., Dias, J.M.A., Garcia,<br />

C., Schimdt, S., Reyss, J.L., 2002. Recent sedimentation and sedimentary budgets on<br />

the western Iberian shelf. Progress In Oceanography, 52: 261-275.<br />

Kaufman, D.S., Ager, T.A., Anderson, N.J., Anderson, P.M., Andrews, J.T., Bartlein, P.J.,<br />

Brubaker, L.B., Coats, L.L., Cwynar, L.C., Duvall., M.L., Dyke, A.S., Edwards, M.E., Eisner,<br />

W.R., Gejewski, K., Geirsdóttir, A., Hu, F.S., Jennings, A.E., Kaplan, M.R., Kerwin, M.W.,<br />

Lozhkin, A.V., MacDonald, G.M., Miller, G.H., Mock, C.J., Oswald, W.W., Otto-Bliesner,<br />

B.L., Porinchu, D.F., Rühland, K., Smol, J.P., Steig, E.J. and Wolfe, B.B., 2004. Holocene<br />

thermal maximum in the Western Arctic (0 to 180W). Quaternary Science Reviews, 23:<br />

529-560.<br />

Keigwin, L. and Lehman, B., 1994. Deep circulation change linked to Heinrich event 1<br />

and Younger Dryas in a middepth North Atlantic core. Paleoceanography, 9: 185-194.<br />

Keigwin, L.D., Sachs, J.P., Rosenthal,Y. and Boyle, E.A., 2005. The 8200 year B.P. event<br />

in the slope water system, western subpolar North Atlantic. Palaeoceanography, 20:<br />

2003, DOI:10.1029/2004PA00174.<br />

Kim, J-H., Schneider, R.R., Müller, P.J. and Wefer, G., 2002. Interhemispheric<br />

comparison of deglacial sea-surface temperature patterns in Atlantic eastern<br />

boundary currents. Earth and Planetary Science Letters, 194: 383-393.<br />

Knudsen, K.L., Jiang, H., Jansen, E., Eiriksson, J., Heinemeier, J. and Seidenkrantz, M.S.,<br />

2004. Environmental changes off North Iceland during the deglaciation and the<br />

Holocene: foraminifera, diatoms and stable isotopes. Marine Micropaleontology, 50:<br />

273–305.<br />

Knutti, R., Flückiger, J., Stocker, T.F. and Timmermann, A., 2004. Strong hemispheric<br />

coupling of glacial climate through freshwater discharge and ocean circulation.<br />

Nature, 430: 851-856.<br />

Kucera, M., Rosell-Melé, A., Schneider, R., Walbroeck, C. and Weilnet, M., 2005.<br />

Mutliproxy approach for the reconstruction of the glacial ocean surface (MARGO).<br />

Quaternary Science Reviews, 24: 813-819.<br />

Lebreiro, S.M., Moreno, J.C., McCave, I.N. and Weaver, P.P.E., 1996. Evidence for<br />

Heinrich layers off Portugal (Tore Seamount: 39°N, 12°W). Marine Geology, 131: 47-56.<br />

Leuschner, D.C., Sirocko, F., Grootes, P.M. and Erlenkeuser, H., 2002. Possible influence<br />

of Zoophycos bioturbation on radiocarbon dating and environmental interpretation.<br />

Marine Micropaleontology, 46: 111–26.<br />

217<br />

F. Naughton, 2007


Lotter, A.F., Eicher, U., Siegenthaler, U. and Birks, H.J.B., 1992. Late-glacial climatic<br />

oscillations as recorded in Swiss lake sediments. Journal of Quaternary Science, 7: 187-<br />

204.<br />

Löwemark, L. and Werner, F., 2001. Dating errors in high-resolution stratigraphy: A<br />

detailed X-ray radiograph and AMS- 14 C study of Zoophycos burrows. Marine<br />

Geology, 177: 191-198.<br />

Löwemark, L., Schönfeld, J., Werner, F. and Schäfer, P., 2004. Trace fossils as a<br />

paleoceanographic tool: evidence from Late Quaternary sediments of the<br />

southwestern Iberian margin. Marine Geology, 204: 27-41.<br />

Magri, D., 1995. Some questions on the late-Holocene vegetation of Europe. The<br />

Holocene, 5: 354-360.<br />

Marshall, J., Kushnir, Y., Battisti, D., Chang, P., Czaja, A., Dickson, R., Hurrell, J.,<br />

McCartney, M., Saravanan, R., Visbeck, M., 2001. North Atlantic climate variability:<br />

Phenomena, impacts and mechanisms. International Journal of Climatology, 21:<br />

1863–1898.<br />

McManus, J. F., Francois, R., Gherardi, J.-M., Keigwin, L. D., Brown-Leger, S., 2004.<br />

Collapse and rapid resumption of Atlantic meridional circulation linked to deglacial<br />

climate changes. Nature 428 ,834-837.<br />

Ménot, G., Bard, E., Rostek, F., Weijers, W.H., Hopmans, E.C., Schouten, S. and<br />

Sinninghe Damsté, J.S., 2006. Early Reactivation of European Rivers During the Last<br />

Deglaciation. Science, 313: 1623-1625.<br />

Migeon, S., Weber, O., Faugeres, J.C. and Saint-Paul, J., 1999. SCOPIX: A new imaging<br />

system for core analysis. Geo-Marine Letters, 18: 251-255.<br />

Morey, A.E., Mix, A.C. and Pisias, N.G., 2005. Planktonic foraminiferal assemblages<br />

preserved in surface sediments correspond to multiple environmental variables.<br />

Quaternary Science Reviews, 24: 925-950.<br />

Naughton, F., Sánchez Goñi, M.F., Desprat, S., Turon, J.-L., Duprat, J., Malaizé, B., Joli,<br />

C., Cortijo, E., Drago T., and Freitas, M.C., 2006. Present-day and past (last 25 000<br />

years) marine pollen signal off western Iberia. Marine Micropaleontology, in press.<br />

Naughton F., Sánchez Goñi, M.F., Duprat, J., Cortijo, E., Malaizé, B., Joli, C., Bard, E.,<br />

Rostek, F. and Turon, J-L., in prep. New insights on the impact of Heinrich events and<br />

LGM in the mid-latitudes of the eastern North Atlantic and in the adjacent continent.<br />

Naughton, F., Bourillet, J.-F., Sánchez Goñi, M.F., Turon, J.-L., Jouanneau, J.-M.,<br />

submitted. Long-term and millennial-scale climate variability in north-western France<br />

during the last 8 850 years. The Holocene.<br />

Oliveira, A., Rocha, F., Rodrigues, A., Jouanneau, J.M., Dias, J.M.A., Weber, O. and<br />

Gomes, C., 2002. Clay mineral from the sedimentary cover from the Northwestern<br />

Iberian shelf. Progress in Oceanography, 52: 233-247.<br />

Ottens, J.J., 1991. Planktonic foraminifera as North Atlantic water mass indicators.<br />

Oceanologia Acta, 14: 123-140.<br />

218<br />

F. Naughton, 2007


Pailler, D. and Bard, E., 2002. High frequency palaeoceanographic changes during<br />

the past 140,000 yr recorded by the organic matter in sediments of the Iberian<br />

margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 181: 431-452.<br />

Peterson, L.C., Haug, G.H., Hughen, K.A. and Röhl, U., 2000. Rapid changes in the<br />

hydrologic cycle of the tropical Atlantic during the last Glacial. Science, 290: 1947-<br />

1951.<br />

Pflaumann, U., Duprat, J., Pujol, C. and Labeyrie, L. D., 1996. SIMMAX: A modern<br />

analog technique to deduce Atlantic sea surface temperatures from planktonic<br />

foraminifera in deep-sea sediments. Paleoceanography, 11: 15-35.<br />

Reille, M., de Beaulieu, J.L., Svobodova, H., Andrieu-Ponel, V. and Goeury, C., 2000.<br />

Pollen analytical biostratigraphy of the last five climatic cycles from a long<br />

continental sequence from the Velay region (Massif Central, France). Journal of<br />

Quaternary Science, 15: 665–685.<br />

Rohling, E.J. and Pälike, H., 2005. Centennial-scale climate cooling with a sudden<br />

cold event around 8,200 years ago. Nature, 434: 975-979.<br />

Rühlemann, C., Mulitza, S., Müller, P.J., Wefer, G. and Zahn, R., 1999. Warming of the<br />

tropical Atlantic Ocean and slowdown of thermohaline circulation during the last<br />

deglaciation. Nature, 402: 511-514.<br />

Sánchez Goñi, M.F., Landais, A. Flückiger, J., Desprat, S., Naughton, F. and Fletcher,<br />

W., in prep. Evidence for major sources of methane in the temperate latitudes (20°N-<br />

40°N) during the last climatic cycle.<br />

Seppä, H. and Poska, A., 2004. Holocene annual mean temperature changes in<br />

Estonia and their relationship to solar insolation and atmospheric circulation patterns.<br />

Quaternary Research, 61: 22-31.<br />

Seppä, H., Hammarlund, D. and Antonsson, K., 2005. Low-frequency and highfrequency<br />

changes of temperature and effective humidity during the Holocene in<br />

South central Sweden: Implications for atmospheric and oceanic forcings of climate.<br />

Climate Dynamics, 25: 285-297.<br />

Shemesh, A., Hodell, D., Crosta, X., Kanfoush, S., Charles, C. and Guilderson, T., 2002.<br />

Sequence of events during the last deglaciation in Southern Ocean sediments and<br />

Antarctic ice cores. Paleoceanography, 17: 599-605.<br />

Stuiver, M. and Reimer, P., 1993. Extended 14C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

Stuiver, M., Reimer, P.J. and Reimer, R.W., 2005. CALIB 5.0. (WWW program and<br />

documentation).<br />

Telford, R.J., Heegaard, E. and Birks, H.J.B., 2004. The intercept is a poor estimate of a<br />

calibrated radiocarbon age. The Holocene, 14: 296-298.<br />

Teller, J.T., Leverington, D.W. and Mann, J.D., 2002. Freshwater outbursts to the oceans<br />

from glacial Lake Agassiz and their role in climate change during the last<br />

deglaciation. Quaternary Science Reviews, 21: 879–887.<br />

Tinner, W. and Lotter, A.F., 2001. Central European vegetation response to abrupt<br />

climate change at 8.2 ka. Geology, 29: 551–554.<br />

219<br />

F. Naughton, 2007


Tinner, W. and Lotter, A.F., 2006. Holocene expansions of Fagus silvatica and Abies<br />

alba in Central Europe: where are we after eight decades of debate? Quaternary<br />

Science Reviews, 25: 526-549.<br />

Turon, J.-L., Lézine, A.-M. and Denèfle, M., 2003. Land-sea correlations for the last<br />

glaciation inferred from a pollen and dinocyst record from the Portuguese margin.<br />

Quaternary Research, 59: 88-96.<br />

Veski, S., Seppä, H. and Ojala, A.E.K., 2004. Cold event at 8200 yr BP recorded in<br />

annually laminated lake sediments in eastern Europe. Geology, 32: 681-684.<br />

Wanner, H., Brönnimann, S., Casty, C., Gyalistras, D., Luterbacher, J., Schmutz, C.,<br />

St<strong>ephe</strong>nson, D.B. and Xoplaki, E., 2001. North Atlantic Oscillation-Concepts and<br />

studies. Surveys in Geophysics, 22: 321– 382, 2001.<br />

Watts, W.A., 1980. Regional variation in the reponse of vegetation to Lateglacial<br />

climatic events in Europe. J.J. Lowe, Gray, J.M. and Robinson, J.E. (Eds), Studies in the<br />

Lateglacial of North West Europe. Pergamon Press, Oxford,1-22 pp.<br />

Wunsch, C., 2003. Greenland-Antarctic phase relations and millennial time-scale<br />

climate fluctuations in the Greenland ice-cores. Quaternary Science Reviews, 22:<br />

1631-1649.<br />

Wunsch, C., 2006. Abrupt climate change: An alternative view. Quaternary Research,<br />

65: 191-203.<br />

Zaragosi, S. Eynaud, F., Pujol, C., Auffret, G.A., Turon, J.-L. and Garlan, T., 2001.<br />

Initiation of the European deglaciation as recorded in the northwestern Bay of Biscay<br />

slope environments (Meriadzek Terrace and Trevelyan Escarpment): a multi-proxy<br />

approach. Earth and Planetary Science Letters, 188: 493-507.<br />

220<br />

F. Naughton, 2007


Capítulo 5| Long-term and millennial-scale climate<br />

variability in north-western France during the last 8 850 years<br />

Variações climáticas de longa e pequena escala no<br />

noroeste de França durante os últimos 8 850 anos<br />

Variabilité climatique orbitale et sub-orbitale dans le nordouest<br />

de la France pendant les derniers 8 850 ans<br />

221<br />

The Holocene<br />

Submitted<br />

F. Naughton a, b , J-F. Bourillet c , M.F. Sánchez Goñi d , J-L. Turon a , J-M. Jouanneau a<br />

a Environnements et Paléoenvironnements Océaniques (UMR CNRS 5805 EPOC),<br />

Université Bordeaux 1, Av. des Facultés, 33405 Talence, France<br />

b Departamento de Geologia, Universidade de Lisboa, Portugal<br />

C IFREMER, Département Géosciences Marines, Laboratoire Environnements<br />

Sédimentaires, Plouzané, France<br />

d Ecole Pratique des Hautes Etudes, Environnements et Paléoenvironnements<br />

Océaniques (UMR CNRS 5805 EPOC), Université Bordeaux 1, Av. des Facultés, 33405<br />

Talence, France<br />

F. Naughton, 2007


Resumo<br />

As variações do coberto vegetal e a estimativa dos parâmetros<br />

climáticos obtidos numa sondagem recolhida na plataforma continental do<br />

noroeste de França (VK03-58Bis), permitiram detectar variações climáticas<br />

de escala orbital e sub-orbital, para os últimos 8850 anos, nesta região. O<br />

padrão geral de diminuição gradual da temperatura de verão, marcada<br />

pelo declínio progressivo da floresta temperada e húmida, é<br />

contemporâneo do arrefecimento progressivo da temperatura na<br />

Gronelândia, assim como, da redução dos valores de insolação de verão<br />

nas latitudes medias até, pelo menos, 2000 anos cal BP. Ao mesmo tempo, a<br />

diminuição da sazonalidade segue o aumento da precessão.<br />

Entre 8739 e 8387 anos cal BP, a floresta de Corylus expande-se em<br />

detrimento do Quercus deciduous, como resposta a uma amplificação do<br />

contraste sazonal, resultante da expansão de gelo marinho durante o<br />

inverno nas altas latitudes do Atlântico Norte, contrariando o padrão geral<br />

de forçamento orbital. Este forte contraste sazonal, resulta da expulsão<br />

drástica de água doce dos lagos de “Agassiz” e “Ojibway” e, da gradual<br />

redução da circulação termohalina no Atlântico Norte (MOC).<br />

Entre 8387 e 8062 anos cal BP, o súbito declínio da floresta de Corylus,<br />

marca o evento frio “8.2 kyr”, no noroeste de França. Este episódio, foi<br />

provavelmente produzido pela redução severa da MOC, a qual provocou<br />

uma diminuição suplementar da temperatura de inverno, na Europa e na<br />

Gronelândia. No entanto, o contraste sazonal permaneceu elevado durante<br />

este evento. O forte contraste sazonal registado entre 8739 e 8062 anos cal<br />

BP reflecte o evento multi-secular "8.6-8.0 kyr» no Atlântico Norte.<br />

Após os estádios finais de expulsão de água dos lagos “Agassiz” e<br />

“Ojibway”, o clima torna-se mais estável. No entanto, estão registados uma<br />

série de ligeiros e rápidos episódios frios os quais, estão associados a um<br />

pequeno arrefecimento no inverno e a um ligeiro aumento da precipitação.<br />

Finalmente, a análise das associações de dinocistos e a quantificação<br />

de gastrópodes do tipo Turritella communis permitiu-nos de detectar<br />

variações regionais tais com: a migração para sul da zona biogeográfica<br />

marinha Boreal entre 8739 e 8479 anos cal BP e, a abertura do canal da<br />

mancha entre 8479 e 8387 anos cal BP.<br />

222<br />

F. Naughton, 2007


Résumé<br />

Les variations du couvert végétal et des paramètres climatiques, mises<br />

en évidence à partir de l’étude d’une carotte prélevée sur la plateforme<br />

continentale nord-ouest française (VK03-58Bis), témoignent de la variabilité<br />

climatique orbitale et sub-orbitale des derniers 8850 ans dans cette région. A<br />

l’échelle orbitale et jusqu’au moins 2000 ans cal BP, la reconstruction<br />

climatique quantitative montrent un refroidissement estival qui parallélise la<br />

régression progressive de la forêt tempérée et humide, la baisse des<br />

températures au Groenland et la diminution de l’insolation d’été des<br />

moyennes latitudes. La diminution de la saisonnalité détectée par le<br />

changement graduelle de la végétation est compatible avec<br />

l’augmentation de la précession.<br />

L’expansion de Corylus au détriment de la forêt de Quercus<br />

caducifolié, entre 8739 et 8387 ans cal BP, est liée au fort contrast saisonnier,<br />

amplifié par l’expansion de la banquise d’hiver dans les hautes latitudes de<br />

l’Atlantique Nord, contrebalançant ainsi le forçage orbital. Ce fort contraste<br />

saisonnier serait le résultat des épisodes terminaux de purges des lacs<br />

d’Agassiz et d’Ojibway et de la réduction graduelle de la circulation<br />

thermohaline de l’Atlantique Nord (MOC-Meridional Overturning Circulation).<br />

Entre 8387 et 8062 ans cal BP, la régression soudaine de la forêt de<br />

Corylus, tout en restant forte la saisonnalité, marque l’événement froid « 8.2<br />

ka » dans le nord-ouest de la France. Cet épisode est probablement lié à la<br />

réduction ultime et sévère de la MOC qui a provoquée une diminution<br />

supplémentaire des températures sur l’Europe et sur le Groenland. Le fort<br />

contraste saisonnier enregistré dans la VK03-58Bis, entre 8739 et 8062 ans cal<br />

BP, correspond au refroidissement pluri-séculaire "8.6-8.0 ka» de l’Atlantique<br />

Nord.<br />

Après les derniers épisodes de purges des lacs d’Agassiz et d’Ojibway,<br />

le climat devient plus stable. Toutefois, nos reconstructions climatiques<br />

montrent des refroidissements d’ordre millénaire caractérisés par une légère<br />

baisse des températures hivernales et par l’augmentation des précipitations.<br />

De plus, l’analyse des assemblages de dinokystes et des occurrences<br />

du gastropode Turritella communis, indiquent des changements régionaux<br />

importants, comme la migration vers le sud de la région biogéographique<br />

223<br />

F. Naughton, 2007


marine Boréale entre 8739-8479 ans cal BP et l’ouverture du Chenal de la<br />

manche entre 8479 et 8387 ans cal BP.<br />

Abstract<br />

Vegetation and quantitative climate reconstructions from a north-<br />

western France shelf core (VK03-58Bis) show orbital and suborbital climate<br />

variability for the last 8850 years in this region. A long-term cooling trend in<br />

summer temperatures, marked by gradual temperate and humid forest<br />

decline, parallels cooling in Greenland and the decrease of mid-latitude<br />

summer insolation reduction until at least 2000 yr cal BP. At long-term scale,<br />

the lowering in seasonal contrast revealed by vegetation changes follows the<br />

increase of precession.<br />

Corylus woodlands spread at the expense of deciduous Quercus<br />

forest, between 8739 and 8387 cal yr BP, linked with the high seasonality<br />

conditions which, counterbalancing the long-term astronomical forcing trend,<br />

were amplified by the north Atlantic high-latitudes winter sea-ice expansion.<br />

High seasonality conditions resulted from the Agassiz and Ojibway final<br />

outburst episodes and consequent gradual reduction of the MOC (Meridional<br />

Overturning Circulation).<br />

Between 8387-8062 cal yr BP, a sudden Corylus woodland decline<br />

marks the 8.2 kyr cold event in north-western France probably triggered by<br />

the severe MOC reduction leading to the additional drop in winter<br />

temperature over Europe and Greenland. Nonetheless, seasonality remains<br />

high during this interval. The high seasonality conditions detected in VK03-<br />

58Bis between 8739-8062 cal yr BP reflects the multi-centennial-scale climate<br />

cooling 8.6-8.0 kyr episode of the North Atlantic.<br />

Following the Agassiz and Ojibway final outburst episodes, climate<br />

became more stable. However, millennial scale climate cooling episodes are<br />

recorded and characterised by weak winter cooling and increases<br />

precipitation.<br />

Furthermore, dinocyst analysis and benthic gastropod Turritella<br />

communis occurrences indicate regional changes such as the southward<br />

migration of the Boreal biogeographical zone between 8739-8479 cal yr BP<br />

224<br />

F. Naughton, 2007


and the subsequent opening of the English Channel at around 8479-8387 cal<br />

yr BP.<br />

225<br />

F. Naughton, 2007


226<br />

F. Naughton, 2007


5. 1 Introduction<br />

For a long time the Holocene interglacial has been considered a<br />

period of stable climate. However, several studies have shown that<br />

superimposed on the orbitally-induced long-term cooling (e.g. Kutzbach and<br />

Gallimore, 1988; Crucifix et al., 2002; Marchal et al., 2002; Renssen et al., 2005;<br />

Lorenz et al., 2006) sub-orbital millennial scale climate variability has affected<br />

this interglacial (e.g. Denton and Karlén, 1973; O’Brien et al., 1995; Bond et al.,<br />

1997; Mayewski et al., 2004).<br />

The most extreme short-lived cold episode noticed in the Greenland<br />

Ice cores (O’Brien et al., 1995; Alley et al., 1997; Muscheler et al., 2004), known<br />

as the “8.2-kyr-BP event” and lasting 100-200 years, has been detected<br />

elsewhere in several climate proxy data from the North Atlantic marine deep-<br />

sea cores (Bond et al., 1997; 2001; Bianchi and McCave; 1999) and from the<br />

European continent (e.g. Von Grafenstein et al., 1998; Klitgaard-Kristensen et<br />

al., 1998; Nesje and Dahl, 2001; Tinner and Lotter, 2001; 2006; Baldini et al.,<br />

2002; Magny et al., 2003; Veski et al., 2004). The causes triggering the “8.2<br />

event” have been strongly debated over the last decade. Some authors<br />

suggest that this event results from changes in solar activity (Denton and<br />

Karlén, 1973; Bond et al., 2001; Van Geel et al., 2003) while others from<br />

freshwater pulses (Von Grafenstein et al., 1999; Barber et al., 1999; Rind et al.,<br />

2001; Alley et al., 2003).<br />

The fact that this event is more prominent in the North Atlantic region,<br />

that it follows two outburst flooding episodes, and that the existing similarities<br />

between reconstructed anomaly patterns and patterns expected following a<br />

North Atlantic freshening seem to favour the freshwater pulse mechanism as<br />

the major trigger for the 8.2 event (Alley and Ágústsdóttir, 2005).<br />

Two recent publications (Rohling and Pälike, 2005 and Ellison et al.,<br />

2006) suggest that the 8.2 kyr event occurred within a long climate cooling<br />

anomaly of multi-centennial-scale, between 8600 and 8000 years ago. This<br />

long-lived episode has been previously noticed by a dust supply increase in<br />

GISP2 (Mayewski et al., 1997); a decrease of sea surface temperatures (SST) in<br />

the North Atlantic (Risebrobakken et al., 2003; Knudsen et al., 2004; Keigwin et<br />

al., 2005) and a decrease of annual temperature from a few northern<br />

227<br />

F. Naughton, 2007


European pollen sequences (Seppä and Poska, 2004) which the authors<br />

usually associate with the short-lived 8.2 kyr event.<br />

Besides the 8.2 kyr, event most of the northern European pollen<br />

sequences from Estonia and Sweden detect a Holocene Thermal Maximum<br />

(HTM) (Seppä and Poska, 2004; Seppä et al., 2005) between 8 000 and 4 000<br />

cal yr BP.<br />

So far, no studies have shown the vegetation response either to<br />

orbitally-induced long-term cooling or to sub-orbital millennial scale climate<br />

variability other than the extreme 8.2 kyr event in western-central Europe<br />

during the Holocene. The aim of this study is therefore to test whether longer<br />

and shorter-term climatic variability, involving the 8.2 kyr event and the 8.6-8.0<br />

kyr episode, has affected western France. Towards this aim we have<br />

performed palynological analyses (pollen and dinocysts) and pollen-derived<br />

quantitative climate reconstructions from a shelf core VK03-58Bis retrieved in<br />

the “Grande Vasière” of the Bay of Biscay. This core gives an integrated<br />

image of the past regional vegetation and, therefore, the climate of western<br />

France. This region is particularly sensitive to hydrological changes of the<br />

North Atlantic Drift (Rahmstorf, 2002).<br />

5. 2 Environmental Setting<br />

The Bay of Biscay presents a 300 km wide continental shelf in the north-<br />

westernmost area and becomes narrow with a steep slope further south (30<br />

km wide) (Fig. V.1). This shelf is composed of two small and one large open-<br />

shelf mud patches: the W and S Gironde shelf mud fields and the “Grande<br />

Vasière” (Allen and Castaing, 1977). According to McCave’s classification the<br />

“Grande Vasière” is a mid-shelf mud belt (McCave, 1972).<br />

The “Grande Vasière” is large (more than 225 km length and 40 km<br />

wide), located between 80 to 110 m water depth and presents an annual<br />

mean sedimentary rate of 0.1-0.2 cm yr -1 (Lesueur et al., 2001) (Fig. V.1). Shelf<br />

upkeep depends essentially on: a) continental supply by n<strong>ephe</strong>loid layers<br />

(Jouanneau et al., 1999; Lesueur et al., 2001); b) wave action (Pinot, 1974)<br />

and hydrology, and c) sea level changes (Lesueur and Klingebiel, 1976). The<br />

“Grande Vasière” rests over two sandy units and consists of a thin (few<br />

228<br />

F. Naughton, 2007


decimetres) Holocene feature of muddy autochthon sand (Bourillet et al.,<br />

2002). The present day spreading of sediments to the shelf is also influenced<br />

by resuspension and redistribution of the sediments during storm episodes and<br />

under the effects of trawling nets (Bourillet et al., 2006).<br />

The shelf is nourished by fine grained sediments released essentially by<br />

the Gironde and Loire rivers and to a lesser extent by the Adour, Vilaine and<br />

Charente (Castaing and Jouaneau, 1987). The Gironde and Loire rivers have<br />

large catchment areas (including the Massif Central and the Pyrenees zones)<br />

recruiting pollen grains from most of the western part of France.<br />

Indeed, previous works on world wide coastal zones with complex<br />

fluvial systems have shown that pollen grains after being produced and<br />

initially dispersed by the wind are mainly transported to the sea by rivers and<br />

streams (Muller, 1959; Bottema and Van Straaten, 1966; Peck, 1973; Heusser<br />

and Balsam, 1977; Naughton et al., in press). Experimental studies on pollen<br />

from the French margin have shown that river systems are mainly responsible<br />

for pollen input into the sea and that the marine pollen signature reflects an<br />

integrated image of the regional vegetation of the adjacent continent<br />

(Turon, 1984). Furthermore, westerly prevailing winds probably impede direct<br />

airborne transport of pollen seaward.<br />

Mean annual precipitation over the catchment area (PANN) varies<br />

from 1000 mm in the westernmost part to 600 mm in the eastern zone. High<br />

altitudinal zones, such as the Massif Central region are characterised by more<br />

than 2200 mm of PANN while the Pyrenees vary from 2000 mm in the western<br />

part to 1000 mm in the eastern zone. Present-day annual temperature is 13°C<br />

in western France (data from French public Agency: “Meteo france”).<br />

The oceanic, mild and humid climate of this region allows the<br />

development of a temperate deciduous and warm mixed forest mainly<br />

composed of deciduous Quercus (Q .pedunculata, Q. pubescens and Q.<br />

sessiflora) with some scattered evergreen Quercus (Q. ilex), cork oak (Q.<br />

suber) as well as elm (Ulmus) and ash (Fraxinus) associations. Littoral zones are<br />

mainly composed of cluster pine (Pinus pinaster) and gorses (Ulex). There are<br />

also beech (Fagus) and hornbeam (Carpinus) woodlands at higher altitudes.<br />

229<br />

F. Naughton, 2007


5. 3 Material and methods<br />

The 2.72 m long core, VK03-58Bis, was retrieved at 96.8 m water depth<br />

in the “Southwest-Glénan” sector of the “La Grande Vasière” mud patch<br />

(47°36’ N and 4°08’ W) using a vibrocorer during the “Vibarmor”<br />

oceanographic cruise (integrated in the “Défi Golfe de Gascogne” Ifremer<br />

programme) (Fig. V.1). The “Glénan” sector is one of the end members of the<br />

“Grande Vasière” and is composed of 3 m of sediments with high<br />

percentages of fine material (greater than 80%).<br />

Sedimentological analysis including micro-granulometry, calcimetry, x-<br />

ray analysis core using SCOPIX image-processing mode (Migeon et al., 1999)<br />

and benthic gastropod Turritella communis counting on the VK03-58Bis shelf<br />

core as well as the final core description were performed by Folliot (2004).<br />

Fig. V.1 | Location of shelf core VK03-58Bis; and deep-sea core MD99-2551 (Ellison et al., 2006).<br />

5. 3. 1 Radiometric dating<br />

Five accelerator mass spectrometer (AMS) 14 C dates on T. communis<br />

were obtained in the Poznan Radiocarbon Laboratory (Poland) (Bourillet et<br />

230<br />

F. Naughton, 2007


al., 2005) indicating that the VK03-58Bis sedimentary sequence covers the last<br />

8 850 years (Tab. V.1 and Fig. V.2). T. communis dated levels from twin cores,<br />

VK03-58 (47°36’ N, 4°08’ W; 97.3 m water depth) and VK03-59Bis (47°38’ N,<br />

4°09’ W; 94.6 m water depth), were correlated with that of core VK03-58Bis for<br />

the age model construction.<br />

All AMS 14 C dated levels were calibrated using CALIB Rev 5.0 program<br />

and the "global" marine calibration dataset (marine 04.14c) (Stuiver and<br />

Reimer, 1993; Hughen et al., 2004; Stuiver et al., 2005). This dataset uses the<br />

global marine age reservoir correction (R) of 400 years. For accommodating<br />

local effects, we have introduced the difference Δr (of about 3 years) in<br />

reservoir age of the Bay of Arcachon (France), the closest area to our core,<br />

as suggested by Stuiver et al. (2005). We used the 95.4% (2 sigma) confidence<br />

intervals and their relative areas under the probability curve as well as the<br />

median probability of the probability distribution (Telford et al., 2004) as<br />

suggested by Stuiver et al. (2005).<br />

Lab<br />

code<br />

POZ-<br />

10166<br />

POZ-<br />

10167<br />

POZ-<br />

10168<br />

POZ-<br />

10170<br />

POZ-<br />

10171<br />

POZ-<br />

6079<br />

POZ-<br />

10172<br />

POZ –<br />

6077<br />

Core-<br />

depth<br />

(cm)<br />

VK03 58Bis<br />

106<br />

VK03 58Bis<br />

149<br />

VK03 58Bis<br />

160<br />

VK03 58Bis<br />

177<br />

VK03 58Bis<br />

226<br />

VK03 59Bis<br />

190<br />

VK03 59Bis<br />

212<br />

VK03-58<br />

201<br />

Material<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

T.<br />

communis<br />

Conv.<br />

AMS 14 C<br />

age BP<br />

Conv.<br />

AMS 14 C<br />

age BP<br />

(-400 yr)<br />

231<br />

error<br />

Weighted<br />

Mean Δr<br />

Arcachon<br />

France<br />

95.4 % (2σ)<br />

Cal BP<br />

age ranges<br />

Cal BP age<br />

median<br />

probability<br />

3820 3420 30 3 3667 BP:3865 BP 3763<br />

7020 6620 30 3 7427 BP:7576 BP 7507<br />

8030 7630 30 3 8391 BP:8576 BP 8479<br />

8170 7770 30 3 8532 BP:8808 BP 8652<br />

8240 7840 30 3 8613 BP:8938 BP 8764<br />

7920 7520 40 3 8298 BP:8476 BP 8377<br />

8200 7800 40 3 8567 BP:8884 BP 8696<br />

8090 7690 50 3 8411 BP:8692 BP 8545<br />

Tab. V.1| Radiocarbon ages from VK03-58Bis and VK03-58 and VK03-59Bis shelf cores.<br />

5. 3. 2 Pollen and dinocyst analyses<br />

42 and 15 samples were collected with a sample spacing of 4 to 8 cm<br />

along the VK03-58Bis sedimentary record for pollen and dinocyst analysis,<br />

respectively. The treatment used for palynological analysis followed the<br />

procedure described by de Vernal et al. (1996), slightly modified at the UMR<br />

CNRS 5805 EPOC (Desprat, 2005).<br />

F. Naughton, 2007


Chemical digestion using cold HCl (at 10%, 25% and 50%) and cold HF<br />

(at 40% and 70%) were applied to eliminate carbonates and silicates. A<br />

Lycopodium spike of known concentration was added to each sample to<br />

calculate pollen concentrations. The residue was sieved through 10 µm nylon<br />

mesh screens (Heusser and Stock, 1984) and mounted in bidistillate glycerine.<br />

Pollen and cysts were identified and counted using a Zeiss microscope<br />

with x550 and x1250 (immersion) magnifications, the last one only applied for<br />

pollen analysis. At least 100 pollen grains (excluding Pinus, aquatic plants and<br />

spores) and at least 15 pollen types were counted. Pinus pollen is usually over-<br />

represented in marine deposits and therefore is often excluded from the main<br />

sum (Heusser and Balsam, 1977; Turon, 1984). However, it is known that the<br />

percentages of this taxa increase seaward although total pollen content<br />

decreases (Muller, 1959; Groot and Groot, 1966; Bottema and Van Straaten,<br />

1966; Koreneva, 1966; van der Kaars and Deckker, 2003). Because the site<br />

location is closed to the present-day coast line we assume that Pinus pollen<br />

percentages are not over-represented in this core and, therefore, Pinus pollen<br />

grains have not been excluded from the main pollen sum.<br />

Pollen percentages of each taxa were calculated based on the main<br />

pollen sum that excludes aquatic plants, pteridophyte spores and<br />

indeterminable pollen. 57 to 424 cysts were counted and interpreted by<br />

comparison with modern dinoflagellate cyst distribution (de Vernal et al.,<br />

1998; Rochon et al., 1999).<br />

5. 3. 3 Pollen-based quantitative climate reconstruction<br />

Quantitative climate reconstruction of north-western of France for the<br />

last 8850 years was obtained by applying the modern analogue technique<br />

(MAT) (Guiot et al., 1989; Guiot; 1990) to the VK03-58Bis pollen sequence. This<br />

method is based on a modern pollen assemblage dataset including 1328<br />

pollen spectra from Europe, Eurasia and North Africa (Peyron et al., 1998;<br />

Peyron et al., 2005), and it selects the 5 modern pollen assemblages closest to<br />

the fossil pollen spectra. These 5 analogues present the smallest chord<br />

distance (Guiot, 1990) representing the best modern analogues for a given<br />

fossil pollen spectrum and, therefore, the best samples for climate parameters<br />

estimate. Climate parameter estimates are obtained by taking a weighted<br />

232<br />

F. Naughton, 2007


average of the values for all selected best modern analogues which<br />

represents the inverse of the chord distance.<br />

Each modern analogue sample is associated with several climate<br />

parameters which have been previously interpolated from meteorological<br />

stations by using an Artificial Neural Network (ANN) technique (Peyron et al.,<br />

1998). The parameters selected for climate reconstruction of north-western<br />

France are: TANN (mean annual temperatures); PANN (mean annual<br />

precipitation) and the difference between the temperature of the warmest<br />

(MTWA) and the coldest (MTCO) months (seasonality). These climate<br />

parameters are understood to play a prominent role on the distribution of the<br />

vegetation and related pollen assemblages (Peyron et al., 2005).<br />

5. 4 Results<br />

5. 4. 1 Lithostratigraphy and age model<br />

VK03-58Bis is characterised by a homogenous silt sequence marked<br />

between 210 and 150 cm by a level containing T. communis (Fig. V.2).<br />

Between 210 and 160 cm this T. communis community presents all the<br />

characteristics of a biocenose: the shells are deposited in life position; both<br />

young and adult specimens are present within the same level; they do not<br />

present any evidence of shelf destruction by transport.<br />

Between 160 and 150 cm, there is an increase in T. communis<br />

abundance, and in contrast with the underlying level they are not in life<br />

position. This indicates a drastic change in the environmental conditions<br />

which probably resulted in their mortality.<br />

The age obtained from the bottom of this layer in VK03-58Bis shelf core<br />

is 7630 yr BP (8479 cal yr BP). This single drastic episode has also been<br />

observed in the twin cores: VK03-58 dated at the bottom (7690 yr BP; 8545 cal<br />

yr BP) and VK03-59Bis (at 4 km of distance) between 7520 and 7700 yr BP<br />

(8377-8550 cal yr BP) (Tab. V.1). Considering the shortness of this drastic T.<br />

communis mortality episode we can assume that this event has been<br />

synchronous in the three cores and, therefore level 150 cm in VK03-58Bis can<br />

be correlated with the top of that layer dated at 7520 (8377 cal yr BP) in core<br />

VK03-59Bis.<br />

233<br />

F. Naughton, 2007


Because there is no sedimentological evidence (no erosional surfaces<br />

from the RX data and continuous grain size decrease) for a hiatus phase after<br />

this drastic episode in our core, we decided to reject the date obtained for<br />

level 149 cm which seems too young (6620 yr BP, 7507 cal yr BP) when<br />

compared with the age limits of the Turritella layer of the twin cores.<br />

5. 4. 2 Evolution of dinocyst assemblages<br />

Dinocyst analysis performed in VK03-58Bis between 262 and 98 cm<br />

shows a unique assemblage essentially composed of: Lingulodinium<br />

machaerophorum, Operculodinium centrocarpum, and several species of<br />

Spiniferites (Spiniferites lazus, Spiniferites bentorii, Spiniferites spp., Spiniferites<br />

ramosus, Spiniferites mirabilis, Spiniferites membranaceus, Spiniferites<br />

delicatus, Spiniferites bulloideus, Spiniferites belerius, Spiniferites elongates).<br />

Spiniferites dominates the dinocyst associations between 262 and 180<br />

cm and is replaced by Lingulodinium machaerophorum between 180 and<br />

160 cm.<br />

A drastic decrease in Lingulodinium machaerophorum is detected<br />

between 160 and 150 cm contemporaneous with the T. communis mortality<br />

episode.<br />

Finally, and above 150 cm, all species are replaced by Lingulodinium<br />

machaerophorum which again completely dominates the dinocyst<br />

assemblages (Fig. V.2).<br />

5. 4. 3 Vegetation succession and quantitative climate reconstruction<br />

Pollen analysis of the VK03-58Bis shelf core records eight main pollen<br />

zones (numbered from the bottom to the top and prefixed by the<br />

abbreviated sequence name VK03-58Bis) (Fig. V.2). The establishment of<br />

these 8 pollen zones has been performed by using qualitative fluctuations of<br />

a minimum of 2 curves of ecologically important taxa (Pons and Reille, 1986).<br />

To delimitate chronologically each pollen zone, we have used interpolated<br />

ages assuming a constant sedimentary rate between two consecutive dated<br />

samples. Fig. V.3 shows the percentage curves of selected pollen taxa<br />

plotted together with the curves of climatic parameter estimates (PANN,<br />

MTCO, MTWA, Seasonality and TANN).<br />

234<br />

F. Naughton, 2007


The first pollen zone (VK03-58Bis-1), 266-245 cm, (7897-7867 yr BP; 8855-<br />

8807 cal yr BP - extrapolated age assuming the same sedimentary rate than<br />

that obtained between 226 and 177 cm) reflects a Pinus and deciduous<br />

Quercus forest with Corylus and Ulmus (Fig. V.2). Quantitative climate<br />

reconstruction shows that TANN and PANN values are 3 to 2°C and 200 mm<br />

lower (10-11°C, 600 mm), respectively, than present day values (13°C, 800<br />

mm) (Fig. V.3). The expansion of deciduous Quercus forest associated with<br />

the slight spread of Corylus, Betula and Ulmus and the gradual contraction of<br />

pine are indicated by VK03-58Bis-2 pollen zone (245-215 cm, 7867-7824 yr BP;<br />

8807-8739 cal yr BP) (Fig. V.2). This pollen zone also suggests the presence of<br />

scattered pockets of Acer, Fraxinus excelsior-type, Alnus and Tilia, Climatic<br />

reconstruction estimates an increase of precipitation (150-200 mm), a<br />

decrease of seasonality ( ΔS (summer-winter) =5°C) and a slight cooling in<br />

summer by 4°C (Fig. V.3). In several French continental sequences such as<br />

those from the Pyrenees and the Massif Central, the first occurrence of Tilia<br />

has been recorded later, at the beginning of the Atlantic period (7500-5000 yr<br />

BP; 8321-5734 cal yr BP) (Reille, 1990b; de Beaulieu et al., 1984). However,<br />

other sequences such as that of the Soucarat in the Eastern Pyrenees records<br />

the appearance of Tilia earlier, at around 7740±180 yr BP (8575 cal yr BP)<br />

(Reille and Andrieu, 1994). Tilia has been also detected earlier (before 7800 yr<br />

BP; 8700 cal yr BP) in several central-European pollen sequences such as<br />

Soppensee and Bibersee (Switzerland), Schleinsee (Germany) (Tinner and<br />

Lotter, 2001; 2006) and in northern European sequences such as those of<br />

Raigastvere, Viitna, Rõuge, Ruila (Estonia) (Seppä and Poska, 2004; Veski et<br />

al., 2004).<br />

The next pollen zone, VK03-58Bis-3 (215-151 cm, 7824-7531 yr BP; 8739-<br />

8387 cal yr BP) reflects the maximum expansion of Corylus associated with the<br />

contraction of the deciduous Quercus forest. This suggests an important<br />

increase of seasonality between 8700 and 8200 cal yr BP that is supported by<br />

climate estimates (Fig. V.2 and Fig. V.3). It is widely known that Corylus<br />

competed against deciduous Quercus trees mostly during the early Holocene<br />

in southern Europe (Tallantire, 2002). Corylus is a light-demanding tree and its<br />

expansion is favoured by forest openings (Bradshaw and Hannon, 2004).<br />

Furthermore, Corylus is considered to be one tolerant climate species<br />

235<br />

F. Naughton, 2007


supporting high seasonality conditions (Tallantire, 2002). In French continental<br />

sequences, the maximum expansion of Corylus has been documented during<br />

the Boreal period (9000-8000 yr BP; 10184-8866 cal yr BP) (de Beaulieu et al.,<br />

1984; Reille and Andrieu, 1991; Reille and Lowe, 1993). Nevertheless, Buzy and<br />

Barbazan pollen diagrams (Reille and Andrieu, 1991) as well as other<br />

sequences from the central and eastern Pyrenees (Biscaye, Lourdes,<br />

Freychinede, le Monge; Landos, Pinet 1, La Moulinasse 4 and Laurenti) (de<br />

Beaulieu et al., 1984; Reille and Andrieu, 1995; Reille, 1990a; Reille and Lowe,<br />

1993) provide evidence for a longer period of Corylus optimum extent. VK03-<br />

58Bis-3 pollen zone also records a Pinus forest re-expansion. Betula and Ulmus<br />

are consistently present and there are sporadic occurrences of Alnus, Fraxinus<br />

excelsior-type and Tilia. Occurrences of Fagus at 8652 and 8448 cal yr BP<br />

reflect a slight decrease in seasonality and temperatures within the period of<br />

high seasonality that characterises VK03-58Bis-3 pollen zone.<br />

VK03-58Bis-4 pollen zone (151-147 cm, 7531-7240 yr BP; 8387-8062 cal yr<br />

BP) is marked by a drastic reduction of Corylus woodlands and an increase of<br />

Pinus forest along with the maintaining of deciduous Quercus forest (Fig. V.2).<br />

Climate estimates show that the onset of seasonality decrease coincides with<br />

the Corylus minimum extent at around 7427 yr BP (8272 cal yr BP) (Fig. V.3).<br />

This episode of Corylus decline has been also observed in several central-<br />

European (Soppensee and Bibersee in Switzerland and Schleinsee in<br />

Germany; Tinner and Lotter, 2001; 2006) and northern European pollen<br />

sequences (Raigastvere, Viitna, Rõuge, Ruila in Estonia and Lake Flarken in<br />

Sweden (Seppä and Poska, 2004; Veski et al., 2004; Seppä et al., 2005).<br />

Corylus deflection has been interpreted as the vegetation response to the<br />

well known 8.2 ka cooling event.<br />

Deciduous Quercus forest attained its maximum expansion in the<br />

following period (VK03-58Bis-5 pollen zone, 147-102 cm, 7240-3289 yr BP; 8062-<br />

3621 cal yr BP) suggesting a change in climate to milder (reduced<br />

seasonality) conditions as the result of MTCO increase (Fig. V.2 and Fig. V.3).<br />

These conditions, together with an increase of precipitation, favoured the<br />

establishment of Alnus, Ulmus, Tilia, Fraxinus excelsior-type and Fagus trees in<br />

western France.<br />

236<br />

F. Naughton, 2007


The next zone, VK03-58Bis-6 (102-55 cm, 3289-1749 yr BP; 3621–1953 cal<br />

yr BP), indicates the slight contraction of deciduous Quercus forest, the<br />

expansion of Fagus and the gradual increase of herbaceous plants. The slight<br />

decrease of MTWA and the increase of MTCO lead to this vegetation<br />

dynamic in which the Fagus spread has been probably favoured by weak<br />

seasonality and high precipitation (Fig. V.3). In almost all the continental<br />

French sequences, Fagus spread occurred between 4500-4000 and 2000<br />

years BP (5000-4400 and 2150 cal yr BP) coinciding with the beginning of the<br />

oak forest decline (Reille and Lowe, 1993; Reille and Andrieu, 1995; Reille et<br />

al., 2000). In our VK03-58 Bis pollen record, the first occurrence of Fagus is<br />

recorded at around 8652 and 8448 cal yr BP. Several occurrences were<br />

detected during and after the Corylus regression episode. The beginning of a<br />

continuous presence of Fagus occurred at around 4352 yr BP (4812 cal yr BP)<br />

just after the Corylus regression although its maximum expansion started later<br />

(3289 yr BP; 3621 cal yr BP). Tinner and Lotter (2001; 2006) based on pollen<br />

analysis from Soppensee and Bibersee (Switzerland) and Schleinsee<br />

(Germany) suggest, as our sequence, that Fagus expands after the episode<br />

of Corylus deflection, favoured by more humid summer conditions and less<br />

extreme seasonality.<br />

VK03-58Bis-7 pollen zone (55-24 cm, 1749-733 yr BP; 1953-852 cal yr BP)<br />

still shows the gradual reduction of deciduous Quercus forest and the<br />

maximum expansion of Poaceae. Fagus is still present in this pollen zone until<br />

1061 yr BP (1207 cal yr BP) (Fig. V.2 and Fig. V.3). The continuous presence of<br />

Cerealia type, Juglans and Castanea testifies to agricultural practices at<br />

around 2000 years ago in western France. In the last pollen zone, VK03-58Bis-8<br />

(upper 24 cm, last 733 yr BP; 852 cal yr BP) there is a strong increase of Pinus,<br />

heathlands and herbaceous plants, mainly Taraxacum and Cyperaceae.<br />

Deciduous Quercus forest decrease and Fagus virtually disappears from this<br />

region.<br />

237<br />

F. Naughton, 2007


Fig. V.2 | Lithology and synthetic pollen diagram against depth (cm). From left to right: radiocarbon and<br />

calibrated ages; lithology (after Folliot, 2004) including T. communis level (represented by small shells);<br />

dinocyst percentages (Operculodnium centrocarpum; Total of Spiniferites and Lingulodinium<br />

machaerophorum); pollen diagram and pollen zones.<br />

238<br />

F. Naughton, 2007


Fig. V.3 | Pollen diagram and quantitative pollen-based climate estimates against depth (cm). From left to<br />

right: calibrated ages; selected pollen taxa from the synthetic pollen diagram (other deciduous trees<br />

include: Fraxinus excelsior-type, Tilia and Ulmus); climate parameters: PANN (mean annual precipitation);<br />

difference between the temperature of the warmest (MTWA) and the coldest (MTCO) months (seasonality)<br />

and TANN (mean annual temperatures). Dashed lines represent maxima (bold) and minima values and the<br />

dark line represents mean values. Grey dashed lines represent the tendency of each curve; pollen zones.<br />

239<br />

F. Naughton, 2007


5. 5 Climate variability in north-western France<br />

5. 5. 1 Long-term cooling pattern and the Holocene thermal maximum<br />

Vegetation changes and pollen-based quantitative climate estimates<br />

permit the detection of a small-amplitude long-term pattern of summer<br />

temperature decrease between 8850 and 2000-1000 yr cal BP.<br />

The long-term cooling is marked by a general trend of temperate and<br />

humid tree decline and by the increase of herbaceous plants. This long-term<br />

cooling is characterised by the gradual decrease in the MTWA (mean<br />

temperature of the warmest month) values (from 20.5° to 17.5° C) (Fig. V.4<br />

and Fig. V.2) coinciding with the general trend of mid-latitude summer<br />

insolation reduction until at least 2000 cal yr BP (Fig. V.4). The continuous<br />

decrease of seasonality follows the gradual increase of the precessional<br />

signal. Nonetheless the weak values in precession between 8855 and 8000 cal<br />

yr BP surprisingly coincides with an interval of particularly high seasonality<br />

suggesting that other mechanisms have probably amplified this precessional<br />

signal (see below). This suggests that long-term vegetation changes in the<br />

north-western France seem to respond directly to the Holocene orbital<br />

induced climatic variability on which human impact on vegetation was<br />

superimposed since at least 2000 cal yr BP. Temperate and humid forest<br />

decrease also mimics the general decreasing trend observed in the δ 18 O-<br />

isotope composition of the NorthGRIP ice-core (Johnsen et al., 2001) (Fig.<br />

V.4). These results are in agreement with the previous suggestion that forest<br />

recession along the Holocene might be the result of natural processes rather<br />

than a consequence of the human impact (Magri, 1995).<br />

Previous studies on sea surface conditions of the North Atlantic and<br />

Mediterranean regions have shown an apparent long-term cooling trend that<br />

was driven by northern high latitudes summer insolation decreases during the<br />

Holocene (Marchal et al., 2002; Andersen et al., 2004; Moros et al., 2004).<br />

Several climate models also suggest an orbitally-induced mechanism as the<br />

main forcing factor for the long-term climatic trend over the Holocene<br />

(Kutzbach and Gallimore, 1988; Crucifix et al., 2002; Weber and Oerlemans,<br />

2003; Renssen et al., 2005). Other authors, Lorenz et al., 2006, have compared<br />

global alkenone-derived sea-surface temperature (SST) data with transient<br />

climate simulations using a coupled atmosphere-ocean general circulation<br />

240<br />

F. Naughton, 2007


model (AOGMC) for the last 7000 yr cal BP (less instable climate period)<br />

suggesting that mid- to late Holocene long-term SST trends were driven by<br />

insolation changes. The general cooling trend of the Holocene starts<br />

generally during or after the well known Holocene thermal maximum (HTM).<br />

However, the Holocene warming that defines the HTM occurred on different<br />

times depending of places (Kaufman et al., 2004). Several studies on both<br />

North-Atlantic marine and Greenland ice cores detect the HTM period at the<br />

beginning of the Holocene (Andrews and Giraudeau, 2002; Marchal et al;<br />

2002; Duplessy et al., 2001; Kaufman et al., 2004; Knudsen et al., 2004; de<br />

Vernal et al., 2005) while others point to a later climatic optimum (Dahl-<br />

Jensen et al., 1998; Bauch et al., 2001; Johnsen et al., 2001; Levac et al., 2001;<br />

Kaplan et al., 2002; Solignac et al., 2004; Kaufman et al., 2004; Keigwin et al.,<br />

2005).<br />

Unfortunately VK03-58Bis shelf core does not cover the entire Holocene<br />

record. However, MTWA values were higher between 8855 and 8000 cal yr BP<br />

than between 8000 and 1000 cal yr BP contrasting with the MTCO (mean<br />

temperature of the coldest month) trends which show lower values during the<br />

late early-Holocene than during the mid- and late-Holocene (Fig. V.3 and Fig.<br />

V.4). This strong seasonal contrast between 8855 and 8000 cal yr BP likely<br />

favoured the development of Corylus woodlands at the expense of<br />

deciduous Quercus forest although MTWA values were high. Mild (lower<br />

seasonality) conditions which allowed the expansion of deciduous Quercus<br />

forest in north-western France occurred roughly between 8000 and 4000 cal<br />

yr BP. This period has been considered as the HTM in the westernmost part of<br />

central Europe because MTWA reconstruction shows higher values than those<br />

from nowadays (Davis et al., 2003). Other pollen-based climate estimates<br />

obtained from several northern European pollen sequences such as Lake<br />

Raigastvere, Lake Viitna, Lake Ruila in Estonia and Lake Flarken in Sweden<br />

(Seppä and Poska, 2004; Seppä et al., 2005) shows between 8000 and 4500<br />

cal yr BP higher TANN (mean annual temperatures) values than present-day,<br />

reflecting the HTM in those regions. In contrast, our climate estimates does not<br />

detect either higher MTWA or TANN than present day values but reduced<br />

seasonality between 8000 and 4500 cal yr BP. Furthermore, the continuous<br />

presence of Fagus within this period also suggests (Tinner and Lotter, 2001)<br />

241<br />

F. Naughton, 2007


esides a weak seasonality, cooler summers and moist conditions.<br />

Interestingly, during the HTM defined in central and northern Europe, the<br />

middle latitudes of Western Europe were not submitted to particular high<br />

temperatures precluding the identification of the HTM in this region during the<br />

last 8850 cal yr BP.<br />

Fig. V.4 | Correlation between vegetation changes, quantitative climate estimates, summer insolation at<br />

45° N and precessional signal (after Berger, 1978) and δ 18O-isotope composition of the NorthGRIP ice-core<br />

(Johnsen et al., 2001) during the Holocene. Temperate and humid trees include: Acer, Alnus, Betula,<br />

Corylus, Cupressaceae, deciduous Quercus, Fagus, Fraxinus excelsior-type, Pinus, Quercus ilex-type, Salix,<br />

Tilia and Ulmus while Brassicaceae, Caryophyllaceae, Asteraceae (including Aster- and Anthemis- types)<br />

and Taraxacum-type, Cyperaceae, Ericaceae and Calluna, Plantago, Poaceae and semi-desert plants<br />

(including Chenopodiaceae, Artemisia and Ephedra) are integrated in the herbaceous plants association.<br />

242<br />

F. Naughton, 2007


5. 5. 2 Sub-orbital climate variability<br />

Superimposed to the orbital induced long-term cooling pattern, pollen<br />

analysis and quantitative climate reconstruction from VK03-58Bis shelf core<br />

detects sub-orbital climatic variability during the last 8855 cal yr BP (Fig. V.5).<br />

Fig. V.5 | Correlation between selected pollen taxa, quantitative climate estimates (PANN, TANN, MTCO,<br />

MTWA and seasonality) and δ 18O-isotope composition of the NorthGRIP ice-core (Johnsen et al., 2001)<br />

during the Holocene. The 8.2 kyr event is represented by the dark grey bar which is superimposed to 8.6-<br />

8.0 kyr event represented by the light grey bar. Dark arrows indicate possible millennial-scale cooling<br />

events during the Holocene.<br />

243<br />

F. Naughton, 2007


5. 5. 2. 1 The multi-centennial-scale climate cooling and the 8.2 ka event<br />

The maximum expansion and subsequent decline of Corylus<br />

woodlands (between 8739-8062 cal yr BP) associated with a high seasonality<br />

episode in north-western France occurred synchronously with the following<br />

event succession: the last stages of the Laurentide Ice sheet decay, the<br />

catastrophic final drainage episodes of the “glacial lakes Agassiz-Ojibway”<br />

(Clarke et al., 2004) into the Hudson Bay, at around 8470 cal yr BP (error range<br />

of 8160- 8740 cal yr BP; Barber et al., 1999), and the consequent 8.2 kyr event<br />

(Teller et al., 2002; Clarke et al., 2004). The introduction of large amounts of<br />

freshwater into the North Atlantic (Alley et al., 1997; Clark et al., 2001) triggers<br />

an important decrease of sea surface temperatures (SST) earlier than the<br />

recorded isotopic signal of the 8.2 kyr event in the Greenland ice cores and<br />

lasting several centuries, between ~8900 to 8000 cal yr BP (Ellison et al., 2006).<br />

This multi-centennial SST cooling detected by the high resolution North<br />

Atlantic deep-sea core MD99-2251 (Fig. V.1) occurred roughly<br />

contemporaneously with the climate cooling defined by Rohling and Pälike<br />

(2005) (~8600 and 8000 cal yr BP) (Ellison et al., 2006). SST cooling (~8600 and<br />

8000 cal yr BP) has also been observed in other regions of the North Atlantic<br />

such as over the Laurentian Fan (Keigwin et al., 2005) and in the north of<br />

Iceland (Knudsen et al., 2004).<br />

The cooling and freshening of the surface ocean, that started at<br />

around 400-500 yr before the drastic 8.2 kyr event, is linked with the beginning<br />

of a long and gradual pattern of reduction in the flow speed of Iceland-<br />

Scotland Overflow water (ISOW), a component of the NADW, which attains<br />

the slowest flow speed at around 8290 yr cal BP and lasted 200 yr,<br />

concomitant with the 8.2 kyr event (Ellison et al., 2006). The introduction of<br />

large amounts of freshwater favoured the reduction of the North Atlantic<br />

Deep Water (NADW) formation (Clark et al., 2001) and the consequent<br />

weakening of the conveyor belt (Barber et al., 1999; Rahmstorf, 2002; Renssen<br />

et al., 2001). This mechanism has a great impact on the spread of winter sea-<br />

ice in the North Atlantic region playing an important role on seasonality<br />

increase (Denton et al., 2005).<br />

We propose that the amplified signal of seasonality in north-western<br />

France has been driven by the final episodes of Agassiz and Ojibway<br />

244<br />

F. Naughton, 2007


outbursts, through the winter sea ice expansion in the high latitudes of the<br />

North Atlantic region triggering the beginning of the maximum spread of<br />

Corylus woodlands (at around 8739 cal yr BP).<br />

The maximum Corylus woodlands (8739-8387 cal yr BP) expansion,<br />

related with colder winters, is almost synchronous with the T. communis level<br />

(8739-8479 cal yr BP) in VK03-58Bis shelf core (Fig. V.2) but also with its decline<br />

(8479 and 8387 cal yr BP). T. communis can locally occur in fine sandy beds of<br />

the south-western French shelf (Glemarec, 1969) although is commonly found<br />

nowadays in the Boreal marine biogeographical zone of the north Atlantic<br />

region, between 50 and 68 °N, further northern north than the Lusitanian<br />

region from where VK03-58Bis core was retrieved (Fig. V.6) (Funder et al.,<br />

2002). It is known that during the warmest phases of the early Eemian the<br />

boreal marine zone migrated further north through the Barents and Kara sea<br />

to the Taymyr (Funder et al., 2002). On the contrary during the early Holocene<br />

and, in particular, between 8739-8479 cal yr BP, when the north Atlantic sea-<br />

ice cover was most likely extended further south as the result of the decrease<br />

in winter temperatures, this Boreal biogeographical zone was probably<br />

deflected several degrees further south allowing the settlement of the T.<br />

communis off north-western France.<br />

Between 8479 and 8387 cal yr BP a drastic environmental change<br />

triggered the T. communis death and the decrease of the dinocyst<br />

Lingulodinium machaerophorum (Fig. V.2). This change can not be due to<br />

the relatively low decrease (1-3° C) of Holocene SST (Bond et al., 1997) and<br />

salinity because both species tolerate great amplitude changes (Funder et<br />

al., 2002; Turon, 1984; Lewis and Hallet, 1997). One regional event such as the<br />

opening of the English Channel (Fig. V.1) (9000-7500 cal yr BP, Lambeck, 1997;<br />

8500-8400 cal yr BP, Jiang et al., 1997; 8600-8500 cal yr BP, Gyllencreutz and<br />

Kissel, 2006) could be the main trigger for T. communis mortality and<br />

Lingulodinium machaerophorum decline. Indeed the opening of the English<br />

Channel contributed to a drastic hydrological, sedimentological as well as<br />

biological change in the north-western France (Bourillet et al., 2005) which<br />

probably affected the benthic and planktonic communities.<br />

245<br />

F. Naughton, 2007


Fig. V.6 | Present day and past marine biogeographical zones in the North-East Atlantic (adapted from<br />

Funder et al., 2002). Bold dashed lines represent the limits of the present-day marine biogeographical<br />

zones in the North-East Atlantic; Grey dashed lines represent: a) the northward displacement of the boreal<br />

southern limit during the early Eemian (Funder et al., 2002) and b) the southward displacement of the<br />

boreal southern limit during the during 8.6-8.0 kyr event (this work).<br />

The 8.2 kyr event is marked in north-western France by the drastic<br />

episode of Corylus forest decline (8387-8062 cal yr BP) (Fig. V.5) as already<br />

observed in central and northern Europe (Tinner and Lotter, 2001, 2006; Seppä<br />

and Poska, 2004; Veski et al., 2004; Seppä et al., 2005). Contemporaneously<br />

with the Corylus forest decline, PANN and MTCO decreases (of about 100<br />

mm, 2°C) and high seasonality remains important in north-western France<br />

(Fig. V.5). Modelling-data comparison (Wiersma and Renssen, 2005) and<br />

pollen-based quantitative climate estimates from Europe (Davis et al., 2003)<br />

also show a temperature reduction by at least 1°C. The drastic MOC<br />

(Meridional Overturning Circulation) reduction at 8.2 kyr BP, associated with<br />

slowest flow of the ISOW (Iceland-Scotland Overflow Water) (Ellison et al.,<br />

2006), has probably amplified the signal of the Greenland isotopic record<br />

contributing to the European temperature decrease which favoured the<br />

decline of the Corylus trees not only in north-western France but also in<br />

Central and Northern Europe. Decreasing seasonality following the 8.2 kyr<br />

246<br />

F. Naughton, 2007


event favoured deciduous Quercus expansion at the expense of Corylus<br />

woodlands.<br />

Furthermore, our data suggests a complex pattern of annual<br />

precipitation in north-western France during the multi-centennial cooling that<br />

encompasses the 8.2 kyr event (Fig. V.3 and Fig. V.5). High annual<br />

precipitations characterise the beginning and the end phases of this cooling<br />

episode bracketing a drier period. Lake level changes in lake Annecy reveal<br />

the same complex pattern around the 8.2 kyr event with two high levels<br />

separated by a low one (Magny et al., 2003). Furthermore these high lake<br />

levels, interpreted by the authors as two episodes of high precipitation, are<br />

associated with relatively low MTWA values (Magny et al., 2003; Magny and<br />

Bégeot, 2004).<br />

5. 5. 2. 2 Other possible millennial scale cooling episodes<br />

After the final episodes of the Agassiz and Ojibway outburst flooding<br />

climate became less instable and therefore millennial scale climatic events<br />

are less evident during the mid- and late- Holocene. Quantitative climate<br />

estimates from VK03-58Bis show a series of small amplitude millennial-scale<br />

variability after the 8.2 kyr event (Fig. V.5). Cooling events are marked in most<br />

cases by an increase of precipitation values and seasonality as well as by a<br />

slight decrease of MTCO values. Because we only have two radiocarbon<br />

dates levels for the last 8000 years it is difficult to correlate our events with<br />

other well dated and world wide cooling episodes (Mayewski et al., 2004).<br />

Nonetheless these events can probably be linked with the SST coolings<br />

detected in the North Atlantic (Bond et al., 1997; 2001) and with some of the<br />

ten high lake levels identified in several mid-European lacustrine records that<br />

occurred after the 8.2 kyr event (Magny, 2004).<br />

5. 6 Conclusions<br />

High resolution pollen analysis and quantitative climate reconstruction<br />

from VK03-58Bis shelf core allow the detection of a small-amplitude long-term<br />

cooling pattern as well as millennial-scale climate variability over the last 8850<br />

years in north-western France:<br />

247<br />

F. Naughton, 2007


- Both the gradual decrease of temperate and humid trees and MTWA (mean<br />

temperature of the warmest month) values follow the general trend of<br />

northern mid-latitude summer insolation reduction until at least 2000 cal yr BP.<br />

The general trend of seasonality decrease follows the gradual increase of<br />

precession;<br />

- The high seasonality conditions of north-western France (between 8739-8062<br />

cal yr BP) was concomitant with the multi-centennial-scale climate cooling<br />

encompassing the 8.2 kyr event. Orbital induced colder winters were likely<br />

amplified by the increase of winter sea ice cover in the high latitudes of the<br />

north Atlantic as the result of the final episodes of Agassiz and Ojibway<br />

outbursts and consequent gradual reduction of the MOC (Meridional<br />

Overturning Circulation). This increase of seasonality favoured the spread of<br />

Corylus woodlands at the expense of the deciduous Quercus forest between<br />

8739-8387 cal yr BP;<br />

- Superimposed on the multi-centennial-scale climate event an extreme<br />

winter cooling triggered the Corylus tree decline (8387-8062 cal yr BP) in north-<br />

western France and has been identified and related to the short-lived 8.2 kyr<br />

cooling event. Although seasonality remains important, winter temperature<br />

over Europe and Greenland dropped due certainly to the final drastic MOC<br />

reduction associated with slowest flow of the ISOW (Iceland-Scotland<br />

Overflow Water);<br />

- We have detected a complex pattern in annual precipitation within the<br />

multi-centennial-scale cooling (between 8739-8062 cal yr BP): a relatively dry<br />

period in north-western France was sandwiched by two episodes of wetness;<br />

- our study also suggests several small amplitude millennial-scale cooling after<br />

the 8.2 kyr event marked by an increase of PANN and seasonality as well as<br />

by a slight decrease of MTCO (mean temperature of the coldest month)<br />

values.<br />

Finally, high occurrence of the marine mollusc T. communis points to the<br />

migration of the Boreal biogeographical zone several degrees further south,<br />

between 8739-8479 cal yr BP, during the first part of the multi-centennial<br />

cooling episode. T. communis death and the decrease of the Lingulodinium<br />

machaerophorum dinocyst, between 8479 and 8387 cal yr BP, was certainly<br />

triggered by the opening of the English Channel.<br />

248<br />

F. Naughton, 2007


References:<br />

Allen, G.P. and Castaing, P., 1977. Carte de répartition des sédiments superficiels sur<br />

le plateau continental du Golfe de Gascogne. Bulletin de l'Institut de Géologie du<br />

Bassin d'Aquitaine. Talence France, 255-260.<br />

Alley, R.B. and Ágústsdóttir, A.M., 2005. The 8 k event: cause and consequences of a<br />

major Holocene abrupt climate change. Quaternary Science Reviews, 24 : 1123–<br />

1149.<br />

Alley, R.B., Mayewski, P.A., Sowers, T., Stuiver, M., Taylor, K.C. and Clark, P.U., 1997.<br />

Holocene climatic instability: a prominent, widespread event 8200 yr ago. Geology,<br />

25: 483-486.<br />

Alley, R.B., Marotzke, J., Nordhaus, W.D., Overpeck, J.T., Peteet, D.M., Pielke, R.A.,<br />

Pierrehumbert, R.T.Jr., Rhines, P.B., Stocker, T.F., Talley, L.D. and Wallace, J.M., 2003.<br />

Abrupt climate change. Science, 299: 2005-2010.<br />

Andersen, A., Koç, N., Jennings, A. and Andrews, J.T., 2004. Nonuniform response of<br />

the major surface currents in the Nordic Seas to Insolation forcing: implications for the<br />

Holocene climate variability. Paleoceanography, 19, 2003, DOI:<br />

10.1029/2002PA000873.<br />

Andrews, J.T. and Giraudeau, J., 2002. Multi-proxy records showing Holocene<br />

oceanographic variability: the inner N. Iceland Shelf (Húnaflói). Quaternary Science<br />

Reviews, 22: 175-193.<br />

Baldini, J.U.L., McDermott, F. and Fairchild, I.J., 2002. Structure of the 8200-year cold<br />

event revealed by a speleothem trace element record. Science, 296: 2203-2206.<br />

Barber, D.C., Dyke, A., Hillaire-Marcel, C., Jennings, A.E., Andrews, J.T., Kerwin, M.W.,<br />

Bilodeau, G., McNeely, R., Southon, J., Morehead, M.D. and Gagnon, J.-M., 1999.<br />

Forcing of the cold event of 8,200 years ago by catastrophic drainage of Laurentide<br />

lakes. Nature, 400: 344-348.<br />

Bauch, H.A., Erlenkeuser, H., Spielhagen, R.F., Struck, U., Matthiessen, J., Thiede, J. and<br />

Heinemeier, J., 2001. A multi-proxy reconstruction of the evolution of deep and<br />

surface waters in the subartic Nordic Seas over the last 30,000 yr. Quaternary Science<br />

Reviews, 20: 659-678.<br />

Berger, A., 1978. Long-term variations of daily insolation and Quaternary climatic<br />

changes. Journal of Atmospheric Sciences, 35, 2362-2367.<br />

Bianchi, G.G. and McCave, I.N., 1999. Holocene periodicity in North Atlantic climate<br />

and deep-ocean flow south of Iceland. Nature, 397: 515-517.<br />

Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P.,<br />

Cullen, H., Hajdas, I. and Bonani, G.l., 1997. A pervasive millenial-scale cycle in North<br />

Atlantic Holocene and Glacial Climates. Science, 278: 1257-1266.<br />

Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M., Showers, W., Hoffmann, S.,<br />

Lotti-Bond, R., Hajdas, I. and Bonani, G., 2001. Persistant solar influence on North<br />

Atlantic climate during the Holocene. Science, 294: 2130-2136.<br />

Bottema, S. and van Straaten, L.M.J.U., 1966. Malacology and palynology of two<br />

cores from the Adriatic sea floor. Marine Geolology, 4: 553-564.<br />

249<br />

F. Naughton, 2007


Bourillet J-F., Loubrieu B., Levron L., Apprioual R. and Weber O., 2002. Reconnaissance<br />

sédimentologique du nord de la Grande Vasière: résultats préliminaires de la mission<br />

Trophal 1. In Ifremer, Ed Colloque du Défi Golfe de Gascogne. Ifremer, Brest, 11-.13<br />

December 2002.<br />

Bourillet, J-F., Dubrulle, C., Goubert E., Jouanneau, J.-M., Cortijo, E., Weber, O.,<br />

Lesueur, P. and Vicaire, O., 2005. La Grande Vasière: architecture, mise en place et<br />

estimation des facteurs de son évolution. Ifremer Ed, Colloque Défi Golfe de<br />

Gascogne. Ifremer, Brest, 22-24 March 2005.<br />

Bourillet J-F., Jouanneau J-M., Macher C., Le Hir P. and Naughton F., 2006. “La<br />

Grande Vasière” mid-shelf mud belt: Holocene sedimentary structure, natural and<br />

anthropogenic impacts, X International Symposium on Oceanography of the Bay of<br />

Biscay, Vigo, Spain, April 19-21 2006, 131-134.<br />

Bradshaw, R.H.W and Hannon, G.E. 2004: The Holocene structure of North-west<br />

European temperate forest induced from palaeoecological data. O. Honnay, K.<br />

Verheyen, B. Bossuyt and M. Hermy (Eds.), Forest biodiversity: Lessons from History for<br />

conservation, 11-25pp.<br />

Castaing, P. and Jouanneau, J.-M., 1987. Les apports sédimentaires actuels d’origine<br />

continentale aux océans. Bulletin de l’Institut Géologique du Bassin d’Aquitaine 41:<br />

53–65.<br />

Clark, P.U., Marshall, S.J., Clarke, G.K.C., Hostetler, S.W., Licciardi, J.M. and Teller, J.T.,<br />

2001. Freshwater forcing of abrupt climate change during the last glaciation.<br />

Science, 293: 283-287.<br />

Clarke, G.K.C., Leverington, D.W., Teller, J.T. and Dyke, A.S., 2004. Paleohydraulics of<br />

the last outburst flood from glacial Lake Agassiz and the 8200 BP cold event.<br />

Quaternary Science Reviews, 23: 389–407.<br />

Crucifix, M., Loutre, M.-F., Tulkens, P., Fichetet, T. and Berger, A., 2002. Climate<br />

evolution during the Holocene: a study with an Earth system model of intermediate<br />

complexity. Climate Dynamics, 19: 43-60.<br />

Dahl-Jensen, D., Monsegaard, K., Gundestrup, N., Clow, G.D., Johnsen, S.J., Hansen,<br />

A.W. and Balling, N., 1998. Past temperatures directly from the Greenland ice sheet.<br />

Science, 282: 268–271.<br />

Davis, B.A.S., Brewer, S., Stevenson, A.C., Guiot J. and Data Contributors., 2003. The<br />

temperature of Europe during the Holocene reconstructed from pollen data.<br />

Quaternary Science Reviews, 22: 1701-1716.<br />

de Beaulieu, J.-L., Pons, A. and Reille, M., 1984. Histoire de la végétation, du climat et<br />

de l'action de l'homme dans le Massif Central français depuis 15.000 ans. Institut<br />

français de Pondichéry, 30: 27-32.<br />

de Vernal, A., Hillaire-Marcel, C. and Darby, D.A., 2005. Variability of sea ice cover in<br />

the Chukchi Sea (western Arctic Ocean) during the Holocene.<br />

Paleoceanography, 20, 4018, DOI: 10.1029/2005PA001157.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3: 16-27.<br />

250<br />

F. Naughton, 2007


de Vernal, A., Rochon, A., Turon, J.-L. and Matthiessen, J., 1998. Organic-walled<br />

dinoflagelatte cysts: Palynological tracers of sea-surface conditions in middle to high<br />

latitude marine environments. Geobios, 30(7): 905-920.<br />

Denton, G.H. and Karlén, W., 1973. Holocene climatic variations: their pattern and<br />

possible cause. Quaternary Research, 3: 155– 205.<br />

Denton, G.H., Alley, R.B., Comer, G.C. and Broecker, W.S., 2005. The role of seasonality<br />

in abrupt climate change. Quaternary Science Reviews, 24: 1159-1182.<br />

Desprat, S., 2005. Réponses climatiques marines et continentales du SW de l'Europe<br />

lors des derniers interglaciaires et des entrées en glaciation. PhD Thesis, Bordeaux 1<br />

University, Bordeaux (France), 282 pp.<br />

Duplessy, J.C., Ivanova, E., Murdmaa, I., Paterne, M. and Labeyrie, L., 2001. Holocene<br />

paleoceanography of the northern Barents Sea and variations of the northward heat<br />

transport by the Atlantic Ocean. Boreas, 30: 2– 16.<br />

Ellison, C.R.W., Chapman, M.R. and Hall, I.R., 2006. Surface and deep ocean<br />

interactions during the cold climate event 8200 years ago. Science, 312: 1929-1932.<br />

Folliot, B., 2004. La Grande Vasière: Etude sédimentologique de deux secteurs<br />

septentrionaux. DEA on Dynamique et Environnements Sédimentaires Lille-Caen-<br />

Rouen, France, 39pp.<br />

Funder, S., Demidov, I. and Yelovicheva, Y., 2002. Hydrography and mollusc faunas of<br />

the Baltic and the White Sea–North Sea seaway in the Eemian. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology, 184: 275-304.<br />

Glemarec, M., 1969. Les peuplements benthiques du plateau continental nord-<br />

Gascogne. PhD Thesis, Paris University, France, 167 pp.<br />

Groot, J.J. and Groot, C.R., 1966. Marine palynology: possibilities, limitations, problems.<br />

Marine Geology, 4: 387-395.<br />

Guiot, J., Pons, A., de Beaulieu, J.L. and Reille, M., 1989. A 140.000 years continental<br />

climate reconstruction from two European pollen records. Nature, 338: 309-313.<br />

Guiot, J., 1990. Methodology of the last climatic cycle reconstruction from pollen<br />

data. Palaeogeography, Palaeoclimatology, Palaeoecology, 80: 49-69.<br />

Gyllencreutz, R. and Kissel, C., 2006. Lateglacial and Holocene sediment sources and<br />

transport patterns in the Skagerrak interpreted from high-resolution magnetic<br />

properties and grain size data. Quaternary Science Reviews, 25: 1247-1263.<br />

Heusser, L.E. and Balsam, W.L., 1977. Pollen distribution in the N.E. Pacific ocean.<br />

Quaternary Research, 7: 45-62.<br />

Heusser, L.E. and Stock, C.E., 1984. Preparation techniques for concentrating pollen<br />

from marine sediments and other sediments with low pollen density. Palynology, 8:<br />

225-227.<br />

Hughen, K.A., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C., Blackwell,<br />

P.G., Buck, C.E., Burr, G., Cutler, K.B., Damon, P.E., Edwards, R.L., Fairbanks, R.G.,<br />

Friedrich, M., Guilderson, T.P., Kromer, B., McCormac, F.G., Manning, S., Bronk Ramsey,<br />

C., Reimer, P.J., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver, M., Talamo, S.,<br />

Taylor, F.W., van der Plicht, J. and Weyhenmeyer, C.E., 2004. Marine04 Marine<br />

Radiocarbon Age Calibration, 0-26 Cal Kyr BP. Radiocarbon, 46: 1059-1086.<br />

251<br />

F. Naughton, 2007


Jiang, H., Björck, S. and Knudsen, K.L., 1997. A palaeoclimatic and<br />

palaeoceanographic record of the last 11000 14 C years from the Skagerrak–Kattegat,<br />

northeastern Atlantic margin. The Holocene, 7: 301–310.<br />

Johnsen, S. Dahl-Jensen, D., Gundestrup, N., Steffensen, J.P., Clausen, H.B., Miller, H.,<br />

Masson-Delmotte, V., Sveinbjornsdottir, A.E. and White, J., 2001. Oxygen isotope and<br />

palaeotemperature records from six Greenland ice-core stations: Camp Century,<br />

Dye-3, GRIP, GISP2, Renland and NorthGRIP. Journal of Quaternary Science, 16: 299-<br />

307.<br />

Jouanneau, J.M., Weber, O., Cremer, M. and Castaing, P., 1999. Fine-grained<br />

sediment budget on the continental margin of the Bay of Biscay. Deep Sea Research<br />

Part II: Topical Studies in Oceanography, 46: 2205-2220.<br />

Kaplan, M., Wolfe, A. and Miller, G., 2002. Holocene environmental variability in<br />

southern Greenland inferred from lake sediments. Quaternary Research, 58: 149–159.<br />

Kaufman, D.S., Ager, T.A., Anderson, N.J., Anderson, P.M., Andrews, J.T., Bartlein, P.J.,<br />

Brubaker, L.B., Coats, L.L., Cwynar, L.C., Duvall., M.L., Dyke, A.S., Edwards, M.E., Eisner,<br />

W.R., Gejewski, K., Geirsdóttir, A., Hu, F.S., Jennings, A.E., Kaplan, M.R., Kerwin, M.W.,<br />

Lozhkin, A.V., MacDonald, G.M., Miller, G.H., Mock, C.J., Oswald, W.W., Otto-Bliesner,<br />

B.L., Porinchu, D.F., Rühland, K., Smol, J.P., Steig, E.J. and Wolfe, B.B., 2004. Holocene<br />

thermal maximum in the Western Arctic (0 to 180W). Quaternary Science Reviews, 23:<br />

529-560.<br />

Keigwin, L.D., Sachs, J.P., Rosenthal,Y. and Boyle, E.A., 2005. The 8200 year B.P. event<br />

in the slope water system, western subpolar North Atlantic. Palaeoceanography, 20:<br />

2003, DOI:10.1029/2004PA00174.<br />

Klitgaard-Kristensen, D., Sejrup, H.P., Haflidason, H., Johnsen, S. and Spurk, M., 1998. A<br />

regional 8200 cal yr BP cooling event in northwest Europe, induced by final stages of<br />

the Laurentide ice sheet deglaciation? Journal of Quaternary Science, 13: 165–169.<br />

Knudsen, K.L., Jiang, H., Jansen, E., Eiriksson, J., Heinemeier, J. and Seidenkrantz, M.S.<br />

,2004. Environmental changes off North Iceland during the deglaciation and the<br />

Holocene: foraminifera, diatoms and stable isotopes. Marine Micropaleontology, 50:<br />

273–305.<br />

Koreneva, E.V., 1966. Marine palynological researches in the U.S.S.R. Marine Geology,<br />

4: 565-574.<br />

Kutzbach, J. E. and Gallimore, R. G., 1988. Sensitivity of a coupled atmosphere/mixed<br />

layer ocean model to changes in orbital forcing at 9000 years BP. Journal of<br />

Geophysical Research, 93: 803–821.<br />

Lambeck, K., 1997. Sea-level change along the French atlantic and channel coasts<br />

since the time of the Last Glacial Maximum. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 129: 1-22.<br />

Lesueur, P., Jouanneau, J-M., Boust, D., Tastet, J-P. and Weber, O., 2001.<br />

Sedimentation rates and fluxes in the continental shelf mud fields in the Bay of Biscay<br />

(France). Continental Shelf Research, 21: 1383-1401.<br />

252<br />

F. Naughton, 2007


Lesueur P. and Klingebiel, A., 1976. Carte des sédiments superficiels du plateau<br />

continental du Golfe de Gascogne (échelle 1/500 000). BRGM-IFREMER Ed Carte<br />

Géologique de la marge continentale française, 23 pp.<br />

Levac, E., de Vernal, A. and Blake W.Jr., 2001. Sea-surface conditions in northernmost<br />

Baffin Bay during the Holocene: palynological evidence. Journal of Quaternary<br />

Science, 16: 353-363.<br />

Lewis, J. and Hallett, R., 1997. Lingulodinium polyedrum (Gonyaulax polyedrum), a<br />

blooming dinoflagellate. Oceanography and Marine Biology: an Annual Review, 35:<br />

97-161.<br />

Lorenz, S. J., Kim, J.-H., Rimbu, N., Schneider, R. R. and Lohmann G., 2006. Orbitally<br />

driven insolation forcing on Holocene climate trends: evidence from alkenone data<br />

and climate modelling. Paleoceanography, 21: 1002, DOI:10.1029/2005PA001152.<br />

Magny, M. and Bégeot, C., 2004. Hydrological changes in the European midlatitudes<br />

associated with freshwater outbursts from Lake Agassiz during the Younger Dryas<br />

event and the early Holocene. Quaternary Research, 61: 181–192.<br />

Magny, M., Bégeot, C., Guiot, J. and Peyron, O., 2003. Contrasting patterns of<br />

hydrological changes in Europe in response to Holocene climate cooling phases.<br />

Quaternary Science Reviews, 22: 1589–1596.<br />

Magny, M., 2004. Holocene climate variability as reflected by mid-European lakelevel<br />

fluctuations, and its probable impact on prehistoric human settlements.<br />

Quaternary International, 113: 65-79.<br />

Magri, D., 1995. Some questions on the late-Holocene vegetation of Europe. The<br />

Holocene, 5: 354-360.<br />

Marchal, O., Cacho, I., Stocker, T.F., Grimalt, J.O., Calvo, E., Martrat, B., Shackleton,<br />

N., Vautravers, M., Cortijo, E., van Kreveld, S., Andersson, C., Koç, N., Chapman, M.,<br />

Sbaffi, L., Duplessy, J.-C., Sarnthein, M., Turon, J.-L., Duprat, J. and Jansen, E., 2002.<br />

Apparent long-term cooling of the sea surface in the northeast Atlantic and<br />

Mediterranean during the Holocene. Quaternary Science Reviews, 21: 455-483.<br />

Mayewski, P.A., Meeker, L.D., Twickler, M.S., Whitlow, S., Yang, Q., Lyons, W.B. and<br />

Prentice, M., 1997. Major features and forcing of high-latitude northern hemisphere<br />

atmospheric circulation using a 110,000-year-long glaciochemical series. Journal of<br />

Geophysical Research, 102: 26345-26366.<br />

Mayewski, P.A., Rohling, E., Stager, C., Karlén, W., Maasch, K., Meeker, L.D., Meyerson,<br />

E., Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F.,<br />

Staubwasser, M. and Schneider, R., 2004. Holocene climate variability. Quaternary<br />

Research, 62: 243-255.<br />

McCave, I.N., 1972. Transport and Escape of Fine-Grained Sediment from Shelf Areas.<br />

In Swift, D.J.P., Duane, D.B.and Pilkey, O.H. Eds. Shelf Sediment Transport Process and<br />

Pattern, Dowden, Hutchinson and Ross, Inc., Stroudsburg, PA, 225-248.<br />

Migeon, S., Weber, O., Faugeres, J.C. and Saint-Paul, J., 1999. SCOPIX: A new imaging<br />

system for core analysis. Geo-Marine Letters, 18: 251-255.<br />

Moros, M., Emeis, K., Risebrobakken, B., Snowball, I., Kuipers, A., McManus, J. and<br />

Jansen, E., 2004. Sea surface temperatures and ice rafting in the Holocene North<br />

253<br />

F. Naughton, 2007


Atlantic: climate influences on northern Europe and Greenland. Quaternary Science<br />

Reviews, 23: 2113-2126.<br />

Muller, J., 1959. Palynology of recent Orinoco delta and shelf sediments.<br />

Micropaleontology, 5: 1-32.<br />

Muscheler, R., Beer, J. and Vonmoos, M., 2004. Causes and timing of the 8200 yr BP<br />

event inferred from the comparison of the GRIP 10 Be and the tree ring Δ 14 C record.<br />

Quaternary Science Reviews, 23: 2101-2111.<br />

Naughton, F., Sanchez Goni, M.F., Desprat, S., Turon, J.-L., Duprat, J., Malaizé, B., Joli,<br />

C., Cortijo, E., Drago, T. and Freitas, M.C., 2006. Present-day and past (last 25 000<br />

years) marine pollen signal off western Iberia. Marine Micropaleontology, in press.<br />

Nesje, A. and Dahl, S.O., 2001. The Greenland 8200 cal yr BP event detected in loss-on<br />

ignition profiles in Norwegian lacustrine sediment sequences. Journal of Quaternary<br />

Science, 16: 155–166.<br />

O’Brien, S.R., Mayewski, P.A., Meeker, L.D., Meese, D.A., Twickler, M.S. and Whitlow,<br />

S.I., 1995. Complexity of Holocene climate as reconstructed from a Greenland ice<br />

core. Science, 270: 1962– 1964.<br />

Peck, R.M., 1973. Pollen budget studies in a small Yorkshire catchment. In Birks, H.J.B.<br />

and West, R.G. Eds. Quaternary Plant Ecology, The 14 th symposium of the British<br />

Ecological Society, University of Cambridge, Blackwell Scientific Publications, 43-60.<br />

Peyron, O., Guiot, J., Cheddadi, R., Tarasov, P., Reille, M., de Beaulieu, J.-L., Bottema,<br />

S. and Andrieu, V., 1998. Climatic Reconstrustion in Europe for 18,000 yr BP from Pollen<br />

Data. Quaternary Research, 49: 183-196.<br />

Peyron, O., Bégeot, C., Brewer, S., Heiri, O., Magny, M., Millet, L., Ruffaldi, P., Van<br />

Campo E. and Yu, G., 2005. Lateglacial climatic changes in Eastern France (Lake<br />

Lautrey) from pollen, lake-levels, and chironomids. Quaternary Research, 64: 197-211.<br />

Pinot, J-P., 1974: Le précontinent Breton entre Penmarch, Belle Île et l’escarpement<br />

continental. Etude géomorphologique. PhD Thesis, Brest University, France, 256 pp.<br />

Pons A. and Reille, M., 1986. Nouvelles recherches pollenanalytiques à Padul<br />

(Granada): La fin du dernier glaciaire et l’Holocene. E. López-Vera (Ed.). Quaternary<br />

climate in Western Mediterranean, University Autónoma de Madrid, 405-420 pp.<br />

Rahmstorf, S., 2002. Ocean circulation and climate during the past 120,000 years.<br />

Nature 419: 207-214.<br />

Reille, M. and Andrieu, V., 1991. Données nouvelles sur l'histoire postglaciaire de la<br />

végétation des Pyrénées occidentales (France). Comptes Rendues de l'Académie<br />

de Sciences, 312: 97-103.<br />

Reille, M. and Andrieu, V., 1994. Vegetation history and human action in Ariege<br />

(Pyrenees, France). Dissertationes Botanicae, 234: 413–422.<br />

Reille, M. and Andrieu, V., 1995. The late Pleistocene and Holocene in the Lourdes<br />

basin, western Pyrenees, France: new pollen analytical and chronological data.<br />

Vegetation History and Archaeobotany, 4: 1-21.<br />

254<br />

F. Naughton, 2007


Reille, M. and Lowe, J.J., 1993. A re-evaluation of the vegetation history of the eastern<br />

Pyrenees (France) from the end of the last glacial to the present. Quaternary Science<br />

Reviews, 12: 47-77.<br />

Reille, M., de Beaulieu, J.L., Svobodova, H., Andrieu-Ponel, V. and Goeury, C., 2000.<br />

Pollen analytical biostratigraphy of the last five climatic cycles from a long<br />

continental sequence from the Velay region (Massif Central, France). Journal of<br />

Quaternary Science, 15: 665–685.<br />

Reille, M., 1990a. Novelles recherches pollenanalytiques à Freychinede, Pyrénées<br />

ariégeoises, France. Laboratoire de Botanique Historique et Palynologie, Marseille,<br />

Unpublished report, November, 1990, 10pp.<br />

Reille, M., 1990b. Novelles recherches pollenanalytiques dans le bassin de Lourdes<br />

(Pyrénées Occidentales, France). Laboratoire de Botanique Historique et Palynologie,<br />

Marseille, Unpublished report, December, 1990, 16pp.<br />

Renssen, H., Goosse, H., Fichefet, T. and Campin, J.M., 2001. The 8.2 kyr BP event<br />

simulated by a global atmosphere–sea–ice–ocean model. Geophysical Research<br />

Letters, 28: 1567-1570.<br />

Renssen, H., Goosse, H., Fichefet, T., Brovkin, V., Driesschaert, E. and Wolk, F., 2005.<br />

Simulating the Holocene climate evolution at northern high latitudes using a coupled<br />

atmosphere-sea ice-ocean-vegetation mode. Climate Dynamics, 24: 23-43.<br />

Rind, D., Chandler, M., Lerner, J., Martinson, D.G. and Yuan, X., 2001. Climate<br />

response to basin-specific changes in latitudinal temperature gradients and<br />

implications for sea ice variability. Journal of Geophysical Research, 106: 20161-20173.<br />

Risebrobakken, B., Jansen, E., Andersson, C., Mjelde, E. and Hevroy, K., 2003. A highresolution<br />

study of Holocene <strong>paleoclimat</strong>ic and paleoceanographic changes in the<br />

Nordic seas. Paleoceanography, 18, 1017, doi: 10.1029/2002PA000764.<br />

Rochon, A., de Vernal, A., Turon, J-L., Matthiessen, J. and Head, M., 1999. Distribution<br />

of dinoflagellate cyst assemblages in surface sediments from the North Atlantic<br />

Ocean and adjacent basins and quantitative reconstruction of sea-surface<br />

parameters. American Association of Stratigraphic Palynologists, Contribution Series<br />

No. 35.<br />

Rohling, E.J. and Pälike, H., 2005. Centennial-scale climate cooling with a sudden<br />

cold event around 8,200 years ago. Nature, 434: 975-979.<br />

Seppä, H. and Poska, A., 2004. Holocene annual mean temperature changes in<br />

Estonia and their relationship to solar insolation and atmospheric circulation patterns.<br />

Quaternary Research, 61: 22-31.<br />

Seppä, H., Hammarlund, D. and Antonsson, K., 2005. Low-frequency and highfrequency<br />

changes of temperature and effective humidity during the Holocene in<br />

South central Sweden: Implications for atmospheric and oceanic forcings of climate.<br />

Climate Dynamics, 25: 285-297.<br />

Solignac, S., de Vernal, A. and Hillaire-Marcel, C., 2004. Holocene sea-surface<br />

conditions in the North Atlantic—contrasted trends and regimes in the western and<br />

eastern sectors (Labrador Sea vs. Iceland Basin). Quaternary Science Reviews, 23:<br />

319–334.<br />

255<br />

F. Naughton, 2007


Stuiver, M. and Reimer, P., 1993. Extended 14C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

Stuiver, M., Reimer, P.J. and Reimer, R.W., 2005. CALIB 5.0. (WWW program and<br />

documentation).<br />

Tallantire, P.A., 2002. The early Holocene spread of hazel (Corylus avellana L.) in<br />

Europe north and west of the Alps: an ecological hypothesis. The Holocene, 12: 81-96.<br />

Telford, R.J., Heegaard, E. and Birks, H.J.B., 2004. The intercept is a poor estimate of a<br />

calibrated radiocarbon age. The Holocene, 14: 296-298.<br />

Teller, J.T., Leverington, D.W. and Mann, J.D., 2002. Freshwater outbursts to the oceans<br />

from glacial Lake Agassiz and their role in climate change during the last<br />

deglaciation. Quaternary Science Reviews, 21: 879–887.<br />

Tinner, W. and Lotter, A.F., 2001. Central European vegetation response to abrupt<br />

climate change at 8.2 ka. Geology, 29: 551–554.<br />

Tinner, W. and Lotter, A.F., 2006. Holocene expansions of Fagus silvatica and Abies<br />

alba in Central Europe: where are we after eight decades of debate? Quaternary<br />

Science Reviews, 25: 526-549.<br />

Turon, J.-L., 1984. Le palynoplancton dans l'environnement actuel de l'Atlantique<br />

nord-oriental. Evolution climatique et hydrologique depuis le dernier maximum<br />

glaciaire. Mémoires de l'Institut de Géologie du bassin d'Aquitaine, 17. Université de<br />

Bordeaux I, Bordeaux, 313 pp.<br />

Van der Kaars, S. and de Deckker, P., 2003. Pollen distribution in marine surface<br />

sediments offshore Western Australia. Review of Palaeobotany and Palynology, 124:<br />

113-129.<br />

Van Geel, B., van der Plicht, J. and Renssen, H., 2003. Major Δ 14 C excursions during<br />

the late glacial and early Holocene: changes in ocean ventilation or solar forcing of<br />

climate change? Quaternary International, 105: 71-76.<br />

Veski, S., Seppä, H. and Ojala, A.E.K., 2004. Cold event at 8200 yr BP recorded in<br />

annually laminated lake sediments in eastern Europe. Geology, 32: 681-684.<br />

Von Grafenstein, U., Erlenkeuser, H., Müller, J., Jouzel, J. and Johnsen, S., 1998. The<br />

cold event 8200 years ago documented in oxygen isotope records of precipitation in<br />

Europe and Greenland. Climate Dynamics, 14: 73-81.<br />

Von Grafenstein, U., Erlenkeuser, H., Brauer, A., Jouzel, J. and Johnsen, S.J., 1999. A<br />

mid-European decadal isotope-climate record from 15,500 to 5000 years B.P.<br />

Science, 284: 1654-1657.<br />

Weber, S.L. and Oerlemans, J., 2003. Holocene glacier variability: three case studies<br />

using an intermediate-complexity climate model. The Holocene, 13: 353-363.<br />

Wiersma, A.P. and Renssen, H., 2005. Model-data comparison for the 8.2 ka BP event:<br />

confirmation of a forcing mechanism by catastrophic drainage of Laurentide Lakes.<br />

Quaternary Science Reviews, 25: 63-88.<br />

256<br />

F. Naughton, 2007


Capítulo 6| HOLOC<strong>EN</strong>E CHANGES IN THE DOURO ESTUARY<br />

(NORTHWESTERN IBERIA)<br />

Mudanças Holocénicas no Estuário do Douro (Noroeste<br />

Ibérico)<br />

Changements au cours de l’Holocène dans l’estuaire du<br />

Douro (Nord-ouest de la Péninsule Ibérique)<br />

Journal of Coastal Research<br />

In press. 2006<br />

F. Naughton a, b, c , M.F. Sánchez Goñi a , T. Drago b , M.C. Freitas c , A. Oliveira d<br />

a Environnements et Paléoenvironnements Océaniques (UMR CNRS 5805 EPOC),<br />

Université Bordeaux 1, Av. des Facultés, 33405 Talence, France<br />

b Centro Regional de Investigação Pesqueira do Sul , Instituto Nacional de<br />

Investigação Agrária e Pescas (INIAP) (IPIMAR-CRIPSUL), Av. 5 de Outubro, 8700-305<br />

Olhão, Portugal<br />

c Departamento e Centro de Geologia -Universidade de Lisboa, Bloco C6, 3º piso,<br />

Campo Grande, 1749-016 Lisboa, Portugal<br />

d Instituto Hidrográfico (IH), Rua das Trinas, 29, 1100 Lisboa, Portugal<br />

257<br />

F. Naughton, 2007


Resumo<br />

O estudo sedimentológico e palinológico de uma sondagem de 20 m<br />

colhida na embocadura do estuário do Douro (noroeste de Portugal) mostra<br />

uma série de variações que ocorreram ao longo do Holocénico. A presença<br />

de uma floresta regional composta por Pinus-Quercus-Alnus, entre 10 720 e 6<br />

530 cal BP, mostra que, o princípio do Holocénico seria caracterizado por um<br />

clima quente e húmido. A presença e o aumento, até ao final deste período<br />

(10 720 e 6 530 cal BP), de associações de foraminíferos marinhos<br />

(plataforma e talude) e de equinodermes, reflecte uma subida gradual do<br />

nível do mar.<br />

A atenuação da subida do nível do mar e a presença de um período<br />

de forte hidrodinamismo do rio permitiram a formação de uma barreira<br />

cascalhenta na embocadura sul do estuário. A variação radical entre<br />

associações polínicas provenientes da vegetação regional (composta<br />

essencialmente por árvores), transportadas pelo rio, e espectros polínicos<br />

provenientes da vegetação local (composto essencialmente por<br />

Ericaceae/Poaceae) ocorreu entre 6 530 e 1 500 cal BP e, é contemporânea<br />

ao estabelecimento dessa barreira de cascalho. Esta mudança de<br />

assinatura polínica sugere que a migração de rio para norte, testemunhada<br />

pela existência de um paleovale cujo eixo se situa a sul do presente canal,<br />

ocorreu após 6 530 cal BP.<br />

Résumé<br />

L’étude sédimentologique et palynologique d’une carotte de 20 m de<br />

long prélevée dans l’embouchure de l’estuaire du Douro (nord-ouest de<br />

Portugal), a permis de détecter une série de variations environnementales au<br />

cours de l’Holocène. Les résultats montrent que le début de l’Holocène (10<br />

720-6 530 ans cal BP) est marqué par un climat chaud et humide, caractérisé<br />

par le développement d’une forêt régionale de Pinus-Quercus-Alnus.<br />

L’augmentation graduelle de la concentration de foraminifères benthiques,<br />

typiques des environnements de plateforme et de talus continental, ainsi que<br />

la présence des échinodermes à la fin de cette période suggère<br />

l’augmentation du niveau marin entre 11 500 et 6 000 ans cal BP.<br />

258<br />

F. Naughton, 2007


L’atténuation de la montée du niveau marin et l’augmentation de<br />

l’hydrodynamisme de la rivière, suggérées par cette étude indique le<br />

développement d’une barrière de gravier dans la partie Sud de cet estuaire<br />

a partir de 6535 ans cal BP<br />

La variation radicale entre des associations polliniques provenant de la<br />

végétation régionale (composées essentiellement d’arbres) et transportées<br />

donc par la rivière, et des spectres polliniques provenant de la végétation<br />

locale (essentiellement Ericaceae et Poaceae) est datée entre 6 530 et 1 500<br />

ans cal BP, ce qui est contemporain de l’établissement de la barrière de<br />

gravier. Ces résultats suggèrent que la migration vers le nord du chenal<br />

principal de la rivière, précédemment démontrée par l’existence d’une<br />

paléo-valley au sud de l’actuel chenal principal, a pu avoir lieu aussi tôt que<br />

6 530 ans cal BP.<br />

Abstract<br />

Holocene changes are recorded by sedimentology and palynology on<br />

a 20 m long core retrieved in the mouth of the Douro estuary (north-western<br />

Portugal). Results show that the early Holocene (10 720-6 530 cal yr BP) was<br />

characterised by a warm and humid climate as testified by a well-established<br />

Pinus-Quercus-Alnus regional forest. Shelf and slope foraminifera assemblages<br />

as well as echinoderms gradually increased towards the end of this period<br />

reflecting the sea-level rise, which occurred between 11 500 and 6 000 cal yr<br />

BP.<br />

A gravel barrier developed in the southern part of the estuary as a<br />

result of sea-level rise attenuation and strong hydrodynamism of the river. A<br />

radical change from regional fluvially transported pollen assemblages (mainly<br />

composed of trees) to pollen spectra derived from local vegetation (mainly<br />

Ericaceae/Poaceae) occurred between 6 530 and 1 500 cal yr BP,<br />

contemporaneously to the settlement of the gravel barrier. This suggests that<br />

the northward migration of the river main channel, already testified by the<br />

existence of a palaeovalley with its axis located southward of the present<br />

main channel, occurred as early as 6 530 cal yr BP.<br />

259<br />

F. Naughton, 2007


260<br />

F. Naughton, 2007


6. 1 Introduction<br />

Coastal ecosystems are highly productive interfaces in permanently<br />

changing processes controlled by global factors (glacio-eustatic, climatic<br />

and sea-level changes) and local factors, such as sand barrier dynamics or<br />

anthropogenic impact. Only the correct interpretation of past ecological and<br />

geomorphological changes will allow us to predict future coastline responses<br />

to these global and local factors.<br />

In the middle of the Holocene, during sea-level rise attenuation, the<br />

geomorphological evolution of some coastal areas depended more on local<br />

factors (Devoy et al., 1996; Zong, 2004), such as sediment availability or land<br />

management practices (Devoy et al., 1996), than on major sea-level and<br />

climate changes. Recent studies from coastal systems in southern Portugal<br />

show that sea-level rise attenuation during the mid-Holocene is partly<br />

responsible for sand barrier formation in coastal depressions (Bao et al., 1999).<br />

As a consequence, a beach-barrier-lagoon environment has been<br />

established and its evolution, after 5000-4000 BP, depends essentially on local<br />

parameters, such as barrier permeability (Freitas et al., 2002, 2003; Freitas,<br />

1995: Freitas and Andrade, 1997).<br />

Our knowledge of coastal evolution in northern Portugal is based<br />

almost exclusively on studies concerning the continental shelf. These studies<br />

reveal sea-level changes (Magalhães, 2000), coast line evolution (Dias et al.,<br />

2000) and sedimentary dynamics (Dias, 1987; Dias et al., 2002a; 2002b; Drago,<br />

1995; Jouanneau et al., 2002; Magalhães, 1999; Oliveira et al, 2002). However,<br />

and up to present only few studies have been carried out on the<br />

geomorphological evolution of the northern coastal systems in response to<br />

global and local Holocene changes (Granja and de Groot, 1996; Granja,<br />

1999).<br />

The aim of this work is to fill this gap by documenting the history of the<br />

Douro estuary during the Holocene. This estuary is repeatedly overflowed by<br />

the Douro river, one of the most important sediment suppliers of the<br />

Portuguese northern coast and adjacent continental shelf (Dias et al., 2002a;<br />

2002b; Drago, 1995; Oliveira et al., 2002; Magalhães, 2000; Thouveny et al.,<br />

2000).<br />

261<br />

F. Naughton, 2007


The ancient course of this river followed a northeast-southwest<br />

direction (Carvalho and Rosa, 1988) but, at present, the river follows an east-<br />

west orientation in the estuary zone. Some authors have tentatively proposed<br />

that this change of flow direction resulted from the development of a sand<br />

barrier in the mouth of the estuary (Ferreira et al., 1989). However, the<br />

chronology of both the barrier formation and the river migration remains<br />

unknown. The causes triggering these changes and their impact on the<br />

estuary geomorphology are also unknown.<br />

This is why our sedimentary and palynological approach focuses on<br />

the southern area of the Douro estuary, in S. Paio bay. We discuss the different<br />

geomorphological processes occurring in this area during the Holocene,<br />

based on texture, mineralogy and composition of sediments (carbonates and<br />

organic matter). Pollen analysis will provide, in turn, useful information on<br />

variations of the river course over the last 10000 years.<br />

6. 2 Environmental Setting<br />

The Douro estuary is located in the western limit of the Douro basin, in<br />

northern Iberia (Fig. VI.1a). This is a narrow funnelled estuary, 2250 m long by<br />

1250 m wide, partially enclosed by a sand barrier formation in the southern<br />

area (Cabedelo), close to Vila Nova de Gaia and the palaeo-valley zone<br />

defined by Carvalho and Rosa (1988) (Fig. VI.1b).<br />

The lithology of the Douro basin is composed essentially of granite,<br />

schist, gneiss and quartzite rocks (Carrington da Costa and Teixeira, 1957);<br />

present-day climate varies from very humid in the west to semi-arid towards<br />

the east. In the western part of the Douro basin, the wet season (October-<br />

March) covers 73% of the total annual precipitation (Loureiro et al., 1986).<br />

The present-day vegetation of the Douro basin clearly illustrates both<br />

Mediterranean and Atlantic influences (Pina Manique, 1957, Costa et al.,<br />

1998, Navarro Andrés and Valle Gutiérrez, 1987). In northeastern Portugal, the<br />

continental climate favours evergreen oaks (Quercus ilex and Quercus<br />

suber), deciduous oaks (Quercus pyrenaica and Quercus faginea) and<br />

Juniperus spp.. Ericaceous and Cistaceous species dominate the understory<br />

vegetation, a consequence of intense anthropogenic activities. The<br />

262<br />

F. Naughton, 2007


vegetation is similar in the eastern part of the Douro basin. However, cork oak<br />

(Q. suber) is absent here. The oceanic influence is particularly important in the<br />

northwest of the basin, where the Quercion occidentale (Quercus robur in<br />

association with Q. suber) predominates (Braun-Blanquet et al., 1956). The<br />

spread of both Pinus pinaster and Eucalyptus globulus has been favoured by<br />

people. The understory vegetation is largely dominated by Ulex, in association<br />

with heathers. The river margins are colonized by Alnus glutinosa, Fraxinus<br />

angustifolia, Ulmus spp., Salix spp. and Populus spp. and the estuary<br />

surroundings are dominated essentially by Poaceae and Ericaceae.<br />

Annual river sediment charge was nearly 1.8x10 6 m 3 under natural<br />

conditions prior to the construction of dams in the last 40 years (0.25 x10 6 m 3 ).<br />

The mean annual draining is 22.400x10 6 m 3 , equivalent to 710 m 3 /s of mean<br />

annual discharge, with a maximum of 3000 m 3 /s and a minimum of 50 m 3 /s<br />

(Loureiro et al., 1986).<br />

Fig. VI.1 | a) Douro estuary localisation in the Iberian Peninsula. b) Core (1, 1B and 2) and surface sampling<br />

sites. Dark circles represent core sites and white circles surface samples sites. Dark lines represent the<br />

palaeoisobathic curves defined by Carvalho and Rosa (1988). Dashed line represents the ancient direction<br />

of the river main channel flow and bold dark line the present day river main channel flow. This<br />

palaeobathymetric map shows the palaeovalley of the Douro river.<br />

263<br />

F. Naughton, 2007


6. 3 Material and methods<br />

Two sediment cores (1 and 1B) were retrieved by rotary perforation in<br />

the southern edge of Douro estuary, near the river mouth (S.Paio Bay in the<br />

Douro estuary) (Fig. VI.1b). Another core, core 2, was retrieved in the sand<br />

barrier, at ~500 m distance from cores 1 and 1B. Core 2 attains -39.70 m OD<br />

(Ordnance Datum) and covers the Lateglacial/Holocene transition.<br />

Sedimentological, geochemical and micropalaeontological (excluding<br />

pollen) analyses are discussed in Drago et al. (in press).<br />

Core 1 reached a depth of 7.90 m (-4.4 m OD), hitting a gravel unit<br />

which stopped drilling. Coring was then moved less than 50 cm from the<br />

original site and a new perforation (Core 1B) reach 19.70 m (-16.20 m OD) of<br />

depth. Correlation of both cores is presented in (Fig. VI.2). The basal rock<br />

appears between 19.70 and 17.40 m (-16.20 m/ -13.90 m OD) and is<br />

constituted by altered granite.<br />

Fig. VI.2 | Lithlogy of the Douro sequence. Depth is represented in Ordnance Datum which is a coordinate<br />

system for heights above mean sea level (optometric heights). Five calibrated ages are also represented<br />

along the two cores.<br />

264<br />

F. Naughton, 2007


The sequence was sub-sampled for different analyses, including<br />

sedimentary characterization (texture, mineralogy of coarser sediments,<br />

organic matter and carbonate content), palynology and AMS radiocarbon<br />

dating.<br />

6. 3. 1 Radiometric dating<br />

Five organic levels (1 cm-thick) from both cores (1 and 1B) (Fig. VI.2)<br />

were chosen according to drastic granulometry changes to be dated by<br />

AMS 14 C (Beta Analytic, USA). Calibration of radiocarbon dates was<br />

effectuated by Beta Analytic by using INTCAL98 (STUIVER et al., 1998) (Tab.<br />

VI.1).<br />

Laboratory<br />

code<br />

Depth (m)<br />

OD<br />

Depth (m) δ 13 C/δ 12 C<br />

265<br />

Date<br />

( 14 C years BP)<br />

Calibrated age<br />

(cal years BP) (2δ)<br />

Beta-164370 -0.60/-0.61 4.14-4.15 -24.4 o/oo 1110 ±40 990<br />

Beta-154312 -1.67/-1.68 5.17-5.19 -26.2 o/oo 1580 ± 40 1500<br />

Beta-154313 -7.14/-7.15 10.64-10.65 -26.6 o/oo 5750 ± 40 6530<br />

Beta-174809 -8.33/-8.34 11.83-11.84 -25.2 o/oo 6050+-60 6890<br />

Beta-154314 -13.90 /13.91 17.40-17.41 -25.6 o/oo 9490 ± 60 10720<br />

Tab. VI.1 | Radiocarbon and calibrated dates from the site under study.<br />

6. 3. 2 Sedimentological analyses<br />

Sixty four samples were selected according to lithological changes<br />

observed during core description.<br />

Sediment texture was determined by using the Flemming classification<br />

(Flemming, 2000). Sand mineralogy was determined by means of<br />

stereomicroscope observation (x500 magnification), following the<br />

methodology described by Dias (1987) and Magalhães (1993, 1999).<br />

The sand samples were separated, by dry-sieving, in five different<br />

fractions (63-125 µm, 125-250 µm, 250-500 µm, 500-1000 µm, 1000-2000 µm). A<br />

minimum of 300 grains was counted for each fraction. Quartz, micas,<br />

aggregates and other terrigenous represent the mineralogical association of<br />

F. Naughton, 2007


fluvial origin while bioclasts (foraminifera and echinoderms) indicate marine<br />

origin.<br />

Gravel study followed the methodology by Dobbkins and Folk (1970).<br />

221 pebbles were randomly selected along the gravel unit and for each one<br />

we measured width (l), length (L), thickness (E) and sphericity (R1) and we<br />

determinate roundness (RK) and effective sphericity (Ee).<br />

Organic matter was calculated by loss on ignition (Craft et al., 1991)<br />

and carbonate content by using a Bernard calcimeter (Hulseman, 1966).<br />

6. 3. 3 Micropalaeontological analyses<br />

The treatment used for palynological analysis follows de Vernal et al.<br />

(1996), slightly modified by the Département de Géologie et Oceanographie<br />

(DGO), University Bordeaux I.<br />

Palynological treatment consists on pollen concentration by chemical<br />

digestion using HCl (at 10%, 25% and 50%) and HF (at 40% and 70%) to<br />

eliminate carbonates and silicates, respectively. Samples with very high<br />

content of organic matter were treated with KOH (at 10%). All samples were<br />

sieved through 10 µm nylon mesh screens and the residue was mounted in<br />

bidistillate glycerine. A Zeiss microscope with x550 and x1250 (immersion)<br />

magnifications was used for pollen observation and counting.<br />

Pollen identifications were achieved via comparison with specialist<br />

atlases (Reille, 1992) as well as with the DGO pollen reference collection. At<br />

least 350 pollen grains (excluding aquatic plants and spores) and 100<br />

Lycopodium grains (exotic pollen introduced during the sample preparation)<br />

were counted in each of the seventy seven samples analysed to obtain<br />

statistically reliable pollen spectra (Rull, 1987; Maher, 1981). Pollen<br />

percentages were calculated based on the main pollen sum which excludes<br />

aquatic plants, spores, indeterminate and unknown pollen. We also<br />

determined arboreal pollen (AP) and Non Arboreal pollen (NAP) percentages<br />

excluding Pinus pollen from the counts. Pinus is often overrepresented in<br />

coastal records (Heusser, 1978).<br />

Present-day pollen percentages supply from Douro river into the study<br />

area was also determinate to help us recognize the fossil assemblage mainly<br />

derived from fluvial transport. This study comprised five surface (Laquasup)<br />

266<br />

F. Naughton, 2007


samples from the left side (southern part) of the estuary, deposited after the<br />

Douro overflow in November 2000 (Fig. VI.1b).<br />

6. 4 Results and Discussion<br />

6. 4. 1 Chronology<br />

AMS 14 C dates obtained from the Douro sequence indicate that this<br />

sequence covers most of the Holocene, from c. 10720 cal yr BP up to present<br />

(Tab. VI.1).<br />

6. 4. 2 Holocene sedimentary processes in the Douro estuary<br />

Results from textural analyses, organic matter and carbonate content<br />

as well as the mineralogy of sand allow the identification of three main<br />

sedimentary units (SED2, SED3 and SED4) (Fig. VI.3). SED1 unit has been<br />

defined in the adjacent core 2 (Drago et al., in press) as representing the<br />

Lateglacial period.<br />

SED2 (-13.90 m / -6.81 m OD depth), c. 10720-6530 cal yr BP, is<br />

characterized essentially by slightly muddy sand between -13.90 m and -9.89<br />

m and by sandy mud between -9.89 m and -6.81 m. The terrigenous content<br />

(quartz, mica and aggregates) is higher (70-100%) than the enclosed<br />

bioclastic grains (0-30%), suggesting that the river sediment supply was<br />

important in the southern part of the estuary. However, carbonate content<br />

and the increase of bioclastic fragments (benthic and planktonic foraminifers<br />

and echinoderms) towards the top of this unit (between -9.89 m and -6.81 m)<br />

suggest an increase of marine influence. The succession from estuarine to<br />

shelf and slope foraminifera assemblages over this period indicates gradual<br />

sea-level rise (Drago et al., in press) and therefore a coast line retreat that<br />

characterizes this time period (Dias et al., 2000; Granja and de Groot, 1996).<br />

Also, the relatively high content of organic matter, towards the top of this unit,<br />

is probably linked to an attenuation of the hydrodynamic behaviour of the<br />

river, favouring sandy mud deposition. The increase of micas supports this<br />

weaker hydrodynamism which is, however, repeatedly interrupted by small<br />

episodes of stronger hydrodynamism, as indicated by the increase of quartz<br />

grains within the fine muddy sand layers.<br />

267<br />

F. Naughton, 2007


Fig. VI.3 | Lithology of the Douro sequence. Depth is represented in Ordnance Datum which is a coordinate<br />

system for heights above mean sea level (optometric heights). Five calibrated ages are also represented<br />

along the two cores.<br />

268<br />

F. Naughton, 2007


SED3 (-6.81 m / -1.67 m OD depth), accumulated c. 6530-1500 cal yr<br />

BP, is composed by a quartzite gravel layer with weak carbonates and<br />

organic matter values. Gravel unit presents 98-100% of ballasts with mean<br />

grain size between 16 mm and 32 mm. Fig. VI.4a and VI.4b show the limit<br />

bands for both high and low energy beach and river environments, based on<br />

roundness and sphericity values plotted against the grain size, proposed by<br />

Dobbkins and Folk (1970). Pebbles from SED3 present roundness values<br />

between 0 and 1 (mean value=0.63) and sphericity values from 0.4 to 1<br />

(mean value=0.69), confirming the different influences applied to these<br />

ballasts. This suggests that the quartzite grain morphology from this unit results<br />

from both fluvial and marine processes. Furthermore, the lithology of these<br />

pebbles is similar to that observed nowadays in the scattered quartzite ridges<br />

along the Douro basin, indicating that they were most likely transported by<br />

the river. However, some of these pebbles display a composite morphology<br />

suggesting that subsequent marine processes influenced the existing fluvial<br />

deposit. Around 6530 cal yr BP (5750 BP), sea-level was probably similar to<br />

present-day levels, as previously shown by Dias et al. (2000) and Granja and<br />

de Groot (1996).<br />

Fig. VI.4 | a) Grain roundness and b) effective sphericity of gravel pebbles plotted against particle size. Two<br />

black lines delimit the different roundness and sphericity averages defined by Dobbkins and Folk (1970),<br />

representing the average limit of grain characterising river and/or low and high-energy beach<br />

environments. Dark squares represent all the measures obtained and Circle represents the value means of<br />

all measures.<br />

269<br />

F. Naughton, 2007


SED4 (-1.67 m - +3.5 m), dated to post c. 1500 cal yr BP, is subdivided<br />

into three sub-zones: SED4.a (-1.67 m / -0.1 m), SED4.b (-0.1 m / +3.43 m),<br />

SED4.c (+3.43 m / +3.5 m) and is composed of sandy mud and slightly sandy<br />

mud layers alternating with sand and slightly muddy sand levels. Terrigenous<br />

content dominates all sub-units, suggesting a lack of marine supply in the<br />

coring site. This is also indicated by the near absence of carbonate content in<br />

this estuarine level. In SED4.a micas and high content of organic matter<br />

essentially compose the sandy mud and slightly sandy mud layers. Sand and<br />

slightly muddy sand levels are, in turn, constituted by quartz and low organic<br />

matter.<br />

SED4b is represented by quartz sand with almost no organic and<br />

carbonate material. The mineralogy of this sand unit is similar to that of the<br />

sand layers included in SED4a, characterized by the dominance of<br />

terrigenous sediments.<br />

Finally, the upper 70 cm of SED4c are formed by an alternation of sand<br />

and muddy sand levels with relatively high values of organic matter and<br />

micas. In comparison with the previous sedimentary units (SED2 and SED3),<br />

SED4 reflects high terrigenous inputs and slight marine influence.<br />

6. 4. 3 Vegetation changes versus variations in pollen catchment area<br />

during the Holocene<br />

Previous studies have shown that an important part of the pollen input<br />

in most estuaries is fluvial in origin (Brush, 1989, Reille, 1990; Suc and Drivaliari,<br />

1991; Chmura et al. 1999; Dupont and Wyputta, 2003). This must be indeed<br />

the case of Douro estuary, since it is located in the western European coast<br />

which is dominated by north-westerly Atlantic winds (Turon, 1984). Our pollen<br />

diagram should, therefore, reflect plant communities that have colonised the<br />

Douro basin over the last 10720 cal years. However, Santos et al. (2001) have<br />

shown that fluctuations in pollen frequencies of coastal sequences are not<br />

exclusively related to vegetation changes but can also derive from local<br />

geomorphological changes. As a result, the pollen diagram from the Douro<br />

coastal sequence could detect changes in pollen catchment area rather<br />

than vegetation changes in the Douro basin.<br />

270<br />

F. Naughton, 2007


The Douro diagram (Fig. VI.5) is divided in three main zones. These<br />

have been numbered from the base towards the top and prefixed by the<br />

name LAQUA. The pollen zones are interrupted by two important pollen hiati,<br />

associated with gravel and sand levels (at -7.03 m / - 1.70 m and 0 m / + 3.43<br />

m) whose aerobic depositional environment prevents pollen preservation.<br />

The first pollen zone, LAQUA1 (-13.90 m / -7.03 m), dated c. 10720-6530<br />

cal yr BP, is marked by high frequencies of arboreal pollen (20-50%), reflecting<br />

the dominance of a mixed Pinus-Quercus forest with Alnus in the Douro valley.<br />

Some scattered open communities of Poaceae-Ericaceae are also present in<br />

the regional vegetation.<br />

LAQUA2 (-1.70 m / 0 m), c. 1500-900 cal BP, is characterized by<br />

minimum percentages of arboreal pollen (5-10%) and maximum values of<br />

Ericaceae, Poaceae and Asteraceae pollen. The floristic composition of this<br />

pollen assemblage is similar to that characterizing present-day local<br />

vegetation.<br />

The upper centimetres of sedimentation are represented by LAQUA3.<br />

The pollen assemblage of this zone is similar to LAQUA2, however, Pinus<br />

percentages are higher resulting from human impact. Historical archives<br />

indicate that pine developed in coastal areas of northern Portugal at the<br />

beginning of the last century (Figueiral, 1995). This botanical event (input of<br />

Pinus) allows us to infer that this zone may represent less than 100 years of<br />

sedimentation.<br />

Two hypotheses may explain the differences observed between<br />

LAQUA1 and LAQUA2: a) change in regional vegetation composition, from<br />

forest to open formation, due to strong human impact in the Douro basin<br />

since 6530 cal yr BP; or, b) a change in pollen supply area.<br />

The first hypothesis considers that zones LAQUA1 and LAQUA2 preserve pollen<br />

grains that were mainly transported by the river from the vegetation<br />

colonising the Douro basin. The second hypothesis proposes that a change in<br />

the pollen source area occurred between LAQUA1 and LAQUA2: pollen was<br />

essentially transported by the river from the regional forests in LAQUA1, while<br />

in LAQUA2 pollen represents the local vegetation. This second hypothesis<br />

implies that a change of the river trajectory occurred after 6530 cal yr BP,<br />

decreasing the deposition of fluvial-transported pollen in the coring area.<br />

271<br />

F. Naughton, 2007


Fig. VI.5| Pollen diagram. From the left to the right: lithology, calibrated ages, arboreal pollen, AP (total of<br />

arboreal pollen), Pinus, NAP (total of non-arboreal pollen), pollen of herbaceous plants and pollen zones.<br />

272<br />

F. Naughton, 2007


To test these hypotheses, our data was compared with other pollen<br />

diagrams from Douro basin, covering the same time period (Peñalba, 1994;<br />

Goméz-Lobo Rodrigues, 1993; Garcia Antón et al., 1995; Allen et al., 1996;<br />

Peñalba et al., 1997; Von Engelbrechte, 1998; Sánchez Goñi and Hannon,<br />

1999). These diagrams show a gradual decrease, from 6000 BP (uncalibrated)<br />

up to the present of pollen tree percentages (in particular deciduous<br />

Quercus), probably due to human impact and/or climate. However, total<br />

arboreal pollen during this period is clearly much more important (AP=30-80%)<br />

in those diagrams than in LAQUA2 zone (AP=5-10%) (Fig VI.5). Therefore, a<br />

vegetation change in the Douro basin from forest to open communities can<br />

not explain the difference observed in the pollen record before and after<br />

6530 cal yr BP.<br />

The second hypothesis, involving a shift in the pollen source area,<br />

seems the most likely explanation. To verify this hypothesis, surface sediments<br />

were collected in the Douro estuary, after the November 2000 overflow, in<br />

order to identify the type and frequencies of present-day pollen taxa,<br />

transported by the Douro and incorporated in the sediments (Fig VI.6). This<br />

surface sediment study shows that arboreal pollen percentages are high<br />

(AP=20-30%), contrary to what is observed in LAQUA2.<br />

Furthermore, the comparison between the modern (LAQUASUP) and<br />

fossil spectra (LAQUA1-3) reflects a similarity between LAQUASUP and<br />

LAQUA1, which means that nowadays, after an overflow, the river transports<br />

a great amount of arboreal pollen from the entire Douro basin (AP=20-30%),<br />

and this despite present-day anthropisation. It is possible that from c. 1500 cal<br />

yr BP onwards the river might have had less importance on pollen supply to<br />

the coring site, and this is probably related to a northward river main channel<br />

migration, between 6530 – 1500 cal yr BP. A palaeo-valley with its axis located<br />

south of the present main channel testifies to this northward deflection of the<br />

river flow (Carvalho and Rosa, 1988) (Fig VI.1b).<br />

273<br />

F. Naughton, 2007


Fig. VI.6| LAQUASUP: surface samples sites and pollen percentages.<br />

274<br />

F. Naughton, 2007


6. 4. 4 Geomorphological changes in the Douro estuary during the<br />

Holocene<br />

10720 – 6530 cal yr BP (Fig. VI.7a)<br />

The first part of this period is marked by the prevalence of the fluvial<br />

influence as shown by the presence of terrigenous components, the almost<br />

absence of foramifera and echinoderms and the presence of regional fluvial-<br />

transported pollen.<br />

Concerning the geomorphology, the present day estuarine zone was<br />

an ancient river course within a northeast-southwest incised valley (Fig. VI.7a),<br />

as also shown by the palaeobathymetric map inferred by Carvalho and Rosa<br />

(1988) (Fig. VI.1b).<br />

The relatively warm and humid climate which characterizes the<br />

beginning of the Holocene, favoured the development Pinus-Quercus forest<br />

with Alnus, as reflected by their percentages contribution to our pollen<br />

spectra during (10720 – 6530 cal yr BP).<br />

The second part of this interval (10720 – 6530 cal yr BP), is characterized<br />

by an increase in marine microfossils towards the top and the presence of<br />

fluvial-transported pollen indicating a period of mixed fluvial and marine<br />

influences in the area, a likely result of sea-level rise and coast line recession.<br />

This is in agreement with the sea-level rise and coast line recession detected<br />

for the NW of Iberia by Dias et al. (2000) and Granja and de Groot (1996).<br />

6530 – 1500 cal yr BP (Fig. VI.7b)<br />

During this period, the accumulation of 5 m of gravel suggests drastic<br />

environmental changes probably due to more frequent torrential rains.<br />

This gravel unit, essentially constituted by pebbles derived from<br />

quartzite outcrops in the Douro basin that exhibit a mixed morphology,<br />

showing that they had been initially transported by the river waters until the<br />

core site and then, reworked by the sea.<br />

At this time, the reduction in the rate of sea-level rise (Dias et al., 2000)<br />

favours sediment deposition in the estuary. This, in turn, contributed to the<br />

formation of a gravel barrier and its northward growth triggered the migration<br />

of the Douro main channel to the north (Fig. VI.7b).<br />

275<br />

F. Naughton, 2007


This river main channel migration is suggested by the change in pollen<br />

origin, from regional fluvially-transported pollen to local pollen vegetation.<br />

Following the gravel barrier formation, local factors seem to play a major role<br />

in the geomorphological development of the Douro estuary.<br />

Sand barriers development has also been recorded further south, in<br />

coastal ecosystems such as Albufeira, Santo André and Melides (Freitas et al.,<br />

2002; 2003).<br />

1500 cal yr BP – present-day (Fig. VI.7c)<br />

The development of the gravel barrier in this area precluded the arrival<br />

of marine material over the last 1500 years, as demonstrated by the lack of<br />

carbonates and bioclast fragments in the estuarine sediments.<br />

The local pollen signature recorded during this period suggests the<br />

weak influence of the Douro river in the southern part of the estuary.<br />

However, after an overflow, the study area receives some detrital<br />

material and pollen grains from the hydrographical basin as showed by the<br />

LAQUASUP assemblages.<br />

The establishment of the gravel barrier in the southern area of the<br />

Douro mouth and river main channel migration to the north had contributed<br />

to the enlargement of the estuarine zone and settlement of present-day<br />

geomorphology (Fig. VI.7b and VI.7c).<br />

Fig. VI.7| Conceptual model of the Holocene geomorphological evolution of the Douro estuary: a)<br />

between 10720 and 6530 cal yr BP, b) from 6530 to 1500 cal yr BP and c) the last 1500 years.<br />

276<br />

F. Naughton, 2007


6. 5 CONCLUSIONS<br />

The multiproxy study of core 1 and 1B from the Douro estuary<br />

documents the Holocene geomorphological history of this coastal area and<br />

identifies the major factors explaining the evolution of this system.<br />

Between 10720 and 6530 cal yr BP, changes in geomorphology of the<br />

Douro estuary were essentially influenced by the global sea-level rise. Gradual<br />

increase of marine influences in the estuary indicates this coastline recession.<br />

Later on, a shift in pollen record shows a change in pollen source area due to<br />

a northward migration of the Douro main channel between 6530 and 1500<br />

cal yr BP. This was likely caused by the formation of a gravel barrier as the<br />

result of both sea-level rise attenuation and high fluvial hydrodynamism<br />

caused by the increase of torrential rains. Local characteristics become a<br />

major agent controlling geomorphology evolution of the Douro estuary,<br />

during the upper part of the Holocene.<br />

This study further highlights the potential of pollen analysis as a proxy to<br />

reconstruct geomorphological changes in coastal environments.<br />

References<br />

Allen, J.R.M., Huntley, B. and Watts, W.A., 1996. The vegetation and climate of<br />

northwest Iberia over the last 14 000 yr. Journal of Quaternary Science, 11: 125-147.<br />

Bao, R., Freitas, M.C. and Andrade, C., 1999. Separating eustatic from local<br />

environmental effects: a late Holocene record of coastal change in Albufeira<br />

lagoon, Portugal. The Holocene, 9: 341-352.<br />

Braun-Blanquet, J., Pinto da Silva, A.R. and Rozeira, A., 1956. Résultats de deux<br />

excursions géobotaniques à travers le Portugal septentrional et moyen. II. Chenaies à<br />

feuilles caduques (Quercion occidentale) et chenaies à feuilles persistentes<br />

(Quercion faginea) au Portugal. Agronomia Lusititana, 18: 167–234.<br />

Brush, G.S., 1989. Rates and patterns of estuarine sedimentation. Limnology and<br />

Oceanography, 34: 1235-1246.<br />

Carrington da Costa, J. and Teixeira, C., 1957. Carta Geológica de Portugal na<br />

escala 1/50.000. Notícia Explicativa da folha 9-C, Porto., 38pp.<br />

Carvalho, A.F. and Rosa, M.M.P., 1988. Localização do Paleovale do Rio Douro. Anais<br />

do Instituto Hidrográfico, 9: 77-82.<br />

Chmura, G.L., Smirnov, A. and Campbell, I.D., 1999. Pollen transport through<br />

distributaries and depositional patterns in coastal waters. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology, 149: 257-270.<br />

277<br />

F. Naughton, 2007


Costa, J.C., Aguiar, C., Capelo, J.H., Lousã, M. and Neto, C., 1998. Biogeografia de<br />

Portugal continental. Quercetea, 0: 5-56.<br />

Craft, C.B., Seneca, E.D. and Broome, S.W., 1991. Loss on ignition and Kjeldahl<br />

Digestion for estimating organic carbon and total nitrogen in estuarine marsh soils:<br />

Calibration with dry combustion. Estuaries, 14: 175-179.<br />

de Vernal, A., Henry, M. and Bilodeau, G., 1996. Techniques de préparation et<br />

d'analyse en micropaléontologie. Les cahiers du GEOTOP, Département des<br />

Sciences de la terre. Québec University, Montréal, 3, pp.16-27.<br />

Devoy, R.J.N., Delaney, C., Carter, R.W.G. and Jennings, S.C., 1996. Coastal<br />

stratigraphies as indicators of environmental changes upon European Atlantic coasts<br />

in the late Holocene. Journal of Coastal Research, 12: 564-588.<br />

Dias, J.M.A., 1987. Dinâmica sedimentar e evolução recente de plataforma<br />

continental Portuguesa setentrional. PhD Thesis, Lisbon University, Lisbon (Portugal),<br />

384 pp.<br />

Dias, J.M.A., Boski, T., Rodrigues, A. and Magalhães, F., 2000. Coast line evolution in<br />

Portugal since the last glacial maximum until present: a synthesis. Marine geology,<br />

170: 177-186.<br />

Dias, J.M.A., Jouanneau, J.M., Gonzalez, R., Araújo, M.F., Drago, T., Garcia, C.,<br />

Oliveira, A., Rodrigues, A., Vitorino, J. and Weber, O., 2002a. Present day sedimentary<br />

processes on the northern Iberian shelf. Progress In Oceanography, 52: 249-259.<br />

Dias, J.M.A., Gonzalez, R., Garcia, C. and Diaz del Rio, V., 2002b. Sediment distribution<br />

patterns on the Galicia-Minho continental shelf. Progress in Oceanography, 52: 215-<br />

231.<br />

Dobbkins, J.E. and Folk, R.L., 1970. Shape Development on TahitiNui. Journal of<br />

Sedimentary Petrology, 40: 1167-1203.<br />

Drago, M.T., 1995. La vasiere Ouest-Douro sur la Plateforme Nord-Portugaise. Role,<br />

Fonctionnement, Evolution. PhD Thesis, Bordeaux I University, Bordeaux (France),<br />

232pp.<br />

Drago, T., Freitas, C., Rocha, F., Moreno, J., Cachão, M., Naughton, F., Fradique, C.,<br />

Araujo, F., Silveira, T., Oliveira, A., Cascalho, J. and Fatela, F., (in press):<br />

Palaeoenvironmental evolution of estuarine systems during the last 14000 years- the<br />

case of Douro estuary (NW Portugal). Journal of Coastal Research.<br />

Dupont, L.M. and Wyputta, U., 2003. Reconstructing pathways of aeolian pollen<br />

transport to the marine sediments along the coastline of SW Africa. Quaternary<br />

Science Reviews, 22: 157-174.<br />

Ferreira, Ó., Dias, J.M.A. and Soares de Carvalho, G., 1989. Síntese dos<br />

conhecimentos sobre a dinâmica sedimentar da zona costeira entre Peniche e o Rio<br />

Minho. Report DISEPLA 8/88, 31pp.<br />

Figueiral, I., 1995. Charcoal analysis and the history of Pinus pinaster (cluster pine) in<br />

Portugal. Review of Palaeobotany and Palynology, 89: 441-454.<br />

278<br />

F. Naughton, 2007


Flemming, B.W., 2000. A revised textural classification of gravel-free muddy sediments<br />

on the basis of ternary diagrams. Continental Shelf Research, 20: 1125-1137.<br />

Freitas, M.C., 1995. A Laguna de Albufeira (Peninsula de Setubal). Sedimentologia,<br />

Morfologia e Morfodinâmica. PhD Thesis, Lisbon University, Lisbon (Portugal), 337pp.<br />

Freitas, M.C. and Andrade, C., 1997. Sistemas Fluvio-lagunares a sul da Tejo: evolução<br />

a escalas temporais diferentes. Colectânea de ideias sobre a zona costeira de<br />

Portugal. Associação Eurocoast-Portugal, pp. 525-545.<br />

Freitas, M.C., Andrade, C. and Cruces, A., 2002. The geological record of<br />

environmental changes in southwestern Portuguese coastal lagoons since the<br />

Lateglacial. Quaternary International, 93-94: 161-170.<br />

Freitas, M.C., Andrade, C., Rocha, F., Tassinari, C., Munhá, J.M., Cruces, A., Vidinha, J.<br />

and Marques da Silva, C., 2003. Lateglacial and Holocene environmental changes in<br />

Portuguese coastal lagoons 1: the sedimentological and geochemical records of the<br />

Santo André coastal area. The Holocene, 13: 433-446.<br />

Garcia Antón, M., Franco Múgica, F., Maldonado Ruiz, J., Morla Juaristi, C. and Sainz<br />

Ollero, H., 1995. Una secuencia polínica en Quintana Redonda (Soria). Evolucíon<br />

holocena del tapiz vegetal en el sistema Ibérico septentrional. Anales Jardin<br />

Botánico Madrid, 52:187-195.<br />

Gómez-Lobo Rodríguez, A., 1993. Historia de la vegetación durante los últimos 15.000<br />

años en los Picos de Urbión (Soria) en base al análisis polínico, Universidad de Alcalá<br />

de Henares, 182 pp.<br />

Granja, H.M., 1999. Evidence for Late Pleistocene and Holocene sea level,<br />

neotectonic and climate control in the coastal zone of northwest Portugal. Geologie<br />

en Mijnbouw, 77: 233-245.<br />

Granja, H.M. and de Groot, T., 1996. Sea-level rise and neotectonism in a Holocene<br />

coastal environment at Cortegaca Beach (NW Portugal): a case study. Journal of<br />

Coastal Research, 12: 160-170.<br />

Heusser, L., 1978. Spores and pollen in the marine realm. Introduction to Marine<br />

Micropalaeontology: New York (Elsevier), Haq, B. and Boersma, A.(Eds.), pp. 327–340.<br />

Hulseman, J., 1966. On the routine analysis of carbonates in unconsolidated<br />

carbonates. Journal of Sedimentary Petrology, 36: 622-625.<br />

Jouanneau, J.M., Weber, O., Drago, T., Rodrigues, A., Oliveira, A., Dias, J.M.A., Garcia,<br />

C., Schimdt, S. and Reyss, J.L., 2002. Recent sedimentation and sedimentary budgets<br />

on the western Iberian shelf. Progress In Oceanography, 52: 261-275.<br />

Loureiro, J.J., Machado, M.L., Macedo, M.E., Nunes, M.N., Botelho, O.F., Sousa, M.L.,<br />

Almeida, M.C. and Martins, J.C., 1986. Monografias hidrológicas dos principias cursos<br />

de água de Portugal continental. Direcção Geral dos Serviços Hidráulicos, Lisbon,<br />

569pp.<br />

Magalhães, F., 1993. A cobertura sedimentar da plataforma e da vertente<br />

continental superior a norte de Espinho. Master’s Thesis, Lisbon University, Lisbon<br />

(Portugal), 191pp.<br />

279<br />

F. Naughton, 2007


Magalhães, F., 1999. A cobertura sedimentar da plataforma continental portuguesa.<br />

Distribuição espacial. Contrastes temporais: Potencialidades económicas. PhD Thesis,<br />

Lisbon University, Lisbon (Portugal), 289pp.<br />

Magalhães, F., 2000. Sea-level in Portugal in the last 18 000 years. III Symposium of the<br />

Continental Iberian Atlantic Margin, pp. 205-206.<br />

Maher, L.J.Jr., 1981. Statistics for microfossil concentration measurements employing<br />

samples spiked with marker grains. Review of Palaeobotany and Palynology, 32: 153-<br />

191.<br />

Navarro Andrés and Valle Gutiérrez. 1987. Castilla Y Leon. La Vegetation de España,<br />

M. Peinado Lorca and S. Rivas-Martinez (Eds), pp.119-161.<br />

Oliveira, A., Rocha, F., Rodrigues, A., Jouanneau, J.M., Dias, J.M.A., Weber, O. and<br />

Gomes, C., 2002. Clay mineral from the sedimentary cover from the Northwestern<br />

Iberian shelf. Progress in Oceanography, 52: 233-247.<br />

Peñalba, C., 1994. The history of the Holocene vegetation in northern Spain from<br />

pollen analysis. Journal of Ecology, 82: 815-832.<br />

Peñalba, M.C., Arnold, M., Guiot, J., Duplessy, J.-C. and de Beaulieu, J.-L., 1997.<br />

Termination of the Last Glaciation in the Iberian Peninsula inferred from the pollen<br />

sequence of Quintanar de la Sierra. Quaternary Research, 48: 205-214.<br />

Pina Manique, J., 1957. Zones écologiques Portugaises. Publications of the Institute of<br />

Applied Biology, XXVI: 19-26.<br />

Reille, M., 1990. Leçons de palynologie et d'analyse pollinique. Centre National de la<br />

Recherche Scientifique, Paris, 206 pp.<br />

Reille, M., 1992. Pollen et spores d'Europe et d'Afrique du Nord. Laboratoire de<br />

botanique historique et palynologie, Marseille, 520 pp.<br />

Rull, V., 1987. A note on pollen counting in paleoecology. Pollen et Spores, 29: 471-<br />

480.<br />

Sánchez Goñi, M.F. and Hannon, G., 1999. High altitude vegetational patterns on the<br />

Iberian Mountain chain (north-central Spain) during the Holocene. The Holocene, 9:<br />

39-57.<br />

Santos, L., Bao, R. and Sánchez Goñi, M.F., 2001. Pollen record of the last 500 years<br />

from the Doniños coastal lagoon (NW Iberian Peninsula): Changes in the pollinic<br />

catchment size versus palaeoecological interpretation. Journal of Coastal Research,<br />

17: 705-713.<br />

Stuiver, M. and Reimer, P., 1993. Extended 14C database and revised CALIB<br />

radiocarbon calibration program. Radiocarbon, 35: 215-230.<br />

Suc, J.-P. and Drivaliari, A., 1991. Transport of bisaccate coniferous fossil pollen grains<br />

to coastal sediments. An example from the earliest Pliocene Orb ria (Languedoc,<br />

southern France). Review of Palaeobotany and Palynology, 70: 247-253.<br />

Thouveny, N., Moreno, E., Delanghe, D., Candon, L., Lancelot, Y. and Shackleton N.J.,<br />

2000. Rock magnetic detection of distal ice-rafted debries: clue for the identification<br />

280<br />

F. Naughton, 2007


of Heinrich layers on the Portuguese margin. Earth and Planetary Science Letters, 180:<br />

61-75.<br />

Turon, J.-L., 1984. Le palynoplancton dans l'environnement actuel de l'Atlantique<br />

nord-oriental. Evolution climatique et hydrologique depuis le dernier maximum<br />

glaciaire. Mémoires de l'Institut de Géologie du bassin d'Aquitaine, 17. Université de<br />

Bordeaux I, Bordeaux, 313 pp.<br />

Von Engelbrechten, S., 1998. Late-glacial and Holocene vegetation and<br />

environmental history of the sierra de Urbión, North-Central Spain. PhD Thesis, Dublin<br />

University, Trinity College (Ireland), 212 pp.<br />

Zong, Y., 2004. Mid-Holocene sea-level highstand along the Southeast Coast of<br />

China. Quaternary International, 117: 55-67.<br />

281<br />

F. Naughton, 2007


282<br />

F. Naughton, 2007


Capítulo 7| Conclusions and perspectives<br />

In order to deeply characterise and understand the response of<br />

vegetation and climate of south-western Europe to the climate variability of<br />

the North Atlantic over the last 30 000 years we have carried out a high<br />

resolution multi-proxy analysis (pollen, planktonic foraminifera assemblages,<br />

planktonic and benthic foraminifera δ 18 O measurements and Uk 37 sea surface<br />

temperature reconstruction) of two marine deep-sea cores (MD99-2331 and<br />

MD03-2697) retrieved in the Galician margin (north-western Iberia). These<br />

marine deep-sea cores located at ~100 Km from the coast preserve a high<br />

quantity of pollen grains from the adjacent landmasses together with several<br />

marine climate indicators and an ice-volume proxy. This has allowed us to<br />

establish a direct correlation between marine and terrestrial proxies and to<br />

identify possible leads and lags in the response of the different Earth reservoirs<br />

to a given climatic change.<br />

Before starting this <strong>paleoclimat</strong>ic study we have verified whether the<br />

pollen preserved in the Iberian margin sediments represent one integrated<br />

image of the regional vegetation colonising the adjacent landmasses. We<br />

have compared modern pollen assemblages from the continent with those<br />

from marine sediments. For this, we have analysed pollen content from<br />

several sedimentary modern samples retrieved in south (Mediterranean<br />

region) and in north-western (Atlantic region) Iberia following two transects<br />

including the estuary, shelf and slope. The resulting marine and coastal<br />

modern pollen signatures have been compared with several present-day<br />

continental samples including mosses, surface sediments and top of peat<br />

bog and lake sequences available in the European Pollen Database.<br />

This comparison showed that the pollen signature from the Iberian<br />

margin is similar to that of the Iberian terrestrial deposits, and, in particular, to<br />

that of the estuarine samples which recruit pollen from the vegetation<br />

colonising the adjacent hydrographic basins. Therefore, western Iberian<br />

margin pollen spectra reflect an integrated image of the regional vegetation<br />

of the adjacent continent. Our study has also showed that marine pollen<br />

283<br />

F. Naughton, 2007


spectra clearly discriminate both the Mediterranean and the Atlantic plant<br />

communities colonising southern and northern Iberian Peninsula, respectively.<br />

Furthermore, it also demonstrate that Pinus pollen are as elsewhere<br />

overrepresented in marine sediments indicating that this pollen must not be<br />

included in the main pollen sum which is usually used for calculate pollen<br />

percentages. Further, this work has allowed us to understand the mechanisms<br />

involved in pollen transfer from Iberia to the ocean.<br />

The western Iberian margin is at present influenced by north-western<br />

prevailing winds which preclude substantial direct airborne transport of pollen<br />

seaward. On the other hand, this margin is closed to several important<br />

hydrographic basins (Tagus, Sado, Douro and Minho) favouring pollen<br />

seaward transfer mainly by fluvial transport. Results from Total Pollen<br />

Concentration (TPC) together with proposed conceptual models of fine<br />

particle dynamics in the Iberian margin have allowed us to establish the<br />

present-day pattern of pollen dispersion in this region. In north-western Iberian<br />

margin, pollen grains released by Douro followed by Minho rivers are<br />

enclosed in n<strong>ephe</strong>loid layers and transported to the shelf until getting<br />

blocked by the rocky outcrops. In winter, during downwelling conditions<br />

pollen grains are then transported polewards, firstly deposited in the Douro<br />

mud patch (S-N direction) then in the Galicia mud patch, and finally they<br />

flow westward to the deep-sea. Only small quantities of pollen grains can be<br />

transported directly to the outer shelf and upper slope under extreme stormy<br />

events. In summer, under upwelling conditions, pollen transfer to the slope<br />

must be restricted to offshore filaments. In the southern Iberian margin, pollen<br />

grains released by the Tagus and to a lesser extent by the Sado river, are<br />

partially deposited in the shelf and transported to the south and seaward by<br />

littoral and oceanic currents probably through the southern canyons during<br />

upwelling conditions.<br />

Besides the verification of the reliability of Iberian margin pollen signatures, our<br />

study reveals that marine pollen sequences from western Iberian margin are a<br />

powerful tool to accurately reconstruct the vegetation response to oceanic<br />

and atmospheric climate changes within a reliable chronological framework.<br />

The comparison of the high resolution pollen analysis from the Galician margin<br />

284<br />

F. Naughton, 2007


composite core (MD99-2331 and MD03-2697) with other marine and terrestrial<br />

pollen sequences has allowed us to document vegetation changes in the<br />

Iberian Peninsula over the last 25 000 years. It also shows that our Galician<br />

margin sequence mainly represent the vegetation of low- and mid-altitude<br />

zones. Further, the direct correlation between marine proxies such as IRD<br />

content, planktonic polar foraminifera percentages and planktonic<br />

foraminifera δ 18 O with pollen from north-western Iberia reveals that<br />

vegetation of north-western Iberia has responded synchronously to the<br />

climatic variability detected elsewhere in the North Atlantic realm during the<br />

Heinrich events 2 and 1 (H2 and H1) (26 000-24 380 cal yr BP e 15 900-18 500<br />

cal yr BP), the Last Glacial Maximum (LGM) (24 300 and 18 500 cal yr BP), the<br />

Bölling Alleröd (B-A) warm period (13 200-15 900 cal yr BP) and the Younger<br />

Dryas (YD) cold event (11 600-13 200 cal yr BP).<br />

Vegetation response to the well known Heinrich events 2 and 1 (H2<br />

and H1) is complex and characterised by two vegetation phases at low and<br />

mid-altitudes of north-western Iberia. The beginning of each Heinrich event is<br />

marked on land by an important pine forest reduction and the expansion of<br />

heathers suggesting that this first phase was cold and wet. Pinus forest<br />

expansion characterising the second phase of each Heinrich event indicates<br />

a less cold episode associated with an increase of dryness as suggested by<br />

the development of semi-desert associations. We have also demonstrated<br />

that the H1 event is the marine equivalent of the Oldest Dryas in the<br />

continent.<br />

The Last Glacial Maximum (LGM), bracketed by H2 and H1 events, was<br />

characterised by the expansion of Pinus in an herbaceous-dominant<br />

environment along with scattered pockets of deciduous trees (Quercus,<br />

Corylus and Alnus) in north-western Iberia. This suggested that not only<br />

southern Iberia but also northern Iberia acted as a refugium zone for<br />

temperate trees, though at a smaller scale. Furthermore, western Iberia was<br />

characterised by prevailing wet conditions during the LGM as suggested by<br />

the expansion of heath communities (Ericaceae including Calluna) in both<br />

north-western and south-western regions.<br />

285<br />

F. Naughton, 2007


The Bölling-Allerød interstadial in our Galician sequence shows a more<br />

rapid and strong expansion of deciduous Quercus at low- and mid-altitude<br />

zones than in the high altitude sites of north-western Iberia. This indicates that<br />

the temperate forest from low and mid-altitudes of north-western Iberia<br />

together with that from southern Iberia region responded more rapidly to the<br />

climate variability of the North Atlantic during the B-A interstadial than in the<br />

northern high altitudes of Iberia.<br />

The Younger Dryas cold event was characterised at low- and mid-<br />

altitude of north-western Iberia by an increase of pioneer species (mainly<br />

Betula), grasses and semi-desert associations (Artemisia and Ephedra) at the<br />

expense of the temperate forest. However, the vegetation reversal<br />

characterizing the Younger Dryas event was less marked in those regions than<br />

in the high altitude sites of northern Iberia certainly due to the maximum<br />

expansion of deciduous forest during the previous B-A interstadial and a<br />

mitigated decrease in temperatures in the lower areas.<br />

The response of deciduous forest to the climate improvement that<br />

characterised the onset of the Holocene at low and mid altitudes of north-<br />

western Iberia seems to lead those observed in the high altitude sites,<br />

although the same succession of trees (Juniperus, Betula, Pinus Quercus,<br />

Corylus and Alnus) was observed in all northern regions. Indeed, this<br />

vegetation succession started during the Younger Dryas in our Galician core<br />

while in the high altitudes of northern Iberia at the beginning of the Holocene.<br />

In southern Iberia deciduous Quercus expansion occurred as in the mid- and<br />

low-altitudes of north-western Iberia rapidly most likely as the result of the<br />

higher density of refugia for temperate trees in these zones during the LGM.<br />

The direct correlation between terrestrial and marine climatic<br />

indicators enables us to tackle the mechanisms responsible for the complex<br />

pattern left by Heinrich events in north-western Iberia which corresponds to<br />

that previously observed in the mid-latitudes of the North Atlantic. Two main<br />

vegetation phases in north-western Iberia were linked to the complex pattern<br />

left by the typical Heinrich events (H4, H2 and H1) in the Iberian margin. The<br />

first phase was marked by a drop of Sea Surface Temperature (SST) (indicated<br />

by the planktonic foraminifera assemblages and δ 18 O analyses together with<br />

Uk 37 measurements) the virtual absence of icebergs in north-western Iberian<br />

286<br />

F. Naughton, 2007


margin and the strong cooling of the adjacent continent revealed by the<br />

Pinus forest decline. This first phase was also characterised by an increase of<br />

moisture conditions in Iberia as showed by the Calluna expansion in concert<br />

with the increase of total pollen concentration in MD99-2331 record. The<br />

second phase, associated with the maximal arrival of icebergs into this<br />

Iberian margin, was characterised by less cold sea surface and atmospheric<br />

conditions (Pinus forest development) and by an increase of dryness<br />

identified by the expansion of semi-desert plants. The impact of H3 in north-<br />

western Iberia was peculiar and associated with prevailing wet conditions<br />

over almost the entire event. Nonetheless, a substantial drop in SST and<br />

atmospheric temperatures (Pinus forest decline) preceded the episode of<br />

maximal icebergs discharges into the Iberian margin as during H4, H2 and H1<br />

events.<br />

We have proposed two main mechanisms underlying this complex<br />

pattern. The extreme atmospheric cold conditions indicated by the<br />

substantial decline of Pinus forest in north-western Iberia during Heinrich<br />

events 4, 3, 2 and 1 were probably caused by the introduction of large<br />

amounts of freshwater via Northern hemisphere icebergs drifting and<br />

consequent melting triggering a shutdown of the Atlantic Meridional<br />

Overturning Circulation (MOC). This MOC shutdown was followed by ocean-<br />

atmosphere rapid reorganizations favouring the fast transfer of cold<br />

conditions into north-western Iberian Peninsula. Superimposed to this<br />

oceanographic mechanism, changes similar to those of prevailing (negative<br />

to positive) North Atlantic Oscillation (NAO) index seems to have played a<br />

crucial role for explaining this complex pattern. Indeed during the first phase,<br />

prevailing negative mode of NAO-like index likely triggered the increase of<br />

winter precipitation in Iberia and enhanced river flow favouring the seaward<br />

pollen transfer. These prevailing conditions generated a SST increase in north-<br />

western Atlantic (at latitudes northern than 45°-50°N) favouring iceberg<br />

melting in the IRD belt preventing their southern migration to the mid-<br />

latitudes. This icebergs melting produced then a drop in sea surface<br />

conditions in the IRD belt region. This prevailing NAO-like negative mode also<br />

produced the drop of SST along the branch between north-western Iberian<br />

margin and the Greater North Sea amplifying the sea surface cooling in<br />

287<br />

F. Naughton, 2007


north-western Iberian margin. During the second phase of the typical Heinrich<br />

events (H4, H2 and H1) the change to prevailing positive mode of the NAO–<br />

like index triggered the strengthening and northward displacement of the<br />

westerlies favouring the increase of dryness in Europe including Iberia<br />

Peninsula. These prevailing conditions produced relatively warm sea surface<br />

conditions at mid-latitudes of the North Atlantic (~ 20°- 40°N) favouring the<br />

southward migration of the icebergs to the mid-latitudes sites including<br />

western Iberian margin. During this episode of maximum arrival of icebergs<br />

into the mid-latitudes of the North Atlantic, sea surface conditions off north-<br />

western Iberia were slightly warmer than the precedent phase showing the<br />

influence of prevailing positive NAO-like index along north-western Iberian<br />

margin and Greater North Sea branch. The prevailing wet conditions during<br />

the atypical H3 in north-western Iberia probably resulted from the maintaining<br />

of reduced westerlies in this region.<br />

The land-sea direct correlation has revealed that during the LGM<br />

period temperate tree expansion was largely reduced when compared with<br />

the previous late MIS 3 D-O interstadials although sea surface conditions were<br />

similar and relatively high. We have proposed three mechanisms for<br />

explaining the decoupling between the ocean surface temperatures and the<br />

temperate forest development: a) albedo increase which enhanced the<br />

cooling produced by low summer insolation in the Northern Hemisphere, b)<br />

high seasonality as a result of substantial winter cooling, and c) weak CO2<br />

concentration. However, western Iberia have been influenced by relatively<br />

wet conditions during the LGM likely as the response to the strengthening of<br />

the Meridional Overturning Circulation (MOC) which was more vigorous than<br />

during the bracketing Heinrich events.<br />

Besides the last glacial period, we were also interested to know about<br />

the millennial-scale climatic oscillations at mid-latitudes of the North Atlantic<br />

and in north-western Iberia which occurred during the last deglaciation<br />

(19 500-7000 yr cal BP), a period characterised by the increase of summer<br />

insolation and the substantial ice volume reduction in the northern high<br />

288<br />

F. Naughton, 2007


latitudes. This period encompasses the end of the LGM, the H1, Bölling-Alleröd<br />

(B-A) interstadial, Younger Dryas and early Holocene (11 500-8 000 cal yr BP).<br />

The substantial atmospheric (deciduous Quercus expansion) and sea-<br />

surface warming following H1 was probably caused by both the increase of<br />

mid-latitude summer insolation and the strengthening of the MOC. Maximum<br />

expansion of deciduous Quercus at around 14 000 cal yr BP occurred<br />

synchronously with the Greenland temperature peak of GIS (Greenland<br />

Interstadial) 1 and with the Meltwater Pulse 1A (MWP 1A). This suggests that<br />

the severe warming of the Northern Hemisphere could be the trigger<br />

mechanism for the drastic melting episode (MWP 1A) of the Laurentide Ice<br />

sheet implying that this event has been probably initiated in the Northern<br />

rather than in the Southern Hemisphere.<br />

Subsequent decrease of deciduous Quercus forest in north-western<br />

Iberia was contemporaneous with sea surface cooling at the mid-latitudes of<br />

the North Atlantic and reflects the Younger Dryas (YD) event in that region.<br />

MOC reduction, instead of shutdown, and the increase of northern mid-<br />

latitude summer insolation favoured the decrease rather than the complete<br />

decline of deciduous Quercus forest, as it is the case during H1, in north-<br />

western Iberia. Besides the cooling, the expansion of semi-desert plants<br />

reflected an increase of dryness in Iberia which was likely the result of<br />

prevailing positive mode of the NAO-like index in the North Atlantic realm.<br />

Concomitant with high values in the Northern mid-latitudes summer<br />

insolation we have detected the Holocene Thermal Maximum (HTM) in north-<br />

western Iberia between 11 700 and 8 200 cal yr BP. At around 8 200 cal yr BP a<br />

slight drop in SST of the eastern North Atlantic mid-latitudes together with a<br />

sudden decrease of deciduous Quercus and Corylus forest in north-western<br />

Iberia marked the 8.2 Kyr event in those regions. This event reflects the<br />

culmination of the successive episodes associated with the Laurentide Ice<br />

sheet decay which enhanced the cooling over Greenland and Europe.<br />

Following the 8.2 ky event the gradual decrease of temperate forest<br />

paralleled mid-latitudes summer insolation decrease showing that this long-<br />

term trend of forest reduction was influenced by orbitally-induced cooling<br />

rather than by human impact.<br />

289<br />

F. Naughton, 2007


The low resolution multi-proxy analysis for the Holocene interval in<br />

MD03-2697 deep-sea core has precluded, however, the identification of the<br />

millennial scale climatic variability of the Holocene in the mid-latitudes of the<br />

North Atlantic and in particular, the multi-centennial cooling event between<br />

~8 600 and 8 000 cal yr BP detected elsewhere. For filling this gap, we have<br />

analysed the pollen content of a 2.72 m long core, Vk03-58Bis, retrieved in<br />

north-western French shelf (47°36’ N and 4°08’ W). Vegetation and<br />

quantitative climatic reconstructions from this core have shown orbital and<br />

suborbital climate variability over the last 8 850 years in north-western France.<br />

Gradual temperate forest decline together with the decrease of MTWA<br />

(mean temperature of the warmest month) values parallels until at least 2 000<br />

yr cal BP the steady long-term cooling of Greenland and the reduction of<br />

mid-latitude summer insolation. Also, the gradual decrease of seasonality<br />

followed, as it was expected, the increase of precession. Counterbalancing<br />

this long-term astronomical forcing trend, Corylus woodlands spread at the<br />

expense of deciduous Quercus forest, between 8 739 and 8 387 cal yr BP,<br />

reflecting particular high seasonality conditions as the result of the<br />

amplification produced by the expansion of sea-ice in North Atlantic high-<br />

latitudes during winter. This increase of winter sea ice cover was triggered by<br />

the final episodes of Agassiz and Ojibway outbursts and consequent gradual<br />

reduction of the MOC. A sudden Corylus woodland decline in north-western<br />

France, between 8 387-8 062 cal yr BP, already noticed in other European<br />

regions, marked the 8.2 kyr cold event. This extremely cold episode was<br />

probably triggered by the final drastic MOC reduction associated with<br />

slowest flow speed of the ISOW (Iceland-Scotland Overflow Water) leading to<br />

an additional drop in winter temperature over Europe and Greenland.<br />

Nonetheless, seasonality remained high during this interval. Indeed the high<br />

seasonality conditions detected in VK03-58Bis between 8 739-8 062 cal yr BP<br />

reflected therefore the multi-centennial-scale climate cooling 8.6-8.0 kyr<br />

episode of the North Atlantic.<br />

Following the Agassiz and Ojibway final outburst episodes, climate<br />

became more stable. However, millennial scale climate cooling episodes are<br />

recorded and characterised by weak winter cooling and increases in<br />

precipitation and seasonality.<br />

290<br />

F. Naughton, 2007


Besides pollen, dinocyst analysis and benthic gastropod Turritella<br />

communis occurrences in VK03-58Bis indicated regional changes such as: a)<br />

the southward migration of the Boreal biogeographical zone between 8 739-8<br />

479 cal yr BP as the result of the southern extension of the North Atlantic sea-<br />

ice cover triggered by the decrease of winter temperatures which favoured<br />

the settlement of the T. communis off north-western France; and b) the<br />

subsequent opening of the English Channel at around 8 479-8 387 cal yr BP<br />

which produced a drastic environmental change triggering the T. communis<br />

death and the decrease of the dinocyst Lingulodinium machaerophorum.<br />

Since the last deglaciation, climate variability has a great impact on<br />

coastal systems evolution. Indeed the increases of high latitude summer<br />

insolation leading to the decrease of ice volume favoured the global sea-<br />

level rise and consequent changes in coastal areas. In order to understand<br />

the impact of global forcing factors such as climate and sea-level changes<br />

on the geomorphology of north-western Iberia coastal systems, we have<br />

applied a multi-proxy study of two cores retrieved in the mouth of the Douro<br />

estuary (north-western Portugal).<br />

The early to mid-Holocene (between 10 720 and 6 530 cal yr BP) was<br />

characterised by a well-established Pinus-Quercus-Alnus forest reflecting<br />

warm and moist conditions in the Douro basin. A gradual increase of shelf<br />

and slope benthic foraminifera assemblages together with echinoderms<br />

reflect a steady sea-level rise. Pollen grains from the regional vegetation were<br />

transported by the river along the ancient river main channel and deposited<br />

in the study area which was at that time a river margin.<br />

At around 6 530 cal yr BP a drastic environmental change is marked by<br />

the deposition of 5 m of gravel in the southern part of the estuary as the result<br />

of high hydrodynamism of the river and sea-level rise attenuation. The<br />

settlement of this gravel barrier between 6530 and 1500 cal yr BP leads to the<br />

northward migration of the river main channel preventing fluvially transported<br />

pollen grains from the regional vegetation of the Douro basin to be deposited<br />

in the study area after 1500 cal yr BP. Indeed, the radical change from<br />

regional fluvially transported pollen assemblages (mainly composed of trees)<br />

to pollen spectra derived from local vegetation (mainly Ericaceae/Poaceae)<br />

291<br />

F. Naughton, 2007


occurred contemporaneously with the settlement of this gravel barrier. The<br />

establishment of the gravel barrier in the southern area of the Douro mouth<br />

and river main channel migration to the north had contributed to the<br />

enlargement of the estuarine zone. Following this, after 1500 cal yr BP, local<br />

factors seemed to have played a major role in the geomorphological<br />

development of the Douro estuary.<br />

Perspectives<br />

After this work, several questions are still unsolved….<br />

It will be necessary to carry out some high resolution multi-proxy study<br />

of the last 30 000 cal yr BP in the mid-latitudes of the western North Atlantic<br />

and at higher latitudes than Iberian margin in the eastern and western North<br />

Atlantic to:<br />

- confirm the impact of the atmospheric mechanism linked to changes<br />

in the prevailing negative to positive mode of NAO-like index, on Heinrich<br />

events;<br />

- understand if the extreme warming detected in north-western Iberia<br />

and Greenland during the Melwater pulse 1A occurred in other North Atlantic<br />

regions to investigate the Northern Hemisphere origin of this last event;<br />

-to detect the multi-centennial episode between 8.6 and 8.0 ky BP<br />

including the short-lived 8.2 ky event, the Holocene Thermal Maximum and<br />

the millennial-scale climate variability of the mid- and late-Holocene in the<br />

North Atlantic.<br />

Finally, it would be also of extreme importance to correlate a multi-<br />

proxy marine deep-sea core of north-western Iberia representing the last<br />

glacial and interglacial transition (LGIT) with an estuarine core from north-<br />

western of Portugal, in order to understand the impact of the North Atlantic<br />

climate variability on the evolution of that coastal system during this<br />

transitional period.<br />

292<br />

F. Naughton, 2007


Conclusões<br />

No intuito de caracterizar e compreender a resposta da vegetação e<br />

do clima do sudoeste da Europa à variabilidade climática que caracteriza a<br />

região do Atlântico Norte, durante os últimos 30 000 anos, foi efectuada uma<br />

análise de proxies múltiplos com alta resolução temporal (pólen, associações<br />

de foraminíferos planctónicos, δ 18 O de foraminíferos planctónicos e<br />

bentónicos e reconstrução da temperatura superficial do oceano baseada<br />

na análise de alcanonas) em duas sondagens marinhas (MD99-2331 e MD03-<br />

2697) recolhidas na margem Galega (noroeste da margem Ibérica). Estas<br />

sondagens foram recolhidas a cerca de 100 Km da costa e possuem uma<br />

grande abudância de grãos de pólen provenientes do continente<br />

adjacente, assim como uma série de indicadores paleoclimáticos marinhos e<br />

um indicador do volume de gelo acumulado nos pólos. A presença de<br />

diferentes tipos de indicadores permite o estabelecimento de uma<br />

correlação directa entre os diferentes sub-sistemas climáticos assim como<br />

identificar possíveis sincronias e assincronias na resposta dos mesmos a um<br />

dado evento climático.<br />

Antes de iniciar este estudo paleoclimático, foi necessário verificar se<br />

os grãos de pólen preservados nos sedimentos da margem ibérica<br />

representavam de facto uma imagem integral da vegetação que coloniza o<br />

continente adjacente. Desta forma, foi efectuada a análise polínica em<br />

várias amostras superficiais colhidas no sudoeste (região Mediterrânica) e<br />

noroeste (região Atlântica) da margem Ibérica ao longo de um trajecto<br />

transversal à actual linha de costa incluindo estuários, plataforma e talude<br />

continental. Estas amostras foram comparados com as assinaturas polínicas<br />

continentais actuais provenientes da base de dados polínicos da Europa<br />

(European Pollen Database).<br />

Neste estudo comparativo foi possível verificar que as assinaturas<br />

polínicas da margem ibérica são semelhantes àquelas detectadas nas<br />

amostras continentais. De facto, as assinaturas polínicas da margem Ibérica<br />

são particularmente semelhantes àquelas obtidas para as zonas estuarinas<br />

as quais recrutam os grãos de pólen provenientes da vegetação que<br />

293<br />

F. Naughton, 2007


coloniza as bacias hidrográficas adjacentes. Os resultados obtidos permitem-<br />

nos assim afirmar que os espectros polínicos da margem Ibérica reflectem<br />

uma imagem integral da vegetação que coloniza o continente adjacente.<br />

Este estudo permitiu ainda demonstrar que os espectros polínicos<br />

marinhos discriminam claramente ambas as comunidades vegetais<br />

mediterrânicas e atlânticas que colonizam as zonas situadas a sul e a norte<br />

da Peninsula Ibérica, respectivamente.<br />

Para além disso, este estudo testemunha, tal como noutras regiões do<br />

mundo, que os grãos de pólen de Pinus encontram-se sobrerepresentados<br />

nos sedimentos marinhos, e que devem por isso ser retirados no somatório de<br />

base o qual é geralmente utilizado no cálculo das percentagens polínicas.<br />

Este trabalho permitiu ainda determinar os tipos de mecanismos<br />

associados à transferência dos grãos de pólen do continente para a<br />

margem Ibérica.<br />

Actualmente esta margem é dominada por ventos vindos de noroeste<br />

os quais impedem o transporte aéreo dos grãos de pólen do continente<br />

para o oceano. Por outro lado, a margem Ibérica localiza-se nas<br />

proximidades de uma série de bacias hidrográficas importantes tais como o<br />

Tejo, o Sado, o Douro e o Minho, as quais facilitam o transporte fluvial destes<br />

grãos para o mar aberto.<br />

A comparação dos resultados obtidos no estudo das concentrações<br />

polínicas totais das amostras de superfície com modelos conceptuais<br />

relacionados com a dinâmica sedimentar das partículas finas ao longo desta<br />

margem permitiu-nos propôr um padrão para a dispersão polínica nesta<br />

região.<br />

No noroeste da Península Ibérica, os grãos de pólen são incorporados<br />

nas camadas nefelóides e transportados para o mar pelos rios Douro e<br />

Minho. Ao chegar à plataforma continental a maioria dos grãos são<br />

bloqueados pelos afloramentos rochosos ai existentes. No inverno, durante<br />

condições de "downwelling”, os grãos de pólen serão transportados para<br />

norte, depositando-se primeiramente no complexo silto argiloso do Douro, no<br />

complexo silto argiloso do Minho, e finalmente para o mar aberto (oeste).<br />

Apenas uma parte ínfima destes grãos é transportada directamente do rio<br />

para o mar aberto, principalmente durante períodos de fortes tempestades.<br />

294<br />

F. Naughton, 2007


Durante o verão, as condições de "upwelling” impedem o transporte de<br />

grãos de pólen para o mar aberto, pelo que apenas uma pequena<br />

quantidade poderá ser transferida através de filamentos transversais.<br />

No sul da margem Ibérica, os grãos de pólen são libertados pelos rios<br />

Tejo e Sado e são parcialmente depositados na plataforma continental. Estes<br />

grãos são posteriormente transportados para sul e em direcção do mar<br />

aberto por correntes litorais e oceânicas durante condições de<br />

“downwelling”, ou ainda através dos canhões submarinos.<br />

Para além da credibilidade das assinaturas polínicas marinhas actuais,<br />

este estudo demonstrou ainda que as sequências polínicas marinhas da<br />

margem oeste Ibérica constituem uma ferramenta indispensável à<br />

correlação directa oceano-continente visto que este tipo de estudos se<br />

baseia numa base cronológica comum. Desta forma, foi efectuada a<br />

comparação entre os resultados obtidos pela análise polínica de alta<br />

resolução temporal, efectuada numa sequência marinha composta (a qual<br />

inclui as sondagens MD99-2331 e MD03-2697) com dados obtidos noutras<br />

sondagens marinhas desta mesma margem e ainda com várias sequências<br />

polínicas continentais. Esta comparação permitiu ainda documentar as<br />

variações do coberto vegetal ao longo da Península Ibérica para os últimos<br />

25 000 anos e mostrar que a sequência marinha Galega apresenta uma<br />

assinatura polínica representante essencialmente das baixas e médias<br />

altitudes do noroeste da Península Ibérica.<br />

A correlação directa entre indicadores paleoclimáticos marinhos<br />

(conteúdo em IRD, percentagens de foraminíferos planctónicos<br />

característicos de ambientes polares e valores em δ 18 O dos foraminíferos<br />

planctónicos) e continentais (pólens) revelou ainda que a vegetação do<br />

noroeste da Península Ibérica respondeu contemporaneamente aos eventos<br />

climáticos detectados no Atlântico Norte, nomeadamente aos eventos de<br />

Heinrich 2 e 1 (H2 e H1) (26 000-24 380 anos cal BP e 15 900-18 500 anos cal<br />

BP), ao último máximo glaciar (LGM) (24 300 e 18 500 anos cal BP), ao evento<br />

quente Bölling Alleröd (B-A) (13 200-15 900 anos cal BP) e ao evento frio Dryas<br />

recente (YD) (11 600-13 200 anos cal BP).<br />

295<br />

F. Naughton, 2007


A resposta da vegetação à variabilidade climática que caracteriza os<br />

eventos de Heinrich 2 e 1 é complexa e caracterizada essencialmente por<br />

duas fases distintas principalmente nas baixas e médias altitudes do noroeste<br />

da Península Ibérica.<br />

O início de cada um dos eventos é caracterizado por uma forte<br />

regressão da floresta de pinheiros (Pinus) e pela expansão de urze<br />

(Ericaceae incluindo Calluna) sugerindo que esta fase inicial seria bastante<br />

fria e húmida. A segunda fase é caracterizada pela re-expansão da floresta<br />

de pinheiros, a qual indica um episódio ligeiramente mais quente do que o<br />

anterior. Para além disso, o desenvolvimento de plantas semi-desérticas<br />

sugere um aumento da aridez durante esta segunda fase. Neste trabalho foi<br />

ainda possível demonstrar que o evento de H1 é o equivalente marinho do<br />

episódio designado por "Oldest Dryas” no continente.<br />

O último máximo glaciar (LGM) foi essencialmente dominado por uma<br />

vegetação herbácea no noroeste da Península Ibérica. Este período é ainda<br />

caracterizado pela expansão pinheiros (Pinus) e presença esporádica da<br />

floresta decídua (Quercus, Corylus e Alnus). A presença esporádica de<br />

árvores temperadas permite-nos inferir que não só o sul mas também a<br />

região norte terá agido como uma zona refugio para essas árvores durante o<br />

LGM. A presença de urze (Ericaceae incluindo Calluna) sugere que o<br />

noroeste foi, tal como o sudoeste da Península Ibérica, dominada por<br />

condições húmidas.<br />

A nossa sequência Galega mostra ainda que a expansão de<br />

carvalhos (Quercus deciduous) ocorreu de forma mais rápida e intensa nas<br />

baixas e médias altitudes do que nas regiões altas do noroeste da Península<br />

Ibérica durante o episódio quente que caracteriza o evento Bölling-Allerød.<br />

Isto sugere que a floresta temperada dessas regiões baixas reagiu, tal como<br />

na zona sul, de forma mais rápida à variabilidade climática do Atlântico<br />

Norte durante esse período interestadial.<br />

O Dryas recente foi caracterizado pelo aumento de espécies<br />

pioneiras (compostas essencialmente por Betula) e pela expansão de<br />

gramíneas e plantas semi-desérticas (Artemisia e Ephedra) em detrimento da<br />

floresta de carvalhos, nas baixas e médias altitudes do noroeste da Península<br />

296<br />

F. Naughton, 2007


Ibérica. Contudo, este evento é ligeiramente menos acentuado nas baixas e<br />

médias do que nas altas altitudes do noroeste Ibérico.<br />

A resposta da vegetação decídua à melhoria climática que<br />

caracteriza o início do Holocénico aparenta ser mais rápida nas baixas e<br />

médias do que nas altas altitudes do noroeste da Península Ibérica. No<br />

entanto, a sucessão de árvores (Juniperus, Betula, Pinus Quercus, Corylus e<br />

Alnus) observada nas várias altitudes da região nortenha é semelhante e<br />

bastante diferente daquela que caracteriza a região sul.<br />

No entanto, a expansão de deciduous Quercus na região sul ocorreu<br />

tal como nas regiões baixas e médias do noroeste da Península Ibérica de<br />

forma mais rápida provavelmente como resposta à forte densidade de<br />

zonas refugio durante o precedente LGM nestas zonas.<br />

A correlação directa entre indicadores paleoclimáticos continentais e<br />

marinhos permitiu-nos propôr dois eventuais mecanismos responsáveis pelo<br />

padrão complexo deixado pelos eventos de Heinrich no noroeste da<br />

margem Ibérica.<br />

Este estudo permitiu constatar a presença de duas fases distintas na<br />

vegetação do noroeste Ibérico as quais parecem estar intimamente ligadas<br />

ao sinal complexo deixado pelos eventos de Heinrich (H4, H2 and H1) ao<br />

longo da margem Ibérica.<br />

A primeira fase é marcada por condições oceânicas de superfície<br />

muito frias (evidenciada pelas associações de foraminíferos planctónicos,<br />

δ 18 O e estimativa da temperatura baseada nas alcanonas) e pela virtual<br />

ausência de icebergues no noroeste da margem Ibérica, assim como por um<br />

arrefecimento continental extremo o qual é revelado pelo forte declínio da<br />

floresta de Pinus. Esta primeira fase é ainda marcada pela expansão Calluna<br />

e por um aumento da concentração polínica total sugerindo um aumento<br />

da humidade na Península Ibérica.<br />

A segunda fase, associada à chegada máxima dos icebergs a esta<br />

margem, é caracterizada por condições oceânicas de superfície e<br />

continentais (expansão de Pinus) menos frias e por um aumento da aridez<br />

(representada pelo desenvolvimento de plantas semi-desérticas).<br />

297<br />

F. Naughton, 2007


O impacto do evento H3 no noroeste da Península Ibérica é peculiar<br />

reflectindo condições húmidas durante quase todo o episódio. No entanto,<br />

tal como acontece nos outros eventos de Heinrich (H4, H2 e H1), a<br />

diminuição substancial da temperatura da massa de água de superfície e<br />

da atmosfera (regressão da floresta de Pinus) precede a chegada máxima<br />

dos icebergs durante este evento atípico.<br />

Ao longo deste trabalho foram propostos dois mecanismos principais<br />

para explicar este padrão complexo. O drástico arrefecimento atmosférico<br />

(declínio da floresta de Pinus) detectado durante os eventos H4, H3, H2 e H1<br />

resulta provavelmente da introdução de grandes quantidades água doce<br />

provenientes da fusão dos icebergues no Atlântico Norte a qual provocou a<br />

interrupção da circulação termohalina do Atlântico Norte (MOC). Esta<br />

interrupção da MOC foi seguida por rápidas reorganizações entre o oceano<br />

e a atmosfera as quais favoreceram a transmissão das condições frias para o<br />

noroeste da Península Ibérica.<br />

Sobreimpostas a este mecanismo oceanográfico, variações<br />

semelhantes aos modos negativo e positivo do índice da Oscilação Norte<br />

Atlântica (NAO-like) parecem ter tido um papel crucial no padrão complexo<br />

deixado pelos eventos típicos de Heinrich no noroeste Ibérico. Durante a<br />

primeira fase, a predominância do modo negativo da NAO-like explicaria o<br />

aumento na precipitação invernal e importantes descargas fluviais nesta<br />

região, as quais favoreceriam o transporte de grãos de pólen do continente<br />

para o oceano. Estas condições provocariam ainda o aumento da<br />

temperatura superficial do Atlântico Norte a latitudes situadas acima dos 45°-<br />

50°N, facilitando a fusão dos icebergues provenientes da calote glaciar da<br />

“Laurentide” na cintura de IRD (IRD belt) impedindo o seu transporte para as<br />

latitudes médias do Atlântico Norte. A fusão dos icebergues produziria ainda<br />

um arrefecimento da massa de água superficial nessa zona. A dominância<br />

do modo negativo da NAO-like produziria ainda uma diminuição da<br />

temperatura da camada de água superficial numa zona estreita próxima da<br />

costa entre o noroeste da margem Ibérica e o Grande Mar do Norte<br />

(Greater North Sea) favorecendo a amplificação das condições frias na<br />

nossa zona de estudo. Durante a segunda fase dos eventos de Heinrich<br />

típicos (H4, H2 e H1), a dominância do modo positivo da NAO-like poderiam<br />

298<br />

F. Naughton, 2007


ter provocado a intensificação e a migração para norte dos ventos de oeste<br />

provocando um aumento da aridez na Europa incluindo a zona da Península<br />

Ibérica. Estas condições produziram condições oceânicas de superfície<br />

relativamente quentes nas médias latitudes do Atlântico Norte (~ 20°- 40°N)<br />

favorecendo a migração dos icebergues para sul e a sua fusão nessa região<br />

(incluindo a margem Ibérica). Apesar dos icebergues atingirem a margem<br />

Ibérica, as condições atmosféricas prevalecentes (NAO-like positivo)<br />

favoreceriam um aquecimento da massa de água superficial entre o<br />

noroeste da margem Ibérica e o Grande Mar do Norte latitudes médias do<br />

Atlântico Norte mascarando por isso o arrefecimento causado pela fusão<br />

dos icebergues na zona de estudo.<br />

A predominância de condições húmidas durante o evento atípico,<br />

H3, poderia ser explicada pela permanência de ventos fracos vindos de<br />

oeste, nesta região.<br />

A comparação directa entre condições no continente e no oceano<br />

revelou ainda que durante o LGM, apesar das condições oceânicas de<br />

superfície serem quentes, a expansão da floresta temperada foi largamente<br />

reduzida quando comparada com os episódios interestadiais que<br />

caracterizam o MIS3 tardio (MIS3-Marine isotopic Stage 3).<br />

Neste trabalho são propostos três mecanismos de forma a explicar<br />

esta disparidade entre as condições oceânicas e atmosféricas tais como a)<br />

o aumento do albedo o qual teria provocado uma amplificação do<br />

arrefecimento produzido pela baixa insolação de verão no Hemisfério Norte,<br />

b) o forte contraste sazonal das altas latitudes do Atlântico Norte como<br />

resultado de um arrefecimento substancial da temperatura de inverno, e c)<br />

a diminuição da concentração de CO2 na atmosfera. No entanto, a<br />

humidade relativa que dominou o oeste da Península Ibérica durante o LGM<br />

parece resultar do aumento da intensidade da MOC em relação aos<br />

eventos precedente e antecedente a este, ou seja, o H2 e o H1.<br />

Para além do último período glaciário interessámo-nos ainda em<br />

compreender a variabilidade climática milenar que ocorreu durante a última<br />

deglaciação (19 500-7000 anos cal BP) nas médias latitudes do Atlântico<br />

299<br />

F. Naughton, 2007


Norte e no noroeste da Península Ibérica. Este período é caracterizado por<br />

um aumento da insolação de verão no Hemisfério Norte e pela forte<br />

redução do volume de gelo acumulado nos pólos. A deglaciação engloba<br />

o final do LGM, o evento H1, o interestadial Bölling-Alleröd (B-A), o Dryas<br />

recente (YD) e inicio do Holocénico (11 500-8 000 anos cal BP).<br />

O aquecimento continental (expansão de Quercus deciduous) e<br />

oceânico que caracteriza o evento Bölling-Alleröd (B-A) foi favorecido pelo<br />

aumento da insolação de verão das latitudes médias do Hemisfério Norte e<br />

pela intensificação da MOC. A expansão máxima de Quercus deciduous, a<br />

qual reflecte um aquecimento continental extremo ocorreu há cerca de<br />

14 000 anos cal BP e é síncrona do pico máximo da temperatura na<br />

Gronelândia e do episódio de subida súbita do nível do mar (MWP 1A). Este<br />

aquecimento severo do Hemisfério Norte poderá ter sido o impulsionador<br />

deste drástico evento designado “Meltwater Pulse 1A”.<br />

A subsequente diminuição da floresta de Quercus no noroeste da<br />

Península Ibérica contemporânea da diminuição da temperatura da massa<br />

de água superficial no oceano adjacente caracteriza o Dryas recente (YD)<br />

nessa região. A redução em vez da interrupção da MOC e o aumento da<br />

insolação de verão das latitudes médias do Hemisfério Norte favoreceram a<br />

redução em vez do total declínio da floresta decídua de Quercus no<br />

noroeste Ibérico. Para além do arrefecimento, a expansão de plantas semi-<br />

desérticas reflecte o aumento da aridez continental. Este incremento das<br />

condições áridas parece resultar da dominância do modo positivo da NAO-<br />

like na região Norte Atlântica.<br />

O máximo térmico do Holocénico (HTM), caracterizado por uma forte<br />

insolação de verão das latitudes médias do Hemisfério Norte, foi detectado<br />

no noroeste da Península Ibérica entre 11 700 e 8 200 anos cal BP e é<br />

contemporâneo da expansão máxima da floresta de Quercus deciduous<br />

durante o actual interglaciário.<br />

Por volta dos 8 200 anos cal BP, o rápido arrefecimento da massa<br />

oceânica de superfície das médias latitudes do Atlântico Norte e a<br />

diminuição da floresta decídua de Quercus e Corylus no noroeste da<br />

Península Ibérica, reflectem o evento frio “8.2 ky” nessa região.<br />

300<br />

F. Naughton, 2007


Este episódio resulta da culminação de uma série de episódios<br />

relacionados com o colapso da calote glaciar da “Laurentide” a qual<br />

provocou a amplificação da diminuição da temperatura na Europa e na<br />

Gronelândia.<br />

Após o evento “8.2 ky” a diminuição gradual da floresta temperada é<br />

contemporânea da diminuição da temperatura induzida pela diminuição<br />

da insolação de verão das latitudes médias do Hemisfério Norte. Isto sugere<br />

que a regressão da floresta temperada parece ter sido mais afectada pelas<br />

variações orbitais do que pelo impacto antrópico.<br />

A baixa resolução das sondagens marinhas profundas para o período<br />

Holocénico impediu a detecção da variabilidade climática milenar durante<br />

o actual interglaciário nas médias latitudes do Atlântico Norte, e em<br />

particular de identificar o evento multi-secular que ocorreu entre ~8 600 e 8<br />

000 anos cal BP. De forma a resolver esta lacuna foi efectuada a análise<br />

polínica de uma sondagem marinha pouco profunda com cerca de 2.72 m<br />

a qual foi recolhida na plataforma continental do noroeste de França (VK03-<br />

58Bis). As variações do coberto vegetal e a estimativa dos parâmetros<br />

climáticos obtidos nesta sondagem, permitiram detectar variações<br />

climáticas de escala orbital e sub-orbital, para os últimos 8 850 anos, nesta<br />

região.<br />

O declínio gradual da floresta temperada juntamente com a<br />

diminuição da temperatura de verão (MTWA-mean temperature of the<br />

warmest month) desde 8 855 até, pelo menos, 2 000 anos cal BP é<br />

contemporâneo do arrefecimento progressivo da temperatura na<br />

Gronelândia, assim como, da redução dos valores de insolação de verão<br />

nas latitudes médias do Hemisfério Norte. Ao mesmo tempo, a diminuição da<br />

sazonalidade segue, tal como seria de esperar, o aumento da precessão.<br />

Entre 8 739 e 8 387 anos cal BP, a floresta de Corylus expande-se em<br />

detrimento do Quercus deciduous, como resposta a uma amplificação do<br />

contraste sazonal, resultante da expansão de gelo marinho invernal nas altas<br />

latitudes do Atlântico Norte, contrariando o padrão geral de forçagem<br />

orbital. Este forte contraste sazonal, resulta da expulsão drástica de água<br />

301<br />

F. Naughton, 2007


doce dos lagos de “Agassiz” e “Ojibway” e, da gradual redução da MOC. O<br />

súbito declínio da floresta de Corylus, entre 8 387 e 8 062 anos cal BP, marca<br />

tal como em outras regiões da Europa, o evento frio “8.2 kyr”, no noroeste de<br />

França. Este episódio, foi provavelmente produzido pela redução severa da<br />

MOC a qual está interligada com um fluxo mínimo da “Iceland-Scotland<br />

Overflow Water” (ISOW) provocando uma diminuição suplementar da<br />

temperatura invernal, na Europa e na Gronelândia. No entanto, apesar do<br />

forte declínio da floresta de Corylus o contraste sazonal permaneceu<br />

elevado durante este evento. De facto, o forte contraste sazonal registado<br />

entre 8 739 e 8 062 anos cal BP reflecte o evento multi-secular "8.6-8.0 kyr”<br />

nesta região.<br />

Após os estádios finais de expulsão de água dos lagos “Agassiz” e<br />

“Ojibway”, o clima torna-se relativamente mais estável. No entanto, estão<br />

registados uma série de ligeiros e rápidos episódios frios os quais, estão<br />

associados a um pequeno arrefecimento invernal e a um ligeiro aumento da<br />

precipitação.<br />

Para além da análise polínica, foi ainda efectuado um estudo das<br />

associações dinocistos e de gastrópodes do tipo Turritella communis o qual<br />

permitiu detectar variações regionais tais com: a) migração para sul da zona<br />

biogeográfica marinha Boreal, entre 8 739 e 8 479 anos cal BP, como<br />

resposta ao aumento da área coberta por gelo marinho no Atlântico Norte<br />

permitindo o estabelecimento das comunidades bentónicas (T. Communis)<br />

no noroeste da margem françesa e, b) a abertura do canal da mancha<br />

entre 8479 e 8387 anos cal BP a qual provocou grandes modificações<br />

ambientais nesta região e consequente mortalidade da T. communis assim<br />

como a forte diminuição das percentagens do Lingulodinium<br />

machaerophorum.<br />

A variabilidade climática tem uma forte influência na evolução dos<br />

sistemas costeiros principalmente desde a última deglaciação. De facto, o<br />

aumento da insolação de verão das altas latitudes do Hemisfério Norte<br />

provocou a diminuição do volume de gelo, e como consequência um<br />

302<br />

F. Naughton, 2007


aumento do nível do mar global o qual teve um forte impacto na evolução<br />

dos sistemas costeiros mundiais.<br />

Na tentativa de compreender o impacto destes mecanismos forçadores<br />

globais, nomeadamente o clima e o nível do mar na evolução<br />

geomorfológica dos sistemas costeiros do noroeste da Península Ibérica, foi<br />

efectuado um estudo multidisciplinar em duas sondagens recolhidas na<br />

embocadura do estuário do Douro (noroeste de Portugal).<br />

A primeira metade do período Holocénico (10 720 e 6 530 anos cal BP)<br />

é caracterizada por uma floresta de Pinus-Quercus-Alnus reflectindo um<br />

clima quente e húmido na bacia do Douro. Durante este período, o<br />

aumento gradual de foraminíferos bentónicos característicos de zonas de<br />

plataforma e talude continental, assim como a presença de equinodermes<br />

sugere o aumento gradual do nível do mar nessa zona. Os grãos de pólen,<br />

representantes da vegetação regional que colonizava a bacia hidrográfica,<br />

foram transportados pelo rio ao longo do antigo canal principal e<br />

depositados na zona de estudo a qual seria nesse momento uma antiga<br />

margem do rio.<br />

Há cerca de 6 530 anos cal BP uma drástica variação ambiental,<br />

associada ao desaceleramento da subida do nível do mar e aumento do<br />

hidrodinamismo do rio, é marcada pela deposição de cerca de 5 m de<br />

cascalho na parte sul deste estuário. O estabelecimento desta barreira<br />

cascalhenta na parte sul deste estuário ocorreu entre 6 530 e 1 500 anos cal<br />

BP e provocou a migração para norte do canal principal impedindo o<br />

transporte e a deposição dos grãos de pólen representantes da vegetação<br />

regional na zona de estudo após 1 500 anos cal BP.<br />

A formação desta barreira e a consequente migração para norte do<br />

canal principal do rio contribuíram para o alargamento da embocadura<br />

deste estuário.<br />

303<br />

F. Naughton, 2007


304<br />

F. Naughton, 2007


Anexos


Anexo A<br />

Climate variability of the last five isotopic interglacials:<br />

direct land-sea-ice correlation from the multiproxy analysis of north<br />

western Iberian margin deep-sea cores.<br />

S. Desprat, M.F., Sánchez Goñi, F., Naughton, J.-L., Turon, J. Duprat, B. Malaizé, E. Cortijo and<br />

J.-P. Peypouquet<br />

In press in The climate of past interglacials. Elsevier publications.


Climate variability of the last five isotopic interglacials:<br />

direct land-sea-ice correlation from the multiproxy analysis of north western<br />

Iberian margin deep-sea cores.<br />

S. Desprat a *, M.F., Sánchez Goñi a , F., Naughton b , J.-L., Turon b , J. Duprat b , B. Malaizé b , E. Cortijo c<br />

and J.-P. Peypouquet a<br />

a Ecole Pratique des Hautes Etudes, Paléoclimatologie et Paléoenvironnements marins, Département<br />

de Géologie et Océanographie, Université Bordeaux 1, Avenue des Facultés, 33405 Talence, France<br />

b Département de Géologie et Océanographie, Université Bordeaux 1, Avenue des Facultés, 33405<br />

Talence, France<br />

c Laboratoire des Sciences du Climat et de l’Environnement, LSCE-Vallée, Bât. 12, avenue de la<br />

Terrasse, F-91198 Gif-Sur-Yvette cedex, France<br />

*corresponding author: s.desprat@epoc.u-bordeaux1.fr


Abstract<br />

The last five marine isotopic interglacials (Marine Isotope Stages 11, 9, 7, 5 and 1) were investigated<br />

in Iberian margin deep-sea cores, using terrestrial (pollen) and marine (planktic foraminifera, benthic<br />

and planktic oxygen isotopes) climatic indicators. This work shows that the climatic variability<br />

detected on the continent is contemporaneously recorded in the ocean, but temperature changes are not<br />

in phase with ice volume variations. The comparison of the different marine isotope stages highlights a<br />

common pattern of climatic dynamic within these interglacials. This dynamic is characterized by three<br />

major climatic cycles, related to orbital cyclicity, on which suborbital climatic fluctuations are<br />

superimposed. Particularly, suborbital events interrupt the deglacial warming associated to<br />

Terminations IV to I and the second major warm period of each isotopic interglacial as well as the<br />

transitions towards glacial marine isotope stages. MIS 7 displays a short first warm period (~8 ka)<br />

followed by a striking cold and dry period succeeded by a new strong warmth. In contrast, MIS 11<br />

presents the longest (~31 ka) period of the last 450,000 years.


1. Introduction<br />

Forecasting the future climatic evolution of the current interglacial period is a great challenge.<br />

Before that, it is necessary to determine the evolution of the past interglacials and evaluate the<br />

response of different components of the Earth’s climatic system. Due to the Earth’s astronomical<br />

configuration, Marine Isotope Stage (MIS) 11 is the best candidate to be the analogue of MIS 1.<br />

However, characterising the climatic evolution over different situations of insolation forcing will allow<br />

us a better understanding of climate dynamics during interglacial periods. The continental<br />

<strong>paleoclimat</strong>ic records covering the last 425,000 years are rare and often fragmentary, and their<br />

chronologies are difficult to establish. This impedes the comparison of the climatic changes detected<br />

on land with those identified in the oceanic and cryospheric realms.<br />

We present, here, the first direct land-sea-ice correlation for the last five isotopic interglacials<br />

(MIS 11, 9, 7, 5 and 1). The main purpose of this work is to document the climatic variability of these<br />

periods and to assess the phase relationship between the responses of the different Earth’s<br />

environments –continent, ocean and ice– to climatic changes in order to discern analogies and<br />

differences between them. For that, a multiproxy study (pollen, assemblages of planktic foraminifera,<br />

and planktic and benthic δ 18 O) was performed from several NW Iberian margin deep-sea cores. By<br />

comparing the last five isotopic interglacials, we will highlight, on the one hand, the similarities of<br />

their climatic dynamic despite different astronomical forcing and, on the other hand, the dissimilarities<br />

concerning duration, warmth magnitude and forest succession of the warm periods.<br />

2. Present-day environmental setting and pollen signal in the Iberian margin<br />

The Iberian margin deep-sea cores were retrieved ~100 km off the Galician coast at ~2,100 m<br />

of water depth (Fig. 1). This site is at present under the influence of the North Atlantic Deep Water.<br />

The north western Iberian climate is considered temperate and humid as a result of the influence of<br />

dominant Atlantic winds over the year. Mean annual temperature is 12.5°C (Mean Temperature of the<br />

Coldest Month, MTCO = 5-12°C; Mean Temperature of the Warmest Month, MTWA = 17-22°C) and<br />

mean annual precipitation is between 1000 and 2000 mm.an -1 (Atlas Nacional de España, 1992). This<br />

region, incised in the north by the Rias Baixas valleys (Galician coast basin) and the Miño-Sil river<br />

(Sil basin) (Atlas Nacional de España, 1993) and crossed in the south by the Douro river, belongs to<br />

the Eurosiberian and sub-Mediterranean regions (Ozenda, 1982). At present, deciduous oak woodlands<br />

(Quercus robur, Q. pyrenaica and Q. petraea), heaths (Ericaceae including Calluna), brooms


(Genista) and gorses (Ulex) dominate the vegetal cover of north western Iberia (Alcara Ariza et al.,<br />

1987).<br />

Studies on present-day pollen deposition in marine sediments show that pollen grains reach<br />

marine sites from the adjacent continent by both fluvial and aeolian transport and subsequent sinking<br />

through the water column (Chmura et al., 1999; Dupont and Wyputta, 2003; Heusser, 1978; Heusser<br />

and Balsam, 1977). Further, it is suggested that cores located near continental regions with well<br />

developed hydrographic basins and prevailing offshore winds, as it is the case of our cores, mainly<br />

recruit pollen from rivers (Dupont and Wyputta, 2003; Heusser, 1978; Turon, 1984). The north<br />

western Iberian rivers, mainly the Douro and Miño and slightly the Rias Baixas, provide sediments to<br />

the shelf area. On the shelf, the fine-particle sediments in suspension are transported northwards by<br />

poleward currents, and some are deposited in the Douro and Galicia Mud Patches (Dias et al., 2002).<br />

Extreme storm events can produce re-suspension of some sediment from the mud patches and transport<br />

of sediments off the shelf can occur (Jouanneau et al., 2002; Vitorino, 2002). Pollen grains belonging<br />

to the fine-particle fraction have a similar behaviour as fine sediments during the sedimentary<br />

processes (Chmura and Eisma, 1995; Muller, 1959). This suggests that pollen grains preserved in our<br />

Iberian margin cores mainly come from the Galician and Douro fluvial basins. The comparison of<br />

marine and continental modern pollen samples with the present-day Iberian vegetation shows that our<br />

marine pollen records represent an integrated image of the regional vegetation of the north western<br />

part of the peninsula (Naughton et al., in prep).<br />

3. Material and method<br />

Records of pollen and classical climate indicators (planktic foraminafera assemblages, benthic<br />

and planktic δ 18 O) for the last five isotopic interglacials derive from three north western Iberian margin<br />

deep-sea cores (MD01-2447, MD03-2697, MD99-2331) (Fig. 1). They were retrieved at the same<br />

coordinates on board of the Marion Dufresne oceanographic ship, using the giant corer CALYPSO. As<br />

shown in figure 2, the intervals corresponding to MIS 11, 9 and 7 were studied in core MD01-2447.<br />

The beginning of stage 9, being unfortunately disturbed in this core, was studied in the twin core<br />

MD03-2697. We present therefore a composite record of MIS 9 built from the correlation of several<br />

marine proxies analysed in the twin cores (lightness L*, CaCO3 content, percentages of N. pachyderma<br />

s., coiling ratio of Globorotalia truncatulinoides) (Desprat, 2005a). MIS 1 and 5 records come from<br />

the third core MD99-2331 (Naughton et al., in prep; Sanchez Goñi et al., 2005).


3.1 Pollen analysis<br />

Each interval corresponding to MIS 11, 9, 7, 5 and 1 was subsampled for pollen analysis at 10<br />

or 5 cm intervals. The sample preparation technique followed the procedure described by de Vernal et<br />

al. (1996) improved at the "Département de Géologie et Océanographie", University Bordeaux I<br />

(Desprat, 2005a). After chemical and physical treatments (cold HCl, cold HF and sieving through 10<br />

µm nylon mesh screens), the residue was mounted unstained in glycerol. Pollen was counted using a<br />

Zeiss Axioscope light microscope at 500 and 1250 (oil immersion) magnifications. A minimum of 100<br />

pollen grains excluding the over-represented Pinus grains in marine deposits (Heusser and Balsam,<br />

1977; Turon, 1984), were counted in each of the 327 samples analysed. The pollen percentages for<br />

each taxon are based on a main pollen sum that excludes Pinus, aquatic plants, pteridophyte spores and<br />

indeterminable pollen. Pinus percentages were calculated from the main sum plus<br />

Pinus. Spores and aquatic pollen percentages were obtained from the total sum (pollen + spores +<br />

indeterminables + unknowns).<br />

3.3 Isotopic analyses<br />

The sampling resolution interval oscillates between 20 and 2 cm for Cibicides wuellerstorfi and<br />

Melonis barleeanus benthic foraminifera and Globigerina bulloides planktic foraminifera. Each<br />

specimen has been picked up within the 250 - 315 µm grain-size fraction, and cleaned with distilled<br />

water. The preparation of each aliquot (4 to 8 specimens, representing a mean weight of 80 µg) has<br />

been done using the Micromass Multiprep autosampler, using an individual acid attack for each<br />

sample. The CO2 gas extracted has been analyzed against NBS 19 standard, taken as an international<br />

reference standard. The isotopic analyses of core MD01-2447 and MD99-2331 have been carried out<br />

at the "Département de Géologie et Océanographie" (Bordeaux 1 University, France), using an Optima<br />

Micromass mass spectrometer, and those of core MD03-2697 were performed at the "Laboratoire des<br />

Sciences du Climat et l’Environnement" (Gif-sur-Yvette, France), using a delta plus Finnigan isotope<br />

mass spectrometer. All the isotopic results are presented versus PDB. The mean external<br />

reproducibility of powdered carbonate standards is ±0.05‰ for oxygen.<br />

The δ 18 O values for Cibicides wuellerstorfi and Melonis barleeanus were adjusted by +0.64 per<br />

mil and +0.36 per mil, respectively, to account for species-dependent departure from isotopic<br />

equilibrium (Duplessy et al., 1984; Graham et al., 1981; Jansen et al., 1988; Shackleton and Opdyke,<br />

1973).


3.4. Chronological framework<br />

The age model of the intervals corresponding to MIS 11 and 9 is based on the graphical<br />

correlation of the benthic δ 18 O curve with the Low Latitude Stack of Bassinot et al. (1994) (Desprat,<br />

2005a; Desprat et al., 2005b). The chronology of MIS 7 section also derived from a graphical<br />

correlation but in this case using the benthic-stack of Martinson et al. (1987) (Desprat et al.,<br />

submitted).<br />

The chronologies of MIS 5 and 1 are, in contrast, independent of the astronomical calibration.<br />

That of the last isotopic interglacial is based on the correlation of the major climatic phases detected in<br />

core MD99-2331 with those identified and dated in the southern Iberian margin core MD95-2042<br />

using the MD95-2042 chronology of Shackleton et al. (2002) (Sánchez Goñi et al., 2005). For the<br />

interval corresponding to the last 25,000 years, the age model was established using the chronology of<br />

the climatic episodes identified in other North Atlantic records and the ages assigned to several welldated<br />

botanical events in the Iberian Peninsula (Naughton et al., in prep).<br />

4. The climatic variability of the last five isotopic interglacials in and off NW Iberia<br />

4.1. General climatic dynamic<br />

During the last five isotopic interglacials, the warm periods in north western Iberia are<br />

characterized by the development of the temperate and humid forest, principally deciduous oak. In<br />

turn, open vegetation dominated by Poaceae and Asteraceae, with some semi-desert plants, or mainly<br />

composed by heathland, expands during cold periods. The recorded vegetation changes indicate that<br />

climate has strongly oscillated during the previous isotopic interglacials.<br />

The climatic evolution detected on the continent parallels the oceanic changes reflected by<br />

marine proxies (planktic foraminifera δ 18 O and percentages). Indeed, each cold episode is marked by<br />

an increase of the percentages of the polar planktic foraminifera N. pachyderma s. and heavier planktic<br />

δ 18 O values. The record of planktic foraminifera assemblages is only available for MIS 11, 9 and 7. It<br />

shows that the tropical and summer subtropical species generally reach their maximal development<br />

during the warm periods detected on the continent.<br />

During each of the Terminations I to IV, an abrupt cold event interrupts the development of the<br />

temperate and humid forest associated to the deglacial warming: Younger Dryas, post-Zeifen stadial,<br />

MD47-7-S1 and MD47-9-S1 (Fig. 3). These cold events have a clear imprint in the oceanic realm, in<br />

particular that of Termination IV which is slightly marked on the continent. These coolings appear of<br />

different magnitude and during Termination II, III and IV, they occur at the onset of minimum ice


volume. For example, MD47-8-I1/MD47-7-S1 cycle shows the strongest amplitude of vegetation<br />

changes. The occurrence of such an episode during Termination V is still unknown because our<br />

sedimentary core does not cover the whole MIS12/MIS11 transition.<br />

As suggested by the long European pollen sequences (Reille et al., 2000; Reille et al., 1998;<br />

Tzedakis et al., 2001), our direct land-sea-ice correlation confirms that each isotopic interglacial is<br />

characterized by three major warm periods on the continent associated to low ice volume, in response<br />

to the astronomical forcing (Table 1, Fig. 3). Indeed, the major forested periods in north western Iberia<br />

are associated to low ice volume contrasting with the open vegetation phases related with ice cap<br />

development. Nevertheless, our direct land-sea-ice correlation puts forward that the ice volume<br />

changes are not synchronous with the temperature shifts on the continent and in the ocean. As<br />

observed by Sánchez Goñi et al. (1999, 2005), Shackleton et al., 2002 and Tzedakis et al. (2004), the<br />

limits of isotopic substages do not correspond to those of the climatic phases detected in western<br />

Iberia. For example, the Eemian in Iberia does not correspond to the entire MIS 5e (Sánchez Goñi et<br />

al., 1999, 2005; Shackleton et al., 2002).<br />

During the isotopic interglacials MIS 11, 9, 7 and 5, the first major warm periods (Vigo,<br />

Pontedevedra, Arousa and Eemian) are marked by a more developed forest than the following ones.<br />

This is particularly true for MIS 11. Indeed, the climate optimum of each stage, detected by the<br />

maximal expansion of oak forest, Mediterranean plants and the maximal contraction of pine<br />

woodlands, occurs during these first major warm periods. These climate optima are contemporaneous<br />

to the ice volume minimum of each stage. However, MIS 7 presents another particularity likely related<br />

to especially strong insolation maximum: the second major warm period (Ribeira) is also marked by<br />

strong expansion of the temperate and humid forest, development of warm planktic foraminifera and<br />

important ice volume decrease. This implies that Ribeira would be, at least, as warm as the Arousa<br />

interglacial. Therefore, MIS 7 displays two climatic optima both associated to low ice volume as<br />

shown by the benthic isotopic record.<br />

The Vigo, Pontedevedra and Eemian interglacials are followed by a strong cold period (MD47-<br />

11-S1, MD97-9-S2 and Mélisey I). In contrast, MIS 7 includes a suborbital cycle (MD47-7-S2/MD47-<br />

7-I2) between the Arousa interglacial and the strong cold period MD47-7-S3.<br />

MD47-7-S3 was the coldest and driest period on the continent of the last five isotopic<br />

interglacials, as indicated by the highest values of grassland and semi-desert taxa of our pollen record.<br />

This phase is also marked by the strongest decrease of sea surface temperatures as shown by the<br />

maximum percentages of N. pachyderma left coiling. This cold episode is also coeval with the largest


ice volume increase (MIS 7.4) of the last past isotopic interglacials, being similar to the glacial<br />

maximum of MIS 8. These unusual very cold conditions and huge ice-sheet enlargement within an<br />

isotopic interglacial, are also recorded by northern North Atlantic ODP sites 983 and 980 (McManus et<br />

al., 1999; Channell et al., 1997). It is remarkable that this episode occurs during the most important<br />

insolation minimum of the last 450,000 years.<br />

Other suborbital events are superimposed to this orbital climatic variability. The Vigo<br />

interglacial is marked by two cool events circa 417 and 402 ka, having a clear imprint on vegetation<br />

and planktic foraminifera. More especially, the temperate and humid forest developments associated to<br />

the second major warm periods of MIS 11, 9 and 5 (Moana, Sanxenxo, St-Germain I) are all<br />

interrupted by a cold episode (MD47-11-S1, MD97-9-S2 and Montaigu, respectively). After our age<br />

model, these cold events are short, between 2 and 4 kyrs. Moreover, during these episodes, the benthic<br />

δ 18 O values become heavier, in particular during MD97-9-S2. This indicates a significant increase of<br />

ice volume during the cold event within MIS 9c, as shown by Tzedakis et al. (2004). These cold<br />

fluctuations are also clearly recorded in the Velay sequence (Reille et al., 2000). In contrast, MIS 7<br />

does not show such a climatic event.<br />

During the interglacial-glacial transitions MIS 7-MIS 6 and MIS 5-MIS 4, another climatic<br />

cycle of minor order is detected. The cooling associated to the MIS 9-MIS 8 transition seems also to be<br />

interrupted by a warm oscillation which needs to be confirmed by supplementary analysis.<br />

Nonetheless, all these warm events are clearly recorded in the oceanic realm by light values of planktic<br />

δ 18 O and an increase of warm planktic percentages.<br />

In sum, in spite of the different astronomical configuration of the last five isotopic interglacials<br />

a common climatic evolution pattern emerges. However, each warm phase is characterized by different<br />

duration, amplitude, and forest succession.<br />

4.2. Warmth amplitudes of the last 450,000 years<br />

At present, discussions are open on the different amplitudes of warmth during the last 450,000<br />

years. The results are often contradictory, depending on the regions and proxies concerned. Some<br />

works suggest that the warmest phase of MIS 11 shows the highest temperatures of the last 500,000<br />

years (Howard, 1997; Droxler and Farrel, 2000; Berstad et al., 2002). However, this idea is challenged<br />

by many works (Hodell et al., 2000; Bauch et al., 2000; Kunz-Pirrung et al., 2002; McManus et al.,<br />

2003) which have shown that MIS 11 was not warmer than today. The deuterium signal of Vostok ice


core records the highest temperatures for MIS 9 (Petit et al., 1999) but the recent results of EPICA-<br />

Dome C ice core do not confirm this idea (EPICA community members, 2004). These new data also<br />

show higher temperatures in Antarctica during MIS 11 than during the Holocene (EPICA community<br />

members, 2004).<br />

As previously shown, in our record, each isotopic stage displays its climate optimum during the<br />

first warm period, excepting MIS 7 which presents a second optimum during Ribeira. The first<br />

optimum of this isotopic interglacial (Arousa) is marked by the highest percentages of temperate and<br />

humid trees. However, the Mediterranean plants are scantily represented during this interval and only<br />

by evergreen Quercus which does not reveal clear Mediterranean conditions. In contrast, although the<br />

Vigo, Ribeira, Eemian and Holocene interglacials are characterized by lower percentages of temperate<br />

and humid forest than those during the Arousa interglacial, they record true Mediterranean species<br />

such as Pistacia, Olea or Cistus. For this reason, it remains difficult to determine what period is the<br />

warmest of the last five climatic cycles in northern Iberia.<br />

The warm planktic foraminifera record of MIS 11, 9 and 7 indicates the highest sea surface<br />

temperatures during MIS 11 climatic optimum. However, the development of these warm foraminifera<br />

is only slightly stronger than during the other climatic optima. Therefore, the Vigo interglacial may be<br />

the warmest period of the last 450,000 years but the difference of temperature does not appear large.<br />

Moreover, the benthic isotopic signal does not display weaker ice volume during Vigo interglacial than<br />

during Pontedevedra, Ribeira, Eemian or Holocene interglacials. The genesis of such a warm<br />

interglacial during MIS 11 remains still a mystery in palaeoclimatology taking into account the weak<br />

insolation forcing.<br />

4.2. Duration of the forest phases<br />

MIS 11 is marked by a long first major warm period, lasting ~31 ka after our age model. The<br />

Vigo interglacial appears two times longer than the Eemian (~16 ka), and at least three times longer<br />

than the Pontedevedra and Arousa interglacials (~11 and 8 ka, respectively). The Holocene began<br />

10,000 years ago and it is already longer than the Arousa interglacial.<br />

On the basis of the pollen analysis of the south western Iberian margin deep-sea core MD01-<br />

2443, Tzedakis et al. (2004) and Roucoux et al. (this volume) suggest a very short forest phase during<br />

MIS 9e, lasting 3,600 years, after which Ericaceae expand. To bypass the difference in duration<br />

inferred from the age models, we have tuned our planktic δ 18 O record to that of MD01-2443. After this<br />

exercise, the resulting duration of the period Pontedevedra is of 13,000 years. It is possible that the


cold/arid event responsible for the shortening of the first warm period of MIS 9 in south western Iberia<br />

is not shown by our sequence due to a too low resolution analysis. However, our results do not suggest<br />

that this abrupt climatic deterioration detected in south western Iberia brings to an end the first MIS 9<br />

forest period in the north western part of the Peninsula. In the same way, the Praclaux sequence<br />

(Massif Central, France) also records a long warm period even if it includes a slight forest reduction<br />

(Tzedakis et al., 2004).<br />

Our observation confirms, despite the uncertainties associated to the age scale of our record,<br />

that the warmest period of MIS 11 would be longer than those of the following isotopic interglacial<br />

stages, and so far three times longer than the Holocene.<br />

4.4. Forest successions<br />

The warm periods of our record are marked by the development of the pioneer trees, principally<br />

Betula, followed by the expansion of deciduous oaks and that of hazel trees. This is in agreement with<br />

the classical vegetational succession during interglacial period in northern Europe, described by Van<br />

der Hammen et al. (1971). This ideal succession sees at the latest stages the expansion of hornbeam,<br />

beech, fir, finishing by the development of the boreal forest with spruce. In the north western Iberian<br />

region, the boreal forest phase is never reached during the warm periods. Moreover, the expansion of<br />

the latecomer trees is different from one period to another:<br />

• Abies strongly developed only at the end of the Vigo interglacial and St Germain Ic and it was only<br />

present at the end of the Eemian and Bueu (the last forested period of MIS 9).<br />

• Carpinus betulus expanded in the second part of all first major warm phases when deciduous<br />

Quercus decreased excepting during the Holocene. However, the hornbeam expansion was very strong<br />

during the Eemian, smaller during the Pontedevedra and Arousa interglacials and very weak during the<br />

Vigo interglacial. During the other warm periods, hornbeam had also its maximal expansion after that<br />

of deciduous oak.<br />

• Fagus never had an important expansion at the end of the first major warm phases. Beech is only<br />

sporadically recorded at the end of the Vigo and Pontedevedra interglacials. In contrast, during the<br />

periods Bueu, Ribeira and Rianxo, Fagus plays an important role in the vegetal cover of the north<br />

western Iberian region, always associated to Carpinus betulus. It is also noteworthy that beech<br />

developed rapidly and strongly at the beginning of the Ribeira interglacial (Desprat et al., submitted).<br />

The different behaviour of these three trees depending on the periods and regions has been<br />

previously noticed by Tzedakis and Bennett (1995) and Tzedakis et al. (2001). Disentangling the


factors responsible for the settlement of tree species in given region and period is a difficult task. The<br />

late expansion of some tree species can be linked: a) to their migration rate in relation with their own<br />

dispersion mechanisms such as reproduction or seed scattering and with their competencies to develop<br />

on more or less mature soils, b) to the distance with the glacial refugial zone, c) to the interspecific<br />

competition, d) the individual response of each specie to a climatic change, and e) also to the direct<br />

effect of the climate (Huntley and Webb, 1989; Huntley, 1996). The climate can also play an indirect<br />

role in changing the interspecific relationships (Lischke et al., 2002). Small differences of climatic<br />

conditions at the beginning of a warm phase can also influence the development of the late-expanding<br />

trees (i.e. Carpinus, Abies, Fagus) (Tzedakis et al., 2001). Moreover, reduced diversity of taxa such as<br />

Fagus, Carpinus and Abies may imply that they are more susceptible to disease or adverse climatic<br />

conditions (Tzedakis et al., 2001). During a short warm period such as the Arousa interglacial, the<br />

virtual absence of beech and fir in north western Iberia may be associated to a too short time for<br />

migrating from faraway refugial areas to Iberian Peninsula, likely in relation with their own migration<br />

mechanisms. Nevertheless, the biotic processes cannot explain the very late arrival of Fagus in north<br />

western Iberia, approximately 25,000 years after the beginning of the Vigo interglacial, since it<br />

developed only 7,000 years after the beginning of the Pontedevedra interglacial. Therefore, the<br />

different timing and magnitude of the expansion of the late succession trees is somehow linked to the<br />

inherent climatic conditions of each warm phase of the different isotopic interglacials, which are<br />

related with the ice-sheet extension and the orbital parameters.<br />

5. Conclusions<br />

This work constitutes a new step in documenting the climatic variability of interglacial isotopic<br />

stages. It provides the first direct land-sea-ice correlation of the last five isotopic interglacials (MIS 1,<br />

5, 7, 9 and 11) from the multi-proxy analysis of three pollen-rich north western Iberian margin cores.<br />

This record puts forward the phasing, previously identified during MIS 5, between changes in oceanic<br />

surface conditions and continental climate during the previous isotopic interglacials. Despite the<br />

differences of astronomical forcing, several similarities between these isotopic stages emerge: a) the<br />

occurrence of three major climatic cycles, related to the orbital cyclicity, b) a climatic optimum during<br />

the first major warm periods, associated to ice volume minimum, c) a suborbital cold event<br />

interrupting the second major warm period, and d) a suborbital climatic instability during the glacialinterglacial<br />

and interglacial-glacial transitions.


The largest insolation oscillations controlling MIS 7 may explain the discrepancies between the<br />

climatic variability of this isotopic interglacial and the observed general scheme: a second major warm<br />

period at least as warm as the first one and preceded by a very cold and dry episode associated to an<br />

unusual important ice volume. Another striking feature of this stage is the very short first warm period,<br />

the Arousa interglacial, which is even shorter than the Holocene. In contrast, MIS 11 presents the<br />

longest warm period (Vigo interglacial) of the last 450,000 years, three times longer than our present<br />

interglacial.<br />

Acknowledgements<br />

Financial support was provided by IPEV and PNEDC French programs. We thank logistics and coring<br />

teams on board of the R/V Marion Dufresne II during the Ginna, Geosciences and Picabia<br />

oceanographic cruises and Marie-Hélène Castera, Karine Charlier, Olivier Ther and Françoise Vinçon<br />

for invaluable technical assistance. This paper is Bordeaux 1 University, UMR-CNRS 5805 EPOC<br />

Contribution n°.<br />

References<br />

Alcara Ariza, F., Asensi Marfil, A., de Bolos y Capdevilla, O., Costa Tales, M., Arco Aguilar, M.,<br />

Diaz Gonzales, T.E., Diez Garretas, B., Fernandez Prieto, J.A., Fernandez Gonzales, F., Izco<br />

Sevillando, J., Loidi Arregui, J., Martinez Parras, J.M., Navarro Andres, F., Ninot I Sugranes, J.M.,<br />

Peinado Lorca, M., Rivas Martinez, S., Sanchez Mata, D., Valle Guitierrez, C., Vigo I Bonada, J.,<br />

Wildpret de la Torre, W., 1987. La vegetación de España. Collection Aula Abierta. Universidad de<br />

Alcala de Henares, 544 pp.<br />

Atlas Nacional de España, C., 1992. Ministerio de Obras Publicas y Transportes, Direccion General<br />

del Instituto Geografico Nacional, Madrid.<br />

Atlas Nacional de España, H., 1993. Ministerio de Obras Publicas, Transportes y Medio Ambiente,<br />

Direccion General del Instituto Geografico Nacional, Madrid.<br />

Bassinot, F.C., Labeyrie, J.D., Vincent, E., Quidelleur, X., Shackleton, N.J., Lancelot, Y., 1994. The<br />

astronomical theory of climate and the age of the Brunhes-Matuyama magnetic reversal. Earth and<br />

Planetary Science Letters 126, 91-108.<br />

Bauch, H.A., Erlenkeuser, H., Helmke, J.P., Struck, U., 2000. A <strong>paleoclimat</strong>ic evaluation of marine<br />

oxygen isotope stage 11 in the high-northern Atlantic (Nordic seas). Global and Planetary Change 24,<br />

27-39.


Berstad, I.M., Lundberg, J., Lauritzen, S.E., Linge, H.C., 2002. Comparison of the climate during<br />

Marine Isotope Stage 9 and 11 inferred from a speleothem isotope record from Northern Norway.<br />

Quaternary Research 58, 361-371.<br />

Channell, J.E.T., Hodell, D.A., Lehman, B., 1997. Relative geomagnetic paleointensity and δ 18 O at site<br />

ODP 983 (Gardar Drift, North Atlantic) since 350 ka. Earth and Planetary Science Letters 153, 103-<br />

118.<br />

Chmura, G.L., Eisma, D., 1995. A palynological study of surface and suspended sediments on a tidal<br />

flat: implications for pollen transport and deposition in coastal waters. Marine Geology 128, 183-200.<br />

Chmura, G.L., Smirnov, A., Campbell, I.D., 1999. Pollen transport through distributaries and<br />

depositional patterns in coastal waters. Palaeogeography, Palaeoclimatology, Palaeoecology 149, 257-<br />

270.<br />

de Vernal, A., Henry, M., Bilodeau, G., 1996. Techniques de préparation et d'analyse en<br />

micropaléontologie. Les cahiers du GEOTOP 3, 16-27.<br />

Desprat, S., 2005a. Réponses climatiques marines et continentales du Sud-Ouest de l'Europe lors des<br />

derniers interglaciaires et des entrées en glaciations. Thèse doctorale, Université Bordeaux I, Talence,<br />

272 pp.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., Duprat, J., Malaizé, B., Peypouquet, J.-P., submitted.<br />

Climatic variability of Marine Isotope Stage 7: direct land-sea-ice correlation from a multiproxy<br />

analysis of a northwestern Iberian margin deep-sea core. Quaternary Science Reviews.<br />

Desprat, S., Sánchez Goñi, M.F., Turon, J.-L., McManus, J.F., Loutre, M.F., Duprat, J., Malaizé, B.,<br />

Peyron, O., Peypouquet, J.-P., 2005b. Is vegetation responsible for glacial inception during periods of<br />

muted insolation changes? Quaternary Science Reviews 24, 1361-1374.<br />

Dias, J.M.A., Jouanneau, J.M., Gonzalez, R., Araujo, M.F., Drago, T., Garcia, C., Oliveira, A.,<br />

Rodrigues, A., Vitorino, J., Weber, O., 2002. Present day sedimentary processes on the northern<br />

Iberian shelf. Progress in Oceanography 52, 249-259.<br />

Droxler, A.W., Farrell, J.W., 2000. Marine Isotope Stage (MIS 11): new insights for a warm future.<br />

Global and Planetary Change 24, 1-5.<br />

Duplessy, J.-C., Shackleton, N.J., Matthews, R.K., Prell, W., Ruddiman, W.F., Caralp, M., Hendy,<br />

C.H., 1984. 13 C record of benthic Foraminifera in the last interglacial ocean: implication for carbon<br />

cycle and the global deep water circulation. Quaternary Research 21, 225-243.<br />

Dupont, L., Wyputta, U., 2003. Reconstructing pathways of aeolian pollen transport to the marine<br />

sediments along the coastline of SW Africa. Quaternary Science Reviews 22, 157-174.


EPICA community members, 2004. Eight glacial cycles from an Antarctic ice core. Nature 429, 623-<br />

628.<br />

Graham, D.W., Corliss, B.H., Bender, M.L., Keigwin, D., 1981. Carbon and Oxygen disequilibria of<br />

recent deep-sea benthic foraminifera. Marine micropaleontology 6, 483-497.<br />

Heusser, L., 1978. Spores and pollen in the marine realm. In: Haq, B.U., Boersma, A. (Eds.),<br />

Introduction to marine micropaleontology. Elsevier, New York, pp. 327-339.<br />

Heusser, L.E., Balsam, W.L., 1977. Pollen distribution in the N.E. Pacific Ocean. Quaternary Research<br />

7, 45-62.<br />

Hodell, D.A., Charles, C.D., Ninnemann, U.S., 2000. Comparison of interglacial stages in the South<br />

Atlantic sector of the southern ocean for the past 450 kyr: implications for Marine Isotope Stage (MIS)<br />

11. Global and Planetary Change 24, 7-26.<br />

Howard, W.R., 1997. A warm future in the past. Nature 388, 418-419.<br />

Huntley, B., 1996. Quaternary paleoecology and ecology. Quaternary Science Reviews 15, 591-606.<br />

Huntley, B., Webb III, T., 1989. Migration: species' response to climatic variations caused by changes<br />

in the Earth's orbit. Journal of Biogeography 16, 5-19.<br />

Jansen, E., Bleil, U., Henrich, R., Kringstad, L., Slettemark, B., 1988. Paleoenvironmental changes in<br />

the Norwegian sea and the northeast Atlantic during the last 2.8 m.y.: deep sea drilling project/ocean<br />

drilling program sites 610,642, 643 and 644. Paleoceanography 3 (5), 563-581.<br />

Jouanneau, J.M., Weber, O., Drago, T., Rodrigues, A., Oliveira, A., Dias, J.M.A., Garcia, C., Schimdt,<br />

S., Reyss, J.L., 2002. Recent sedimentation and sedimentary budget on the western Iberian shelf.<br />

Progress in Oceanography 52, 261-275.<br />

Kunz-Pirrung, M., Gersonde, R., Hodell, D.A., 2002. Mid-Brunhes century-scale diatom sea surface<br />

temperature and sea ice records from the Atlantic sector of the Southern Ocean (ODP Leg 177, sites<br />

1093, 1094 and core PS2089-2). Palaeogeography, Palaeoclimatology, Palaeoecology 182, 305-328.<br />

Lischke, H., Lotter, A.F., Fischlin, A., 2002. Untangling a Holocene pollen record with forest model<br />

simulations and independent climate data. Ecological Modelling 150, 1-21.<br />

Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C., Shackleton, N.J., 1987. Age dating<br />

and orbital theory of the Ice Ages: Development of a high-resolution 0 to 300,000-year<br />

chronostratigraphy. Quaternary Research 27, 1-29.<br />

McManus, J., Oppo, D., Cullen, J., Healey, S., 2003. Marine isotope stage 11 (MIS 11): analog for<br />

Holocene and future climate. In: Droxler, A.W., Poore, R.Z., Burckle, L.H. (Eds.), Earth's climate and


orbital eccentricity: the marine isotope stage 11 question. Geophysical Monograph. American<br />

Geophysical Union, Washington, DC, pp. 69-85.<br />

McManus, J.F., Oppo, D.W., Cullen, J.L., 1999. A 0.5-million-year record of millennial-scale climate<br />

variability in the North Atlantic. Science 283, 971-975.<br />

Muller, J., 1959. Palynology of recent Orinoco delta and shelf sediments. Micropaleontology 5, 1-32.<br />

Ozenda, P., 1982. Les végétaux dans la biosphère. Doin, Paris, 431 pp.<br />

Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, J.-M., Basile, I., Bender, M., Chappellaz, J.,<br />

Davis, M., Delaygue, G., Delmotte, M., Kotyakov, V.M., Legrand, M., Lipenkov, V.Y., Lorius, C.,<br />

Pépin, L., Ritz, C., Saltzman, E., Stievenard, M., 1999. Climate and atmospheric history of the past<br />

420,000 years from the Vostok ice core, Antarctica. Nature 399, 429-436.<br />

Reille, M., Andrieu, V., de Beaulieu, J.-L., Guenet, P., Goeury, C., 1998. A long pollen record from<br />

Lac du Bouchet, Massif Central, France: for the period ca. 325 to 100 ka BP (OIS 9c to OIS 5e).<br />

Quaternary Science Reviews 17, 1107-1123.<br />

Reille, M., de Beaulieu, J.-L., Svobodova, V., Andrieu-Ponel, V., Goeury, C., 2000. Pollen analytical<br />

biostratigraphy of the last five climatic cycles from a long continental sequence from the Velay region<br />

(Massif Central, France). Journal of Quaternary Science 15 (7), 665-685.<br />

Roucoux, K.H., Tzedakis, P.C., de Abreu, L., Shackleton, N.J. (this volume). Fine tuning the land–<br />

ocean correlation for the late Middle Pleistocene of southern Europe. The climate of the past<br />

interglacials.<br />

Sánchez Goñi, M.F., Loutre, M.F., Crucifix, M., Peyron, O., Santos, L., Duprat, J., Malaizé, B., Turon,<br />

J.-L., Peypouquet, J.-P., 2005. Increasing vegetation and climate gradient in western Europe over the<br />

Last Glacial Inception (122-110 ka): models-data comparison. Earth and Planetary Science Letters<br />

231, 111-130.<br />

Sánchez Goñi, M.F., Eynaud, F., Turon, J.-L., Shackleton, N.J., 1999. High resolution palynological<br />

record off the Iberian margin: direct land-sea correlation for the Last Interglacial complex. Earth and<br />

Planetary Science Letters 171, 123-137.<br />

Shackleton, N.J., Chapman, M., Sanchez-Goñi, M.F., Pailler, D., Lancelot, Y., 2002. The classic<br />

Marine Isotope Substage 5e. Quaternary Research 58, 14-16.<br />

Shackleton, N.J., Opdyke, N.D., 1973. Oxygen isotope and paleomagnetic stratigraphy of Equatorial<br />

Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 10 5 year and 10 6 year scale.<br />

Quaternary Research 3, 39-55.


Turon, J.-L., 1984. Le palynoplancton dans l'environnement actuel de l'Atlantique nord-oriental.<br />

Evolution climatique et hydrologique depuis le dernier maximum glaciaire. Mémoires de l'Institut de<br />

Géologie du bassin d'Aquitaine 17. Université de Bordeaux I, Bordeaux, 313 pp.<br />

Tzedakis, P.C., Bennett, K.D., 1995. Interglacial vegetation succession: a view from southern Europe.<br />

Quaternary Science Reviews 14, 967-982.<br />

Tzedakis, P.C., Andrieu, V., de Beaulieu, J.-L., Birks, H.J.B., Crowhurst, S., Follieri, M.,<br />

Hooghiemstra, H., Magri, D., Reille, M., Sadori, L., Shackleton, N.J., Wijmstra, T.A., 2001.<br />

Establishing a terrestrial chronological framework as a basis for biostratigraphical comparisons.<br />

Quaternary Science Reviews 20, 1583-1592.<br />

Tzedakis, P.C., Roucoux, K.H., de Abreu, L., Shackleton, N.J., 2004. The duration of forest stages in<br />

southern Europe and interglacial climate variability. Science 306, 2231-2235.<br />

van der Hammen, T., Wijmstra, T.A., Zagwijn, W.H., 1971. The floral record of the Late Cenozoic of<br />

Europe. In: Turekian, K.K. (Ed.), The Late Cenozoic glacial ages. Yale University press, New Haven,<br />

pp. 391-424.<br />

Vitorino, J., 2002. Winter dynamics on the northern Portuguese shelf. Part2: bottom boundary layers<br />

and sediment dispersal. Progress in Oceanography 52, 155-170.<br />

Figure 1: Location of the studied deep-sea cores MD01-2447, MD99-2331, MD03-2697.


Figure 2: Lightness record of the three Iberian margin deep-sea cores. Pollen analysis has been<br />

performed in the intervals represented by grey areas. Hatched area corresponds to the disturbed<br />

interval in core MD01-2447.


Figure 3: Direct correlation of continental and marine proxies from Iberian margin deep-sea cores.<br />

From the left to the right: 1) Synthetic pollen diagram; 2) Percentages of warm planktic foraminifera<br />

(only for MIS 7, 9 and 11) and Neogloboquadrina pachyderma left coiling; 3) Planktic δ 18 O curve; 4)<br />

Benthic δ 18 O curves. For the last 25,000 years, the benthic isotopic data of core MD99-2331 being not<br />

available, we present those of core MD01-2447. The correlation between both cores has been<br />

performed using different marine proxies (lightness, CaCO3 content, abundance and coiling ratio of<br />

Globorotalia hirsuta and G. truncatulinoides and percentages of N. pachyderma s.) (Desprat, 2005a);<br />

5) June insolation at 65°N. Blue areas indicate cold periods.


Marine Isotope<br />

Isotopic events<br />

Stages<br />

MIS 1 1.1 Holocene<br />

Iberian major warm<br />

periods<br />

MIS 5 5.1 St Germain II<br />

5.3 St Germain I<br />

5.5 Eemian<br />

MIS 7 7.1 Rianxo<br />

7.3 Ribeira<br />

7.5 Arousa<br />

MIS 9 9a Bueu<br />

9c Sanxenxo<br />

9e Pontedevedra<br />

MIS 11 11.1 Cangas<br />

11.23 Moana<br />

11.3 Vigo<br />

Table 1: Major warm periods detected in north western Iberia during the last 425,000 years versus<br />

marine isotopic stratigraphy. The time lags between the boundaries of isotopic substages and stages<br />

and those of forested phases are not indicated in this table.


Anexo B<br />

Paleoenvironmental evolution of estuarine systems during the<br />

last 14000 years – the case of Douro estuary (NW Portugal)<br />

T. Drago, M. C.Freitas, F.Rocha, J.Moreno, M.Cachão, F.Naughton, C.Fradique, F.Araújo,<br />

T.Silveira, A.Oliveira , J.Cascalho, F.Fatela<br />

In press in Journal of Coastal Research


Paleoenvironmental evolution of estuarine systems during the last 14000<br />

years – the case of Douro estuary (NW Portugal)<br />

T. Drago †, M. C.Freitas ‡, F.Rocha∞, J.Moreno‡, M.Cachão‡, F.Naughton†§, C.Fradique£, F.AraújoΦ,<br />

T.Silveira†, A.Oliveira Ψ, J.Cascalho£, F.Fatela‡<br />

† INIAP, ‡ Dep. e Centro ∞ Departamento § DGO UMR- £ Museu Nacional Φ Instituto Ψ Instituto<br />

IPIMAR, Geol. Univ. Lisboa, de Geociências, CNRS 5808, História Natural, Tecnológico e Hidrográfico,<br />

CRIPSUL Lisboa, 1749-016, Univ. Aveiro, Univ. Bordeaux I, Lisboa, 1600-250, Nuclear, Lisboa, 1100,<br />

Olhão, 8700- Portugal<br />

Aveiro, 3810- Talence, 33405, Portugal<br />

Sacavém Portugal<br />

305, Portugal cfreitas@fc.ul.pt 193, Portugal, France,<br />

crfradique@fc.ul.pt 2686-953 anabela.oliveira<br />

tdrago@ipimar mcachao@fc.ul.pt frocha@geo.ua.pt f.naughton@epoc jpcascalho@fc.ul.pt Portugal @hidrografico.pt<br />

.pt<br />

ffatela@fc.ul.pt<br />

.u-bordeaux1.fr<br />

faraujo@itn.pt<br />

ABSTRACT<br />

DRAGO, T.; FREITAS, M. C.; ROCHA, F. ; MOR<strong>EN</strong>O, J.; CACHÃO, M.; NAUGHTON, F.; FRADIQUE, C.;<br />

ARAÚJO, F.; SILVEIRA, T.; OLIVEIRA, A.; CASCALHO, J.; FATELA, F., 2004. Paleoenvironmental<br />

evolution of estuarine systems during the last 14000 years – The case of Douro estuary (NW Portugal)<br />

(Proceedings of the 8th International Coastal Symposium), pg – pg. Itajaí, SC – Brazil, ISSN 0749-0208<br />

The multidisciplinary study (sedimentology: texture, carbonates, organic matter, elemental composition,<br />

mineralogy of sand and clay; micropaleontology: foraminifera, calcareous nannoplankton and pollen – and<br />

radiocarbon dating) of three cores recovered from the Douro estuary infill has allowed to reconstruct the<br />

paleoenvironmental guidelines of this area for the last 14000 years. The major environmental changes recognized<br />

in the geological record of this area are principally associated with the Holocene transgression and the succession<br />

of different climatic types. The transgressive sequence is expressed through several environmental changes,<br />

related with a succession of different facies: a continental one till circa 9834 BP, followed by an alternated<br />

marine/continental facies with an intensification of marine signature till 5750BP; particularly after 6050BP, the<br />

environment is typically marine. After that and till present, it is followed by a regressive sequence (although<br />

within a still positive eustatic trend).<br />

ADITIONAL INDEX WORDS: environmental changes, estuarine systems, multi-proxies, sea-level rise<br />

INTRODUCTION<br />

Among depositional environments, estuarine systems are<br />

privileged locals to study the continental and marine influences<br />

that occurred in the recent past. The study of their sedimentary<br />

record can be a useful tool in the understanding of the<br />

environmental changes affecting the continental margin,<br />

particularly since the last deglaciation and following sea level rise.<br />

The present-day knowledge on the evolution pattern for the<br />

Portuguese continental margin since the last millennia is based on<br />

the work carried out by DIAS (1987) on the sedimentary dynamics<br />

and recent evolution of the shelf and paleoenvironmental<br />

reconstructions of southwestern coastal lagoons such as Santo<br />

André and Melides (BAO et al., 1999; CRUCES et al., 1999;<br />

FREITAS et al., 2002). The aims of this paper are thus, to present<br />

the data resulting from multi-proxy studies of the Douro estuary<br />

sedimentary record, in order to propose a general evolutionary<br />

model for the last 14000 years and to provide additional<br />

information on the paleoenvironmental evolution of the NW<br />

Portuguese coast.<br />

STUDY AREA<br />

The Douro estuary is located in the northwestern Portuguese<br />

coast, south of Oporto. It is elongated almost W-E, funnel-shaped<br />

with 2,25km length, 1,25m maximum width and a mean depth of<br />

about 5m. The estuary is partially sheltered from the ocean by a<br />

sand spit, rooted in the southern margin, known as Cabedelo<br />

(Fig.1). This is a common feature of all northern Portuguese<br />

estuaries (Ave, Cávado, Lima, Minho), with the barrier differing<br />

in length (in the case of Douro, it can attain 1km long) and<br />

resulting from local reversal of littoral drift. Its morphology is<br />

quite variable, dependant on the marine and fluvial hydrodynamic<br />

conditions. In winter, during extreme floods, this barrier can be<br />

overtopped, breached or even drowned for short periods that can<br />

last some days or weeks.<br />

Douro basin is located in the rainiest region of Portugal. It<br />

presents a climate varying from humid to semi-arid; the humid<br />

season occurs in October-March with 73% of total precipitation.<br />

The annual mean discharge is 710m 3 /s, with a maximum of<br />

3000m 3 /s and a minimum of 50 m 3 /s (LOUREIRO et al., 1986).<br />

METHODS<br />

This work is based on the study of three cores, which were<br />

obtained by rotary drilling in the Douro estuary, one in the barrier<br />

(core 2), known as Cabedelo and two in the back-barrier (cores 1<br />

and 1B, 50 cm apart), inside de S. Paio Bay (Fig. 1).<br />

Core 1 (+3.5/-4.4m height), 1B (-6.6/-16.2m height) and 2<br />

(-15.16/-39.66m height) can be considered as representative of the<br />

complete sedimentary sequence present in the southern margin of


Douro estuary. Cores 1B and 2 reached basement, consisting of<br />

weathered granite.<br />

N<br />

Porto<br />

Cabedelo<br />

Figure 1: Location of Douro estuary and coring sites.<br />

2<br />

1+1B<br />

After visual description the cores were sub-sampled for several<br />

analyses, which included sedimentology (texture, mineralogy of<br />

fine-63μm- fractions, organic matter,<br />

carbonate and water content), geochemistry and<br />

micropaleontology (foraminifera, calcareous nannoplankton and<br />

pollen associations). Nine organic sediment samples were dated<br />

by 14 C AMS at Beta Analytic Inc., USA (Fig.2).<br />

Textural analysis was performed by means of the traditional<br />

sieving methods and gravel-free sediments were classified<br />

according to FLEMMING (2000). The mineralogy of sand was<br />

determined by means of a stereomicroscope observation following<br />

DIAS (1987). Microscope counting of the heavy minerals species,<br />

separated using Na-polytungstate in the medium to very fine<br />

fractions, was made according to the line method (GALEHOUSE,<br />

1969). The study of gravel followed the methodology suggested<br />

by DOBBKINS and FOLK (1970). The organic matter (OM) and<br />

carbonates contents have been determined following CRAFT et al.<br />

(1991) and HULSEMAN (1966), respectively. Mineralogical<br />

analysis of both the fine (


The basal levels of SED 1 show a heavy mineral suite resulting<br />

from the direct erosion of the bedrock being mainly composed by<br />

biotite (20%), andalusite (27%) and apatite (6%). The majority of<br />

the mineral grains have euhedral forms, reflecting an incipient<br />

transport. The medium and upper levels of this unit are basically<br />

composed by biotite (40%), amphibole (12%) andalusite (6%),<br />

tourmaline and garnet (both with 3%.).<br />

In SED 2 unit the heavy mineral assemblage consists of biotite<br />

(45%), amphibole (10%), andalusite (4%), tourmaline and garnet<br />

(together with 3%). The levels with high (>50%) content of biotite<br />

match the muddy sand/sandy mud sequences in this unit.<br />

In SED 3, heavy minerals appear only in core 2 (less coarse),<br />

revealing the presence of biotite (49%), amphibole (7%) and<br />

rounded grains of garnet (3%), andalusite (4%) and tourmaline<br />

(2%).<br />

The SED 4 unit has a heavy mineral suite composed by biotite<br />

(32%), amphibole (18%) and andalusite (5%). It should be<br />

referred the occurrence of a distinct heavy mineral suite that<br />

characterizes the upper layers of SED4 in cores 1 and 2. This<br />

distinct suite was used by FRADIQUE et al. (this issue) to define a<br />

sub-unit - SED 4A. It is composed of rounded grains of garnet<br />

(15%), andalusite (13%), tourmaline (8%) and staurolite (5%).<br />

The vertical abundance of marine biogenic components (namely<br />

molluscs, foraminifera, equinoderms, etc.) and plant roots have<br />

opposite behaviour: the former are totally absent in SED1 and in<br />

the lower half of SED 2 (in core 2) while the latter are abundant,<br />

reaching maxima of 34.5% (Fig. 2); further upcore in SED 2, plant<br />

remains are virtually absent and marine biogenic components are<br />

abundant. The molluscs are the most represented biogenic marine<br />

group: they become more important when root content is less<br />

significant (in core 2). Higher values of molluscs usually associate<br />

with maxima in benthic foraminifera and “other biogenics” (e.g.<br />

between -28.11 and -27.45m and at -11.69m). A marked increase<br />

in biogenic component is verified from -8.69m till -8.20m (Fig.2).<br />

SED3 and SED4 are virtually barren in these components: only<br />

a discreet presence of molluscs at -1.69m is verified (in core 1).<br />

Mineralogical composition of fine and clay<br />

fractions<br />

Mineralogical characteristics of fine and clay fractions are quite<br />

different in the several sedimentary units.<br />

SED1 unit may be divided in three subzones:<br />

- from the base until -37.66m, the fine fraction mineralogy is<br />

characterized by an almost silicilastic association, constituted by<br />

quartz and feldspars (and therefore, denoting a low mineralogical<br />

maturity), and by a discrete presence of phyllosilicates and rare<br />

carbonates. The clay minerals association is rich in kaolinite and<br />

illite (but showing high K/I ratio), although with a relatively high<br />

expression of smectite or chlorite (Fig.2);<br />

- between -37.66m and -34.07m, quartz increases<br />

simultaneously with feldspars decrease (and a discrete increment<br />

in phyllosilicates) whereas clay fraction is characterized by a<br />

decrease of kaolinite (which leads to the predominance of illite)<br />

and by a much higher increase of chlorite in relation to smectite;<br />

- between -34.07m and -31.16m, a relative increase in quartz<br />

related to the decrease of phylosilicates and feldspars can be<br />

observed. In clay fraction, kaolinite drastically decrease, illite<br />

(very degraded) becomes much more important and smectite<br />

acquire higher importance than chlorite (rare and even absent in<br />

some samples).<br />

This last subzone continues through the lowest layers of SED2<br />

(core2). This unit is mainly characterized (from -31.16m to<br />

-20.66m) by the gradual increase of feldspars and the presence of<br />

Paleoenvironmental evolution Douro estuary (NW Portugal)<br />

micas (and more discreetly, carbonates) in the fine fraction, and by<br />

the decrease of illite (now more ordered) in relation to kaolinite,<br />

whereas smectite decrease comparatively to chlorite. After<br />

-22.74m, C/D ratio increases and K/I decreases. In core 1B, SED2<br />

is characterized by the presence of some detrital minerals such as<br />

quartz, feldspars and phyllossilicates, while the predominant<br />

mineral in the clay fraction is illite, followed by kaolinite<br />

discretely accompanied by chlorite. Smectite gradually increases<br />

in relation to the progressive decrease of chlorite.<br />

SED3 is characterized by a large increase in quartz (particularly<br />

evident for core 2) and a very discrete decrease of chlorite in clay<br />

fraction (core 2); kaolinite increases and smectite decreases,<br />

relatively to SED2 unit, more clearly at the bottom of this unit<br />

(particularly for Core 1) but these tendencies reverse topwards<br />

(core 1) (Fig.2).<br />

SED4 contains the lowest values (almost null) of C/D ratio and<br />

the highest values of FD/CD ratio. In the clay fractions, kaolinite<br />

decreases in relation to illite whereas smectite increases in relation<br />

to chlorite. Illite is once again predominant in a relatively complex<br />

assemblage, illite plus kaolinite plus chlorite plus vermiculite plus<br />

smectite; topwards (core 1), smectite decreases very significantly,<br />

at the same time that other clay minerals increase – such as<br />

vermiculite, mixed-layers and chlorite.<br />

Geochemical analysis<br />

The elemental analysis of core 2 shows that SED1 is<br />

characterized by higher values of Al, K, Si, Ti and Zr, comparable<br />

to the continental crust composition, which can be attributed to<br />

continental input (Fig.2). The systematically low Ca (~0.2-0.5%)<br />

and Sr contents are due to the widespread occurrence of weathered<br />

granitic rocks in the region. However, a significant and<br />

unexpected increase in Ca, Sr (non-common in sediments with a<br />

granitic origin) and Zr was observed at the base of SED1, which is<br />

probably related to some local enrichment in specific minerals (e.g<br />

apatite) in non-weathered basement rock. In particularly, Al, K,<br />

Rb contents should be related to the existence of feldspars and/or<br />

micas of granitic rocks. Besides, the gradually decrease of these<br />

elements till the middle of SED2, negatively correlated with Si<br />

(that increases upwards -32.56m) can be associated with the<br />

decrease in terrigenous input (in agreement with other proxies). It<br />

is worth noting the “anomalous” increasing concentrations in Ca<br />

and Sr found in the upper sections of core 2 (from -25.36m<br />

upwards), probably due to the presence of biogenic marine<br />

organisms.<br />

Micropaleontological analysis<br />

Foraminiferal assemblages mainly record estuarine/brackish,<br />

estuary mouth/marine and outer shelf/slope environments (Fig.2).<br />

The earliest foraminifera occurrence is noticed at -27.77/<br />

-27.66m, with abundant Haynesina germanica, followed by<br />

Elphidium gunteri, Cibicides lobatulus and Amonia spp., which<br />

suggest a brackish environment. After a barren interval of circa<br />

2.5m, foraminifera reappear in the fossil record with H.germanica<br />

(with lower abundance than the levels cited above) or E.gunteri,<br />

followed by A. tepida and some species typical of estuary mouth<br />

and of major depths, suggesting a brackish intertidal, low salinity<br />

estuarine environment. The facies then changes to a subtidal<br />

environment with normal to moderate (-22.39 and -21.48m), or<br />

moderate to low salinity (-21.26 and -20.53m), typical of an<br />

estuarine mouth under marine influence as it is suggested by the<br />

presence of H. germanica E. gunteri and A. tepida occurring<br />

conjointly with Cassidulina laevigata, C. lobalutus and<br />

Gavelinopsis praegeri.


Core 1<br />

SED4<br />

1110±40BP<br />

(-0.65m)<br />

1580±40BP<br />

(-1.80m)<br />

SED3<br />

Core 1B<br />

5750±40BP<br />

(-7.15m)<br />

6050±60BP<br />

(-8.34m)<br />

SED2<br />

9490±60BP<br />

(-13.91m)<br />

Core 2<br />

SED4<br />

SED3<br />

8930±60BP<br />

(-20.59m)<br />

9230±50BP<br />

(-22.75m)<br />

SED2<br />

10310±80BP<br />

(-32.33m)<br />

SED1<br />

13730±90BP<br />

(-39.07m)<br />

2.8<br />

1.8<br />

0.8<br />

-0.2<br />

-1.2<br />

-2.2<br />

-3.2<br />

-4.2<br />

-5.2<br />

-6.2<br />

-7.2<br />

-8.2<br />

-9.2<br />

-10.2<br />

-11.2<br />

-12.2<br />

-13.2<br />

-14.2<br />

-15.2<br />

-16.2<br />

-15,7<br />

-16,7<br />

-17,7<br />

-18,7<br />

-19,7<br />

-20,7<br />

-21,7<br />

-22,7<br />

-23,7<br />

-24,7<br />

-25,7<br />

-26,7<br />

-27,7<br />

-28,7<br />

-29,7<br />

-30,7<br />

-31,7<br />

-32,7<br />

-33,7<br />

-34,7<br />

-35,7<br />

-36,7<br />

-37,7<br />

-38,7<br />

-39,7<br />

Gravel<br />

Gravely sand<br />

Sand<br />

Slightly muddy sand<br />

Fine sediments<br />

(


Between -12.7 and -9.38m, a dominance of H. germanica,<br />

followed by E. gunteri, A.tepida, C.praegeri and C. lobatulus,<br />

indicates a brackish intertidal estuarine environment, of low, and<br />

punctually, moderate salinity; between -9.26 and -8.85m, C.<br />

lobalutus and G.praegeri become dominant, probably in relation<br />

with an intertidal-subtidal estuarine mouth environment with<br />

normal to moderate salinity and clear marine influence (Fig.2).<br />

It is only upwards -8.35m and till the end of SED2 that some<br />

middle continental shelf species as Cassidulina crassarossencis<br />

Globocassidulina subglobosa and Bolivina difformis become more<br />

important. On the other hand, the dominance of C. lobalutus,<br />

G..praegeri and B.pseudoplicata, together with the low abundance<br />

of H. germanica (5,000<br />

CAI) and diversity, is compatible with a clear marine environment<br />

in which Helicosphaera carteri, E. huxleyi, Gephyrocapsa<br />

ericsonii, G. oceanica and G. muellerae dominate together with<br />

Coccolithus pelagicus in lesser degree (Fig.2). H. carteri can<br />

exceed 1,000 CAI, equivalent to more than 80% of the<br />

assemblage.<br />

Pollen analysis<br />

Pollen analysis from core 1B, shows the presence of a Pinus-<br />

Quercus-forest with Alnus and some Poaceae-Ericacea open<br />

communities, in the Douro basin, between 9490 BP and 5750 BP.<br />

This forest reflects a temperate and humid climate in the first part<br />

of the Holocene. Pollen preserved in this unit was transported by<br />

the river (NAUGHTON, 2002; NAUGHTON et al., 2002 and<br />

NAUGHTON et al., submitted).<br />

In core 1 the pollen grains are only present in the muddy sand<br />

layers of SED4. These layers are dominated by pollen of<br />

Ericaceae, Poaceae and Asteraceae. This pollen assemblage<br />

represents the signature of the local vegetation characterising<br />

nowadays the surroundings of the coring point.<br />

DISCUSSION – EVOLUTION MODEL<br />

Overall results from this study allowed to distinguish several<br />

environmental changes related with sea-level rise (in particularly,<br />

with the Holocene transgression), regional climate changes<br />

occurred since the Lateglacial and some local forcing factors.<br />

Continental environment (A) – circa 13730-9834<br />

BP<br />

The basal sedimentary record indicates a continental<br />

environment with terrestrial facies that includes SED1 and the<br />

lower half of SED2 (Fig.2). This terrestrial signature is supported<br />

by the total absence of biogenic sand components (namely<br />

molluscs), the significant abundance of plant remains and the total<br />

absence of foraminifera and calcareous nannoplankton. The suite<br />

of heavy minerals, together with the high contents of Al, Ca, Sr<br />

and Zr and the clay mineral association corroborates this<br />

interpretation suggesting evidences of direct erosion of outcroping<br />

basement. This section seems to have been deposited in a<br />

changing climate: it evolved from temperate to subtropical<br />

conditions (from base to -37.66m) to milder ones (-37.66m/<br />

-34.07m) and finally to temperate conditions (but with colder<br />

periods) with an increase of precipitation (-34.07/-31.16m).<br />

SED1 sequence probably corresponds to a typical fluvial<br />

environment, in the dependence of the Douro river. It was<br />

deposited when the shoreline was far away from the coring sites:<br />

circa 13km (between 13-11ka; sea level at -30m) and 20km<br />

(between 11-10ka; sea level at -60m) according to DIAS (1987).<br />

The sedimentation becomes finer in the lower part of SED2 unit,<br />

probably in response to the approximation of the base level.<br />

Marine/Continental environment (B): 9834-<br />

6050BP<br />

The following facies is characterized by alternating marine/<br />

continental influences and corresponds to the upper half of SED 2<br />

(in core 2) and SED2 (in core 1B) (Fig.2). This facies is<br />

characterized by increasing marine influence, as suggested by the<br />

general increase of estuary mouth/marine and continental shelf<br />

foraminifera (especially expressed in core 1B). However, drastic<br />

decrease in abundance or even interruption in the foraminiferal<br />

record, molluscs and nannoplankton can be observed at specific<br />

levels (Fig.2). These perturbations, are simultaneous with<br />

augmentation in terrigenous indicators, especially plant remains.<br />

Thus, these intervals were attributed to continental inputs, which<br />

seem to have been short-lived (some of them may correspond to<br />

intervals as short as 130-200 years, considering the radiocarbon<br />

dates and a uniform sedimentation rate) and disturbed the<br />

transgressive trend. They can represent either regressive periods<br />

or, more probably, episodes of important fluvial sediment input.<br />

The earliest marine signal is represented by a single <strong>ephe</strong>meral<br />

episode, with brackish foraminifera assemblage and some peaks of<br />

molluscs and “other biogenics”, noticed between -28.11m and<br />

-27.45m and corresponding to an interpolated time period between<br />

circa 9834 and 9756 BP. However, it is after a threshold circa<br />

-25m that the signal of sea level incursions persists, as shown by<br />

the abundance and associations of molluscs and foraminifera.<br />

The marine signature is also supported by the increase in C/D<br />

ratio and decrease in K/I, particularly after -22.74m, when<br />

augmentations in Ca and Sr are also observed.<br />

Besides, the presence of nannoplankton at this specific level<br />

(corresponding to circa 9230 BP) indicate that sea level had<br />

already exceeded this elevation by this time. According to DIAS<br />

(1987), after 11-10ka, when sea level was circa -60m, a rapid rise<br />

took place, reaching approximately -20m at 8ka. However, this<br />

new data seem to indicate that this level was reached at an earlier<br />

stage (circa 9ka), which may imply a higher sea level rise rate than<br />

previously proposed.<br />

The pollen assemblages are in agreement with the climatic<br />

amelioration that characterizes the Holocene as it was deduced by<br />

pollen analysis. In particular, XRD data of core 2 put in evidence a


climate evolution towards more temperate conditions, with less<br />

precipitation and, therefore, less favorable to hydrolysis<br />

development in source areas. In core 1B, hydrolysis seems to<br />

increase progressively, towards a warmer and more humid<br />

climate, with seasonal contrast, simultaneously with less<br />

hydrodinamism.<br />

Marine environment (C): – 6050-5750BP<br />

A clear marine environment upwards -8.35m is very well<br />

expressed (and corresponds to the topmost layers of core 1B),<br />

through the foraminifera assemblage (effective presence of<br />

continental shelf species and minor importance of brackish<br />

environment species), nannoplankton diversity and abundance<br />

(Fig.2) and by the evident increase of C/D ratio. These data<br />

confirms that the coastline stood landward of the core location,<br />

when sea level was circa 7m below present day level, representing<br />

thus the maximum of the Holocene transgression. However, core<br />

1B site was probably related to a partially confined environment,<br />

probably by the precocious development of a sandy bar as it seems<br />

to be suggested by the observed abnormal opportunistic<br />

development of H.carteri which denote some ecological<br />

particularities (CACHÃO et al., 2002).<br />

This environment can be considered an extension of the<br />

preceding one, and some of its characteristics, as the climatic<br />

ones, remained unchanged from the B to the C environment.<br />

Barrier environment (D): post 5750 – before 1580<br />

BP<br />

After 5750 BP, a clear change in the environment is expressed<br />

by the deposition of a gravel layer - SED3 (Fig.2). This unit seems<br />

to have had a fluvial origin and subsequently a littoral reworking<br />

(NAUGHTON, 2002; NAUGHTON et al., submitted), which seems to<br />

be supported by the rounded, spherical and ellipsoidal grains of<br />

garnet, andalusite and tourmaline in this unit of Core 2. This<br />

gravel layer may represent a gravel barrier, which formed post<br />

5750 and before 1580 BP (NAUGHTON, 2002, NAUGHTON et al.,<br />

submitted) and thus, contemporaneous of similar features (though<br />

sandy) found in the SW Portuguese coastal lagoons of Albufeira,<br />

Melides and Santo André (FREITAS and ANDRADE, 2001).<br />

The extension of this gravel barrier allowed the northward<br />

relocation of the Douro channel (NAUGHTON, 2002, NAUGHTON et<br />

al., submitted), leaving its former position fossilized as a<br />

paleovalley located south of the present-day talweg (CARVALHO<br />

and ROSA, 1988). This could be responsible for the observed<br />

change in the pollen signature from regional to local before 5750<br />

and post 1580BP, respectively.<br />

Climatic proxies seem to indicate an increase in temperature<br />

and precipitation, favouring a strong hydrolysis in source areas<br />

and an intense fluvial transport as it was deduced from<br />

mineralogical data.<br />

Barrier/back-barrier environment (E): 1580BP –<br />

present day<br />

A barrier/back-barrier environment is evident in SED 4, with a<br />

brackish signal suggested by micropaleontological data and a<br />

small percentage (


(estuarine?) life of Helicosphaera carteri (preliminary results).<br />

Journal Nannoplankton Research, 24(2), 76-78.<br />

CACHÃO, M. and MOITA, T., 2000. Coccolithus pelagicus, a<br />

productivity proxy related to moderate fronts off Western<br />

Iberia. Mar. Micropaleontology, 39, 131-155.<br />

CARVALHO, A.F. and ROSA, M.M.P., 1988. Localização do<br />

Paleovale do Rio Douro. Anais do Inst. Hidrográfico, 9, 77-<br />

82.<br />

CRAFT, C. B; S<strong>EN</strong>ECA, E. D. and BROOME, S. W., 1991. Loss<br />

on ignition and Kjeldahl Digestion for estimating organic<br />

carbon and total nitrogen in Estuarine marsh soils: Calibration<br />

with dry combustion. Estuaries, 14, nº 2, pp. 175-179<br />

CRUCES, A.; FREITAS, C. and ANDRADE, C., 1999. Factores<br />

forçadores globais versus locais na evolução tardiglaciar e<br />

holocénica de sistemas lagunares do litoral SW portugês.<br />

Encontros Geomorf. , Univ. Coimbra, 105-113.<br />

De VERNAL, A.; H<strong>EN</strong>RY, M. and BILODEAU, G., 1996.<br />

Techniques de préparation et d’analyse en micropaléontologie.<br />

Les Cahiers du GEOTOP, nº3<br />

DIAS, J.M.A., 1987. Dinâmica sedimentar e evolução recente da<br />

plataforma continental portuguesa setentrional. PhDthesis,<br />

Univ. Lisboa (unpublished), 500 p.<br />

DOBBKINS, J.E, and FOLK, R.L., 1970. Shape Development on<br />

TahitiNui. Journal of Sedim. Petrology, vol.40, 4, 1167-1203.<br />

ELLIS, B.F. and MESSINA, A.R., 1995. Catalogue of<br />

foraminifera. American Museum of Natural History,<br />

Micro.Press, New York, 4 vol..<br />

FLEMMING, B.W., 2000. A revised textural classification of<br />

gravel-free muddy sediments on the basis of ternary diagrams.<br />

Cont. Shelf Res., 20, 1125-1137.<br />

FRADIQUE, C; CASCALHO, J;, DRAGO, T.; ROCHA, F. and<br />

SILVEIRA, T., (this issue)- The meaning of heavy minerals in<br />

the recent sedimentary record of the Douro estuary (Portugal);<br />

Int. Coastal Symposium.<br />

FREITAS, M.C., and ANDRADE, C., 2001. Marcadores<br />

ambientais na evolução dos espaços lagunares do SW<br />

português durante o Holocénico: a formação das barreiras<br />

arenosas e a influência antrópica “Proceedings of V Reunião<br />

Quat. Ibérico”, 113-116.<br />

FREITAS, M.C.; ANDRADE, C. and CRUCES, A., 2002. The<br />

geological record of environmental changes in southwestern<br />

Portuguese coastal lagoons since the Lateglacial. Quat. Int.<br />

93-94: 161-170.<br />

FREITAS, M.C; ANDRADE, C. and DRAGO, T., 2003.<br />

Holocene sea-level rise in western Portugal based on the study<br />

of coastal sedimentary sequences. 4 th Symposium on the<br />

Iberian Atalntic Margin, nº19(2a) pp 104-105.<br />

GALEHOUSE, J.S., 1969. Counting grain mounts: number<br />

percentage versus number frequency. Journal of Sed.<br />

Petrology, 39, pp.812-815<br />

HULSEMAN, J., 1966. On the routine analysis of carbonates in<br />

unconsolidated carbonates. Journal of Sed. Petrology, 36,<br />

pp.622-625.<br />

LOEBLICH, A.R. and TAPPAN, H., 1988. Foraminiferal Genera<br />

and their classification. Van Nostrand Reinold Company, 1,<br />

970p<br />

LOUREIRO, J.J; MACHADO, M.L; MACEDO, M.E; NUNES,<br />

M.N; BOTELHO, O.F; SOUSA, M.L; ALMEIDA, M.C and<br />

MARTINS, J.C., 1986. Monografias hidrológicas dos<br />

principais cursos de água de Portugal continental. Direcção<br />

Geral dos Serviços Hidráulicos, Lisboa, 569p.<br />

MELLINGER, R. M., 1979. Quantitative X-ray diffraction<br />

analysis of clay minerals. An evaluation. Saskatchewan<br />

Research Council, Canada, SRC Report G-79: 1-46<br />

Paleoenvironmental evolution Douro estuary (NW Portugal)<br />

NAUGHTON, F., 2002. Evolução Morfodinâmica do Estuário do<br />

Douro com base no registo polínico e sedimentológico. Master<br />

Thesis, Lisbon University, 131 p. (not published).<br />

NAUGTHON F.; ROCHA F.; DRAGO T. and SANCHEZ GOÑI<br />

M.F., 2002. Holocene Environmental Changes in Douro<br />

Estuary (NW Portugal) in Proceedings of “Asamblea<br />

Hispano-Portuguesa de Geodesia y Geofisica; Tomo II, S11,<br />

pp.1027-1030.<br />

NAUGHTON, F; SANCHEZ GONI, M. F; DRAGO, T.,<br />

FREITAS, M. C. and OLIVEIRA, A. (submited). Holocene<br />

geomorphological changes in the Douro estuary (NW Iberia).<br />

Journal of Quaternary Science<br />

PEVEAR, D.R. and MUMPTON, D.R., 1989. Quantitative<br />

mineral analysis of clays. CMS Workshop Lectures, 1. The<br />

Clay minerals Society, Colorado (USA).<br />

REILLE, M., 1992. Pollen et Spores d’Europe et d’Afric du Nort.<br />

Marseille: Laboratoire de Botanique historique et<br />

Palynologie, 520pp.<br />

ROCHA, F., 1993. Argilas Aplicadas a Estudos<br />

Litoestratigráficos e Paleoambientais na Bacia Sedimentar de<br />

Aveiro. PhD Thesis, Univ. Aveiro, 399pp. (unpublished<br />

SCHULTZ, L.G., 1964. Quantitative interpretation of<br />

mineralogical composition from X-ray and chemical data for<br />

the Pierre Shale. United States Geological Survey Professional<br />

Paper 391-C,1-31.<br />

THOREZ, J., 1976. Practical identification of clay minerals. G.<br />

Lelotte. Bruxelles. 99p.<br />

VIDINHA, J.M.; ROCHA, F.; ANDRADE, C. and GOMES, C.,<br />

1998. Mineralogical characterization of the fine fraction of the<br />

beach and dune sediments situated between Espinho and<br />

Torreira (Portugal). Geostatistical approach. Cuaternario y<br />

Geomorfología, 12 (3-4), 49-56.<br />

ACKNOWLEDGEM<strong>EN</strong>TS<br />

This work was undertaken as a part of the project <strong>EN</strong>VI-<br />

CHANGES-PLE/12/00 funded by Fundação para a Ciência e a<br />

Tecnologia (Portugal). The authors thank to C. Andrade for his<br />

field work assistance and enlightening discussions and R. Taborda<br />

and M. Fernanda Sanchez Gõni for the carefully revision of this<br />

paper.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!