15.08.2013 Views

Amandine Merlin - Ecobio - Université de Rennes 1

Amandine Merlin - Ecobio - Université de Rennes 1

Amandine Merlin - Ecobio - Université de Rennes 1

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 : 4343 ANNÉE 2011<br />

THÈSE / UNIVERSITÉ DE RENNES 1<br />

sous le sceau <strong>de</strong> l’<strong>Université</strong> Européenne <strong>de</strong> Bretagne<br />

pour le gra<strong>de</strong> <strong>de</strong><br />

DOCTEUR DE L’UNIVERSITÉ DE RENNES 1<br />

Mention : Biologie<br />

Ecole doctorale Vie- Agro- Santé<br />

présentée par<br />

<strong>Amandine</strong> <strong>Merlin</strong><br />

préparée à l’unité <strong>de</strong> recherche 6553 <strong>Ecobio</strong><br />

Ecosystèmes, Biodiversité, Evolution<br />

UFR Sciences <strong>de</strong> la Vie et <strong>de</strong> l’Environnement<br />

Importance <strong>de</strong>s<br />

interactions biotiques<br />

et <strong>de</strong>s contraintes<br />

environnementales<br />

dans la structuration<br />

<strong>de</strong>s communautés<br />

végétales : le cas <strong>de</strong>s<br />

marais atlantiques et<br />

<strong>de</strong>s pelouses<br />

méditerranéennes.<br />

Soutenance prévue à <strong>Rennes</strong><br />

le 30 septembre<br />

<strong>de</strong>vant le jury composé <strong>de</strong> :<br />

Marie-Laure Navas<br />

PR Montpellier Sup Agro / rapporteur<br />

Eric Tabacchi<br />

CR <strong>Université</strong> Paul Sabatier / rapporteur<br />

Emmanuel Corcket<br />

MC <strong>Université</strong> Bor<strong>de</strong>aux 1 / examinateur<br />

Nicolas Gross<br />

CR Centre d’Etu<strong>de</strong>s Biologiques <strong>de</strong> Chizé/<br />

examinateur<br />

Jan-Bernard Bouzillé<br />

Pr <strong>Université</strong> <strong>Rennes</strong> 1 / directeur <strong>de</strong> thèse<br />

François Mesléard<br />

DR centre <strong>de</strong> recherches Tour du Valat, Pr associé<br />

IUT 2 d’Avignon / co-directeur <strong>de</strong> thèse


- Résumé -<br />

Le stress est connu pour être structurant <strong>de</strong>s communautés végétales. Cependant, son<br />

effet sur les mécanismes et facteurs régissant les patrons <strong>de</strong> distribution <strong>de</strong>s espèces et la<br />

structure <strong>de</strong>s communautés est encore mal compris.<br />

L’objectif <strong>de</strong> cette thèse est <strong>de</strong> comprendre les mécanismes structurant les patrons<br />

d’abondance <strong>de</strong>s espèces et les communautés végétales le long <strong>de</strong> gradients<br />

environnementaux, en quantifiant le stress à (i) l’échelle <strong>de</strong> l’individu et à (ii) l’échelle <strong>de</strong> la<br />

communauté. Deux modèles biologiques ont été considérés autour <strong>de</strong>squels s’articulent <strong>de</strong>ux<br />

problématiques <strong>de</strong> recherche. La première problématique vise à comprendre l’importance<br />

respective <strong>de</strong>s inondations et <strong>de</strong> la compétition dans les patrons d’abondance d’espèces dans<br />

les prairies humi<strong>de</strong>s du Marais poitevin, en considérant le concept <strong>de</strong> niche <strong>de</strong>s espèces. La<br />

secon<strong>de</strong> problématique est consacrée à l’étu<strong>de</strong> <strong>de</strong> l’effet <strong>de</strong>s conditions <strong>de</strong> salinité du sol sur<br />

les réponses <strong>de</strong>s pelouses xéro-halophiles <strong>de</strong> Camargue au pâturage et à la variabilité <strong>de</strong> la<br />

pluviosité. Des approches expérimentales, démographiques et fonctionnelles ont été utilisées<br />

pour traiter ces <strong>de</strong>ux problématiques <strong>de</strong> recherche.<br />

L’approche expérimentale a montré que la compétition conduit à une ségrégation<br />

spatiale <strong>de</strong>s niches <strong>de</strong>s espèces le long du gradient d’inondation, en fonction <strong>de</strong> leur <strong>de</strong>gré <strong>de</strong><br />

tolérance à l’inondation et leur capacité à répondre à la compétition. A l’échelle <strong>de</strong> l’espèce,<br />

l’importance <strong>de</strong> la compétition est, en géneral, prédite par l’intensité <strong>de</strong> la contrainte abiotique<br />

liée au régime hydrique. L’approche démographique a confirmé le rôle structurant <strong>de</strong>s<br />

inondations, en particulier par son impact sur les capacités <strong>de</strong> survie et <strong>de</strong> colonisation <strong>de</strong>s<br />

espèces. Cette approche a permis <strong>de</strong> quantifier la compétition in natura et a ainsi montré que<br />

l’importance <strong>de</strong> la compétition augmente lorsque le <strong>de</strong>gré d’inondation diminue. L’analyse<br />

<strong>de</strong>s patrons <strong>de</strong> distribution <strong>de</strong> traits aériens à l’échelle <strong>de</strong> la communauté (approche<br />

fonctionnelle) a permis <strong>de</strong> montrer que l’assemblage <strong>de</strong>s espèces est la résultante <strong>de</strong> processus<br />

déterministes. Les rôles structurants du patron d’inondation et <strong>de</strong> la compétition dans les<br />

patrons d’abondance <strong>de</strong>s espèces et la structure <strong>de</strong>s communautés le long du gradient<br />

d’inondation sont ainsi démontrés par l’ensemble <strong>de</strong>s approches conduites.<br />

Pour les communautés végétales méditerranéennes, <strong>de</strong>s relevés <strong>de</strong> terrain ont montré<br />

que sous <strong>de</strong>s conditions <strong>de</strong> salinité faible le pâturage était un facteur important contrôlant la<br />

structure <strong>de</strong>s communautés méditerranéennes tandis qu’une forte salinité du sol limitait les<br />

effets <strong>de</strong> la pluviosité et du pâturage sur la structure <strong>de</strong> la communauté. Ce travail a montré<br />

que les espèces dominantes jouent un rôle important dans la réponse <strong>de</strong> ces communautés au<br />

pâturage et à la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité.<br />

D’une manière générale, le niveau <strong>de</strong> contrainte a un effet sur les patrons d’abondance<br />

<strong>de</strong>s espèces et sur la structure <strong>de</strong>s communautés et influence l’importance d’autres filtres<br />

écologiques.<br />

Mots-clés : Importance <strong>de</strong> la compétition, Intensité <strong>de</strong> la compétition, Intensité du stress,<br />

Niches écologiques, Inondations, Règles d’assemblage, Traits fonctionnels, Démographie,<br />

Pâturage


- Abstract –<br />

Stress is known to structure plant communities. However, its effect on the mechanisms<br />

and factors governing patterns of species distribution and community structure is poorly<br />

un<strong>de</strong>rstood.<br />

The objective of this thesis is to un<strong>de</strong>rstand the mechanisms structuring patterns of<br />

species abundance and plant communities along environmental gradients by quantifying the<br />

stress at (i) the individual level and (ii) the community level. Two biological mo<strong>de</strong>ls have<br />

been consi<strong>de</strong>red for which two research axes have been <strong>de</strong>veloped. The first axis aims at<br />

un<strong>de</strong>rstanding the relative importance of flooding and competition in the patterns of species<br />

abundance in wet grasslands of the Marais Poitevin, consi<strong>de</strong>ring the concept of niche of<br />

species. The second axis is <strong>de</strong>voted to the study of the effect of soil salinity conditions on the<br />

responses of xero-halophytic grasslands of the Camargue to grazing and rainfall variability.<br />

Experimental, <strong>de</strong>mographic and functional approaches were used to address these two axes.<br />

The experimental approach has shown that competition leads to a spatial segregation<br />

of species niches along the gradient of flooding, according both to their tolerance to flooding<br />

and to their ability to respond to competition. At the species level, the importance of<br />

competition is generally predicted by the intensity of abiotic stress related to water regime.<br />

The <strong>de</strong>mographic approach has confirmed the structuring role of flooding, especially through<br />

its impact on the species survival and colonization abilities. This approach allowed to quantify<br />

the competition in natura and has shown that the importance of competition increases as the<br />

<strong>de</strong>gree of flooding is reduced. The analysis of distribution patterns of functional traits at the<br />

community level (functional approach) has shown that the species assembly is the outcome of<br />

<strong>de</strong>terministic processes. The structuring role of flooding and competition in the patterns of<br />

species abundance and community structure throughout the flooding gradient is <strong>de</strong>monstrated<br />

by all the approaches initiated.<br />

For Mediterranean plant communities, field surveys have shown that un<strong>de</strong>r conditions<br />

of low salinity, grazing was an important factor controlling Mediterranean community<br />

structure whereas high soil salinity limited the effects of rainfall and grazing on community<br />

structure. This work has shown that the dominant species play an important role in the<br />

response of these communities to grazing and to the variability of rainfall patterns.<br />

In general, the stress level has an effect on patterns of species abundance and<br />

community structure and influences the importance of other ecological filters.<br />

Key-words: Competition Importance, Competition Intensity, Stress Intensity, Ecological<br />

Niches, Flooding, Assembly Rules, Functional Traits, Demography, Grazing


- Remerciements -<br />

Je tiens à commencer ces remerciements en exprimant ma plus profon<strong>de</strong> reconnaissance et ma<br />

gratitu<strong>de</strong> à l’ensemble <strong>de</strong>s coordinateurs <strong>de</strong> ce projet, Anne Bonis, Jan-Bernard Bouzillé et<br />

François Mesléard pour m’avoir accordée leur confiance et m’avoir guidée au cours <strong>de</strong> cette<br />

thèse.<br />

Merci à Jan-Bernard, qui a toujours réussi à trouver les bons mots pendant mes pério<strong>de</strong>s <strong>de</strong><br />

doute.<br />

Je tiens à remercier les membres du jury qui ont accepté <strong>de</strong> juger ce travail.<br />

Un grand merci aux membres (passés et présents) <strong>de</strong> l’équipe Forbio, que j’ai eue la joie <strong>de</strong><br />

côtoyer, et qui m’ont offert leur ai<strong>de</strong> à plusieurs reprises: Marie-Lise, Jean-Clau<strong>de</strong>, Ahmed,<br />

Benoît, pour ne citer qu’eux.<br />

Merci à Olivier et Guillaume pour leur ai<strong>de</strong> lors <strong>de</strong> la mise en place <strong>de</strong>s manips et la récolte<br />

<strong>de</strong>s données.<br />

Un merci tout particulier à Fabienne qui a toujours pris soin <strong>de</strong> moi.<br />

J’exprime toute ma reconnaissance et mes sincères remerciements à Nicole Yavercowski et<br />

Loïc Willm qui gèrent l’expérimentation et récoltent les données, ainsi que les membres <strong>de</strong> la<br />

Tour du Valat qui, à <strong>de</strong> nombreuses reprises, ont fournis une ai<strong>de</strong> terrain.<br />

Je remercie Le National Environmental Research Institute (NERI), Silkeborg, Danemark, et<br />

en particulier Christian Damgaard pour m’avoir accueillie chaleureusement à plusieurs<br />

reprises durant ma thèse.<br />

Merci au Collège Doctoral International <strong>de</strong> Bretagne pour m’avoir allouée une bourse <strong>de</strong><br />

mobilité.<br />

Merci aux personnes ayant participé à mes comités <strong>de</strong> thèse : Catherine Grimaldi, Arnaud<br />

Elger, Cyrille Violle, ainsi que mon tuteur <strong>de</strong> thèse, Jacques Baudry.<br />

Je remercie également à tous les stagiaires qui ont apporté leur ai<strong>de</strong> durant les pério<strong>de</strong>s <strong>de</strong><br />

terrain dans le Marais Poitevin ou en Camargue.<br />

Je remercie Marie-Lise, Véro, Cécile & Co., avec qui je ne partage pas seulement le goût <strong>de</strong> la<br />

recherche, mais aussi et surtout celui <strong>de</strong>s moments <strong>de</strong> détente autour d’un bon verre et surtout<br />

au son d’une bonne musique.<br />

Mention spéciale à Marie-Lise qui a toujours été présente pour moi : tout simplement, merci.<br />

Merci à Véro pour les corrections d’anglais !<br />

Merci à Helle, Kai et Line pour m’avoir accueillie chez eux comme un membre <strong>de</strong> leur propre<br />

famille.<br />

Je remercie ma famille et Alexis pour leur soutien sans faille, je leur dédie ce travail.


TABLE DES MATIERES<br />

INTRODUCTION GENERALE 1<br />

1. Les filtres écologiques 3<br />

1.1. Le concept <strong>de</strong> niche 4<br />

1.2. La variabilité <strong>de</strong>s conditions environnementales comme filtre écologique 6<br />

2. L’organisation <strong>de</strong> la végétation le long <strong>de</strong> gradients environnementaux 8<br />

2.1. Les différents types <strong>de</strong> gradient 8<br />

2.2. Les concepts <strong>de</strong> continuum et <strong>de</strong> communauté 9<br />

2.3. Le consensus actuel 10<br />

3. Les interactions compétitives le long <strong>de</strong> gradients environnementaux 11<br />

3.1. Les composantes <strong>de</strong> la compétition : intensité et importance 11<br />

3.2. Grime vs Tilman: un débat qui n’a donc plus lieu d’être 13<br />

3.3. La hiérarchie compétitive 14<br />

3.4. La nécessité <strong>de</strong> caractériser la contrainte 15<br />

3.5. Importance du niveau d’organisation 16<br />

3.6. Les méthodologies d’étu<strong>de</strong> <strong>de</strong> la compétition 16<br />

3.6.1. L’utilisation d’indices 16<br />

3.6.2. Les mesures <strong>de</strong> performance 17<br />

4. Les traits fonctionnels : outils d’étu<strong>de</strong> <strong>de</strong>s règles d’assemblage 18<br />

OBJECTIFS 23<br />

MODELES BIOLOGIQUES 29<br />

1. La microtopographie, génératrice d’une variété <strong>de</strong> communautés 29<br />

1.1. Les prairies humi<strong>de</strong>s du Marais Poitevin 29<br />

1.2. Les pelouses xéro-halophiles <strong>de</strong> Camargue 33<br />

2. La caractérisation <strong>de</strong> la variable environnementale d’intérêt 34<br />

2.1. Les patrons d’inondation dans les prairies humi<strong>de</strong>s du Marais Poitevin 34<br />

2.2. Les patrons <strong>de</strong> pluviosité en Camargue 39<br />

PARTIE 1: LA COMPETITION DANS LES MILIEUX INONDES 43<br />

Chapitre 1: Detecting changes in the intensity and importance of competition for grasslands<br />

species along a flooding gradient 45<br />

Chapitre 2: Fundamental niche versus realized niche: assessment of the importance of<br />

competition along a flooding gradient<br />

Part1: Fundamental niches of twelve wetland species in response to the variation of<br />

flooding conditions 71


Chapitre 3: Fundamental niche versus realized niche: assessment of the importance of<br />

competition along a flooding gradient<br />

Part 2: Tra<strong>de</strong>-off between competitive ability and tolerance of stress at the species level<br />

driving species distribution along the flooding gradient 93<br />

PARTIE 2: LA DEMOGRAPHIE COMME OUTIL D’ETUDE DES FILTRES ECOLOGIQUES 121<br />

Chapitre 4: Effect of environment on species competitive effect and on importance of<br />

competition along an elevation gradient 123<br />

Chapitre 5: The <strong>de</strong>mography of space occupancy: measuring plant colonization and survival<br />

probabilities using repeated pin-point measurements 145<br />

PARTIE 3: LA PART DES FILTRES ABIOTIQUES ET BIOTIQUES DANS LA STRUCTURE DES<br />

COMMUNAUTES VEGETALES 163<br />

Chapitre 6: Detecting community assembly patterns in floo<strong>de</strong>d and productive<br />

grasslands 165<br />

Chapitre 7: Stress level influences the effect of grazing regime and of variability of rainfall<br />

patterns on xero-halophytic communities 185<br />

DISCUSSION GENERALE 213<br />

1. La caractérisation <strong>de</strong>s contraintes 214<br />

1.1. Deux contraintes rencontrées le long du gradient d’inondation 215<br />

1.2. Les traits reliés à la tolérance aux inondations 216<br />

2. Variation <strong>de</strong> la compétition le long du gradient d’inondation 217<br />

2.1. Intensité <strong>de</strong> la compétition le long du gradient d’inondation 217<br />

2.1.1. Ségrégation spatiale <strong>de</strong>s niches le long du gradient d’inondation en réponse à<br />

la présence <strong>de</strong> compétiteurs 217<br />

2.1.2. Importance <strong>de</strong> la survie et <strong>de</strong> la colonisation pendant les inondations pour être<br />

compétitif 220<br />

2.2. Importance <strong>de</strong> la compétition le long du gradient d’inondation 221<br />

2.2.1. Importance <strong>de</strong> la compétition prédite par le « strain » 221<br />

2.2.2. Variation <strong>de</strong> l’importance <strong>de</strong> la compétition – apport <strong>de</strong> l’approche<br />

démographique 222<br />

2.3. Intensité et importance <strong>de</strong> la compétition le long du gradient d’inondation 223<br />

3. Les traits fonctionnels comme outils d’étu<strong>de</strong> <strong>de</strong>s règles d’assemblage 223<br />

4. Le stress à l’échelle <strong>de</strong> la communauté module les effets du pâturage dans les<br />

milieux méditerranéens 226<br />

5. Conclusion & Perspectives 228<br />

5.1. La mesure <strong>de</strong> la compétition en conditions naturelles 229<br />

5.2. Vers le développement <strong>de</strong>s tests <strong>de</strong>s règles d’assemblage 229


5.3. De la facilitation observée ? 230<br />

5.4. Une étu<strong>de</strong> plus précise <strong>de</strong> l’effet <strong>de</strong>s patrons <strong>de</strong> pluviosité dans les mileux<br />

méditerranéens 231<br />

BIBLIOGRAPHIE 233<br />

ANNEXE 1 : Liste <strong>de</strong>s abréviations <strong>de</strong>s espèces présentes dans les prairies humi<strong>de</strong>s du Marais<br />

Poitevin 245<br />

ANNEXE 2 : Liste <strong>de</strong>s abréviations <strong>de</strong>s espèces présentes dans pelouses xéro-halophiles <strong>de</strong><br />

Camargue 246<br />

ANNEXE 3 : Protocole <strong>de</strong> l’étu<strong>de</strong> <strong>de</strong> la compétition en conditions contrôlées 248<br />

ANNEXE 4 : Clonal growth strategies along flooding and grazing gradients in Atlantic coastal<br />

meadows 250


INTRODUCTION GÉNÉRALE<br />

A<br />

l’heure où les activités humaines induisent <strong>de</strong> nombreux changements<br />

globaux, la préservation <strong>de</strong> la biodiversité est au cœur <strong>de</strong>s préoccupations<br />

actuelles. Les activités humaines auraient déjà conduit à l’extinction entre 5 à<br />

20% <strong>de</strong>s espèces dans différents groupes d’organismes (Chapin et al. 2000).<br />

Les changements globaux sont <strong>de</strong> quatre types : les changements d’usage <strong>de</strong>s terres, les<br />

changements <strong>de</strong>s cycles biogéochimiques, les changements climatiques et l’introduction<br />

d’espèces. Ces changements affectent la distribution <strong>de</strong>s organismes altérant par conséquence<br />

les processus écosystémiques et les capacités <strong>de</strong> résilience <strong>de</strong>s écosystèmes aux changements<br />

environnementaux (Chapin et al. 2000).<br />

Comprendre les conséquences <strong>de</strong> la perte <strong>de</strong> diversité sur le fonctionnement <strong>de</strong>s<br />

écosystèmes et conserver la biodiversité nécessitent <strong>de</strong> caractériser les mécanismes régissant<br />

les patrons <strong>de</strong> diversité et donc <strong>de</strong> structure <strong>de</strong>s communautés végétales (Briske et al. 2005).<br />

La communauté végétale apparaît être un bon niveau d’organisation choisi pour étudier les<br />

patrons <strong>de</strong> distribution <strong>de</strong>s espèces car une communauté représente un ensemble d’espèces<br />

interagissant entre-elles mais aussi avec leur environnement. Etudier la structuration <strong>de</strong>s<br />

communautés le long <strong>de</strong> gradients environnementaux permet <strong>de</strong> caractériser les mécanismes<br />

régissant la structure <strong>de</strong>s communautés et in fine, <strong>de</strong> prédire les réponses <strong>de</strong>s espèces aux<br />

changements environnementaux en cours et/ou à venir.<br />

La compréhension <strong>de</strong>s patrons <strong>de</strong> diversité, d’abondance et <strong>de</strong> composition en espèces<br />

<strong>de</strong>s communautés et <strong>de</strong>s processus sous-jacents se situe au cœur <strong>de</strong> l’écologie <strong>de</strong>s<br />

communautés (Vellend 2010). De nombreuses théories écologiques ont été développées pour<br />

expliquer la coexistence <strong>de</strong>s espèces : les théories se distinguent par l’importance accordée<br />

aux quatre processus fondamentaux impliqués dans la coexistence <strong>de</strong>s espèces : les<br />

interactions interspécifiques, la dérive, la spéciation et la dispersion (Fig. 1).<br />

- Les interactions interspécifiques que les individus <strong>de</strong> différentes espèces peuvent<br />

expérimenter en se rencontrant dans un endroit et à un moment donnés. Le signe <strong>de</strong><br />

l’interaction (négatif, neutre ou positif) dépend alors <strong>de</strong> la valeur sélective <strong>de</strong>s individus en<br />

interaction.<br />

1


- La dérive écologique correspond aux changements <strong>de</strong>s abondances relatives <strong>de</strong>s<br />

espèces correspondant à la dynamique <strong>de</strong>s populations, <strong>de</strong>s fluctuations démographiques par<br />

une balance entre recrutement et mortalité inhérents à <strong>de</strong>s processus aléatoires.<br />

- La spéciation décrit la capacité <strong>de</strong>s espèces à s’adapter aux conditions abiotiques,<br />

pouvant amener à la création <strong>de</strong> nouvelles espèces.<br />

- La dispersion décrit le mouvement <strong>de</strong>s individus à travers l’espace entre<br />

communautés, et les échanges individus entre communautés.<br />

Fig. 1 : Lien entre les quatre processus fondamentaux : la sélection, la dérive, la spéciation et<br />

la dispersion et les patrons générés créés par d’innombrables manières à travers la boîte noire<br />

<strong>de</strong> l’écologie <strong>de</strong>s communautés. Extrait <strong>de</strong> Vellend (2010).<br />

Les règles d’assemblage ont été développées pour hiérarchiser l’importance <strong>de</strong> ces<br />

quatre processus fondamentaux. Ces règles d’assemblage peuvent être étudiées le long <strong>de</strong>s<br />

gradients environnementaux où différents facteurs ou filtres écologiques interviennent et<br />

pouvant affecter la performance <strong>de</strong>s espèces (Van Eck et al. 2004). En particulier, le long <strong>de</strong><br />

ces gradients peuvent varier <strong>de</strong> façon simultanée <strong>de</strong>s contraintes environnementales et les<br />

interactions compétitives. Comprendre l’importance <strong>de</strong>s contraintes abiotiques (facteurs<br />

<strong>de</strong> stress ou <strong>de</strong> perturbation) et celle <strong>de</strong>s interactions compétitives dans la distribution<br />

<strong>de</strong>s espèces, leur succès et in fine la structure <strong>de</strong>s communautés végétales le long <strong>de</strong><br />

gradients environnementaux sont l’objet <strong>de</strong> ce travail <strong>de</strong> recherche.<br />

2


1. Les filtres écologiques<br />

Les règles d’assemblage ont été développées pour comprendre l’assemblage <strong>de</strong>s<br />

espèces et <strong>de</strong>s traits au sein d’une communauté via la détermination et la quantification <strong>de</strong><br />

l’importance <strong>de</strong> filtres abiotiques et biotiques agissant à partir d’un pool d’espèces (Keddy<br />

1992 ; Diaz et al. 1998 ; Weiher & Keddy 1999 ; Lavorel & Garnier 2002). La communauté<br />

est composée d’espèces issues initialement d’un même pool régional, qui ont pu passer au<br />

travers <strong>de</strong> plusieurs types <strong>de</strong> filtres (Belyea & Lancaster 1999 ; Lortie et al. 2004) (Fig. 2).<br />

Tout d’abord le filtre biogéographique apparaît à une échelle large (continentale,<br />

biogéographique). Il dépend <strong>de</strong> processus stochastiques. Les espèces sont sélectionnées<br />

suivant leurs capacités à se disperser sur <strong>de</strong> longues distances.<br />

Le filtre abiotique regroupe l’ensemble <strong>de</strong>s facteurs physico-chimiques qui agissent<br />

<strong>de</strong> l’échelle régionale à locale. Les espèces sont sélectionnées suivant leur tolérance<br />

physiologique à ces facteurs. Ainsi seules les espèces capables <strong>de</strong> tolérer les conditions du<br />

milieu peuvent potentiellement être constitutives <strong>de</strong> la communauté.<br />

Le filtre biotique regroupe les interactions plante-plante dans la communauté<br />

végétale, et également les interactions plantes-organismes (prédation, parasitisme,<br />

mycorhizes…). Ce filtre agit par conséquent à une échelle très locale : ce filtre correspond<br />

aux interactions existantes au sein <strong>de</strong>s communautés végétales. Ces interactions peuvent<br />

modifier les conditions abiotiques d’un milieu, expliquant les interactions possibles entre les<br />

filtres biotique et abiotique.<br />

Ce modèle conceptuel hiérarchise ces différents filtres pour une meilleure<br />

compréhension <strong>de</strong> leurs effets sur les probabilités <strong>de</strong> présence <strong>de</strong>s espèces et sur les<br />

probabilités <strong>de</strong> <strong>de</strong>venir dominantes. Néanmoins, ces filtres agissent simultanément sur le pool<br />

d’espèces (Cingolani et al. 2007).<br />

3


Fig. 2 : Hiérarchisation <strong>de</strong>s filtres déterminant la composition et la structure <strong>de</strong>s communautés<br />

végétales (extrait <strong>de</strong> Lortie et al. 2004).<br />

De nombreuses théories ont été proposées pour expliquer l’assemblage <strong>de</strong>s espèces au<br />

sein <strong>de</strong>s communautés. L’objectif ici n’est pas d’en faire une liste exhaustive, mais plutôt <strong>de</strong><br />

se placer dans le cadre théorique le mieux adapté pour étudier l’importance <strong>de</strong> la compétition<br />

par rapport aux filtres abiotiques. En considérant la littérature, le concept <strong>de</strong> niche <strong>de</strong>s<br />

espèces apparaît être un cadre adéquat.<br />

1.1. Le concept <strong>de</strong> niche<br />

Le concept <strong>de</strong> niche est une théorie déterministe dans laquelle la structure <strong>de</strong>s<br />

communautés végétales reflète les exigences <strong>de</strong>s espèces au regard <strong>de</strong>s conditions<br />

environnementales et <strong>de</strong>s ressources, et également au regard <strong>de</strong>s limitations imposées par les<br />

interactions interspécifiques et intraspécifiques.<br />

La niche d’Hutchinson (Hutchinson 1957) est définie par un hyper-volume à n<br />

dimensions où chaque axe définit un type <strong>de</strong> conditions environnementales et/ou <strong>de</strong><br />

ressources. Au sein <strong>de</strong> cet hyper-volume, les populations d’une espèce peuvent maintenir un<br />

taux <strong>de</strong> croissance positif. Il a ainsi introduit la notion <strong>de</strong> niche fondamentale <strong>de</strong>s espèces,<br />

espace occupé par une espèce en l’absence <strong>de</strong> toute influence d’espèces; et la niche réalisée<br />

correspondant à l’espace occupé par l’espèce en présence d’autres espèces. La niche réalisée<br />

est moins large que la niche fondamentale car la compétition interspécifique réduit la fitness<br />

4


<strong>de</strong> l’espèce (taux <strong>de</strong> survie, fécondité, croissance…). Néanmoins, il ne faut pas oublier que les<br />

interactions interspécifiques peuvent être positives (facilitation) permettant aux espèces <strong>de</strong><br />

subsister dans <strong>de</strong>s conditions peu favorables, impliquant une niche réalisée plus large que la<br />

niche fondamentale.<br />

Le concept <strong>de</strong> niche développé d’Hutchinson décrit la niche comme étant un ensemble<br />

<strong>de</strong> conditions environnementales dans lesquelles un organisme peut subsister. Chase &<br />

Liebold (2003) propose <strong>de</strong> considérer en plus <strong>de</strong> la proposition d’Hutchinson, celle d’Elton<br />

(1927) considérant la niche d’un organisme comme étant son impact sur les autres organismes<br />

et l’environnement, <strong>de</strong>ux points <strong>de</strong> vue qui sont pour eux complémentaires.<br />

D’après le concept <strong>de</strong> niche, <strong>de</strong>ux espèces ne peuvent pas partager indéfiniment une<br />

même ressource limitante, elles ne peuvent donc pas occuper une même niche. Suivant le<br />

principe d’exclusion compétitive (Gause 1934), l’espèce la plus compétitive est amenée à<br />

exclure la moins compétitive. Ce concept est fortement dépendant d’un autre : le concept <strong>de</strong><br />

«limiting similarity» (MacArthur & Levins, 1967 ; Chesson 2000 ; Chase & Liebold 2003)<br />

disant que <strong>de</strong>ux espèces similairement proches, c’est-à-dire ayant <strong>de</strong>s stratégies d’acquisition<br />

et d’utilisation <strong>de</strong>s ressources similaires, entreront plus en compétition que <strong>de</strong>ux autres<br />

espèces présentant <strong>de</strong>s stratégies un peu plus contrastées.<br />

Tilman (1982; 1984) a développé le modèle du « Resource Ratio Hypothesis» pour<br />

expliquer pourquoi l’exclusion compétitive a lieu et dans quelles conditions les espèces<br />

coexistent. Dans ce modèle, les espèces sont caractérisées par <strong>de</strong>s exigences minimales pour<br />

<strong>de</strong>ux ressources, R1* et R2*, nécessaire à leur persistance dans un site. La différence entre ces<br />

exigences minimales mais également dans les taux <strong>de</strong> consommation <strong>de</strong> ces ressources sont<br />

les paramètres déterminant <strong>de</strong> leur coexistence. Une espèce tolérant le plus faible niveau <strong>de</strong><br />

ressource, et diminuant le niveau <strong>de</strong> cette ressource, c’est-à-dire présentant un R* faible, sera<br />

la plus compétitive. Pour ces <strong>de</strong>ux ressources, les espèces peuvent coexister quand une espèce<br />

est compétitive pour l’une <strong>de</strong> ces <strong>de</strong>ux ressources. Cette coexistence <strong>de</strong> <strong>de</strong>ux espèces est<br />

possible en raison d’un compromis entre les capacités compétitives <strong>de</strong>s espèces pour les<br />

ressources, car les espèces ne peuvent pas être compétitives pour toutes les ressources (Fig.<br />

3).<br />

5


Fig. 3 : Domaine <strong>de</strong> coexistence théorique <strong>de</strong> <strong>de</strong>ux espèces A et B qui interagissent pour <strong>de</strong>ux<br />

ressources. L’espèce A présente <strong>de</strong>s exigences minimales pour la ressource 2, l’espèce B<br />

présente <strong>de</strong>s exigences minimales pour la ressource 1. Des isoclines sont définies indiquant si<br />

une espèce exclut l’autre en raison d’une plus forte compétitivité en raison <strong>de</strong> besoins plus<br />

faibles pour la ressource, ou si elles coexistent grâce à un compromis entre les capacités<br />

compétitives <strong>de</strong>s espèces pour les <strong>de</strong>ux ressources.<br />

In natura, plus <strong>de</strong> <strong>de</strong>ux espèces peuvent coexister et interagir ensemble qui car ces<br />

espèces ont besoin <strong>de</strong> ressources similaires (Grubb 1977), il faut donc adapter le modèle <strong>de</strong><br />

Tilman en y intégrant un plus grand nombre d’espèces. Cependant même si le principe<br />

d’exclusion explique l’exclusion <strong>de</strong> certaines espèces, d’autres facteurs sont responsables <strong>de</strong><br />

la coexistence <strong>de</strong>s autres espèces, comme l’hétérogénéité spatiale et temporelle <strong>de</strong>s<br />

ressources.<br />

1.2. La variabilité <strong>de</strong>s conditions environnementales comme filtre écologique<br />

Les concepts précé<strong>de</strong>nts, que ce soit le concept <strong>de</strong> niche ou le modèle du « Resource<br />

Ration Hypothesis » ont été proposés suivant <strong>de</strong>s conditions environnementales stables.<br />

Néanmoins, les organismes sont soumis à une variabilité qui leur est intrinsèque et également<br />

à une variabilité extrinsèque produite par l’hétérogénéité spatiale et/ou temporelle <strong>de</strong>s<br />

ressources (Chase & Liebold 2003).<br />

La variabilité <strong>de</strong>s conditions environnementales peut influencer <strong>de</strong> manière importante<br />

les paramètres démographiques et les capacités compétitives <strong>de</strong>s espèces. Tenir compte <strong>de</strong>s<br />

paramètres qui témoignent <strong>de</strong> la dynamique <strong>de</strong>s populations au sein <strong>de</strong> leurs communautés,<br />

c’est-à-dire <strong>de</strong>s probabilités <strong>de</strong> recrutement et <strong>de</strong> survie <strong>de</strong>s espèces est par conséquent<br />

important (Harper 1977 ; Salguero-Gomez & <strong>de</strong> Kroon 2010). L’importance actuelle <strong>de</strong> ces<br />

6


étu<strong>de</strong>s <strong>de</strong> dynamique <strong>de</strong> populations a d’ailleurs fait l’objet d’un numéro spécial <strong>de</strong> Journal of<br />

Ecology (« Advances in plant <strong>de</strong>mography using matrix mo<strong>de</strong>ls », Mars 2010) permettant <strong>de</strong><br />

montrer les avancées dans l’étu<strong>de</strong> <strong>de</strong> la réponse <strong>de</strong>s populations <strong>de</strong> plantes où la variabilité <strong>de</strong><br />

l’environnement est maintenant mieux prise en compte.<br />

Les étu<strong>de</strong>s utilisant <strong>de</strong>s données démographiques pour comprendre l’importance <strong>de</strong>s<br />

filtres environnementaux sont nécessaires car elles tiennent compte du volet dynamique utile<br />

à toute compréhension du fonctionnement <strong>de</strong>s écosystèmes (Enright et al. 1995). De récents<br />

modèles permettent d’étudier l’effet <strong>de</strong> ces différents filtres environnementaux, notamment sur<br />

la variabilité <strong>de</strong>s conditions environnementales, sur différents sta<strong>de</strong>s du cycle <strong>de</strong> vie <strong>de</strong>s<br />

plantes. Ce type d’approche a été développé dans cette thèse proposant un modèle utilisant<br />

une métho<strong>de</strong> répétable au cours du temps et utilisant <strong>de</strong>s données récoltées in natura.<br />

L’importance <strong>de</strong>s variations du recrutement, <strong>de</strong> la mortalité et également <strong>de</strong> la<br />

dispersion sont les paramètres sur lesquels se fon<strong>de</strong> la théorie neutre (Hubbell 2001) où les<br />

processus déterministes comme par exemple les préférences d’habitats, la différenciation <strong>de</strong>s<br />

niches n’apparaissent pas pertinents pour expliquer la structure <strong>de</strong>s communautés. Longtemps<br />

perçues comme opposées, le concept <strong>de</strong> niche et la théorie neutre doivent être perçues comme<br />

complémentaires:<br />

“The next challenge ahead is to integrate niche and neutral theories<br />

that is to add more processes in neutral theories and more<br />

stochasticity in niche theories.”<br />

Chave (2004)<br />

Il s’agit donc <strong>de</strong> comprendre comment les processus déterministes et stochastiques<br />

peuvent interagir sur l’assemblage <strong>de</strong>s communautés (Heil 2004). Gewin (2006) synthétise les<br />

discussions actuelles sur le sujet via le recensement d’étu<strong>de</strong>s, notamment à partir du jeu <strong>de</strong><br />

données d’Hubbell (2001), montrant que les processus aléatoires ne peuvent seuls expliquer<br />

les diversités en espèces rencontrées. De récentes étu<strong>de</strong>s montrent que la considération <strong>de</strong><br />

processus stochastiques peut mieux expliquer la « limiting similarity » (Chesson 2000) et<br />

ainsi augmenter la différenciation <strong>de</strong>s niches <strong>de</strong>s espèces (Tilman 2004) ou <strong>de</strong> groupes<br />

d’espèces (Swilck & Ackerly 2005). Le succès <strong>de</strong>s processus déterministes par rapport aux<br />

processus stochastiques est fortement dépendant <strong>de</strong>s conditions environnementales (Lortie et<br />

7


al. 2004) et plus particulièrement <strong>de</strong> leur patron <strong>de</strong> distribution temporelle (Westoby et al.<br />

1989 ; Potts et al. 2005), ainsi que <strong>de</strong> l’échelle d’étu<strong>de</strong> considérée (Lortie et al. 2004). Heil<br />

(2004) et Cadotte (2007) proposent d’ailleurs <strong>de</strong> considérer différentes échelles spatiales où<br />

les mécanismes neutres et ceux fondés sur la théorie <strong>de</strong>s niches peuvent agir simultanément et<br />

ainsi expliquer la coexistence <strong>de</strong>s espèces.<br />

La part <strong>de</strong>s processus déterministes expliquant la structure <strong>de</strong>s communautés<br />

végétales peut être étudiée en testant ces processus déterministes face à un modèle neutre qui<br />

est une approche générant aléatoirement <strong>de</strong>s assemblages d’espèces (Gotelli & McCabe<br />

2002). Ce type d’approche a été utilisé dans cette thèse pour déterminer l’importance <strong>de</strong>s<br />

processus déterministes dans l’assemblage <strong>de</strong>s espèces le long d’un gradient d’inondation.<br />

2. L’organisation <strong>de</strong> la végétation le long <strong>de</strong> gradients environnementaux<br />

2.1. Les différents types <strong>de</strong> gradient<br />

Smith 1989) :<br />

Les gradients environnementaux peuvent être <strong>de</strong> trois types (Austin 1980 ; Austin &<br />

- les gradients <strong>de</strong> ressources, correspondant à la variation <strong>de</strong> ressources directement<br />

nécessaires à la croissance <strong>de</strong>s plantes.<br />

- les gradients directs, correspondant aux facteurs environnementaux qui vont impacter<br />

directement l’acquisition <strong>de</strong>s ressources et par conséquent la croissance <strong>de</strong>s plantes.<br />

- les gradients indirects, correspondant aux facteurs qui influencent <strong>de</strong> manière<br />

indirecte la croissance <strong>de</strong>s plantes car influencent eux-mêmes les facteurs directs.<br />

Pour la compréhension <strong>de</strong>s processus agissant sur l’assemblage <strong>de</strong>s espèces le long <strong>de</strong><br />

gradients environnementaux, il est nécessaire <strong>de</strong> se positionner par rapport à l’un <strong>de</strong> ces trois<br />

types <strong>de</strong> gradients. En effet, l’organisation <strong>de</strong> la végétation dépend du type <strong>de</strong> ressource, <strong>de</strong> la<br />

quantité <strong>de</strong> ressource disponible pour lesquelles les plantes interagissent, et également <strong>de</strong><br />

l’hétérogénéité <strong>de</strong> ces ressources (Chase & Liebold 2003). Les attendus en terme <strong>de</strong> résultante<br />

<strong>de</strong> la compétition peuvent par conséquent être différents suivant si on se trouve sous une<br />

contrainte abiotique <strong>de</strong> type ressource (disponibilité en nutriments) ou <strong>de</strong> type indirect.<br />

Néanmoins, il peut être parfois difficile <strong>de</strong> se positionner par rapport à ces propositions<br />

8


lorsque qu’un gradient <strong>de</strong> ressource coïnci<strong>de</strong> avec un gradient indirect (Kotowski et al. 2006).<br />

Les gradients d’inondations sont généralement décrits comme <strong>de</strong>s gradients indirects car ils<br />

correspon<strong>de</strong>nt en fait à un gradient altitudinal où la durée d’inondation varie. Néanmoins,<br />

certains auteurs remettent en cause ce classement car les inondations limitent la disponibilité<br />

en oxygène directement nécessaire les plantes (Blom & Voesenek 1996) et proposent <strong>de</strong><br />

considérer ce gradient comme ressource (Keddy et al. 1994).<br />

2.2. Les concepts <strong>de</strong> continuum et <strong>de</strong> communauté<br />

Deux concepts ont été développés pour expliquer l’organisation <strong>de</strong> la végétation le<br />

long <strong>de</strong> gradients environnementaux : le concept <strong>de</strong> la communauté développé par Clements<br />

(1936) et le concept du continuum développé par Whittaker (1951) et Curtis (1959). Le<br />

concept <strong>de</strong> la communauté propose que les communautés sont <strong>de</strong>s associations d’espèces<br />

fortement structurées et répétables avec <strong>de</strong>s limites finies. Une communauté peut être<br />

représentée par <strong>de</strong>s groupes d’espèces qui ne se chevauchent pas le long d’un gradient<br />

environnemental (Fig. 4A) (Collins et al. 1993 ; Austin 2005). Le concept du continuum<br />

propose quant à lui que les communautés végétales changent graduellement le long <strong>de</strong><br />

gradients environnementaux complexes au point où aucune association d’espèces ne peut être<br />

i<strong>de</strong>ntifiée, car chacune d’elles a une distribution le long du gradient qui lui est propre (Fig.<br />

4B) (Collins et al. 1993 ; Austin, 2005).<br />

Fig. 4 : Patrons <strong>de</strong> végétation hypothétiques le long d’un gradient environnemental<br />

représentant : A : le concept <strong>de</strong> communauté où les associations sont très structurées le long<br />

du gradient environnemental; B : le concept <strong>de</strong> continuum où chaque espèce a une distribution<br />

le long du gradient qui lui est propre (extrait <strong>de</strong> Austin 2005).<br />

9


2.3. Le consensus actuel<br />

La vision d’une organisation <strong>de</strong>s patrons <strong>de</strong> végétation sous forme <strong>de</strong> communautés et<br />

sous forme <strong>de</strong> continuum apparaissent maintenant moins incompatibles. En effet, les espèces<br />

composant les communautés végétales présentant <strong>de</strong>s réponses à l’environnement qui peuvent<br />

être différentes entre elles, argument en faveur du concept <strong>de</strong> continuum <strong>de</strong> l’organisation <strong>de</strong><br />

la végétation. Néanmoins, certaines espèces sont plus souvent observées en commun à<br />

certaines localisations du gradient permettant <strong>de</strong> construire <strong>de</strong>s typologies et donc <strong>de</strong> décrire<br />

<strong>de</strong>s communautés. Un patron intermédiaire à ces <strong>de</strong>ux points <strong>de</strong> vue est l’actuel consensus,<br />

avec d’espèces sténoèces à la base <strong>de</strong> chaque communauté et <strong>de</strong>s espèces euryèces qui leur<br />

sont communes (Austin 2005) (Fig. 5).<br />

Fig. 5 : Patron hypothétique d’organisation <strong>de</strong> la végétation le long d’un gradient<br />

environnemental caractérisé par <strong>de</strong>s espèces présentant <strong>de</strong>s gammes <strong>de</strong> distribution restreintes<br />

permettant <strong>de</strong> la distinction <strong>de</strong>s communautés, toutes trois ayant chacune <strong>de</strong>s espèces<br />

communes présentant une large distribution.<br />

Le choix du niveau d’organisation ainsi que les métho<strong>de</strong>s statistiques associées<br />

influencent la capacité à vali<strong>de</strong>r les hypothèses d’assemblage <strong>de</strong>s espèces le long <strong>de</strong> gradients<br />

environnementaux. D’une manière générale, le concept <strong>de</strong> communauté est démontré à l’ai<strong>de</strong><br />

d’analyses multivariées comme les métho<strong>de</strong>s d’ordination comme l’Analyse Canonique <strong>de</strong>s<br />

Correspondances (Ter Braak 1986) ou <strong>de</strong>s métho<strong>de</strong>s <strong>de</strong> classification comme TWINSPAN<br />

(Kent & Coker 1992) déterminant <strong>de</strong>s associations d’espèces. Les relevés floristiques groupés<br />

vont nous permettre <strong>de</strong> dégager les espèces caractérisant chaque communauté, les espèces<br />

dominantes et les espèces compagnes. L’organisation <strong>de</strong> la végétation sous forme <strong>de</strong><br />

continuum quant à lui est généralement mise en évi<strong>de</strong>nce à l’ai<strong>de</strong> <strong>de</strong> régressions <strong>de</strong> type<br />

modèles linéaires généralisés (GLM) ou modèles additifs généralisés (GAM), permettant <strong>de</strong><br />

10


déterminer les courbes <strong>de</strong> réponses <strong>de</strong>s espèces montrant ainsi <strong>de</strong>s changements continus le<br />

long <strong>de</strong> gradients environnementaux (Austin & Gaywood 1994). Ces métho<strong>de</strong>s constituent<br />

<strong>de</strong>s <strong>de</strong>scripteurs <strong>de</strong> la niche écologique <strong>de</strong>s espèces.<br />

3. Les interactions compétitives le long <strong>de</strong> gradients environnementaux<br />

La compétition a été largement étudiée via notamment <strong>de</strong>s approches expérimentales :<br />

Goldberg & Barton (1992) ont recensés <strong>de</strong> nombreuses expérimentations traitant<br />

majoritairement <strong>de</strong> l’effet <strong>de</strong> la compétition sur la fitness <strong>de</strong>s espèces cibles. Plus récemment,<br />

les étu<strong>de</strong>s se sont intéressées au rôle <strong>de</strong> la compétition sur les patrons d’abondance <strong>de</strong>s<br />

espèces et sur la structure <strong>de</strong>s communautés. Il reste néanmoins à l’heure actuelle <strong>de</strong>s<br />

imprécisions qui obscurcissent la compréhension du rôle <strong>de</strong> la compétition le long <strong>de</strong><br />

gradients environnementaux. De façon fondamentale, le concept d’importance <strong>de</strong> la<br />

compétition a été largement négligé dans la littérature (Craine 2007 ; Brooker & Kikvidze<br />

2008) et sa prise en compte permet <strong>de</strong> résoudre certains débats.<br />

3.1. Les composantes <strong>de</strong> la compétition : intensité et importance<br />

Distinguer le rôle <strong>de</strong> l’intensité <strong>de</strong> la compétition <strong>de</strong> celui <strong>de</strong> l’importance <strong>de</strong> la<br />

compétition est essentielle pour comprendre les patrons d’abondance <strong>de</strong>s espèces et <strong>de</strong><br />

structure <strong>de</strong>s communautés végétales le long <strong>de</strong> gradients environnementaux.<br />

L’intensité <strong>de</strong> la compétition correspond à la réduction <strong>de</strong> performances due à la<br />

présence <strong>de</strong> compétiteurs (Wel<strong>de</strong>n & Slauson 1986 ; Brooker et al. 2005 ; Brooker 2006). Des<br />

espèces fortement compétitives, que ce soit dans leur capacité à réduire la ressource<br />

disponible pour les voisins (effet compétitif) et/ou dans leur capacité à répondre à la réduction<br />

du niveau <strong>de</strong>s ressources par les voisins (réponse à la compétition) (Goldberg 1990 ; Goldberg<br />

& Landa 1991), induiront une compétition intense par rapport à <strong>de</strong> moins bons compétiteurs.<br />

Cette variation d’intensité <strong>de</strong> la compétition est schématisée sur la figure 6, où la compétition<br />

est plus intense pour les conditions environnementales A et B que pour les conditions C et D.<br />

L’importance <strong>de</strong> la compétition correspond à la réduction <strong>de</strong> performances due à la<br />

présence <strong>de</strong> compétiteurs, par rapport à la réduction <strong>de</strong> performances due à d’autres<br />

processus/conditions environnementales (Wel<strong>de</strong>n & Slauson 1986 ; Brooker et al. 2005 ;<br />

Brooker 2006). L’importance <strong>de</strong> la compétition est supposée forte sous <strong>de</strong>s conditions<br />

11


environnementales optimales pour l’espèce, mais faible pour <strong>de</strong>s conditions qui lui sont plus<br />

contraignantes. Sur la figure 6, l’importance <strong>de</strong> la compétition est plus forte pour les<br />

conditions environnementales B et C, car la réduction <strong>de</strong> performances <strong>de</strong> l’espèce s’explique<br />

majoritairement par ce processus.<br />

Même si Wel<strong>de</strong>n & Slauson (1986) affirment qu’une importance <strong>de</strong> la compétition et<br />

intensité <strong>de</strong> la compétition ne sont pas nécessairement reliés, une intensité <strong>de</strong> la compétition<br />

forte serait un pré-requis pour que la compétition soit importante (Bartelheimer et al. 2010).<br />

En effet, l’importance <strong>de</strong> la compétition sera élevée si la variation <strong>de</strong> l’intensité <strong>de</strong> la<br />

compétition le long <strong>de</strong> gradients est la cause <strong>de</strong> la variation <strong>de</strong> la structure <strong>de</strong>s communautés.<br />

Pour <strong>de</strong>s conditions environnementales égales, toute variation <strong>de</strong> l’intensité <strong>de</strong> la compétition<br />

fera varier proportionnellement l’importance <strong>de</strong> la compétition (Kikvidze et al. 2011). Mais<br />

les autres conditions environnementales n’étant pas constantes in natura, ces <strong>de</strong>ux<br />

composantes ne sont pas systématiquement corrélées (Wel<strong>de</strong>n & Slauson 1986 ; Kikvidze et<br />

al. 2011).<br />

Par exemple, une relation entre l’importance et l’intensité <strong>de</strong> la compétition n’est<br />

toujours mise en évi<strong>de</strong>nce : en effet, Lamb & Cahill (2008) montrent une intensité <strong>de</strong> la<br />

compétition racinaire très forte mais qui n’est pas importante, donc non structurante <strong>de</strong>s<br />

prairies à fétuque, milieux contraints par la disponibilité <strong>de</strong>s nutriments et <strong>de</strong> l’eau. La figure<br />

6 schématise cette idée autour <strong>de</strong> la relation possible entre intensité <strong>de</strong> la compétition et<br />

importance <strong>de</strong> la compétition. Pour une même intensité <strong>de</strong> compétition (par exemple sous les<br />

conditions environnementales A et B), l’importance <strong>de</strong> la compétition n’est pas la même : cet<br />

argument est en faveur d’une non-relation <strong>de</strong> ces <strong>de</strong>ux concepts. Néanmoins, pour une<br />

réduction similaire <strong>de</strong>s performances <strong>de</strong> l’espèce par les conditions environnementales<br />

(comme par exemple entre les conditions B et C), l’importance <strong>de</strong> la compétition apparaît être<br />

plus importante là où l’intensité <strong>de</strong> la compétition est la plus forte.<br />

Encore peu d’étu<strong>de</strong>s s’intéressent à la variation <strong>de</strong> l’importance <strong>de</strong> la compétition le<br />

long <strong>de</strong> gradients environnementaux ainsi que sa relation avec l’intensité <strong>de</strong> la compétition,<br />

limitant la compréhension du rôle <strong>de</strong> la compétition dans la structure <strong>de</strong>s communautés<br />

végétales (e.g. Greiner la Peyre et al. 2001 ; Gaucherand et al. 2006).<br />

12


Fig. 6: Représentation <strong>de</strong>s notions d’intensité et d’importance <strong>de</strong> la compétition. Pour <strong>de</strong>ux<br />

intensités <strong>de</strong> compétition différentes (forte dans la partie gauche et faible dans la partie droite)<br />

réduisant différemment les performances <strong>de</strong> l’espèce (différence entre niche fondamentale et<br />

niche réalisée : C), l’importance <strong>de</strong> la compétition correspond à la réduction <strong>de</strong> performances<br />

<strong>de</strong> l’espèce par la compétition et par tout autre processus suivant ce rapport (C/C+S). Lorsque<br />

S augmente, le résultat <strong>de</strong> ce rapport est faible : l’importance <strong>de</strong> la compétition est alors faible<br />

(conditions A et D). En revanche, quand S diminue, le rapport est élevé signifiant une<br />

compétition importante (conditions B et C) (adapté <strong>de</strong> Wel<strong>de</strong>n & Slauson 1986).<br />

3.2. Grime vs Tilman : un débat qui n’a donc plus lieu d’être<br />

Face à l’intérêt actuel <strong>de</strong> distinguer l’intensité <strong>de</strong> la compétition et <strong>de</strong> l’importance <strong>de</strong><br />

la compétition, il est maintenant reconnu que Grime (1977) a discuté non pas <strong>de</strong> l’intensité <strong>de</strong><br />

la compétition, mais <strong>de</strong> l’importance <strong>de</strong> la compétition, différence déjà mise en exergue par<br />

Wel<strong>de</strong>n & Slauson (1986) et Goldberg (1990).<br />

Cependant, <strong>de</strong> nombreuses étu<strong>de</strong>s ont entretenu cette confusion pendant <strong>de</strong><br />

nombreuses années concernant le point <strong>de</strong> vue <strong>de</strong> Grime : les étu<strong>de</strong>s se sont axées sur la<br />

variation <strong>de</strong> l’intensité <strong>de</strong> la compétition, soit augmentant avec la productivité suivant alors la<br />

thèse <strong>de</strong> Grime (1977), soit restant constante le long du gradient avec une bascule entre plus<br />

<strong>de</strong> compétition aérienne dans les milieux productifs et plus <strong>de</strong> compétition souterraine dans<br />

les milieux peu productifs (Tilman 1988). Certaines étu<strong>de</strong>s ont ainsi vérifié le modèle <strong>de</strong><br />

Grime (Pennings & Callaway 1992; Gau<strong>de</strong>t & Keddy 1995; Briones et al. 1998; Choler et al.<br />

2001) ; d’autres le modèle <strong>de</strong> Tilman (Berendse et al. 1992 ; Wilson 1993; Rea<strong>de</strong>r et al. 1994;<br />

Casper & Jackson 1997). Plusieurs citations en introduction d’articles traitant <strong>de</strong> la<br />

compétition témoignent et soutiennent cette confusion pour le débat Grime/Tilman:<br />

13


“Competition intensity may increase with productivity (Grime 1973,<br />

1974, 1979, […] Campbell and Grime 1992), remain constant as<br />

productivity changes (Newman 1973, Tilman 1982, 1988, Grubb<br />

1985) or be <strong>de</strong>pen<strong>de</strong>nt on the ratio of resource supply to <strong>de</strong>mand,<br />

which may be un-related to productivity (Taylor et al. 1990).”<br />

Twolan-Strutt & Keddy (1996)<br />

“A long-running <strong>de</strong>bate in plant ecology concerns whether the<br />

intensity of competitive interactions, or the <strong>de</strong>gree to which neighbors<br />

reduce plant growth, changes along a gradient of productivity,<br />

resource supply or non-resource stress (Grime 1973, 2001; Newman<br />

1973; Tilman 1988;[…]Rajaneimi 2003). In nutrient-poor or<br />

otherwise abiotically ‘stressed’ habitats, Grime (1973) asserted that<br />

the intensity of competition should be weakest; whereas Newman<br />

(1973) countered that competition should be intense (particularly<br />

belowground) in unproductive environments (Tilman 1988; Grace<br />

1991; Rajaneimi 2003; Craine 2005).”<br />

Fraser & Miletti (2008)<br />

Ce débat, lié à un problème <strong>de</strong> sémantique, doit maintenant être dépassé pour<br />

progresser dans la compréhension du rôle <strong>de</strong> la compétition dans la structuration <strong>de</strong>s<br />

communautés. Il est nécessaire d’étudier ces <strong>de</strong>ux composantes (intensité et importance) pour<br />

déterminer leur réelle opérationnalité (Grace 1995).<br />

3.3. La hiérarchie compétitive<br />

L’intensité <strong>de</strong> la compétition a été beaucoup plus étudiée que l’importance <strong>de</strong> la<br />

compétition (Brooker et al. 2005). Son étu<strong>de</strong> permet <strong>de</strong> classer les espèces en terme <strong>de</strong><br />

hiérarchie compétitive au sein d’une communauté (Keddy & Shipley 1989 ; Goldberg &<br />

Landa 1991). Le long <strong>de</strong> gradients environnementaux, la hiérarchie compétitive peut<br />

potentiellement varier et a le potentiel d’expliquer l’organisation <strong>de</strong>s communautés végétales<br />

car elle peut être un facteur important déterminant <strong>de</strong> l’abondance <strong>de</strong>s espèces (e.g. Gau<strong>de</strong>t &<br />

Keddy 1995 ; Fraser & Keddy 2005). L’effet <strong>de</strong>s changements <strong>de</strong> l’environnement sur la<br />

hiérarchie compétitive est largement étudié et les résultats apparaissent assez variables dans<br />

la littérature. Certaines étu<strong>de</strong>s mettent en évi<strong>de</strong>nce un maintien <strong>de</strong> la hiérarchie compétitive le<br />

long <strong>de</strong>s gradients, que ce soit le long <strong>de</strong> gradients d’inondations ou <strong>de</strong> disponibilité <strong>de</strong> la<br />

ressource en eau (e.g. Shipley et al. 1991 ; Keddy et al. 2002 ; Lenssen et al. 2004). Dans ce<br />

cas, la compétition n’est pas un processus majeur expliquant les patrons d’abondance<br />

d’espèces le long <strong>de</strong> ces gradients. En revanche d’autres étu<strong>de</strong>s mettent en évi<strong>de</strong>nce un<br />

14


changement dans la hiérarchie compétitive expliquant les patrons <strong>de</strong> structuration <strong>de</strong>s espèces<br />

(e.g. Keddy 1990 ; Wisheu & Keddy 1992). Les espèces moins compétitives sont alors<br />

observées dans les milieux les plus contraints. Cette variation <strong>de</strong> la hiérarchie compétitive<br />

permet <strong>de</strong> conclure à un rôle important <strong>de</strong> la compétition dans la structuration <strong>de</strong>s<br />

communautés. C’est donc indirectement que l’importance <strong>de</strong> la compétition dans la<br />

structuration <strong>de</strong>s communautés végétales a été étudiée.<br />

Une part <strong>de</strong> la diversité <strong>de</strong>s résultats entre les étu<strong>de</strong>s peut s’expliquer par le critère<br />

choisi pour quantifier l’aptitu<strong>de</strong> compétitive <strong>de</strong>s espèces : certaines considèrent l’effet<br />

compétitif d’autres la réponse à la compétition. Le long <strong>de</strong> gradients, l’effet compétitif tend à<br />

être constant tandis que la réponse à la compétition montre une plus gran<strong>de</strong> sensibilité au<br />

changement <strong>de</strong> l’environnement (Keddy et al. 1994 ; Wang et al. 2010). Les patrons <strong>de</strong> la<br />

réponse à la compétition expliqueraient, mieux que les patrons d’effet compétitif, les patrons<br />

d’abondance <strong>de</strong>s espèces le long <strong>de</strong> gradients (Howard & Goldberg 2001). Néanmoins, la<br />

variation <strong>de</strong> la réponse à la compétition le long <strong>de</strong> gradients environnementaux est encore peu<br />

comprise (Fraser & Miletti 2008). En effet, il existe différents types <strong>de</strong> réponses à la<br />

compétition en réponse à la présence <strong>de</strong> compétiteurs qui peuvent varier suivant les espèces et<br />

le milieu (Keddy et al. 1998) : certaines espèces présentent une stratégie <strong>de</strong> conservation <strong>de</strong>s<br />

ressources, d’autres au contraire, présentent la stratégie inverse <strong>de</strong> prélèvement rapi<strong>de</strong> <strong>de</strong>s<br />

ressources ; d’autres enfin décale le prélèvement <strong>de</strong>s ressources pour éviter cette compétition.<br />

3.4. La nécessité <strong>de</strong> caractériser la contrainte<br />

La caractérisation <strong>de</strong> la contrainte est une étape requise pour étudier la compétition,<br />

car le type <strong>de</strong> contrainte, c’est-à-dire (i) une ressource directement utilisable pour la<br />

croissance <strong>de</strong> la plante ou non, (ii) la distribution temporelle et spatiale <strong>de</strong> la contrainte qui<br />

peuvent influencer la compétition (Novoplanski & Goldberg 2001 ; Craine 2005). Les<br />

différents types <strong>de</strong> contraintes étudiées dans la littérature peuvent expliquer, au moins pour<br />

partie, la large palette <strong>de</strong> résultats obtenus concernant la variation <strong>de</strong> la compétition le long <strong>de</strong><br />

gradients environnementaux. En effet, Grime (2007) explique que les différentes<br />

interprétations du rôle <strong>de</strong> la ressource sujette à la compétition sont l’un <strong>de</strong>s éléments centraux<br />

<strong>de</strong> la controverse Grime-Tilman.<br />

Les résultats sont variables quant à la variation <strong>de</strong> l’intensité <strong>de</strong> la compétition le long<br />

<strong>de</strong> gradients. L’intensité peut être réduite avec l’augmentation <strong>de</strong> la contrainte, que ce soit le<br />

long <strong>de</strong> gradients <strong>de</strong> salinité et d’inondation (Pennings et al. 2005), ou le long d’un gradient<br />

15


<strong>de</strong> disponibilité en nutriments (Fraser & Keddy 2005). L’intensité <strong>de</strong> la compétition peut<br />

également être constante le long <strong>de</strong> gradients <strong>de</strong> disponibilité en nutriments et en eau (Corcket<br />

et al. 2003; Gaucherand et al. 2006; Carlyle et al. 2010) ou <strong>de</strong> gradients <strong>de</strong> salinité (Greiner la<br />

Peyre et al. 2001) ; la part <strong>de</strong> la compétition racinaire et <strong>de</strong> la compétition aérienne peut<br />

cependant varier en fonction <strong>de</strong> la disponibilité en ressources (Bartelheimer et al. 2010 ;<br />

Marion 2010). Comme les espèces peuvent présenter différentes réponses à la présence <strong>de</strong><br />

compétiteurs, il est important <strong>de</strong> savoir précisément quelle est la contrainte et quelle est la<br />

stratégie <strong>de</strong> l’espèce face à cette contrainte.<br />

3.5. Importance du niveau d’organisation<br />

Dans la littérature, on trouve fréquemment le terme « gradient <strong>de</strong> stress ». Or le stress<br />

est généralement défini à l’échelle <strong>de</strong> la communauté via <strong>de</strong>s mesures <strong>de</strong> productivité, suivant<br />

la définition du stress proposée par Grime (1979). Dans ce cadre, les différentes tolérances<br />

physiologiques <strong>de</strong>s espèces (niches fondamentales) le long d’un gradient environnemental, ne<br />

sont pas considérées. Les effets <strong>de</strong> la compétition à l’échelle <strong>de</strong> l’espèce ne sont cependant<br />

pas forcément transférables à l’échelle <strong>de</strong> la communauté dans la mesure où les différentes<br />

espèces <strong>de</strong> la communauté peuvent présenter <strong>de</strong>s niches fondamentales contrastées (Goldberg<br />

1994 ; Greiner la Peyre et al. 2001) : la contrainte doit être examinée à l’échelle <strong>de</strong> l’espèce.<br />

Cette nécessité <strong>de</strong> déterminer la contrainte à l’échelle <strong>de</strong> l’espèce a été introduite par Wel<strong>de</strong>n<br />

& Slauson (1986) : ils ont ainsi différencié le « strain » comme étant la réponse aux facteurs<br />

abiotiques au niveau spécifique et même au niveau individuel, tandis que le stress étant une<br />

mesure <strong>de</strong> la contrainte au niveau <strong>de</strong> la communauté et pouvait être approchée par le niveau<br />

<strong>de</strong> productivité à l’échelle <strong>de</strong> la communauté.<br />

3.6. Les méthodologies d’étu<strong>de</strong> <strong>de</strong> la compétition<br />

3.6.1. L’utilisation d’indices<br />

Les expérimentations in situ et en conditions contrôlées sont les approches<br />

majoritairement utilisées pour étudier la compétition (Damgaard 2004). Cela a permis, en<br />

parallèle, <strong>de</strong> développer <strong>de</strong> nombreux indices visant à quantifier la compétition. Weigelt &<br />

Jocliffe (2003) ont recensé <strong>de</strong> nombreux indices visant à quantifier notamment l’intensité <strong>de</strong><br />

la compétition, comme le Log Response Ratio (Hedges et al. 1999) ou encore le Relative<br />

16


Competition Intensity (Grace 1995). Plus récemment, <strong>de</strong>s indices pour quantifier l’importance<br />

<strong>de</strong> la compétition ont également été développés, comme Cimp (Brooker et al. 2005), Relative<br />

Neighbour Importance (Kikvidze & Armas 2010) ou encore Iimp (Seifan et al. 2010). Même<br />

s’il est indispensable <strong>de</strong> réaliser <strong>de</strong>s expérimentations pour comprendre le rôle <strong>de</strong> la<br />

compétition dans la structuration <strong>de</strong>s communautés grâce au contrôle d’un ou plusieurs<br />

facteurs (Kikvidze & Brooker 2010), le grand nombre d’indices utilisés dans la littérature<br />

limitent la généralisation <strong>de</strong>s résultats et la comparaison <strong>de</strong>s étu<strong>de</strong>s entre elles. De plus, le<br />

nombre limité <strong>de</strong> compétiteurs choisis dans les approches expérimentales limite<br />

l’extrapolation <strong>de</strong>s résultats à <strong>de</strong>s conditions réalistes <strong>de</strong> terrain (Goldberg 1996 ;<br />

Novoplanski & Goldberg 2001 ; Perkins et al. 2007 ; Engel & Weltzin 2008). En outre, dans<br />

les approches expérimentales en conditions contrôlées, les traitements expérimentaux choisis<br />

ne peuvent pas refléter tous les facteurs présents, leur potentielle interaction et leur variation<br />

au cours du temps affectant <strong>de</strong> surcroît les performances <strong>de</strong>s espèces (Wilson 2007).<br />

3.6.2. Les mesures <strong>de</strong> performance<br />

Les expérimentations in situ et en conditions contrôlées impliquent généralement <strong>de</strong>s<br />

mesures <strong>de</strong> biomasse et <strong>de</strong> survie <strong>de</strong>s espèces, récoltées à la fois en monocultures et en<br />

mixtures en fin d’expérimentations (Aarssen & Keogh 2002). Néanmoins, le résultat <strong>de</strong><br />

l’intensité <strong>de</strong> la compétition entre <strong>de</strong>ux espèces peut être différent suivant la mesure <strong>de</strong><br />

performance considérée, comme cela a été démontré entre l’analyse <strong>de</strong> données <strong>de</strong> biomasse<br />

et <strong>de</strong> survie (Goldberg & Novoplanski 1997 ; Goldberg 1999 ; Liancourt et al. 2005). De plus,<br />

les différents sta<strong>de</strong>s <strong>de</strong> développement <strong>de</strong>s plantes montrent différentes capacités compétitives<br />

et différentes capacités à tolérer la contrainte (Grubb 1977). En effet, il a été montré une<br />

variation <strong>de</strong> la hiérarchie compétitive d’espèces suivant le sta<strong>de</strong> mesuré : la germination, la<br />

croissance ou encore la survie (Howard & Goldberg 2001). Ces différents effets <strong>de</strong>s<br />

interactions compétitives sur les différents sta<strong>de</strong>s <strong>de</strong> développement auront <strong>de</strong>s conséquences<br />

sur les performances <strong>de</strong>s adultes (Fayolle et al. 2009). Cependant, peu d’étu<strong>de</strong>s ont testé les<br />

effets <strong>de</strong>s interactions sur différents sta<strong>de</strong>s <strong>de</strong> développement permettant ainsi <strong>de</strong> décrire le<br />

cycle <strong>de</strong> vie entier <strong>de</strong>s espèces ; alors que comprendre la relative importance <strong>de</strong> chaque<br />

composant <strong>de</strong> fitness est nécessaire pour comprendre la structuration <strong>de</strong>s communautés<br />

végétales (Howard & Goldberg 2001).<br />

17


Des approches utilisant <strong>de</strong>s données démographiques, via le développement <strong>de</strong><br />

modèles prenant en compte les différents composants du cycle <strong>de</strong> vie <strong>de</strong>s plantes comme les<br />

analyses LTRE (Life Table Response Experiments) sont utilisées pour évaluer l’influence <strong>de</strong><br />

la compétition sur les différents taux démographiques (e.g. Gustafsson & Ehrlen 2003;<br />

Fréville & Silvertown 2005). Dans cet objectif, <strong>de</strong>s modèles sont maintenant disponibles et<br />

paramétrisés pour estimer l’effet <strong>de</strong> la compétition entre espèces en conditions naturelles<br />

(Damgaard et al. 2009) représentant ainsi une voie à suivre pour évaluer la compétition in<br />

natura.<br />

Les modèles utilisant <strong>de</strong>s données démographiques sont actuellement développés pour<br />

évaluer la compétition. Ces approches sont prometteuses car celles-ci peuvent être conduites<br />

sous <strong>de</strong>s conditions naturelles et considèrent différentes mesures <strong>de</strong> performances <strong>de</strong>s<br />

plantes. Ce type d’approche a été choisi dans ce travail pour mesurer l’importance <strong>de</strong> la<br />

compétition le long d’un gradient environnemental d’inondation.<br />

4. Les traits fonctionnels : outils d’étu<strong>de</strong> <strong>de</strong>s règles d’assemblage<br />

Les traits fonctionnels représentent <strong>de</strong>s caractéristiques physiologiques,<br />

morphologiques ou phénologiques mesurés à l’échelle <strong>de</strong> l’individu reflétant l’adaptation <strong>de</strong>s<br />

plantes à leur milieu et leur réponse aux facteurs environnementaux (McGill et al. 2006 ;<br />

Violle et al. 2007). Depuis l’article séminal <strong>de</strong> McGill et al. (2006), les traits fonctionnels<br />

prennent une place importante dans l’étu<strong>de</strong> <strong>de</strong> la compréhension <strong>de</strong>s mécanismes <strong>de</strong><br />

coexistence <strong>de</strong>s espèces, <strong>de</strong> structure <strong>de</strong>s communautés et du fonctionnement <strong>de</strong>s écosystèmes<br />

(Lavorel & Garnier 2002). Ils permettent <strong>de</strong> s’affranchir <strong>de</strong> la taxonomie (Lavorel et al.<br />

1997). Ils permettent <strong>de</strong> prédire la réponse <strong>de</strong>s plantes à un filtre abiotique donné comme par<br />

exemple l’inondation (Blom & Voesenek 1996; Casanova & Brock 2000; Fraser & Karnesis<br />

2005), ou même la réponse <strong>de</strong>s plantes à la compétition pour la lumière par exemple (Keddy<br />

& Shipley 1989; Violle et al. 2006) et peuvent donc être considérés comme <strong>de</strong>s <strong>de</strong>scripteurs<br />

<strong>de</strong>s niches fondamentale et réalisée <strong>de</strong>s espèces (McGill et al. 2006 ; Violle & Jiang 2009)<br />

La considération du concept <strong>de</strong> niche dans l’approche fonctionnelle sous-entend que<br />

les différents filtres abiotiques et biotiques opèrent une sélection suivant les valeurs <strong>de</strong>s traits<br />

(Keddy 1992 ; McGill et al. 2006). L’étu<strong>de</strong> <strong>de</strong>s patrons <strong>de</strong> distribution <strong>de</strong>s traits <strong>de</strong>s espèces<br />

18


permettrait <strong>de</strong> déterminer l’importance <strong>de</strong>s différents processus responsables <strong>de</strong> ces patrons et<br />

donc d’étudier les règles d’assemblage <strong>de</strong>s espèces (Weiher & Keddy 1995). Dans ce sens,<br />

ont été introduites les notions <strong>de</strong> convergence et la divergence (MacArthur & Levins 1967).<br />

Face à un filtre abiotique, les espèces sont sélectionnées suivant leur tolérance physiologique<br />

à ces conditions environnementales : théoriquement, les espèces <strong>de</strong>vraient présenter <strong>de</strong>s<br />

valeurs <strong>de</strong> traits similaires en réponse à cette condition environnementale (Weiher & Keddy<br />

1995, 1998). On parle <strong>de</strong> convergence, <strong>de</strong> sous-dispersion ou encore d’« habitat filtering »<br />

(Fig. 7c). Par contraste, <strong>de</strong>ux espèces fonctionnellement proches qui interagissent ne peuvent<br />

pas coexister indéfiniment en raison du concept <strong>de</strong> la « limiting similarity ». Pour coexister et<br />

ainsi réduire l’intensité <strong>de</strong> leur interaction, les espèces doivent présenter une divergence<br />

compétitive <strong>de</strong>s traits d’exploitation <strong>de</strong>s ressources (Weiher & Keddy 1995, 1998), on<br />

s’attend à une sur-dispersion ou <strong>de</strong> « niche-differentiation » (Fig. 7b), permettant aux espèces<br />

coexistantes <strong>de</strong> diminuer l’intensité <strong>de</strong> leur interaction. Comme cela a été précisé dans la<br />

partie 1.2., les filtres agissent simultanément in natura : un patron intermédiaire à ceux <strong>de</strong> la<br />

divergence et convergence fonctionnelle est attendu (Fig. 7d). Pour tester l’effet <strong>de</strong> ces <strong>de</strong>ux<br />

processus déterministes dans la structuration <strong>de</strong>s communautés végétales, les données sont<br />

confrontées à un modèle nul qui considère une équivalence fonctionnelle entre les espèces<br />

(Fig. 7a).<br />

Fig. 7 : Représentation <strong>de</strong>s patrons hypothétiques <strong>de</strong>s traits (extrait <strong>de</strong> Weiher & Keddy<br />

1998). a : équivalence fonctionnelle entre toutes les espèces ; b : effet du filtre biotique<br />

sélectionnant <strong>de</strong>s espèces présentant <strong>de</strong>s valeurs <strong>de</strong> traits dispersés ; c : effet du filtre<br />

abiotique sélectionnant <strong>de</strong>s espèces présentant <strong>de</strong>s valeurs <strong>de</strong> traits similaires ; d : effet<br />

conjugué <strong>de</strong>s filtres abiotiques et biotiques résultant un patron <strong>de</strong> distribution <strong>de</strong>s traits<br />

intermédiaires aux <strong>de</strong>ux précé<strong>de</strong>nts.<br />

19


L’étu<strong>de</strong> <strong>de</strong> la diversité fonctionnelle est centrale dans la compréhension <strong>de</strong>s règles<br />

d’assemblage, car elle permet <strong>de</strong> relier la structure <strong>de</strong>s communautés au fonctionnement <strong>de</strong>s<br />

écosystèmes (Diaz & Cabido 2001 ; Diaz et al. 2007). Elle est définie comme étant l’étendue<br />

et les valeurs <strong>de</strong>s traits <strong>de</strong>s espèces influençant différents aspects du fonctionnement <strong>de</strong>s<br />

écosystèmes (Tilman 2001). La diversité fonctionnelle apparaît prometteuse pour étudier les<br />

règles d’assemblage dans les gradients car elle est influencée par les processus déterministes<br />

(Stubbs & Wilson 2004 ; Fukami et al. 2005 ; Grime 2006). La mesure <strong>de</strong> la diversité<br />

fonctionnelle est d’ailleurs le sujet <strong>de</strong> nombreux travaux récents proposant <strong>de</strong>s métho<strong>de</strong>s<br />

différentes pour la quantifier (e.g. Petchey &Gaston 2002 ; <strong>de</strong> Bello et al. 2009, 2010 ; Mason<br />

et al. 2003 ; Mouillot et al. 2005, 2007 ; Ricotta 2005), qui reste un champ <strong>de</strong> recherche<br />

ouvert.<br />

Les tests <strong>de</strong>s hypothèses <strong>de</strong> convergence et divergence en lien avec les processus<br />

déterministes le long <strong>de</strong> gradients environnementaux a fait l’objet d’étu<strong>de</strong>s récentes (e.g.<br />

Weiher et al. 1998 ; Ackerly & Cornwell 2007 ; Cornwell & Ackerly 2009 ; Jung et al. 2010)<br />

testant ces hypothèses <strong>de</strong> convergence et <strong>de</strong> divergence fonctionnelles face à un modèle nul.<br />

L’avantage <strong>de</strong> cette métho<strong>de</strong> est qu’elle est quantitative car les <strong>de</strong>grés <strong>de</strong> convergence et <strong>de</strong><br />

divergence fonctionnelle sont quantifiés. Globalement, cette métho<strong>de</strong> est intéressante pour<br />

étudier la distribution et l’abondance <strong>de</strong>s espèces car elle lie les stratégies écologiques, les<br />

théories d’assemblage <strong>de</strong>s espèces et la diversité fonctionnelle (McGill et al. 2006 ; Ackerly<br />

& Cornwell 2007 ; Violle & Jiang 2009). Les recherches sont encore en plein développement,<br />

elles nécessitent d’être approfondies car <strong>de</strong>s effets <strong>de</strong> l’échelle d’étu<strong>de</strong> (Weiher et al. 1998 ;<br />

Mokany & Roxburgh 2010), <strong>de</strong>s traits étudiés (Grime 2006 ; Schamp et al. 2008 ; Mokany &<br />

Roxburgh 2010) et <strong>de</strong> la mesure <strong>de</strong> diversité fonctionnelle choisie (Stubbs & Wilson 2004)<br />

sont mis en évi<strong>de</strong>nce, limitant la généralisation <strong>de</strong>s résultats.<br />

20


OBJECTIFS DE LA THESE<br />

L<br />

’approche déterministe visant à caractériser les effets <strong>de</strong>s filtres abiotiques et<br />

biotiques sur la structure et la dynamique <strong>de</strong>s communautés végétales a fait l’objet<br />

<strong>de</strong> nombreuses étu<strong>de</strong>s. Comme mentionné dans l’introduction, les interactions<br />

compétitives occupe une place majeure dans la compréhension <strong>de</strong> la structuration<br />

<strong>de</strong>s communautés végétales.<br />

Néanmoins, les différents points présentés précé<strong>de</strong>mment soulignent <strong>de</strong>s manques<br />

limitant la compréhension <strong>de</strong>s mécanismes agissant sur les patrons d’abondance d’espèces et<br />

<strong>de</strong> structure <strong>de</strong>s communautés le long <strong>de</strong> gradients environnementaux : i) le manque <strong>de</strong><br />

caractérisation fine <strong>de</strong> la variable environnementale majeure influençant les performances <strong>de</strong>s<br />

espèces, ii) le manque <strong>de</strong> distinction entre l’intensité <strong>de</strong> la compétition et l’importance <strong>de</strong> la<br />

compétition limitant la compréhension du rôle <strong>de</strong> la compétition dans la structuration <strong>de</strong>s<br />

communautés le long <strong>de</strong> gradients environnmentaux, iii) le manque <strong>de</strong> variabilité <strong>de</strong>s<br />

méthodologies utilisées pour caractériser l’importance <strong>de</strong> différents filtres environnementaux<br />

variables dans le temps et l’espace, couplant à la fois <strong>de</strong>s approches expérimentales et <strong>de</strong>s<br />

approches utilisant <strong>de</strong>s données récoltées in natura.<br />

suivantes :<br />

Face à ce constat, nous avons donc étudié les <strong>de</strong>ux problématiques <strong>de</strong> recherche<br />

- la quantification <strong>de</strong> la part <strong>de</strong>s interactions compétitives relativement aux<br />

facteurs environnementaux, c’est-à-dire l’importance <strong>de</strong> la compétition, pour<br />

comprendre la structuration <strong>de</strong>s communautés végétales.<br />

- l’évaluation <strong>de</strong> l’interaction <strong>de</strong> différents filtres environnementaux sur la structure<br />

<strong>de</strong>s communautés végétales, caractérisés par <strong>de</strong>s niveaux <strong>de</strong> productivité différentes et<br />

soumises à une variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité.<br />

Deux type <strong>de</strong> milieux correspondant à <strong>de</strong>ux types <strong>de</strong> climats contrastés ont été choisis: les<br />

prairies humi<strong>de</strong>s du Marais Poitevin et les pelouses xéro-halophiles <strong>de</strong> Camargue. La<br />

première problématique a été traitée en considérant le modèle <strong>de</strong> prairies humi<strong>de</strong>s du Marais<br />

Poitevin. Ces prairies sont productives et fertiles : dans ce contexte, il est attendu que les<br />

interactions entre plantes soient intenses. Les inondations, se produisant <strong>de</strong> manière régulière<br />

chaque année dans ce système (Amiaud et al. 1998). Cette occurrence régulière <strong>de</strong>s<br />

23


inondations permet l’étu<strong>de</strong> <strong>de</strong> l’interaction entre la contrainte abiotique et les interactions<br />

plantes-plantes. Les pelouses xéro-halophiles sont <strong>de</strong>s milieux peu fertiles et peu productifs,<br />

soumis à une forte variabilité <strong>de</strong>s conditions climatiques et en particulier <strong>de</strong> la ressources en<br />

eau : elles feront l’objet d’étu<strong>de</strong> <strong>de</strong> la secon<strong>de</strong> problématique. Cette variabilité <strong>de</strong>s patrons <strong>de</strong><br />

pluviosité a été appréciée via un jeu <strong>de</strong> données pluri-annuelles, en y analysant son poids<br />

respectif sur la structure <strong>de</strong>s communautés ainsi que sur la démographie <strong>de</strong>s espèces pérennes<br />

dominantes par rapport à différentes modalités <strong>de</strong> gestion pastorale.<br />

Ce travail est structuré en trois parties.<br />

La première partie est axée sur l’évaluation <strong>de</strong> la compétition le long d’un gradient<br />

environnemental d’inondation dans les prairies humi<strong>de</strong>s du Marais poitevin. L’étu<strong>de</strong> <strong>de</strong> la<br />

compétition est centrée espèce, et mobilise le concept <strong>de</strong> niche. Différentes méthodologies ont<br />

été appliquées pour étudier la compétition :<br />

- Le premier chapitre présente les patrons <strong>de</strong> variation <strong>de</strong>s <strong>de</strong>ux composantes <strong>de</strong> la<br />

compétition que sont l’intensité et l’importance, et ce en fonction <strong>de</strong> l’intensité <strong>de</strong> la<br />

contrainte à l’échelle <strong>de</strong> l’espèce. En se référant aux définitions <strong>de</strong> chaque composante<br />

<strong>de</strong> la compétition, nous avons testé les prédictions quant à leur patron <strong>de</strong> variation<br />

théorique le long d’un gradient <strong>de</strong> durée et <strong>de</strong> profon<strong>de</strong>urs d’inondation ont été testées.<br />

Cette étape requise pour comprendre la variation <strong>de</strong> chaque composante est pourtant<br />

souvent éludée (Brooker & Kikvidze 2008). Pour ce faire, nous avons fait l’hypothèse<br />

que l’importance <strong>de</strong> la compétition dépend du niveau <strong>de</strong> contraintes subies par les<br />

espèces cibles alors que l’intensité <strong>de</strong> la compétition ne dépend pas l’intensité <strong>de</strong> cette<br />

contrainte.<br />

- Les seconds et troisièmes chapitres concernent l’évaluation <strong>de</strong> l’importance <strong>de</strong> la<br />

compétition dans les patrons d’abondance d’espèces le long d’un gradient<br />

d’inondation, en testant l’hypothèse d’un tra<strong>de</strong>-off entre les capacités compétitives <strong>de</strong>s<br />

espèces et leur tolérance à la contrainte. La <strong>de</strong>scription <strong>de</strong>s niches fondamentales <strong>de</strong>s<br />

espèces est étudiée expérimentalement par <strong>de</strong>s transplantations d’individus à<br />

différentes positions le long du gradient in situ. La comparaison entre les niches<br />

fondamentales et les niches réalisées <strong>de</strong>s espèces permettra d’évaluer le rôle <strong>de</strong> la<br />

compétition dans les patrons d’abondance <strong>de</strong>s espèces (3 ème chapitre).<br />

24


Dans la secon<strong>de</strong> partie, l’importance <strong>de</strong>s différents filtres écologiques est évaluée à<br />

partir du développement <strong>de</strong> modèles utilisant <strong>de</strong>s données démographiques (survie,<br />

colonisation, croissance) reflétant la dynamique <strong>de</strong>s populations étudiées. Les données<br />

récoltées in natura expriment les conditions auxquelles les populations sont réellement<br />

soumises. Ces modèles sont développés dans les <strong>de</strong>ux sites d’étu<strong>de</strong>.<br />

- Le chapitre 4 concerne l’analyse <strong>de</strong> l’effet <strong>de</strong> l’environnement sur les effets<br />

compétitifs d’espèces présentant <strong>de</strong>s distributions différentes le long d’un gradient<br />

d’inondation, ainsi que l’analyse <strong>de</strong>s patrons <strong>de</strong> variation <strong>de</strong> l’importance <strong>de</strong> la<br />

compétition. Le modèle est développé à partir <strong>de</strong> relevés floristiques suivant la<br />

métho<strong>de</strong> <strong>de</strong>s points contact, permettant <strong>de</strong> tenir compte <strong>de</strong>s probabilités <strong>de</strong> survie, <strong>de</strong><br />

colonisation <strong>de</strong>s espèces et <strong>de</strong> leur croissance. Le développement <strong>de</strong> ce type<br />

d’approche permet <strong>de</strong> compléter la quantification <strong>de</strong> l’importance réalisée à partir<br />

d’indices (Kikvidze & Brooker 2010).<br />

- Le chapitre 5 concerne la proposition d’une métho<strong>de</strong> d’analyse du succès écologique<br />

<strong>de</strong>s espèces pérennes. Le modèle, testé sur les pelouses xero-halophyles<br />

méditerranéennes, prédit l’occupation <strong>de</strong> l’espace par ces espèces à partir <strong>de</strong> leurs<br />

probabilités <strong>de</strong> survie et <strong>de</strong> colonisation estimées à partir <strong>de</strong> points contacts. Le<br />

modèle permet d’évaluer l’élasticité <strong>de</strong> ces paramètres démographiques sous<br />

différentes conditions environnementales : la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité, la<br />

gestion par le pâturage, pour <strong>de</strong>s communautés végétales soumis à différents niveau <strong>de</strong><br />

stress salins.<br />

La troisième partie vise à comprendre les rôles respectifs <strong>de</strong>s filtres biotiques et<br />

abiotiques dans la répartition <strong>de</strong>s espèces végétales et les patrons <strong>de</strong> végétation et <strong>de</strong><br />

structure <strong>de</strong>s communautés, à la fois dans les prairies humi<strong>de</strong>s du Marais Poitevin et dans<br />

les pelouses xéro-halophiles <strong>de</strong> Camargue.<br />

- Le chapitre 6 a pour problématique la compréhension <strong>de</strong>s règles d’assemblage<br />

responsables <strong>de</strong> la coexistence <strong>de</strong>s espèces le long du gradient d’inondation. Dans<br />

cette étu<strong>de</strong>, les hypothèses <strong>de</strong> d’habitat filtering et <strong>de</strong> niche differentiation sont testées<br />

face à un modèle nul. Ces hypothèses ont été testées en mesurant la distribution <strong>de</strong><br />

<strong>de</strong>ux traits fonctionnels. Ces traits fonctionnels, choisis suivant leur pertinence dans la<br />

25


éponse aux filtres écologiques étudiés, ont été mesurés in situ le long du gradient<br />

d’inondation.<br />

- Le chapitre 7 concerne la détermination <strong>de</strong>s rôles respectifs <strong>de</strong> différents facteurs<br />

environnementaux sur la structure et la composition <strong>de</strong>s communautés végétales<br />

annuelles méditerranéennes. Ces communautés sont soumises notamment à une forte<br />

variabilité inter et intra-annuelle <strong>de</strong>s patrons <strong>de</strong> pluviosité (hétérogénéité <strong>de</strong> la<br />

ressource en eau), à une gestion par pâturage et à une salinité variable du sol. L’effet<br />

du niveau <strong>de</strong> stress sur les réponses au pâturage et à la variabilité <strong>de</strong>s patrons <strong>de</strong><br />

pluviosité. Cette analyse a été réalisée à partir <strong>de</strong> huit années <strong>de</strong> relevés floristiques.<br />

26


MODELES BIOLOGIQUES<br />

D<br />

eux sites ont été étudiés : il s’agit <strong>de</strong>s <strong>de</strong>ux plus gran<strong>de</strong>s zones humi<strong>de</strong>s <strong>de</strong><br />

France : la Camargue et le Marais Poitevin. Ces <strong>de</strong>ux sites présentent <strong>de</strong>s<br />

intérêts écologique et patrimonial exceptionnels, avec notamment une richesse<br />

floristique remarquable s’organisant le long d’une microtopographie, résultant<br />

une variation <strong>de</strong>s conditions environnementales et <strong>de</strong> communautés végétales.<br />

1. La microtopographie, génératrice d’une variété <strong>de</strong> communautés<br />

1.1. Les prairies humi<strong>de</strong>s du Marais Poitevin<br />

Les prairies humi<strong>de</strong>s du Marais Poitevin sont caractérisées par un gradient micro-<br />

topographique d’amplitu<strong>de</strong> d’environ 40-45 cm décrivant un gradient <strong>de</strong> durée d’inondation<br />

(Fig. 8). Les inondations sont récurrentes chaque année, c’est-à-dire qu’elles apparaissent à la<br />

fin <strong>de</strong> l’automne jusqu’au printemps <strong>de</strong> l’année suivante. Malgré cette régularité, les dates <strong>de</strong><br />

mise en eau et <strong>de</strong> fin d’inondations sont néanmoins variables : elles dépen<strong>de</strong>nt <strong>de</strong>s patrons <strong>de</strong><br />

pluviosité. Les conditions d’inondations en hauteur d’eau et en durée le long <strong>de</strong> cette micro-<br />

topographie sont variables : elles peuvent influencer par conséquent l’expression <strong>de</strong>s espèces<br />

suivant leur tolérance physiologique à ces conditions environnementales.<br />

Fig. 8 : Profil micro-topographique durant la saison d’inondation avec la baisse inondée au<br />

centre, inondée jusqu’à six mois par an ; et à gauche les niveaux topographiques les plus hauts<br />

non inondés.<br />

29


Une Analyse Factorielle <strong>de</strong>s Correspondances (AFC) <strong>de</strong> relevés floristiques réalisés le<br />

long <strong>de</strong> cette microtopographie nous révèle l’existence <strong>de</strong> trois communautés végétales (Fig.<br />

9).<br />

Dans les niveaux topographiques les plus hauts est décrit la communauté mésophile<br />

composée majoritairement <strong>de</strong> Lolium perenne, Carex divisa, Bromus commutatus, Cynosurus<br />

cristatus et Hor<strong>de</strong>um secalinum. Cette communauté végétale est décrite dans les replats non<br />

inondés, sauf années exceptionnelles caractérisées par une forte pluviosité.<br />

La communauté méso-hygrophile est décrite au niveau intermédiaire (pentes). Elle<br />

est composée en gran<strong>de</strong> majorité <strong>de</strong> Juncus gerardii, espèce qui peut être associée à <strong>de</strong>s<br />

espèces comme Hor<strong>de</strong>um marinum, Plantago coronopus ou encore Leontodon taraxacoï<strong>de</strong>s.<br />

Cette communauté végétale est soumise à <strong>de</strong>s conditions d’inondation qui sont variables, <strong>de</strong><br />

l’ordre <strong>de</strong> quelques semaines dépendantes <strong>de</strong> la quantité <strong>de</strong>s pluies. Localement, le sol <strong>de</strong><br />

cette communauté végétale peut être salé, une caractéristique qui est en lien avec la nappe<br />

d’imbibition salée dans ce secteur et qui semble favorisée par le piétinement et le pâturage du<br />

bétail (Amiaud et al.1998; Bonis et al. 2005), en conséquence une végétation halotolérante<br />

peut être observée.<br />

La communauté hygrophile est décrite dans les niveaux topographiques les plus bas<br />

ou dépressions inondables. Cette communauté végétale est composée d’Eleocharis palustris,<br />

Juncus articulatus, Agrostis stolonifera, Oenanthe fistulosa ou encore Glyceria fluitans. Les<br />

inondations peuvent durer jusqu’à 6 mois dans ces dépressions, avec une hauteur d’eau au<br />

<strong>de</strong>ssus du sol pouvant atteindre jusqu’à 50 cm.<br />

Nous avons cherché à limiter l’influence <strong>de</strong> la salinité du sol afin d’éviter que ce<br />

facteur ne se surimpose au facteur inondation qui nous intéresse. Pour cela, nous avons<br />

travaillé le long <strong>de</strong> toposéquences où la salinité du sol était faible, salinité mesurée à l’ai<strong>de</strong><br />

d’un conductimètre.<br />

30


Fig. 9 : Analyse Factorielle <strong>de</strong>s Correspondances (AFC) <strong>de</strong> l’ensemble <strong>de</strong>s relevés floristiques<br />

(n=261 relevés). a) Carte factorielle <strong>de</strong>s espèces (le co<strong>de</strong> <strong>de</strong>s espèces est donné en Annexe 1)<br />

b) Carte factorielle <strong>de</strong>s relevés avec la distinction <strong>de</strong>s trois communautés végétales.<br />

La composition en espèces peut varier au sein <strong>de</strong> chacune <strong>de</strong>s trois communautés<br />

végétales (Marion 2010). En effet, certains relevés floristiques apparaissent être<br />

intermédiaires à <strong>de</strong>ux communautés. Une analyse <strong>de</strong> la distribution <strong>de</strong>s espèces le long du<br />

gradient micro-topographique indique que certaines espèces ont <strong>de</strong>s amplitu<strong>de</strong>s <strong>de</strong><br />

distributions larges, d’autres plus restreintes (Fig. 10). La distribution <strong>de</strong>s espèces a été<br />

analysée par l’Outlying Mean In<strong>de</strong>x (OMI) qui est un outil performant pour séparer les niches<br />

réalisées <strong>de</strong>s espèces (Dolé<strong>de</strong>c et al. 2000) et pour déterminer les variables environnementales<br />

déterminantes dans cette ségrégation <strong>de</strong>s niches (Thuillier et al. 2004). OMI permet une<br />

interprétation directe <strong>de</strong> la séparation <strong>de</strong>s niches <strong>de</strong>s espèces car la métho<strong>de</strong> calcule la<br />

déviation <strong>de</strong> la distribution d’une espèce par rapport aux conditions moyennes du milieu.<br />

Ainsi est représenté la position moyenne <strong>de</strong> l’espèce mais également la variabilité <strong>de</strong>s habitats<br />

occupés par chaque espèce, qui est considérée comme l’amplitu<strong>de</strong> <strong>de</strong> la niche réalisée <strong>de</strong>s<br />

espèces.<br />

31


Fig 10: Position <strong>de</strong> la niche réalisée <strong>de</strong>s espèces et <strong>de</strong> la largeur <strong>de</strong> leur niche sur l’axe <strong>de</strong><br />

l’OMI (gradient micro-topographique). Les bars horizontales correspon<strong>de</strong>nt à ± SD et sont<br />

interprétés comme les largeurs <strong>de</strong> niche. Les lignes verticales en bas du graphique<br />

correspon<strong>de</strong>nt à la position topographique <strong>de</strong>s relevés le long du gradient (le co<strong>de</strong> <strong>de</strong>s espèces<br />

est donné en Annexe 1).<br />

Une ségrégation <strong>de</strong>s niches <strong>de</strong>s espèces est décrite le long du gradient micro-<br />

topographique, avec <strong>de</strong>s espèces présentant <strong>de</strong>s amplitu<strong>de</strong>s <strong>de</strong> niche faibles comme Bal<strong>de</strong>llia<br />

ranunculoï<strong>de</strong>s (BALRAN) et Mentha pulegium (MENPUL) dans les zones inondées ou<br />

encore Juncus gerardii (JUNGER), Parapholis strigosa (PARSTR) dans les zones<br />

intermédiaires du gradient. La préférence <strong>de</strong> Bal<strong>de</strong>llia ranunculoï<strong>de</strong>s, Mentha pulegium ou<br />

encore Oenanthe fistulosa (OENFIS) pour les habitats longuement inondés a été démontré<br />

lors d’une précé<strong>de</strong>nte étu<strong>de</strong> visant à prédire les patrons <strong>de</strong> distribution théoriques à l’ai<strong>de</strong> <strong>de</strong><br />

régressions <strong>de</strong>s espèces (Violle et al. 2007). Des espèces ont, en revanche, <strong>de</strong> larges<br />

amplitu<strong>de</strong>s <strong>de</strong> niche réalisée comme par exemple Agrostis stolonifera (AGRSTO), Rumex<br />

conglomeratus (RUMCON) ou Galium <strong>de</strong>bile (GALDEB).<br />

Ces analyses multivariées montrent que l’organisation <strong>de</strong> la végétation se traduit par le<br />

rattachement <strong>de</strong>s espèces à <strong>de</strong>s communautés mais reflétant également un continuum <strong>de</strong> leur<br />

répartition le long du gradient <strong>de</strong> durée d’inondation. Nous sommes donc concrètement au<br />

cœur du consensus actuel <strong>de</strong> l’organisation <strong>de</strong> la végétation le long <strong>de</strong> gradients<br />

32


environnementaux à mi-chemin entre les concepts <strong>de</strong> communauté et <strong>de</strong> continuum (voir<br />

Introduction 2.2). En un endroit donné, il est ainsi possible <strong>de</strong> recenser <strong>de</strong>s espèces euryèces<br />

et sténoèces pour lesquelles la question <strong>de</strong> la variation <strong>de</strong>s capacités compétitives se pose.<br />

1.2. Les pelouses xéro-halophiles <strong>de</strong> Camargue<br />

Tout comme pour les prairies humi<strong>de</strong>s du Marais Poitevin, une micro-topographie est<br />

observée dans les systèmes méditerranéens d’une amplitu<strong>de</strong> maximale <strong>de</strong> 50cm. Le long <strong>de</strong><br />

cette micro-topographie, l’humidité du sol et la salinité varient en raison <strong>de</strong> la présence d’une<br />

nappe souterraine salée à faible profon<strong>de</strong>ur. Deux communautés végétales se développent le<br />

long <strong>de</strong> ce gradient (Fig. 11). Les pelouses strictes (Directive Habitats co<strong>de</strong> 6220-2) occupent<br />

les niveaux topographiques les plus élevés (Fig. 12). Elles sont composées d’une large<br />

proportion d’espèces annuelles : Brachypodium distachyon, Scorpiurus muricatus, Psilurus<br />

aristatus, Evax pygmaea, Filago vulgaris, Galium murale, Euphorbia exigua, Crepis sancta<br />

et quelques espèces pérennes comme Brachypodium phoenicoi<strong>de</strong>s et Crepis vesicaria. Une<br />

végétation plus halotolérante se développe dans les niveaux topographiques inférieurs. Cette<br />

végétation est dominée par les espèces suivantes : Halimione portulacoi<strong>de</strong>s, Bromus<br />

hor<strong>de</strong>aceus, Bupleurum semicompositum, Hymenolobus procumbens, Parapholis filiformis,<br />

Parapholis incurva, Plantago coronopus correspond aux prés salés méditerranéens (Directive<br />

Habitats co<strong>de</strong> 1310-4).<br />

Fig. 11 : Analyse Factorielle <strong>de</strong>s Correspondances (AFC) <strong>de</strong> l’ensemble <strong>de</strong>s relevés<br />

floristiques (n=63 relevés). a) Carte factorielle <strong>de</strong>s espèces (le co<strong>de</strong> <strong>de</strong>s espèces est donné en<br />

Annexe 2) b) Carte factorielle <strong>de</strong>s relevés et distinction <strong>de</strong>s <strong>de</strong>ux communautés végétales.<br />

33


Parmi les espèces sont communes aux <strong>de</strong>ux communautés végétales, certaines<br />

montrent <strong>de</strong>s abondances équivalentes entre communautés (Dactylis hispanica, Geranium<br />

molle, Medicago polymorpha ou Polypogon maritimus) ou <strong>de</strong> forts contrastes d’abondance<br />

entre communautés (Bromus madritensis, Bellis annua ou Plantago lagopus).<br />

Fig. 12 : Pelouses strictes au maximum <strong>de</strong> floraison <strong>de</strong> la pâquerette annuelle Bellis annua,<br />

fin mars-début avril.<br />

2. La caractérisation <strong>de</strong> la variable environnementale d’intérêt<br />

2.1. Les patrons d’inondation dans les prairies humi<strong>de</strong>s du Marais Poitevin<br />

Une caractérisation fine du gradient a été mise en place afin d’acquérir <strong>de</strong>s données<br />

précises sur la durée, l’intensité et la fréquence <strong>de</strong>s inondations. Des piézomètres ont été mis<br />

en place à différents points le long <strong>de</strong> différentes toposéquences (Fig. 13), équipés <strong>de</strong> son<strong>de</strong>s<br />

automatiques enregistrant les battements <strong>de</strong> la nappe d’eau du sol sur une année complète et<br />

sur un pas <strong>de</strong> temps horaire. Ce travail s’accompagne <strong>de</strong> mesures topographiques à l’ai<strong>de</strong> d’un<br />

théodolite mesurant l’altitu<strong>de</strong> en différents points du site (Fig. 14), complétant par la même<br />

occasion un même travail précé<strong>de</strong>mment effectué (Violle et al. 2007) dans le but d’extrapoler<br />

l’intensité <strong>de</strong>s inondations en différents points et <strong>de</strong> réaliser un modèle numérique <strong>de</strong> terrain.<br />

Enfin, nous avons fait appel à <strong>de</strong>s géomètres pour placer nos mesures topographiques dans un<br />

système géo-référencé.<br />

34


Fig. 13 : Mise en place <strong>de</strong>s piézomètres à l’automne 2008 dans une dépression inondable.<br />

Fig. 14 : Station fixe du théodolite servant aux mesures topographiques.<br />

Avec les enregistrements automatiques <strong>de</strong> cinq son<strong>de</strong>s placées le long d’une<br />

toposéquence, il est possible <strong>de</strong> suivre la dynamique <strong>de</strong> la nappe d’eau au cours du temps. Sur<br />

la figure 15, les mesures <strong>de</strong> niveau d’eau réalisées par ces 5 son<strong>de</strong>s sont représentées par <strong>de</strong>s<br />

couleurs différentes : <strong>de</strong> la courbe rouge reflétant les battements <strong>de</strong> la nappe d’eau dans les<br />

niveaux topographiques les plus hauts jusqu’à la courbe bleu foncé reflétant les battements <strong>de</strong><br />

la nappe d’eau dans le point le plus bas du gradient sur la pério<strong>de</strong> <strong>de</strong> janvier à juin 2009. Le<br />

long <strong>de</strong> la toposéquence, les battements <strong>de</strong> la nappe sont globalement synchronisés entre les<br />

cinq points topographiques, avec <strong>de</strong>s diminutions du niveau d’eau et <strong>de</strong>s remontées<br />

synchrones suite à <strong>de</strong>s pluies. Durant le printemps 2009, <strong>de</strong>ux dates <strong>de</strong> remise en eau ont été<br />

observées à quelques jours d’intervalle.<br />

35


Fig. 15 : Hydrographe <strong>de</strong> la saison <strong>de</strong> croissance 2009 montrant la dynamique temporelle <strong>de</strong><br />

la nappe d’eau dans le système mesurée en cinq points d’une toposéquence donnée (du rouge<br />

pour la son<strong>de</strong> placée au niveau du replat non inondé au bleu foncé pour la son<strong>de</strong> placée dans<br />

la dépression inondable). Les fluctuations <strong>de</strong> la nappe d’eau sont synchrones au cours du<br />

temps.<br />

Les inondations provoquent une anoxie au niveau racinaire en raison l’engorgement<br />

du sol. Après la fin <strong>de</strong>s inondations, un stress hydrique peut également être une contrainte<br />

forte pour les plantes. Gowing et al. (1998) ont proposé une métho<strong>de</strong> pour quantifier<br />

l’exposition <strong>de</strong>s plantes à ces <strong>de</strong>ux contraintes en chaque point du gradient : elle consiste en la<br />

détermination <strong>de</strong> <strong>de</strong>ux seuils, le premier correspond au seuil à partir duquel la zone racinaire<br />

commence à <strong>de</strong>venir engorgée (seuil d’engorgement), le second à un seuil à partir duquel les<br />

plantes ressentent le manque d’eau (seuil <strong>de</strong> déficit hydrique). Les <strong>de</strong>ux seuils sont déterminés<br />

suivant les caractéristiques du sol : la porosité, la conductivité hydraulique et le point <strong>de</strong><br />

flétrissement. A partir <strong>de</strong> chaque seuil est déterminé le SEV (Sum Exceedance Value)<br />

représentant le <strong>de</strong>gré à partir duquel la nappe d’eau dans le sol se trouve au <strong>de</strong>ssus du seuil<br />

d’engorgement (SEV for aeration) et en <strong>de</strong>ssous du seuil où le stress hydrique apparaît (SEV<br />

for soil drying). Le « SEV for aeration » mesure donc l’exposition <strong>de</strong>s plantes à la contrainte<br />

anoxie et le « SEV for soil drying » mesure la contrainte <strong>de</strong>s plantes au stress hydrique.<br />

Il a été déterminé que le seuil pour le calcul du « SEV for aeration » se trouve à<br />

19.1cm au <strong>de</strong>ssous <strong>de</strong> la surface du sol et celui pour le calcul du « SEV for soil drying » à 42.2<br />

cm (Fig. 16). Ces <strong>de</strong>ux indices sont calculés durant la saison <strong>de</strong> croissance et sont exprimés en<br />

cm.jour -1 ou en m.semaine -1 . De récentes étu<strong>de</strong>s ont démontré l’utilité <strong>de</strong> ces indices pour<br />

expliquer la ségrégation <strong>de</strong>s niches d’espèces le long d’un gradient hydrologique (Silvertown<br />

36


et al. 1999; Silvertown 2004) car ces mesures sont réalistes <strong>de</strong>s conditions auxquelles les<br />

plantes sont soumises.<br />

Fig. 16 : Hydrographe <strong>de</strong> la saison <strong>de</strong> croissance 2009 montrant la dynamique temporelle <strong>de</strong><br />

la nappe d’eau dans le système et <strong>de</strong> visualiser la pério<strong>de</strong> où les plantes sont exposée à une<br />

anoxie du sol est observée (zone bleue au <strong>de</strong>ssus du seuil référence d’engorgement à -19.1cm)<br />

et la pério<strong>de</strong> où les plantes sont exposées à la sécheresse (zone beige en <strong>de</strong>ssous du seuil<br />

référence <strong>de</strong> déficit hydrique à -42.2 cm).<br />

Grâce à ces mesures topographiques géoréférencées pour lesquelles les SEV ont été<br />

déterminés, il est possible <strong>de</strong> réaliser <strong>de</strong>s cartes prédictives <strong>de</strong>s conditions d’anoxie et <strong>de</strong><br />

déficit en eau, à l’ai<strong>de</strong> <strong>de</strong> métho<strong>de</strong>s d’interpolation spatiale (Fig. 17). Ainsi il est possible par<br />

exemple <strong>de</strong> visualiser les zones où l’intensité <strong>de</strong> l’anoxie est forte <strong>de</strong> celles où elle est faible,<br />

reflétant la micro-topographie.<br />

37


Fig. 17 : Carte prédictive du « SEV for aeration » au cours <strong>de</strong> la saison <strong>de</strong> croissance 2009<br />

réalisée à partir d’une interpolation spatiale. La métho<strong>de</strong> a discriminé 9 intensités <strong>de</strong><br />

contraintes allant <strong>de</strong> 0 cm.jour -1 , anoxie inexistante, à 7.08 cm.jour -1 , anoxie la plus forte dans<br />

cette zone cartographiée.<br />

38


Les gradients d’inondations peuvent se surimposer à d’autres variables<br />

environnementales, elles aussi variant le long <strong>de</strong> gradients environnementaux comme par<br />

exemple la disponibilité en nutriments. Dans notre site d’étu<strong>de</strong>, l’analyse du taux <strong>de</strong><br />

minéralisation <strong>de</strong> l’azote, qui est un proxy <strong>de</strong> la quantité d’azote disponible pour les plantes<br />

(Rossignol 2006), a montré que la disponibilité en azote ne varie pas le long du gradient<br />

d’inondation (Rossignol, unpubl.). Nous n’avons donc pas <strong>de</strong> gradient <strong>de</strong> productivité en plus<br />

du gradient d’inondation.<br />

2.2. Les patrons <strong>de</strong> pluviosité en Camargue<br />

La pluviosité annuelle moyenne est <strong>de</strong> l’ordre <strong>de</strong> 600 mm. Les occurrences <strong>de</strong>s pluies<br />

sont surtout automnales et printanières. Pour caractériser le niveau <strong>de</strong> contraintes subi par les<br />

plantes, la pluviosité - évapotranspiration (P-ETP) est quantifiée : il s’agit d’une mesure qui<br />

reflète la disponibilité <strong>de</strong> la ressource en eau pour les plantes (Scarnati et al. 2009). Le bilan<br />

hydrique est généralement déficitaire dès le début du mois <strong>de</strong> février et ce déficit se poursuit<br />

jusqu’en septembre : par conséquent le bilan hydrique annuel est par lui-même déficitaire. Ce<br />

déficit hydrique entraine <strong>de</strong>s remontées capillaires <strong>de</strong> la nappe salée et ainsi <strong>de</strong>s salinisations<br />

<strong>de</strong> sol en surface.<br />

Les patrons <strong>de</strong> P-ETP montrent une forte variabilité inter-annuelle, avec par exemple<br />

un printemps très sec en 1997 succédant à un printemps très humi<strong>de</strong> en 1996 (Fig. 18). La<br />

variabilité est également intra-annuelle : par exemple le mois <strong>de</strong> mars 2001 fut plus sec que le<br />

mois d’avril 2001.<br />

Fig. 18: Pluviosité - Evapotranspiration mensuelle sur une pério<strong>de</strong> <strong>de</strong> trois mois montrant à la<br />

fois la variabilité inter et intra-annuelle sur une pério<strong>de</strong> <strong>de</strong> 20 ans.<br />

39


Afin d’évaluer le rôle potentiel <strong>de</strong>s conditions hydriques (patrons <strong>de</strong> pluviosité) sur la<br />

structure et la composition <strong>de</strong>s communautés végétales, une expérimentation in situ a été mise<br />

en place à l’hiver 2009 visant à simuler <strong>de</strong>s précipitations printanières et/ou automnales,<br />

possible grâce à l’ajout d’eau sur le terrain.<br />

Des quadrats permanents ont été soumis à quatre traitements : (i) un traitement témoin<br />

non arrosé reflétant les conditions <strong>de</strong> pluviosité naturelles, (ii) un traitement arrosage au cours<br />

du printemps, (iii) au cours <strong>de</strong> l’automne et (iiii) au printemps et en automne. Les résultats ne<br />

seront pas présentés dans ce manuscrit, l’expérimentation est toujours en cours.<br />

40


Partie 1 :<br />

La compétition dans les milieux inondés<br />

43


- Chapitre 1 -<br />

Detecting changes in the intensity and importance of competition<br />

for grasslands species along a flooding gradient.<br />

Soumis à Journal of Vegetation Science<br />

45


Chapitre I<br />

Detecting changes in the intensity and importance of competition for<br />

grasslands species along a flooding gradient<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1, 2 , Jan-Bernard Bouzillé 1 , François Mesléard 2, 3 , Anne Bonis 1<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, F-35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc, F-13200, France<br />

3 UMR CNRS-IRD 6116 Institut Méditerranéen d’Ecologie et <strong>de</strong> Paléoécologie, <strong>Université</strong><br />

d’Avignon IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09, France<br />

Abstract:<br />

Questions: Do competition importance and competition intensity <strong>de</strong>crease with the intensity<br />

of the species' strain? Do competitor performance and similarity with the competitor predict<br />

competition intensity? Does competition importance increase with competition intensity?<br />

Location: Greenhouse, <strong>Rennes</strong> University, Brittany, France<br />

Methods: Twelve wetland species were grown in monocultures and mixtures with the<br />

presence of one competitor in five experimental flooding treatments that controlled the<br />

duration of flooding and water <strong>de</strong>pth. The intensity of the strain, i.e. the intensity of the<br />

constraint at the species level, was quantified for each species by the log response ratio<br />

(lnRR(strain)). The intensity and importance of competition were quantified for each species<br />

using the LnRR(interaction) and Cimp indices. The functional similarity between the target and<br />

competitor was measured by the Gower similarity in<strong>de</strong>x for Specific Leaf Area (SLA) and<br />

plant height.<br />

Results: The twelve species experienced different strain types and intensities: long and <strong>de</strong>ep<br />

flooding conditions and the absence of flooding. For all species, competition importance was<br />

significantly related to strain intensity. By contrast, no general relationship was found<br />

between competition intensity and strain intensity, except for two species. The functional<br />

similarity between the target species and competitor did not predict competitive interactions.<br />

Together, competition intensity and importance vary wi<strong>de</strong>ly and in<strong>de</strong>pen<strong>de</strong>ntly along the<br />

flooding gradient.<br />

Conclusion: This study emphasizes the need to distinguish competition intensity and<br />

importance because they represent two facets of competition.<br />

Key-words: lnRR(strain), Cimp, lnRR(interaction), functional similarity, wetland species,<br />

strain<br />

Nomenclature: Tutin, T.G. et al. 1964-1980, Flora Europea, Cambridge University Press,<br />

Cambridge.<br />

47


Chapitre I<br />

48


Introduction<br />

Chapitre I<br />

The effect of competition within plant communities is generally measured without<br />

consi<strong>de</strong>ring the effect of the environment on plant success. This has led to a poor<br />

un<strong>de</strong>rstanding of the role of competition along environmental gradients, as both the abiotic<br />

environment and competition are important factors with respect to the structuring of plant<br />

communities (Keddy 1990; Greiner la Peyre et al. 2001; Brooker et al. 2005). The importance<br />

of competition, i.e. the role of competition in plant success relative to the role of the abiotic<br />

factor, is still rarely measured. However, together with the measurement of the intensity of<br />

competition (effect of competition per se), it is required for a better un<strong>de</strong>rstanding of species<br />

abundance patterns and the variation in the plant community structure along environmental<br />

gradients (Violle et al. 2010; Brooker & Kikvidze 2008).<br />

Competition intensity is <strong>de</strong>fined as the reduction of the fitness of a species due to the<br />

presence of neighbors (Wel<strong>de</strong>n & Slauson 1986). The intensity of competition experienced by<br />

the target species is positively related to the performance of its competitor (e.g. Gerry &<br />

Wilson 1995; Corcket et al. 2003; Liancourt et al. 2005; Jung et al. 2009). Along an<br />

environmental gradient, it can therefore be expected that any stress that locally reduces the<br />

performance of the competitor will thus reduce the intensity of the competition with other<br />

species. The pattern of variation in competition intensity along an environmental gradient<br />

may thus be related to the performance of the competitor. The functional approach of<br />

vegetation patterns (Mc Gill et al. 2006; Ackerly & Cornwell 2007) suggests that, in relation<br />

to the principle of limiting similarity (MacArthur & Levins 1967), two species with close<br />

values in some of their key functional traits are expected to experience intense competition<br />

(Johansson & Keddy 1991; Wilson 2007; Zhang et al. 2008). This prediction remains poorly<br />

studied experimentally (Elmendorf & Moore 2007) with some results fitting the expectation<br />

(e.g. Elmendorf & Moore 2007), whereas other results do not (e.g. Resetarits 1995).<br />

Competition importance is <strong>de</strong>fined as the reduction of fitness due to competition<br />

relative to the reduction of fitness by any other process or condition (Wel<strong>de</strong>n & Slauson<br />

1986). The importance of competition will thus correspond to the relative contribution of<br />

competition to the total reduction in the fitness of an individual relative to its physiological<br />

optimum (Wel<strong>de</strong>n & Slauson 1986). We can therefore expect that the <strong>de</strong>viation from the<br />

physiological optimum of a species in response to competition may correctly assess the<br />

importance of competition (Brooker et al. 2005; Jung et al. 2009; Carlyle et al. 2010;<br />

Kikvidze & Brooker 2010; Violle et al. 2010). Accordingly, when the environmental<br />

49


Chapitre I<br />

conditions are very close to the physiological optimum of a species, the proportion of the<br />

fitness reduced by competition is expected to be high. Symmetrically, the reduction of fitness<br />

will be mainly due to environmental conditions in a situation where they are further away<br />

from the physiological optimum of the studied species (Carlyle et al. 2010). It is therefore<br />

expected that there will be a negative relationship at the species level between competition<br />

importance and the intensity of the constraint (e.g. Greiner la Peyre et al. 2001; Gaucherand et<br />

al. 2006).<br />

The aim of this study was to experimentally <strong>de</strong>termine the patterns of variation in<br />

competition intensity and importance along an experimental flooding gradient which ranged<br />

from long floo<strong>de</strong>d conditions up to a never floo<strong>de</strong>d condition. The 12 species studied are<br />

typical of natural wet grasslands, representative of the three plant communities <strong>de</strong>scribed in<br />

natura along a flooding gradient that varies in flooding duration and water <strong>de</strong>pth. In this study<br />

we will: (i) quantify the intensity of stress at the individual level along a flooding gradient, i.e.<br />

the intensity of the strain (Suding et al. 2003; Gross et al. 2010), comparing species growth<br />

un<strong>de</strong>r optimal and non-optimal environmental conditions to assess the variation of the<br />

competition along environmental gradients as advocated by Elmendorf & Moore (2007); (ii)<br />

measure the competition intensity along the experimental flooding gradient for 12 species in<br />

mixture with one generalist competitor using the in<strong>de</strong>x <strong>de</strong>veloped by Hedges et al. (1999);<br />

(iii) measure the importance of competition in the success of the 12 studied species using the<br />

in<strong>de</strong>x <strong>de</strong>veloped by Brooker et al. (2005) by measuring the variation in plant success due to<br />

the presence of competitors relative to the variation in plant success due to the change in the<br />

flooding regime; and (iv) investigate if the functional similarity among the species in mixture<br />

with respect to the Specific Leaf Area (SLA) and plant height predict the variation in the<br />

intensity of their interaction along the flooding gradient, as both traits were found to vary in<br />

response to the flooding conditions (Blom & Voesenek 1996; Mommer et al. 2006; Colmer &<br />

Voesenek 2009) and the biotic environment (Violle et al. 2011).<br />

Four hypotheses were tested: (i) competition intensity varies in<strong>de</strong>pen<strong>de</strong>ntly from the<br />

intensity of strain experienced by the target species; (ii) competition intensity is higher when<br />

the growth performance of the competitor is high, and when the traits of both the target and<br />

competitor species show high similarity; (iii) competition importance increases with the<br />

<strong>de</strong>crease in strain intensity experienced by the target species; and (iv) competition importance<br />

varies in<strong>de</strong>pen<strong>de</strong>ntly with competition intensity.<br />

50


Methods<br />

Plant material<br />

Chapitre I<br />

Wet grasslands of the Marais Poitevin located on the French Atlantic coast (46°28’N,<br />

1°13’W) are characterized by an elevation gradient of approximately 40 to 45 cm, along<br />

which the flooding duration and water <strong>de</strong>pth vary (Violle et al. 2007). Twelve wetland<br />

species, presenting different life forms, were selected according to their field distribution<br />

along the flooding gradient (Violl et al. 2007). Four mesophilous species are typical of upper<br />

locations that are never floo<strong>de</strong>d: Carex divisa, Cynosurus cristatus, Hor<strong>de</strong>um secalinum and<br />

Lolium perenne (one sedge and three grasses). Three meso-hygrophilous species, Juncus<br />

gerardii, Bellis perennis and Leontodon autumnalis (one rush and two forbs), are found on the<br />

intermediate slopes enduring a variable duration of flooding during some weeks per year.<br />

Four hygrophilous species are found in the lower parts of the elevation gradient where<br />

flooding occurs for several months per year: Glyceria fluitans, Juncus articulatus, Mentha<br />

pulegium and Trifolium fragiferum (one grass, one rush and two forbs). In natura, four<br />

species only occur over a very restricted portion of the gradient (C. cristatus, G. fluitans, J.<br />

articulatus and M. pulegium). The seven other species are wi<strong>de</strong>spread over a large portion of<br />

the gradient (personal observation).<br />

Experimental <strong>de</strong>sign<br />

The effect of competition on the studied species was assessed by comparing their<br />

performances in monocultures and in mixtures with A. stolonifera, in five experimental<br />

hydrological treatments. Six tillers of each of the 12 target species were transplanted from the<br />

field to 3-L pots in the beginning of March 2009, with a potting soil corresponding to a<br />

mixture of gar<strong>de</strong>n soil and compost. Tillers of similar size and stage were transplanted.<br />

Monocultures and mixtures were submitted to five experimental treatments representing the<br />

range of flooding duration and water <strong>de</strong>pth observed in natura: (i) a non-floo<strong>de</strong>d treatment<br />

(NF); (ii) a floo<strong>de</strong>d treatment characterized by a water-<strong>de</strong>pth of 10 cm above the soil surface<br />

for a short duration (10-S); (iii) 10 cm above the soil surface for a long duration (10-L); (iv)<br />

20 cm above the soil surface for a short duration (20-S); and (v) 20 cm above the soil surface<br />

for a long duration (20-L). The duration of flooding lasted nine weeks for the short duration<br />

treatment and 13 weeks for the long duration treatment.<br />

51


Chapitre I<br />

Agrostis stolonifera was used as a standard neighbor as it occurs in situ all along the<br />

studied gradient. This species was also consi<strong>de</strong>red as a target species.<br />

The experiment was conducted at the experimental gar<strong>de</strong>n of the University of <strong>Rennes</strong><br />

(France) in outdoor tanks, with four replicate tanks for each flooding treatment. For each<br />

treatment, six replicates per species were randomly assigned to the tanks. Before applying the<br />

experimental treatments, the species were submitted to an acclimatization phase in<br />

waterlogged conditions for one week (from 9-15 March 2009). The water level was then<br />

progressively adjusted to 10 cm and 20 cm according to the treatment and was kept stable<br />

from mid-March up until 4May 2009 for the short duration treatments (10-S and 20-S) and 3<br />

June 2009 for the long duration treatments (10-L and 20-L). At the end of the experiment, the<br />

aboveground biomass was harvested, cleaned, and oven-dried at 70°C for 72 h, and then<br />

weighed.<br />

Plant traits<br />

Specific Leaf Area (SLA), i.e. the ratio of fresh leaf area to leaf dry mass, and plant<br />

height were measured for each individual in monocultures in every experimental treatment.<br />

The functional traits are related to species flooding responses allowing the improvement of<br />

gas diffusion (Blom & Voesenek 1996; Lenssen et al. 1998; Mommer et al. 2004). They are<br />

also related to the species' competitive ability (Weiher et al. 1999), indicating the resource<br />

acquisition strategy of the species. Following Weiher et al. (1999) and Cornelissen et al.<br />

(2001), the SLA was measured on the youngest fully grown leaf in the light using a leaf area<br />

meter (Li-Cor, Lincoln, NE, USA). Plant height was measured as the difference between the<br />

height of the highest photosynthetic leaf and the base of the plant. Height was not measured<br />

for species with roset life forms (e.g. Bellis perennis and Leontodon autumalis).<br />

Calculation of indices approaching strain intensity, competition intensity and competition<br />

importance<br />

The measure of individual stress: strain<br />

Strain was estimated by comparing the biomass of the species in monoculture and in<br />

optimal conditions with the biomass in non-optimal conditions (Suding et al. 2003, Gross et<br />

al. 2010). This was measured by the log response ratio (lnRR), where:<br />

lnRR (strain) = ln (individual biomass production in non-optimal conditions without<br />

neighbors/mean biomass production in optimal conditions)<br />

52


Chapitre I<br />

Negative values for lnRR (strain) indicate that the growth performance of the target species is<br />

<strong>de</strong>pressed by the environmental conditions. For each species, the optimal treatment<br />

correspon<strong>de</strong>d to the experimental condition in which survival and high biomass production<br />

were the highest.<br />

Intensity of competition<br />

The intensity of competition was characterized by the lnRR (interaction) in or<strong>de</strong>r to<br />

quantify the reduction of individual performances due to the presence of neighbors (Hedges et<br />

al. 1999), where:<br />

lnRR (interaction) = ln (individual biomass production with neighbors/mean biomass<br />

production without neighbors)<br />

A negative value for the log response ratio indicates competition and a positive value<br />

indicates facilitation.<br />

Importance of competition<br />

The in<strong>de</strong>x of competition importance Cimp (Brooker et al. 2005) was used to quantify<br />

the importance of competition relative to the impact of all other environmental factors, where:<br />

Cimp= (Biomasscomp-Biomasstarget)/(MaxBiomasstarget-y)<br />

Biomasscomp is the probability of survival of the target species in the presence of the<br />

competitor A. stolonifera, Biomasstarget is the probability of survival without a competitor,<br />

MaxBiomass is the maximum value of the probability of survival without a competitor among<br />

the hydrological treatments and y is the smallest value of either Biomasscomp or Biomasstarget.<br />

This in<strong>de</strong>x varies between -1 and +1. The contribution of competition relative to the<br />

hydrological conditions increases when Cimp tends toward -1. The contribution of the abiotic<br />

factor relative to competition increases when Cimp tends toward +1.<br />

Functional similarity between the target species and neighbor<br />

The functional similarity was measured by the Gower similarity in<strong>de</strong>x (Gower 1971),<br />

which is calculated as follows:<br />

with<br />

53


Chapitre I<br />

where n= number of functional traits and x = the functional trait. sijk is a measure of<br />

disagreement between species j and k, the competitor A. stolonifera, for variable i. This in<strong>de</strong>x<br />

was calculated for the SLA and height for each pair of the 12 species used as target species<br />

and A. stolonifera, in each replicate and in each experimental treatment. This in<strong>de</strong>x varies<br />

between 0 and 1. A value of 0 indicates a strong similarity between the target species and the<br />

competitor, whereas a value of 1 indicates no similarity in the trait value among the two<br />

species (Podani & Schmera 2006). To avoid an effect of the similarity in<strong>de</strong>x chosen, the<br />

functional similarity was also calculated with the Eucli<strong>de</strong>an distance which is more<br />

commonly used in the literature.<br />

Data analyses<br />

For all species and all water treatments, the tank effect (n=4 per treatment) was not<br />

significant (two-way ANOVA, P>0.05): accordingly, all repetitions were thereafter combined<br />

for analysis.<br />

Effect of the hydrological treatments<br />

Strain – A one-way ANOVA was run for each species to test whether the strain intensity<br />

was different between the experimental treatments, followed by post-hoc Tukey HSD tests<br />

after verifying normality. The effect of strain on the aboveground biomass production of the<br />

plant-target was investigated using linear regression, taking all the experimental treatments<br />

into account.<br />

Competition intensity – A one-way ANOVA was run for each species to test whether the<br />

competition intensity varied between the experimental treatments, followed by post-hoc<br />

Tukey HSD tests after verifying normality. Moreover, two-way ANOVAs were performed to<br />

test the effect of the experimental treatments and the effect of the species on competition<br />

intensity and competition importance. The two-way ANOVA showed a significant interaction<br />

between the species x hydrological treatment term, indicating a shift in the competitive<br />

hierarchy among the species with hydrological treatments (Lenssen et al. 2004).<br />

Effect of strain intensity on competition intensity and importance<br />

The relationships between, on one hand, competition intensity and importance and, on<br />

the other hand, strain intensity were investigated for each of the 12 species using linear<br />

54


Chapitre I<br />

regressions. Because some experimental treatments impact species survival to various extents,<br />

the number of replicates varied between species, and was low for B. perennis, C. divisa and C.<br />

cristatus: it was not possible to use this method for these three species.<br />

Effect of neighbor performances on competition<br />

The relationships between the aboveground biomass of the neighbor A. stolonifera and<br />

both competition intensity and importance were investigated using linear regressions. These<br />

relationships were first tested in<strong>de</strong>pen<strong>de</strong>ntly for each species and then a second time for all<br />

species combined.<br />

Relationship between competition intensity and importance<br />

The relationship between competition importance and intensity was investigated using<br />

linear regression. This relationship was first tested in<strong>de</strong>pen<strong>de</strong>ntly for each species, with the<br />

dataset including all experimental treatments together and then a second time combining all<br />

species and all experimental treatments together.<br />

Relationship between competition intensity and the <strong>de</strong>gree of similarity with the<br />

competitor<br />

The relationship between the <strong>de</strong>gree of similarity between the neighbor and target<br />

species (measured for SLA and plant height) and competition intensity was investigated for<br />

each species separately using linear regressions.<br />

Statistical tests were calculated with the software R (R version 2.7.2., 2008, R<br />

Foundation for Statistical Computing, Vienna, Austria) and the VEGAN package (Oksanen et<br />

al. 2008) was used to calculate Gower’s in<strong>de</strong>x.<br />

Results<br />

Differential patterns of strain<br />

The 12 species showed contrasted strain intensity <strong>de</strong>pending on the hydrological<br />

conditions (Fig. 1). The range of tolerance to the flooding treatments varied strongly between<br />

the species and <strong>de</strong>pen<strong>de</strong>d on the plant community they are related to in situ. Three of the four<br />

55


Chapitre I<br />

mesophilous species did not survive the long flooding duration treatment (C. cristatus, L.<br />

perenne un<strong>de</strong>r the 10-L and 20-L treatments and C. divisa un<strong>de</strong>r the 20-L treatment;<br />

Appendix S1) and H. secalinum was highly constrained by long flooding. As for the meso-<br />

hygrophilous species, various levels of tolerance to flooding were observed: from no survival<br />

(B. perennis, Appendix S1), growth performance very strongly diminished (L. autumnalis) or<br />

only slightly impacted (J. gerardii). Among the four hygrophilous species, the growth of three<br />

species (G. fluitans, J. articulatus and M. pulegium) was constrained by the absence of<br />

flooding whereas their survival was only slightly impacted by all the experimental treatments<br />

(Appendix S1). One hygrophilous species (T. fragiferum) was found to be highly constrained<br />

by a long duration of flooding.<br />

Fig. 1: Strain intensity, evaluated by biomass measurements, for the species in each floo<strong>de</strong>d<br />

treatment. Similar letters indicate no significant differences in the mean between the floo<strong>de</strong>d<br />

treatments from the one-way ANOVA analysis followed by Tukey HSD tests. The optimal<br />

treatment chosen for the calculation of strain was different according to the species: for most<br />

species, this was the non-floo<strong>de</strong>d treatment (NF); for A. stolonifera and M. pulegium: 10-S;<br />

for J. aticulatus: 10-L; for G. fluitans: 20-L.<br />

56


Intensity of competition along the experimental gradient<br />

Chapitre I<br />

Interaction with A. stolonifera negatively impacted the survival rate for all species, but<br />

to varying extents, <strong>de</strong>pending on the species and the experimental treatment (Appendix S2).<br />

For all species combined together, competition intensity was positively and linearly related to<br />

the biomass of the neighbor A. stolonifera (F=20.1; P


Chapitre I<br />

Competition intensity was not significantly related to the strain intensity experienced<br />

by the target species for ten of the 12 species studied. By contrast, for A. stolonifera,<br />

competition intensity increased with the strain intensity whereas for T. fragiferum, the reverse<br />

pattern was found (Fig 3).<br />

Fig. 3: Negative relationships between strain intensity for the target species and competition<br />

intensity for T. fragiferum and positive relationship for A. stolonifera- no significant<br />

relationships were found for all nine other species.<br />

Functional similarity did not predict the intensity of competition<br />

Competition intensity was not significantly related to the similarity between the target<br />

species and the competitor, except for C. cristatus (F=8.93; adjusted r²= 0.665; P=0.05) and<br />

T. fragiferum (F=9.78; adjusted r²= 0.37; P=0.007) which were similar to A. stolonifera in<br />

terms of plant height (data not shown). Only these two significant regressions were<br />

highlighted with the measure of similarity using the Eucli<strong>de</strong>an distance (data not shown).<br />

Competition importance along the experimental gradient<br />

The importance of competition varied between the treatments and species (two-way<br />

ANOVA; F=3.638; P


Chapitre I<br />

Fig. 4: Contribution of the competition compared to hydrological conditions (importance of<br />

competition): increased with the reduction of the strain intensity. The regression was<br />

significant at 5%. Not studied for B. perennis, C. cristatus and C. divisa due to not enough<br />

replicates (low survival rate).<br />

No relationship between competition intensity and importance<br />

For all species together, competition intensity was not related to competition<br />

importance (F=0.733; P>0.05; data not shown).<br />

Discussion<br />

The main purpose of this study was to investigate the variation in the intensity and<br />

importance of competition in relation to the intensity of strain that each species sustained. As<br />

expected, competition intensity was not found to vary with the strain intensity for ten of the<br />

12 studied species. Competition intensity appears to be strongly related to the biomass of the<br />

competitor, with no effect of the trait similarity among the target and competitor, with only<br />

one exception (plant height similarity between A. stolonifera and T. fragiferum). By contrast<br />

and as expected, competition importance increases when the strain intensity of strain<br />

diminishes. It could be conclu<strong>de</strong>d that competition intensity and importance vary<br />

in<strong>de</strong>pen<strong>de</strong>ntly along the flooding gradient.<br />

59


Species specific strain<br />

Chapitre I<br />

Fundamental niches regarding flooding conditions were found to be contrasted among<br />

the 12 species studied, as <strong>de</strong>monstrated for other wetland flora by Shipley et al. (1991),<br />

Baruch (1994) and Jung et al. (2009). The success of species related to mesophilous and<br />

meso-hygrophilous communities appears to be negatively impacted by long duration flooding<br />

and water <strong>de</strong>pth. As expected, hygrophilous species are rather constrained by the absence of<br />

flooding, except T. fragiferum.<br />

While the main contrasts with respect to flooding tolerance occurred among species<br />

connected to different plant communities, our results <strong>de</strong>monstrate that, within a given<br />

community type (i.e. the mesophilous, meso-hygrophilous or hygrophilous community)<br />

species could however show contrasted <strong>de</strong>grees of tolerance to hydrological treatments and<br />

strain appears to be highly species-specific (Gross et al. 2010). The greater tolerance of C.<br />

divisa to experimental flooding compared to the other mesophilous species and of J. gerardii<br />

compared to the other meso-hygrophilous species fits well with their occasional occurrence in<br />

situ in floo<strong>de</strong>d environments. Similarly, the low tolerance to long duration flooding of C.<br />

cristatus fits well with their distributions, which are restricted to non-floo<strong>de</strong>d locations<br />

(Violle et al. 2007).<br />

Competition importance related to the strain intensity of the target species<br />

The importance of competition <strong>de</strong>creases with the increase in strain intensity that each<br />

species sustained, i.e. either the absence of flooding or exposure to a long duration of<br />

flooding; this was observed for all the species studied regardless of their life forms (grasses,<br />

sedges or forbs). This supports the findings of other studies focused on other types of<br />

constraints: the level of productivity (Brooker et al. 2005; Gaucherand et al. 2006) with either<br />

disturbance (Carlyle et al. 2010) or soil salinity (Greiner la Peyre et al. 2001). The<br />

relationship between competition importance and strain intensity for a species thus appears to<br />

be a general one among the species and types of constraints. This was actually expected as, by<br />

<strong>de</strong>finition, the importance of competition represents the balance between the effect of<br />

competition and the effect of the abiotic factor (strain) on species performance. Our findings<br />

<strong>de</strong>monstrate that when species are close to their optimal conditions (the absence of flooding<br />

60


Chapitre I<br />

or a long duration of flooding), competition has a great role in explaining their ecological<br />

success.<br />

Intensity of competition along experimental gradient<br />

As we hypothesized, competition intensity appears unrelated to the strain intensity<br />

experienced by the species, for all studied species except T. fragiferum. The intensity of<br />

competition measured the effect of the presence of competitors per se. The performance of the<br />

generalist competitor, A. stolonifera, should then predict the intensity of competition, and this<br />

was in fact what was found for all the species studied. The performance of the competitor A.<br />

stolonifera may directly impact the plant target performance by shading effect or indirectly,<br />

by reducing the nutrient resource availability. Accordingly, the similarity among plants in<br />

mixture with respect to the SLA and plant height similarity was expected to predict the<br />

competition intensity, as these two functional traits constitute a good proxy of the resource<br />

uptake and light capture ability of the plants (Weiher et al. 1999; Lavorel & Garnier 2002;<br />

Garnier et al. 2004).<br />

However, for ten of the 12 studied species, there does not appear to be any significant<br />

relationship between functional similarity in those two traits and the intensity of their<br />

interaction with A. stolonifera. Only C. cristatus and T. fragiferum showed a <strong>de</strong>crease in their<br />

competitive response with increasing trait similarity with A. stolonifera with respect to plant<br />

height. This does not invalidate the limiting similarity hypothesis as only two traits were<br />

consi<strong>de</strong>red in this study given that the species niche is multidimensional, with a set of traits<br />

associated to each dimension of the niche. Moreover, the duration of the experiment may have<br />

also been too short for such a pattern to emerge as reported in other experiments (Resetarits<br />

1995; Perkins et al. 2007; Wang et al. 2010).<br />

The pattern in competition intensity across the hydrological treatments varied between<br />

the species. The intensity of competition was constant among the treatments for a first group<br />

of species (i.e. L. perenne, C. divisa, B. perennis, G. fluitans and J. articulatus). This kind of<br />

consistency in the intensity of competition for species has already been reported along<br />

productivity gradients (Brooker et al. 2005; Gaucherand et al. 2006; Carlyle et al. 2010). This<br />

constant pattern in competition intensity across a gradient could hi<strong>de</strong> a switch in the balance<br />

between shoot and root competition (Tilman 1988), indicating that more work is required to<br />

distinguish above-ground and below-ground competition intensity in our system. For another<br />

group of species (i.e. H. secalinum, J. gerardii and M. pulegium), the intensity of competition<br />

61


Chapitre I<br />

varied across the hydrological treatments, with different patterns specific to each species.<br />

Such variety in a specific response pattern was already reported by Wilson & Tilman (1995)<br />

and Jung et al. (2009). In<strong>de</strong>ed, the competitive response, which was consi<strong>de</strong>red here along the<br />

experimental treatments, was already reported to vary wi<strong>de</strong>ly according to the environment<br />

and species i<strong>de</strong>ntity (Greiner Peyre et al. 2001; Brooker et al. 2005; Jung et al. 2009; Carlyle<br />

et al. 2010).<br />

Importance and intensity: two facets of competition<br />

Numerous studies have focuses on how competition varies along gradients, which has<br />

also led to much confusion with regards to its measurement (Brooker et al. 2005; Brooker &<br />

Kikividze 2008). Our study helps clarify the expectations about the variation of competition<br />

along environmental gradients by showing that, following Greiner la Peyre et al. (2001),<br />

Gaucherand et al. (2006), Lamb & Cahill (2008), Zhang et al. (2008) and Lamb et al. (2009),<br />

competition intensity and importance varied in<strong>de</strong>pen<strong>de</strong>ntly from each other along the<br />

experimental gradient. Our study, like the re-analysis of Rea<strong>de</strong>r et al.'s data (1994) by<br />

Brooker et al. (2005), also shows that competition intensity and competition importance<br />

correspond to two different facets of competition.<br />

On one hand, the importance of competition measures how much competition shapes<br />

the structure of plant communities. On the other hand, the intensity of competition reflects the<br />

success of species' strategies to coexist with competitors in a given environment and then<br />

measure the species' competitive ability. Lamb et al.'s study (2009) provi<strong>de</strong>s clarification on<br />

the distinction between the intensity and importance of competition. Based on a mesocosm<br />

study controlling the level of fertility, they showed that, even if root competition is intense<br />

between species in mixture, such negative interactions do not directly shape the plant<br />

community, i.e. the do not directly explain the structure of their communities.<br />

Conclusion<br />

Until recently, the shaping role of competition along gradients was generally assessed<br />

through the measurement of competitive hierarchy changes along environmental gradients<br />

(e.g. Bertness 1991; Gau<strong>de</strong>t & Keddy 1995; Costa et al. 2003; Lenssen et al. 2004;<br />

Bartelheimer et al. 2010). This approach was also used in our study. A recent and<br />

62


Chapitre I<br />

complementary approach relies on the recent <strong>de</strong>velopment of indices for measuring the<br />

importance of competition (Brooker et al. 2005; Kikvidze & Armas 2010; Seifan et al. 2010).<br />

Moreover, competition mo<strong>de</strong>ls have been <strong>de</strong>veloped that take variation among the different<br />

stages of the plant life-cycle into account (Frekleton et al. 2009; Damgaard & Fayolle 2010),<br />

which may be differently affected by competition (Howard & Goldberg 2001). This<br />

information is nee<strong>de</strong>d in or<strong>de</strong>r to predict the importance of competition on population<br />

dynamics during a period of time that is as close to reality as possible rather than in a short-<br />

term experiment.<br />

It has become clear that community ecologists must now focus on both components of<br />

competition, i.e. its intensity and importance, especially now that the novel methodological<br />

tools are available for this kind of work.<br />

Acknowledgements<br />

The authors would like to thank the persons who provi<strong>de</strong>d technical support for this study, in<br />

particular Alwin Bleomelen for his contribution toward data collection. This work is a<br />

contribution to GDR 2574 ‘TRAITS’. This study was supported by a doctoral research grant<br />

from the Ministère <strong>de</strong> l’Enseignement Supérieur et <strong>de</strong> la Recherche.<br />

63


References<br />

Chapitre I<br />

Ackerly D.D. & Cornwell W.K. 2007. A trait-based approach to community assembly:<br />

partitioning of species trait values into within- and among-community components. Ecology<br />

Letters, 10: 135-145.<br />

Bartelheimer, M., Gowing D. & Silvertown, J. 2010. Explaining hydrological niches: the<br />

<strong>de</strong>cisive role of below-ground competition in two closely related Senecio species. Journal of<br />

Ecology, 98: 126–136.<br />

Baruch, Z. 1994. Responses to drought and flooding in tropical forage grasses. Plant and Soil,<br />

164: 87–96.<br />

Bertness, M.D. 1991. Interspecific Interactions among high marsh perennials in a New<br />

England salt marsh. Ecology, 72(1): 125-137.<br />

Blom, C. & Voesenek, L. 1996. Flooding: the survival strategies of plants. Trends in Ecology<br />

& Evolution, 11: 290–295.<br />

Brooker R. & Kikividze Z. 2008. Importance: an overlooked concept in plant interaction<br />

research. Journal of Ecology, 96: 703-708.<br />

Brooker, R., Kikvidze, Z., Pugnaire, F.I., Callaway, R.M., Choler, P., Lortie, C.J. & Michalet,<br />

R. 2005. The importance of importance. Oikos, 109: 63-70.<br />

Carlyle, C. N., Fraser, L.H. & Turkington, R. 2010. Using three pairs of competitive indices<br />

to test for changes in plant competition un<strong>de</strong>r different resource and disturbance levels.<br />

Journal of Vegetation Science, 21: 1–10.<br />

Colmer, T.D. & Voesenek, L. 2009. Flooding tolerance: suites of plant traits in variable<br />

environments. Functional Plant Biology, 36: 665–681.<br />

Corcket, E., Liancourt, P., Callaway, R.M. & Michalet, R. 2003. The relative importance of<br />

competition for two dominant grass species as affected by environmental manipulations in the<br />

field. Ecoscience, 10: 186–194.<br />

Cornelissen, J.H.C., Lavorel, S., Garnier, E., Diaz, S., Buchmann, N., Gurvich, D.E., Reich,<br />

P.B., Steege, H., Morgan, H.D., Van Der Heij<strong>de</strong>n, M.G.A. & others. 2001. A handbook of<br />

protocols for standardised and easy measurement of plant functional traits worldwi<strong>de</strong>.<br />

Australian Journal of Botany, 51: 335–380.<br />

Costa, C.S., Marangoni, J.C. & Azevedo, A.M. 2003. Plant zonation in irregularly floo<strong>de</strong>d<br />

salt marshes: relative importance of stress tolerance and biological interactions. Journal of<br />

Ecology, 91: 951–965.<br />

Damgaard, C. & Fayolle, A. 2010. Measuring the importance of competition: a new<br />

formulation of the problem. Journal of Ecology, 98: 1–6.<br />

Elmendorf, S.H. & Moore, K.A. 2007. Plant competition varies with community composition<br />

in an edaphically complex landscape. Ecology, 88(10): 2640–2650.<br />

Freckleton, R.P., Watkinson, A.R. & Rees, M. 2009. Measuring the importance of<br />

competition in plant communities. Journal of Ecology, 97: 379–384.<br />

Garnier, E., Cortez, J., Billès, G., Navas, M.L., Roumet, C., Debussche, M., Laurent, G.,<br />

Blanchard, A., Aubry, A., Bellmann, A., Neill, C. & Toussaint, J.P. 2004. Plant functional<br />

markers capture ecosystem properties during secondary succession. Ecology, 85(9): 2630-<br />

2637.<br />

Gaucherand, S., Liancourt, P. & Lavorel, S. 2006. Importance and intensity of competition<br />

along a fertility gradient and across species. Journal of Vegetation Science, 17: 455–464.<br />

64


Chapitre I<br />

Gau<strong>de</strong>t, C.L. & Keddy, P.A. 1995. Competitive performance and species distribution in<br />

shortline plant communities: a comparative approach. Ecology, 76: 280–291.<br />

Gerry, A.K. & Wilson, S.D. 1995. The Influence of initial size on the competitive responses<br />

of six plant species. Ecology, 76(1): 272-279.<br />

Greiner La Peyre, M.K., Grace, J.B., Hahn, E. & Men<strong>de</strong>lssohn, I.A. 2001. The importance of<br />

competition in regulating plant species abundance along a salinity gradient. Ecology, 82: 62–<br />

69.<br />

Gross, N., Liancourt, P., Choler, P., Suding, K.N. & Lavorel, S. 2010. Strain and vegetation<br />

effects on local limiting resources explain the outcomes of biotic interactions. Perspectives in<br />

Plant Ecology, Evolution and Systematics, 12: 9–19.<br />

Hedges L.V., Gurevitch J. & Curtis P.S. 1999. The meta-analysis of response ratios in<br />

experimental ecology. Ecology, 80: 1150-1156.<br />

Howard, T. G. & Goldberg, D. E. 2001. Competitive response hierarchie for germination,<br />

growth, and survival and their influence on abundance. Ecology, 82(4): 979–990.<br />

Johansson, M.E. & Keddy, P.A. 1991. Intensity and asymmetry of competition between plant<br />

pairs of different <strong>de</strong>grees of similariy: an experimental study on two guilds of wetland plants.<br />

Oikos, 60: 27-34.<br />

Jung, V., Mony, C., Hoffmann, L. & Muller, S. 2009. Impact of competition on plant<br />

performances along a flooding gradient: a multi-species experiment. Journal of Vegetation<br />

Science, 20: 433–441.<br />

Keddy, P.A. 1990. Competitive hierarchies and centrifugal organization in plant communities.<br />

pp. 265-89: In J. Grace and D. Tilman (eds.) Perspectives on Plant Competition, Aca<strong>de</strong>mic<br />

Press, New York.<br />

Kikvidze, Z. & Brooker, R. 2010. Towards a more exact <strong>de</strong>finition of the importance of<br />

competition – a reply to Freckleton et al. (2009). Journal of Ecology, 98: 719–724.<br />

Lamb, E.G. & Cahill, F.J. 2008. When Competition Does Not Matter: Grassland Diversity<br />

and Community Composition. The American Naturalist, 171(6): 777-787.<br />

Lamb, E.G., Kembel, S.W. & Cahill, J.F. 2009. Shoot, but not root, competition reduces<br />

community diversity in experimental mesocosms. Journal of Ecology, 97: 155–163.<br />

Lavorel S. & Garnier E. 2002 Predicting changes in community composition and ecosystem<br />

functioning from plant traits: revisiting the Holy Grail. Functional Ecology, 16: 545- 556<br />

Lenssen, J.P.M., Ten Dolle, G.E. & Blom, C. 1998. The effect of flooding on the recruitment<br />

of reed marsh and tall forb plant species. Plant Ecology, 139: 13–23.<br />

Lenssen, J.P., van <strong>de</strong> Steeg, H.M. & <strong>de</strong> Kroon, H. 2004. Does disturbance favour weak<br />

competitors? Mechanisms of changing plant abundance after flooding. Journal of Vegetation<br />

Science, 15: 305–314.<br />

Liancourt, P., Callaway, R.M. & Michalet, R. 2005. Stress tolerance and competitiveresponse<br />

ability <strong>de</strong>termine the outcome of biotic interactions. Ecology, 86: 1611–1618.<br />

MacArthur, R.H. & Levins, R. 1967. Limiting similarity convergence and divergence of<br />

coexisting species. American Naturalist, 101: 377-387.<br />

McGill, B.J., Enquist, B.J., Weiher, E. & Westoby, M. 2006. Rebuilding community ecology<br />

from functional traits. Trends in Ecology & Evolution, 21: 178–185.<br />

Mommer, L., Pe<strong>de</strong>rsen, O. & Visser, E.J.W. 2004. Acclimation of a terrestrial plant to<br />

submergence facilitates gas exchange un<strong>de</strong>r water. Plant, Cell and Environment, 27: 1281–<br />

1287.<br />

65


Chapitre I<br />

Mommer, L., Lenssen, J.P., Huber, H., Visser, E.J. & D.E. Kroon, H. 2006. Ecophysiological<br />

<strong>de</strong>terminants of plant performance un<strong>de</strong>r flooding: a comparative study of seven plant<br />

families. Journal of Ecology, 94: 1117–1129.<br />

Oksanen, J., Kindt, R., Legendre, P., O'Hara, B., Simpson, GL., Solymos, P., Stevens, MHH.<br />

& Wagner, H. 2008. Vegan: Community Ecology Package. R package version 1.15-1.<br />

http://cran.r-project.org/, http://vegan.r-forge.r-project.org<br />

Perkins, T.A., Holmes, W.R. & Weltzin, J.F. 2007. Multi-species interactions in competitive<br />

hierarchies: New methods and empirical test. Journal of Vegetation Science, 18: 685–692.<br />

Podani, J. & Schmera, D. 2006. On <strong>de</strong>ndrogram-based measures of functional diversity. Oikos<br />

115: 179-185.<br />

Rea<strong>de</strong>r, R. J., Wilson, S. D., Belcher, J. W. et al. 1994. Plant competition in relation to<br />

neighbor biomass: an intercontinental study with Poa pratensis. Ecology 75: 1753-1760.<br />

Resetarits, W.J. 1995. Limiting similarity and the intensity of competitive effects on the<br />

mottled sculpin, Cottus bairdi, in experimental stream communities. Oecologia, 104: 31–38.<br />

Seifan, M., Seifan, T. & Tielbörger, K. 2010. Facilitating an importance in<strong>de</strong>x. Journal of<br />

Ecology, 98: 356–361.<br />

Shipley, B., Keddy, P. A. & Lefkovitch, L. P. 1991. Mechanism producing plant zonation<br />

along a water <strong>de</strong>pth gradient: a comparison with the exposure gradient. Canadian Journal of<br />

Botany, 69: 1420-1424.<br />

Suding, N. K., Goldberg, D.E. & Hartman, K.M. 2003. Relationships among species traits:<br />

separating levels of response and i<strong>de</strong>ntifying linkages to abundance. Ecology, 84: 1–16.<br />

Tilman, D. 1988. Plant strategies and the dynamics and structure of plant communities.<br />

Princeton University Press, Princeton, New Jersey, USA.<br />

Violle, C., Cu<strong>de</strong>nnec, C., Plantgenest, M., Damgaard, C., Le Cœur, D., Bouzillé, J.B. &<br />

Bonis, A. 2007. Indirect assessment of flooding duration as a driving factor of plant diversity<br />

in wet grasslands. Proceedings of Symposium S7 held during the Seventh IAHS Scientific<br />

Assembly at Foz do Iguacu, Brazil,. IAHS Publication 303, 2006, pp. 334<br />

Violle, C., Pu, Z. & Jiang, L. 2010. Experimental <strong>de</strong>monstration of the importance of<br />

competition un<strong>de</strong>r disturbance. Proceedings of the National Aca<strong>de</strong>my of Sciences, 107:<br />

12925-12929.<br />

Violle, C., Bonis, A., Plantegenest, M., Cu<strong>de</strong>nnec, C., Damgaard, C., Marion, B., Le Cœur, D.<br />

& Bouzillé, J.B. 2011. Plant functional traits capture species richness variations along a<br />

flooding gradient. Oikos, 120(3): 389–398.<br />

Wang, P., Stieglitz, T., Zhou, D.W. & Cahill, J.F. 2010. Are competitive effect and response<br />

two si<strong>de</strong>s of the same coin, or fundamentally different? Functional Ecology, 24: 196–207.<br />

Weiher, E., Werf, A., Thompson, K., Ro<strong>de</strong>rick, M., Garnier, E. & Eriksson, O. 1999.<br />

Challenging Theophrastus: a common core list of plant traits for functional ecology. Journal<br />

of Vegetation Science, 10: 609–620.<br />

Wel<strong>de</strong>n, C.W. & Slauson, W.L. 1986. The intensity of competition versus its importance: an<br />

overlooked distinction and some implications. Quarterly Review of Biology, 61: 23–44.<br />

Wilson, S.D. & Tilman, D. 1995. Responses of eight old-field plant species in four<br />

environments. Ecology, 76(4): 1169-1180.<br />

Wilson, S.D. 2007. Competition, resources, and vegetation during 10 years in native<br />

grassland. Ecology, 88: 2951–2958.<br />

66


Chapitre I<br />

Zhang, J., Cheng, G., Yu, F., Kräuchi, N. & Li, M.H. 2008. Intensity and importance of<br />

competition for a grass (Festuca rubra) and a legume (Trifolium pratense) vary with<br />

environmental changes. Journal of Integrative Plant Biology, 50: 1570–1579.<br />

Supporting information<br />

Appendix S1. Species survival along the flooding gradient without competitors<br />

Appendix S2. Species survival along the flooding gradient with competitors<br />

Appendix S3. Competitive hierarchy along the flooding gradient<br />

Supporting information<br />

Appendix S1<br />

Patterns of species survival without a competitor indicated that survival <strong>de</strong>creased with an<br />

increase in water <strong>de</strong>pth and duration for ten species. The survival rate was constant along the<br />

experimental treatments for M. pulegium and G. fluitans.<br />

67


Appendix S2<br />

Chapitre I<br />

Species survival was reduced in the presence of a competitor (A. stolonifera), especially for L.<br />

perenne, C. divisa, H. secalinum and T. fragiferum.<br />

Appendix S3<br />

The competitive hierarchy changed across the hydrological treatments: the species position<br />

varied between the treatments. In the non-floo<strong>de</strong>d treatments, the hierarchy of the species<br />

position was as follows: T. fragiferum, C. divisa, C. cristatus, J. gerardii, M. pulegium, L.<br />

perenne, G. fluitans, L. autumnalis, H. secalinum, B. perennis, A. stolonifera and J.<br />

articulatus.<br />

In the 10-S treatment, the hierarchy was: C. divisa, G. fluitans, B. perennis, J. articulatus, T.<br />

fragiferum, L. autumnalis, M. pulegium, A. stolonifera, H. secalinum, J. gerardii, L. perenne<br />

In the 10-L treatment: T. fragiferum, L. autumnalis, M. pulegium, G. fluitans, J. articulatus, J.<br />

gerardii, A. stolonifera.<br />

In the 20-S treatment, the hierarchy was: M. pulegium, G. fluitans, L. autumnalis, C. divisa,<br />

H. secalinum, J. gerardii, A. stolonifera, B. perennis, J.articulatus, T. fragiferum, L. perenne<br />

In the 20-L treatment: L. autumnalis, G. fluitans, T. fragiferum, M. pulegium, J. gerardii, J.<br />

articulatus, H. secalinum, A. stolonifera.<br />

68


- Chapitre 2 -<br />

Fundamental niche versus Realized niche: Assessment of the<br />

importance of competition along a flooding gradient.<br />

Part 1: Fundamental niches of twelve wetland species in response to the<br />

variation of flooding conditions.<br />

En préparation<br />

71


Chapitre II<br />

Fundamental niche versus Realized niche: Assessment of the importance of<br />

competition along a flooding gradient.<br />

Part 1: Fundamental niches of twelve wetland species in response to the<br />

variation of flooding conditions.<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1,2 , Jan-Bernard Bouzillé 1 , François Mesléard 2,3 & Anne Bonis 1<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, F-35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc, F-13200, France<br />

3 UMR CNRS-IRD 6116 Institut Méditerranéen d’Ecologie et <strong>de</strong> Paléoécologie, <strong>Université</strong><br />

d’Avignon IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09, France<br />

Abstract:<br />

Stress is a fuzzy notion in literature concerning its characterization and its<br />

quantification: biomass measurements are often used to <strong>de</strong>fine the stress but they still appear<br />

insufficient. In this study, survival and biomass data are paired to quantify the intensity of the<br />

stress at individual level, i.e. strain, along a flooding gradient and then <strong>de</strong>termine whether<br />

fundamental niches of species, presenting different distribution patterns along the gradient,<br />

are contrasted.<br />

An in situ removal experiment was run along a flooding gradient: twelve wetland<br />

species were grown at 6 locations along the gradient and in monocultures. Analyses of<br />

survival and biomass patterns using the log response ratio (strain), GLM and Kaplan-Meier<br />

analysis, allow a precise <strong>de</strong>scription of fundamental niches. Shoot elongation and Specific<br />

Leaf Area (SLA) are used to appreciate species responses to the range of conditions.<br />

Fundamental niches differed significantly amongst species along the flooding gradient.<br />

Two kinds of constraints were <strong>de</strong>scribed: the increase in duration of flooding (for 9 species)<br />

and the absence of flooding (for 2 species). Only one species was not affected by the range of<br />

environmental conditions. Species success was generally more related to the length of shoot<br />

than to SLA.<br />

Our methods pairing survival and biomass data, few used in literature, quantified<br />

precisely the strain and the fundamental niche of species along the flooding gradient, useful to<br />

un<strong>de</strong>rstand species abundance patterns in natura.<br />

Key-words: strain, survival, biomass, SLA, shoot elongation<br />

73


Chapitre II<br />

74


Introduction<br />

Chapitre II<br />

The concept of stress is not precise and leads to some <strong>de</strong>bates and controversies<br />

between authors (Elmendorf & Moore 2007), reflecting the difficulty to comprehend it: the<br />

discussion between Körner (2003, 2004) and Lortie (2004) is certainly the most well known.<br />

For the former, the term “stress” is excessively used. He disagrees with the i<strong>de</strong>a that life<br />

conditions are often stressful for organisms, expressed as the “imperfect world” by Lortie et<br />

al. (2004). The lack of precision in literature on the study scale (species versus community)<br />

and on the kind of stress (resource versus non-resource) may be responsible for this<br />

confusion.<br />

In spite of the proposition of Wel<strong>de</strong>n & Slauson (1986) to differentiate the “stress” at the<br />

community level and the “strain” at the species level, this distinction was used seldom since<br />

Suding et al. (2003) followed by Gross et al. (2010). Their proposition to quantify the strain<br />

using controlled experiments un<strong>de</strong>r highly favorable conditions represents an original way to<br />

<strong>de</strong>termine the constraint at the species level and then characterize more precisely fundamental<br />

niches of species, quantitatively comparable between each others. Moreover, the concept of<br />

niche was poorly used these last years, although so essential to un<strong>de</strong>rstand processes acting on<br />

species success (Silvertown 2004). Körner’s and Lortie’s points of view may thus be tested<br />

through the quantification of strain.<br />

Along a flooding gradient, two hydrological stresses may be observed at the species<br />

level. In lower parts of the gradient, waterlogging which is a non-resource stress, leads to a<br />

reduction of oxygen diffusion in the root zone and may be paired with submergence affecting<br />

light quality and quantity available for species; whereas drought, a resource stress, occurs in<br />

higher parts of the gradient (Blom & Voesenek 1996; Braendle & Crawford 1999; Casanova<br />

& Brock 2000; Barber et al. 2004; Bartelheimer et al. 2010).<br />

Species responses to the variations of hydrological conditions are studied a lot. The<br />

production of aerenchyma and of adventitious roots was largely <strong>de</strong>scribed: these<br />

morphological changes facilitate gas exchanges among individuals un<strong>de</strong>r oxygen-<strong>de</strong>ficient<br />

conditions (Wample & Reid 1975; Laan et al. 1989; Baruch 1994; Blom & Voesenek 1996;<br />

Grimoldi 1999; Bouma et al. 2000; Visser et al. 2000; Insausti et al. 2001). Besi<strong>de</strong>s the<br />

difficulty to measure these morphological changes, these responses are observed along a wi<strong>de</strong><br />

range of hydrological conditions, i.e. from short durations of waterlogged conditions to <strong>de</strong>ep<br />

and prolonged submerged conditions (Colmer & Voesenek 2009): they are not sufficient to<br />

distinguish and explain the effects of waterlogging and submergence on species<br />

75


Chapitre II<br />

performances. In addition, other plant traits are well known to express species adaptations to<br />

variations of hydrological conditions: the shoot elongation and the Specific Leaf Area (SLA)<br />

(Blom & Voesenek 1996; Loreti & Oesterheld 1996; Clevering 1998; Lenssen et al. 1998;<br />

Insausti et al. 2001; Mommer et al. 2004; Voesenek et al. 2004; Violle et al. 2011), which<br />

respond strongly to submergence and allow the distinction between the effects of waterlogged<br />

and floo<strong>de</strong>d conditions (Colmer & Voesenek 2009).<br />

Functional traits can be used to predict species performance, measured either by<br />

biomass production, survival or reproduction. However, these plant traits were frequently<br />

related to vegetative biomass (Wright & Westoby 2001; Huber et al. 2009).When SLA and<br />

plant height are related to vegetative biomass, it is not easy to disentangle a response of a<br />

variation of hydrological conditions to an allometric relationship among individuals. On the<br />

other hand, patterns of survival, representing the first measure of species tolerance to a<br />

constraint (Lenssen et al. 1998; Lenssen et al. 2004; Mommer et al. 2006), are poorly used<br />

because of the difficulty to follow plant across time (Violle & Jiang, 2009). Nevertheless, the<br />

relationship between plant survival and plant traits is essential to un<strong>de</strong>rstand how variations of<br />

traits act on species success and thus consi<strong>de</strong>r plant traits as surrogate of species performances<br />

(Violle & Jiang 2009).<br />

The objective of this study is to <strong>de</strong>termine precisely the fundamental niches of twelve<br />

wetland species characterized by in situ different abundance patterns along a flooding<br />

gradient. Few studies have sought to characterize finely fundamental niches of species by<br />

quantifying the environmental gradient and using both different measures of performance that<br />

are the analyses of survival and strain patterns though biomass measurements. SLA and shoot<br />

elongation will be measured to test whether they can be substituted to species performances.<br />

We aimed particularly at answering the following questions:<br />

1) Do species, characterized by contrasted in situ patterns, present contrasted fundamental<br />

niches along the flooding gradient?<br />

2) Are SLA and shoot elongation linked to species success appreciated by plant survival<br />

and plant biomass?<br />

76


Material and methods<br />

Plant material<br />

Chapitre II<br />

Twelve wetland species, collected from the floo<strong>de</strong>d meadow of the Marais Poitevin on<br />

the French Atlantic coast (46°28’N; 1°13’W), were selected according to their distribution<br />

patterns along an elevation gradient <strong>de</strong>scribing a flooding gradient, along which duration of<br />

flooding and water <strong>de</strong>pth vary (Violle et al. 2007) (Table 1). Three compartments were<br />

<strong>de</strong>scribed along the elevation gradient: the higher parts which were never floo<strong>de</strong>d and<br />

characterized by the presence of mesophilous species Carex divisa, Cynosurus cristatus,<br />

Hor<strong>de</strong>um secalinum and Lolium perenne; the intermediate slopes enduring a variable duration<br />

of flooding during some weeks per year <strong>de</strong>pending on climate and characterized by meso-<br />

hygrophilous species: Juncus gerardii, Bellis perennis and Leontodon autumnalis; and the<br />

lower parts of the gradient where flooding occurred up to 6 months per year and characterized<br />

by hygrophilous species: Glyceria fluitans, Juncus articulatus, Mentha pulegium and<br />

Trifolium fragiferum. One species presented a large distribution along the overall gradient<br />

(Agrostis stolonifera).<br />

Table 1: Position and maximal abundance of the twelve studied species along the flooding<br />

gradient and their family.<br />

Species Family Position along Maximal<br />

the gradient abundance (%)<br />

Agrostis stolonifera Poaceae All elevations 100<br />

Bellis perennis Asteraceae Slopes 15<br />

Carex divisa Cyperaceae Higher elevations 50<br />

Cynosurus cristatus Poaceae Higher elevations 22<br />

Glyceria fluitans Poaceae Lower elevations 91<br />

Hor<strong>de</strong>um secalinum Poaceae Higher elevations 22<br />

Juncus articulatus Juncaceae Lower elevations 56<br />

Juncus gerardii Juncaceae Slopes 67<br />

Leontodon autumnalis Asteraceae Slopes 5<br />

Lolium perenne Poaceae Higher elevations 65<br />

Mentha pulegium Lamiaceae Lower elevations 52<br />

Trifolium fragiferum Fabaceae Lower elevations 41<br />

A precise characterization of the waterlogging constraint experienced by plants was<br />

neglected before the work of Gowing et al. (1998) <strong>de</strong>veloping the Sum Exceedance Value for<br />

aeration (SEV for aeration): this quantification was realized in this study (See appendix). The<br />

77


Chapitre II<br />

range of SEV for aeration observed along the studied topographical sequence was from 0 to<br />

8.27cm.day -1 .<br />

Assessment of species’ tolerance to flooding<br />

Field removal and transplant experiment<br />

To <strong>de</strong>termine the range of tolerance to hydrological conditions, a removal experiment<br />

was established in September 2008 before the beginning of flooding season and was run until<br />

June 2009, in the floo<strong>de</strong>d grasslands located in the Marais Poitevin. At 6 locations along a<br />

topographical sequence, three plots were created by a removal of vegetation and by a cutting<br />

of vegetation and roots around the edge of plots, repeated regularly to avoid re-growth of<br />

vegetation. Within each plot, 2 individuals of each species were transplanted randomly and<br />

spaced by 20 cm to avoid any interaction. Each transplant was transplanted at similar size and<br />

life stage. A total of 432 individuals were transplanted into the field experiment.<br />

Measurements of the elevation of plots using a Trimble (Trimble M3,Ohio, USA) were<br />

realized in or<strong>de</strong>r to <strong>de</strong>termine the intensity of the constraint experienced by transplants at each<br />

location related to flooding duration and water <strong>de</strong>pth (Violle et al. 2007). From the location at<br />

the highest part of the gradient to the location at the lowest part of the gradient, we quantified<br />

the following range of SEV for aeration: 0, 0.44, 0.77, 1.42, 4.86 and 8.27cm.day -1<br />

(corresponding to 0, 0.031, 0.054, 0.1, 0.341 and 0.579m.week -1 ). Moreover for the two<br />

lowest locations, submergence occurred with a <strong>de</strong>pth of 5cm and 18cm respectively.<br />

Highly favorable controlled conditions<br />

To measure the stress at the individual level, i.e. the strain sensu Wel<strong>de</strong>n & Slauson<br />

(1986), target of the 12 species were grown un<strong>de</strong>r highly favorable controlled conditions<br />

(experimental gar<strong>de</strong>n, <strong>Rennes</strong>, France). For 10 species, highly favorable conditions<br />

correspond to non limiting water conditions without the presence of interspecific competition.<br />

In contrast, G. fluitans and J. articulatus are hygrophilous species that need long period of<br />

flooding during their growth cycle. However, G. fluitans tolerates high water level compared<br />

to J. articulatus (<strong>Merlin</strong> et al. in prep). Thus, even for a long period of flooding, treatment of<br />

20cm of water was chosen as a reference for G. fluitans and treatment of 10 cm of water to J.<br />

articulatus. One tiller of each species per pot (3L), with a potting soil with a mixture of<br />

78


Chapitre II<br />

gar<strong>de</strong>n soil and compost, was grown early March 2009 and run until July 2009 with 6<br />

replicates per species. Each transplant was transplanted at similar size and life stage.<br />

Plant performance survey<br />

For the field experiment, individuals’ survival was surveyed all along the growing<br />

season, i.e. from mid-February to the end of the experiment at the end of May:<br />

presence/absence of each individual was noted at 0, 20, 29, 36, 42, 50, 57, 71 and 98 days.<br />

Survival was noted, and individual aboveground biomass was harvested after 9 months of<br />

experiment for the field experiment and after 4 months of experiment for the controlled<br />

experiment. Aboveground biomass was dried for 72 h at 60°C before weighing.<br />

Measurements of plant functional traits<br />

Two functional traits were measured to study species response to hydrological<br />

conditions: Specific Leaf Area (SLA) and shoot length. They were measured at the end of the<br />

experiment for each living individual at each location of the field experiment. Following the<br />

standardized protocols (Weiher et al. 1999; Cornelissen et al. 2001), SLA was measured on<br />

the youngest fully grown leaf in the light using a leaf area meter (Li-Cor, Lincoln, Nebraska,<br />

USA). Shoot length was measured as the difference between the elevation of the highest<br />

photosynthetic leaf and the base of the plant, but was not measured for Bellis perennis and<br />

Leontodon autumalis as they present a roset life-form.<br />

Data analysis<br />

The “plot effect” in the in situ experiment was tested for all species and for the 6<br />

locations: the non significant results indicated that the 6 transplants of a species transplanted<br />

at each location can be consi<strong>de</strong>red as 6 replicates (two-way ANOVA, data not shown).<br />

Survival pattern<br />

The <strong>de</strong>termination of the fundamental niches of each species along the flooding<br />

gradient was realized using GLM regression for binary data at the end of the experiment. Chi-<br />

square tests from analysis of variance were realized to test significance of regressions.<br />

Moreover, to assess flooding tolerance of each species, we estimated the median lethal time<br />

79


Chapitre II<br />

LT50, i.e. the number of days after which 50% of individuals had died (Lenssen et al. 2004;<br />

Van Eck et al. 2004; Mommer et al. 2006). The Kaplan-Meier survival analysis followed by<br />

log-rank test was realized to test the patterns of survival of individual species between the 6<br />

locations along the 10 weeks of survey during the flooding period. Such a survival dynamics<br />

analysis across time is poorly used in community ecology (Ozinga et al. 2007), but is essential<br />

to characterize species response patterns. In this analysis, un<strong>de</strong>r the null hypothesis, i<strong>de</strong>ntical<br />

survival patterns along time were expected across the different hydrological conditions, i.e.<br />

between locations.<br />

Measure of strain using biomass data<br />

Following the proposition of Gross et al. (2010), strain intensity was approached by<br />

the comparison of species growth un<strong>de</strong>r non-limiting conditions without competitors in the<br />

controlled experiment, with growth without competitors in the in situ experiment. The log<br />

response ratio was calculated following this formula:<br />

lnRR (strain) = ln (individual biomass production in the field without competitors/mean<br />

biomass production un<strong>de</strong>r non-limiting conditions)<br />

A negative value indicates that target species performance is <strong>de</strong>pressed compared with non-<br />

limiting conditions whereas a positive value indicates increased performances of species<br />

un<strong>de</strong>r the studied conditions.<br />

One-way ANOVA was realized to test the effect of the hydrological conditions on the<br />

species aboveground biomass production, followed by post-hoc Tukey HSD tests after a log-<br />

transformation to verify the hypotheses of normality and homogeneity of variance. A two-<br />

way ANOVA was performed to test the effect of the different topographical locations and the<br />

effect of species on strain intensity after the verification of normality and homogeneity of<br />

variance.<br />

Linear regressions were realized between each measure of performance and intensity<br />

of strain for each species to investigate whether strain affects species performances (survival<br />

and biomass).<br />

80


Relationship between plant traits and species success<br />

Chapitre II<br />

To <strong>de</strong>termine whether plant traits were related to species performances, linear<br />

regressions were performed between SLA, shoot length and survival in a first time, and<br />

between SLA, shoot length and aboveground biomass in a second time. For these analyses,<br />

functional traits data were transformed to fit the hypotheses of normality and homogeneity of<br />

variance.<br />

Statistical tests were calculated with the software R (R version 2.7.2., 2008, R<br />

Foundation for Statistical Computing, Vienna, Austria).<br />

Results<br />

Different ranges of tolerance to hydrological conditions<br />

Effect of hydrological conditions on survival<br />

Four responses groups were <strong>de</strong>scribed concerning the patterns of survival along the<br />

gradient estimated by the GLM (Fig. 1). The first group correspon<strong>de</strong>d to a complete mortality<br />

of individuals at the end of the experiment at one or two floo<strong>de</strong>d locations of the experiment:<br />

two species were inclu<strong>de</strong>d in this group, Cynosurus cristatus and Bellis perennis. In<strong>de</strong>ed,<br />

Kaplan-Meier analysis <strong>de</strong>monstrated that 0 and 20 days after the beginning of the growing<br />

season, half of the transplanted individuals were already <strong>de</strong>ad for these floo<strong>de</strong>d locations<br />

(Table 2). The second group concerned species survival patterns characterized by a reduction<br />

of survival probability in response to the increase of hydrological conditions tested by Chi-<br />

square tests, but without a complete mortality at the end of the experiment: Carex divisa,<br />

Lolium perenne, Hor<strong>de</strong>um secalinum, Juncus gerardii, Leontodon autumnalis and Trifolium<br />

fragiferum. Their survival patterns across the flooding period were significantly different for<br />

the last location characterized by <strong>de</strong>ep flooding for these species, while Leontodon autumnalis<br />

and Trifolium fragiferum showed early median lethal time intervene for the two floo<strong>de</strong>d<br />

locations (Table 2). The third group was only composed of one species, Mentha pulegium,<br />

characterized by an equivalent probability of survival all along the gradient. The last group<br />

was composed of 3 species presenting a reduction of survival probability in the highest parts<br />

of the gradient: Agrostis stolonifera, Glyceria fluitans and Juncus articulatus, confirmed by<br />

81


Chapitre II<br />

differential survival patterns among locations <strong>de</strong>monstrated for Juncus articulatus and<br />

Glyceria fluitans (Table 2).<br />

Fig 1: Fundamental niches of species predicted by GLM in response to the variation of<br />

hydrological conditions, quantified through the measure of the SEV for aeration (from 0<br />

cm.day -1 to 8.27cm.day -1 ), with the results of the chi-square tests indicating the <strong>de</strong>gree of<br />

significance of regressions.<br />

82


Chapitre II<br />

Table 2: Number of days after which 50% of individuals had died, the LT50, at the 6<br />

topographical locations (i<strong>de</strong>ntified from the driest (1) to the wettest (6)) from mid-February,<br />

<strong>de</strong>termined by Kaplan-Meier analysis followed by log-rank tests. In bold: significance at<br />

P


Chapitre II<br />

Table 3: Results of two-way ANOVA testing the effect of species, the topographical locations<br />

and their interaction on the values of strain.<br />

Df Mean Square F value P-value<br />

Species 11 24.71 28.06 < 2.2e-16<br />

Topographical locations 5 6.62 7.52 1.34e-6<br />

Species*Topographical locations 52 5.01 5.69 < 2.2e-16<br />

Residuals 254 0.88<br />

Fig 2: Comparisons of strain intensity (ln RR(strain)) between topographical locations for<br />

each species. Statistical analyses for each location (i<strong>de</strong>ntified from the driest (1) to the wettest<br />

(6)) were performed in<strong>de</strong>pen<strong>de</strong>ntly. Results of ANOVA, P-values and Tukey HSD post-hoc<br />

tests are indicated (different letters indicated a significant difference).<br />

The intensity of strain was significantly related to the survival for the majority of<br />

species (P


Relationship between plant traits and plant success<br />

Chapitre II<br />

Returning the response groups inferred from patterns of strain intensity along the<br />

gradient of flooding, we noted that for species of group 4 (G. fluitans and J. articulatus), the<br />

length of the stem was positively related to their survival pattern (G. fluitans: r²=0.838,<br />

P=0.007; J. articulatus: r²=0.956, P=0.001) and the amount of biomass (G. fluitans: r²=0.297,<br />

P=0.001; J. articulatus: r²=0.575, P=6.82e-6). Survival of species from the second group (L.<br />

perenne, M. pulegium and T. fragiferum) was positively related to the length of stem<br />

(respectively r²=0.907, P=0.002; r²=0.771, P=0.014; r²=0.983, P=7.04e-5); biomass<br />

production of M. pulegium was related to SLA (r²= 0.1956, P=0.012). Success of species from<br />

the first group (survival and biomass production) (B. perennis, H. secalinum and L.<br />

autumnalis) was not related to SLA and stem length except for the survival pattern of H.<br />

secalinum, related to SLA (r²=0.613, P=0.040) and stem length (r²=0.862, P=0.005). For the<br />

group 3 consisting A. stolonifera, J. gerardii, C. divisa and C. cristatus, it was more difficult<br />

to generalize the results. Success, especially survival of A. stolonifera and C. divisa were<br />

related to SLA (r²=0.73, P=0.019; r²=0.627, P=0.036) and length stem (respectively r²=0.975,<br />

P=0.0001; r²=0.771, P=0.013). Survival of J. gerardii was positively related to SLA<br />

(r²=0.860, P=0.005); biomass of C. cristatus related to length of stem (r²=0.369, P=0.002).<br />

Discussion<br />

Few studies have attempted to <strong>de</strong>fine precisely the fundamental niches of species by<br />

quantifying the environmental gradient studied and combining survival and biomass data.<br />

Biomass is the measure of performance preferentially used to <strong>de</strong>fine the <strong>de</strong>gree of the<br />

tolerance to a constraint. Even if the propositions from Suding et al. (2003) and Gross et al.<br />

(2010) represent an interesting approach to quantifying the strain from biomass data allowing<br />

comparisons between species (see the part: Contrasted fundamental niches along the flooding<br />

gradient: habitats’ preferences), it is necessary to combine this information with species<br />

survival patterns, the first measure of tolerance to a constraint (Lenssen et al. 1998; Lenssen et<br />

al. 2004; Mommer et al. 2006). In this way, species growth strategies expressed to tolerate the<br />

constraint can be approximated.<br />

85


Contrasted fundamental niches along the flooding gradient: habitat preferences<br />

Chapitre II<br />

Two kinds of strain were emphasized along the flooding gradient. Then species<br />

presented different fundamental niches along the flooding gradient. Moreover, each species<br />

experiences and tolerates the hydrological conditions differently resulting in habitat<br />

preferences (Shipley 1991; Baruch 1994; Jung et al. 2009; Bartelheimer et al. 2010). The<br />

strain is thus species-specific (Körner 2003; Liancourt et al. 2005; Gross et al. 2010).<br />

Interestingly, our results support Lortie et al. (2004)’s point of view: individuals growing in<br />

their natural habitat are limited by abiotic environmental conditions compared to the optimal<br />

conditions, a result also found by Gross et al. (2010). This indicates that in natura, individual<br />

growth is not only limited by the flooding conditions, but perhaps also by the effect of<br />

flooding on resources.<br />

The increase of flooding conditions represents the constraint for nine species, for<br />

which a reduction in survival and/or in aboveground biomass is observed (Baruch 1994;<br />

Grimoldi 1999; Fraser & Karnesis 2005). Among these species, different <strong>de</strong>grees of tolerance<br />

are observed. Two species, Bellis perennis and Cynosurus cristatus, are intolerant to<br />

submergence with a complete mortality of individuals whatever the <strong>de</strong>pth of flooding. For the<br />

other seven species (Carex divisa, Hor<strong>de</strong>um secalinum, Lolium perenne, Juncus gerardii,<br />

Leontodon autumnalis, Mentha pulegium and Trifolium fragiferum), performances are<br />

reduced with the increase of flooding conditions: the range of tolerance is wi<strong>de</strong> for these<br />

species compared to the two previous species, as mortality was not total in the floo<strong>de</strong>d parts<br />

of the gradient. But these seven species characterized by similar responses curves along the<br />

gradient were constrained differently by flooding. In<strong>de</strong>ed, the analysis of the median lethal<br />

time informs precisely on the pattern of survival across time. Among these species,<br />

Leontodon autumnalis and Trifolium fragiferum tolerate less the submergence than the other<br />

species as their mortality occurred more quickly and for the two floo<strong>de</strong>d locations whereas<br />

mortality occurred only for the last location for the other species. In addition, Carex divisa,<br />

Hor<strong>de</strong>um secalinum, Juncus gerardii, Lolium perenne and Mentha pulegium tolerate longer<br />

flooding with 10 cm high water around than flooding with higher water level. As low water<br />

<strong>de</strong>pth implies shorter duration of flooding compared to high water <strong>de</strong>pth (Casanova & Brock<br />

2000), Leontodon autumnalis and Trifolium fragiferum tolerate (less exten<strong>de</strong>d) flooding, an<br />

effect of duration of flooding <strong>de</strong>monstrated on other species (Fraser & Karnesis 2005).<br />

The second strain is represented by the absence of flooding and is found to reduce<br />

survival and aboveground biomass of Glyceria fluitans and Juncus articulatus. These species<br />

86


Chapitre II<br />

clearly show a preference for floo<strong>de</strong>d habitats. In<strong>de</strong>ed, un<strong>de</strong>r hydrological conditions lower<br />

than the 4cm.day -1 threshold from which submergence occurred, survival of both species<br />

<strong>de</strong>creases, <strong>de</strong>monstrating the low tolerance of the absence of submergence.<br />

Only one species, Agrostis stolonifera, shows a large tolerance to the hydrological<br />

conditions (large fundamental niche) without negative effects on survival and biomass<br />

production. The pattern of strain reveals that the range of hydrological conditions, i.e. from<br />

low anoxia to <strong>de</strong>ep flooding has an equal intensity on its performances.<br />

Species strategy along the gradient<br />

Pairing biomass and survival analyses allow the <strong>de</strong>termination of species strategy, and<br />

especially stress-tolerance strategy (Grime, 1977), <strong>de</strong>fined as a constant survival probability<br />

all along the gradient and a reduction of biomass production in stressful parts of the gradient<br />

resulting in a strategy of resources conservation. Moreover, the species strategy appears<br />

essential in the outcome of competition (Choler et al. 2001; Corcket et al. 2003; Liancourt et<br />

al. 2005; Maestre et al. 2009; Bowker et al. 2010). In<strong>de</strong>ed these performances may behave<br />

differently along the gradient (Van Eck et al. 2004; Fraser & Karnesis 2005; Mommer et al.<br />

2006). It is the case for Mentha pulegium which presents contrasted patterns of survival and<br />

biomass along the range of hydrological conditions <strong>de</strong>scribing a stress-tolerance strategy.<br />

Shoot length more related to species success than SLA<br />

The success of the species measured by their survival and their biomass indicates that<br />

the tolerance of species to the range of hydrological conditions along the gradient may be<br />

largely explained by variations of shoot length rather than by variations of SLA.<br />

For species having a preference for floo<strong>de</strong>d habitats, i.e. Glyceria fluitans and Juncus<br />

articulatus, the positive relationship between species success and shoot length was expected<br />

and <strong>de</strong>monstrated. In<strong>de</strong>ed the shoot elongation is a response to flooding in or<strong>de</strong>r to improve<br />

gas diffusion among individuals (Blom & Voesenek 1996; Loreti & Oesterheld 1996;<br />

Clevering 1998; Lenssen et al. 1998; Insausti et al. 2001; Mommer et al. 2004; Voesenek et<br />

al. 2004; Mommer et al. 2006). This increase in shoot elongation is not paired with an<br />

increase in SLA, a response also known to improve gas exchanges (Blom & Voesenek 1996;<br />

Mommer et al. 2004; Mommer et al. 2006). This may be un<strong>de</strong>rstood by two in<strong>de</strong>pen<strong>de</strong>nt<br />

clusters in response to flooding expressed by species <strong>de</strong>monstrated by Mommer et al. (2006):<br />

species can either respond to flooding by an increase in traits in relation to stem such as the<br />

87


Chapitre II<br />

increase in stem elongation and in aerenchyma proportion, or by leaf traits such as the<br />

variation in SLA and in chlorophyll content.<br />

The positive relationship between shoot length and survival is also observed for<br />

species having preferences for dry habitats, i.e. where the intensity of the anoxia is low, as<br />

Lolium perenne or Mentha pulegium. In<strong>de</strong>ed in this part of this gradient, water resource is<br />

available during the beginning of the growing season, this high availability of resources allow<br />

a good growth of species. For similar reasons, the successes of Carex divisa, Cynosurus<br />

cristatus, Hor<strong>de</strong>um secalinum and Juncus gerardii in dry habitats are related to SLA and to<br />

shoot length: the simultaneous variations of SLA and shoot length reflect an efficient strategy<br />

of resource use (Weiher et al. 1999; Jung et al. 2010), with a high ability of capture of<br />

resources, especially light and space, for high competitive species (Gau<strong>de</strong>t & Keddy 1995).<br />

Interestingly, species responses to floo<strong>de</strong>d and dry habitats are similar i.e. the<br />

elongation of the stem, even if the causalities of these responses are different. The increase<br />

may be due on the one hand to the availability of resource for species preferring dry habitats,<br />

and on the other hand to the increase of flooding conditions for species preferring floo<strong>de</strong>d<br />

habitats.<br />

Conclusion<br />

Our study helps precisely characterizing the fundamental niches of twelve species<br />

using the quantification of strain and the analysis of survival and biomass patterns according<br />

to a range of hydrological conditions, compared to other studies <strong>de</strong>monstrating fundamental<br />

niches of wetlands species (Fraser & Karnesis 2005). Two constraints are <strong>de</strong>monstrated along<br />

the gradient: the increase of the duration of flooding and the absence of flooding, which<br />

differently affect species. Moreover, the species success related to the tolerance to the<br />

constraint may be expressed by the plant height, thus consi<strong>de</strong>red as a surrogate to species<br />

performances along the flooding gradient.<br />

The knowledge of fundamental niches is essential for the un<strong>de</strong>rstanding of processes<br />

acting on species distribution along the gradient and especially the importance of competition,<br />

which will be studied in the second part of this paper.<br />

88


Literature cited<br />

Chapitre II<br />

Barber, K.R., Leeds-Harrison, P.B., Lawson, C.S. & Gowing, D.J.G. (2004) Soil aeration<br />

status in a lowland wet grassland. Hydrological Processes, 18, 329–341.<br />

Bartelheimer, M., Gowing, D. & Silvertown, J. (2010) Explaining hydrological niches: the<br />

<strong>de</strong>cisive role of below-ground competition in two closely related Senecio species. Journal of<br />

Ecology, 98, 126–136.<br />

Baruch, Z. (1994) Responses to drought and flooding in tropical forage grasses. Plant and<br />

Soil, 164, 87–96.<br />

Blom, C. & Voesenek, L. (1996) Flooding: the survival strategies of plants. Trends in<br />

Ecology & Evolution, 11, 290–295.<br />

Bowker, M.A., Soliveres, S. & Maestre, F.T. (2010) Competition increases with abiotic stress<br />

and regulates the diversity of biological soil crusts. Journal of Ecology, 98, 551–560.<br />

Braendle, R. & Crawford, R.M. (1999) Plants as amphibians. Perspectives in Plant Ecology,<br />

Evolution and Systematics, 2, 56–78.<br />

Casanova, M.T. & Brock, M.A. (2000) How do <strong>de</strong>pth, duration and frequency of flooding<br />

influence the establishment of wetland plant communities? Plant Ecology, 147, 237–250.<br />

Choler, P., Michalet, R. & Callaway, R.M. (2001) Facilitation and competition on gradients in<br />

alpine plant communities. Ecology, 82, 3295–3308.<br />

Clevering, O.A. (1998) Effects of litter accumulation and water table on morphology and<br />

productivity of Phragmites australis. Wetlands Ecology and Management, 5, 275–287.<br />

Colmer, T.D. & Voesenek, L. (2009) Flooding tolerance: suites of plant traits in variable<br />

environments. Functional Plant Biology, 36, 665–681.<br />

Cornelissen, J.H.C., Lavorel, S., Garnier, E., Diaz, S., Buchmann, N., Gurvich, D.E., Reich,<br />

P.B., Steege, H., Morgan, H.D., Van Der Heij<strong>de</strong>n, M.G.A. & others. (2003) A handbook of<br />

protocols for standardised and easy measurement of plant functional traits worldwi<strong>de</strong>.<br />

Australian Journal of Botany, 51, 335–380.<br />

Fonseca, C.R., Overton, J.M., Collins, B. & Westoby, M. (2000) Shifts in trait-combinations<br />

along rainfall and phosphorus gradients. Journal of Ecology, 88, 964–977.<br />

Fraser, L.H. & Karnezis, J. P. (2005) A comparative assessment of seedling survival and<br />

biomass accumulation for fourteen wetland plant species grown un<strong>de</strong>r minor water-<strong>de</strong>pth<br />

differences. Wetlands, 25, 520-530.<br />

Goldberg, D.E., Rajaniemi, T., Gurevitch, J. & Stewart-Oaten, A. (1999) Empirical<br />

approaches to quantifying interaction intensity: competition and facilitation along<br />

productivity gradients. Ecology, 80, 1118–1131.<br />

Gowing, D.J.G., Youngs, E.G., Gilbert, J.C. & Spoor G. (1998). Predicting the effect of<br />

change in water regime on plant communities. In Hydrology in a Changing Environment, vol.<br />

I, Wheater H, Kirby C (eds). John Wiley: Chichester; 473–483.<br />

Grime, J.P. (1977) Evi<strong>de</strong>nce for the existence of three primary strategies in plants and its<br />

relevance to ecological and evolutionary theory. American naturalist, 111, 1169–1194.<br />

Grimoldi, A.A., Inausti, P., Roitman, G.G. & Soriano, A. (1999) Responses to flooding<br />

intensity in Leontodon taraxacoi<strong>de</strong>s. New Phytologist, 141, 119-128.<br />

89


Chapitre II<br />

Gross, N., Liancourt, P., Choler, P., Suding, K.N. & Lavorel, S. (2010) Strain and vegetation<br />

effects on local limiting resources explain the outcomes of biotic interactions. Perspectives in<br />

Plant Ecology, Evolution and Systematics, 12, 9–19.<br />

Huber, H., Jacobs, E. & Visser, E.J. (2009) Variation in flooding-induced morphological traits<br />

in natural populations of white clover (Trifolium repens) and their effects on plant<br />

performance during soil flooding. Annals of Botany, 103, 377.<br />

Insausti, P., Grimoldi, A.A., Chaneton, E.J. & Vasellati, V. (2001) Flooding induces a suite of<br />

adaptive plastic responses in the grass Paspalum dilatatum. New Phytologist, 152, 291–299.<br />

Jung, V., Mony, C., Hoffmann, L. & Muller, S. (2009) Impact of competition on plant<br />

performances along a flooding gradient: a multi-species experiment. Journal of Vegetation<br />

Science, 20, 433–441.<br />

Körner, C. (2003) Limitation and stress-always or never? Journal of Vegetation Science, 14,<br />

141–143.<br />

Körner, C. (2004) Individuals have limitations, not communities—A response to Marrs,<br />

Weiher and Lortie et al. Journal of Vegetation Science, 15, 581–582.<br />

Laan, P., Berrevoets, M.J., Lythe, S., Armstrong, W. & Blom, C. (1989) Root morphology<br />

and aerenchyma formation as indicators of the flood-tolerance of Rumex species. The journal<br />

of ecology, 77, 693–703.<br />

Lenssen, J.P.M., Ten Dolle, G.E. & Blom, C. (1998) The effect of flooding on the recruitment<br />

of reed marsh and tall forb plant species. Plant Ecology, 139, 13–23.<br />

Lenssen, J.P., van <strong>de</strong> Steeg, H.M. & <strong>de</strong> Kroon, H. (2004) Does disturbance favour weak<br />

competitors? Mechanisms of changing plant abundance after flooding. Journal of Vegetation<br />

Science, 15, 305–314.<br />

Liancourt, P., Callaway, R.M. & Michalet, R. (2005) Stress tolerance and competitiveresponse<br />

ability <strong>de</strong>termine the outcome of biotic interactions. Ecology, 86, 1611–1618.<br />

Lortie, C.J., Brooker, R.W., Kikvidze, Z. & Callaway, R.M. (2004) The value of stress and<br />

limitation in an imperfect world: A reply to K\örner. Journal of Vegetation Science, 15, 577–<br />

580.<br />

Maestre, F.T., Callaway, R.M., Valladares, F. & Lortie, C.J. (2009) Refining the stressgradient<br />

hypothesis for competition and facilitation in plant communities. Journal of Ecology,<br />

97, 199–205.<br />

Mommer, L., Pe<strong>de</strong>rsen, O. & Visser, E.J.W. (2004) Acclimation of a terrestrial plant to<br />

submergence facilitates gas exchange un<strong>de</strong>r water. Plant, Cell & Environment, 27, 1281–<br />

1287.<br />

Mommer, L., Lenssen, J.P., Huber, H., Visser, E.J. & De Kroon, H. (2006) Ecophysiological<br />

<strong>de</strong>terminants of plant performance un<strong>de</strong>r flooding: a comparative study of seven plant<br />

families. Journal of Ecology, 94, 1117–1129.<br />

Ozinga, W.A., Hennekens, S.M., Schaminée, J.H., Smits, N.A., Bekker, R.M., Römermann,<br />

C., Klimeš, L., Bakker, J.P. & Groenendael, J.M. (2007) Local above-ground persistence of<br />

vascular plants: Life-history tra<strong>de</strong>-offs and environmental constraints. Journal of Vegetation<br />

Science, 18, 489–497.<br />

Shipley, B., Keddy, P.A. & Lefkovitch, L.P. (1991) Mechanisms producing plant zonation<br />

along a water <strong>de</strong>pth gradient: a comparison with the exposure gradient. Canadian Journal of<br />

Botany, 69, 1420–1424.<br />

90


Chapitre II<br />

Silvertown, J. (2004) Plant coexistence and the niche. Trends in Ecology & Evolution, 19,<br />

605–611.<br />

Suding, N. K., Goldberg, D.E. & Hartman, K.M. (2003) Relationships among species traits:<br />

separating levels of response and i<strong>de</strong>ntifying linkages to abundance. Ecology, 84, 1–16.<br />

Van Eck, W., Van <strong>de</strong> Steeg, H.M., Blom, C. & De Kroon, H. (2004) Is tolerance to summer<br />

flooding correlated with distribution patterns in river floodplains? A comparative study of 20<br />

terrestrial grassland species. Oikos, 107, 393–405.<br />

Violle, C., Bonis, A., Plantegenest, M., Cu<strong>de</strong>nnec, C., Damgaard, C., Marion, B., Le Coeur,<br />

D. & Bouzillé, J.B. (2011) Plant functional traits capture species richness variations along a<br />

flooding gradient. Oikos, 120(3), 389-398.<br />

Violle, C. & Jiang, L. (2009) Towards a trait-based quantification of species niche. Journal of<br />

Plant Ecology, 2, 87.<br />

Voesenek, L., Rijn<strong>de</strong>rs, J., Peeters, A.J.M., Van <strong>de</strong> Steeg, H.M. & De Kroon, H. (2004) Plant<br />

hormones regulate fast shoot elongation un<strong>de</strong>r water: from genes to communities. Ecology,<br />

85, 16–27.<br />

Wample, R.L. & Reid, D.M. (1975) Effect of aeration on the flood-induced formation of<br />

adventitious roots and other changes in sunflower (Helianthus annuus L.). Planta, 127, 263–<br />

270.<br />

Weiher, E., Werf, A., Thompson, K., Ro<strong>de</strong>rick, M., Garnier, E. & Eriksson, O. (1999)<br />

Challenging Theophrastus: a common core list of plant traits for functional ecology. Journal<br />

of Vegetation Science, 10, 609–620.<br />

Wel<strong>de</strong>n, C.W. & Slauson, W.L. (1986) The intensity of competition versus its importance: an<br />

overlooked distinction and some implications. Quarterly Review of Biology, 61, 23–44.<br />

Wright, I.J. & Westoby, M. (2001) Un<strong>de</strong>rstanding seedling growth relationships through<br />

specific leaf area and leaf nitrogen concentration: generalisations across growth forms and<br />

growth irradiance. Oecologia, 127, 21–29.<br />

91


Chapitre II<br />

92


- Chapitre 3 -<br />

Fundamental niche versus Realized niche: Assessment of the<br />

importance of competition along a flooding gradient.<br />

Part 2: Tra<strong>de</strong>-off between competitive ability and tolerance to stress at the species<br />

level driving species distribution along a flooding gradient<br />

En préparation<br />

93


Chapitre III<br />

Fundamental niche versus Realized niche: Assessment of the importance of<br />

competition along a flooding gradient.<br />

Part 2: Tra<strong>de</strong>-off between competitive ability and tolerance to stress at the<br />

species level driving species distribution along a flooding gradient<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1,2 , Jan-Bernard Bouzillé 1 , François Mesléard 2,3 & Anne Bonis 1<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, F-35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc, F-13200, France<br />

3 UMR CNRS-IRD 6116 Institut Méditerranéen d’Ecologie et <strong>de</strong> Paléoécologie, <strong>Université</strong><br />

d’Avignon IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09, France<br />

Abstract<br />

The variation in the importance of competition is still poorly un<strong>de</strong>rstood along<br />

environmental gradients. In this study, we examined the role of competition across a flooding<br />

gradient using the niche concept, with the hypothesis of a tra<strong>de</strong>-off between competitiveability<br />

of species and the tolerance to stress responsible for species distribution along the<br />

gradient.<br />

This work involved twelve wetland species presenting different levels of flooding<br />

tolerance. Comparisons between fundamental and realized niches of species, predicted by<br />

generalized mo<strong>de</strong>ls, were realized to assess the competitive ability of species and <strong>de</strong>termine<br />

which species are submitted to the competitive exclusion. The importance of competition was<br />

assessed using the in<strong>de</strong>x of competition importance Cimp.<br />

Five species were exclu<strong>de</strong>d in the stressed parts of the gradient with a displacement of<br />

their niche’s optima observed, whereas no optima displacement was observed for six other<br />

species suggesting a high competitive ability. Only one species presented a large niche width<br />

in presence of neighbors all along the flooding gradient. At individual level, the importance of<br />

competition was related to the intensity of the strain experienced for six species.<br />

The tra<strong>de</strong>-off between competitive ability of species and the tolerance to stress at<br />

species level is responsible for species distribution and their abundance patterns along the<br />

flooding gradient, with the competition importance predicted by the intensity of the<br />

constraint. This experiment shows that niche concept is useful for un<strong>de</strong>rstanding the assembly<br />

rules of communities along a constraint gradient and provi<strong>de</strong>d testable hypothesis.<br />

Key-words: response curves, niche shift, optima displacement, Cimp<br />

95


Chapitre III<br />

96


Introduction<br />

Chapitre III<br />

The study of the competition importance is crucial to <strong>de</strong>termine the roles of the abiotic<br />

and biotic filters in structuring plant communities. However its assessment along<br />

environmental gradients is seldom studied and still poorly un<strong>de</strong>rstood (Goldberg 1994;<br />

Sammul et al. 2000; Greiner La Peyre et al. 2001; Corcket et al. 2003; Gaucherand et al.<br />

2006; Zhang et al. 2008; Jung et al. 2009; Damgaard & Fayolle 2010).<br />

Many studies have been conducted to <strong>de</strong>termine whether the distribution of plant<br />

species along environmental gradients resulted from a tra<strong>de</strong>-off between the competitive<br />

abilities of species and their tolerance to stress (Keddy 1989; Wisheu & Keddy 1992; Lenssen<br />

et al. 1999; Brose & Tielbörger 2005; Jung et al. 2009), with species highly competitive<br />

excluding less competitive from more stressful areas and then increasing their tolerance to<br />

stress. In contrast, very few studies have inclu<strong>de</strong>d the concepts of intensity and importance of<br />

competition relative to this tra<strong>de</strong>-off, thus limiting the un<strong>de</strong>rstanding of their patterns of<br />

variation along gradients. By <strong>de</strong>fining the intensity of competition as a process and the<br />

importance of the competition as a product (Wel<strong>de</strong>n & Slauson 1986), we un<strong>de</strong>rstand that<br />

competitive exclusion of species, allowing the <strong>de</strong>scription of the tra<strong>de</strong>off, represents the<br />

product of the interaction between species, corresponding to the process.<br />

Few studies reported the competitive exclusion of species, viewable by a niche shift.<br />

The <strong>de</strong>monstration of this tra<strong>de</strong>-off requires the knowledge of width of fundamental niches<br />

(Sommer & Worm 2002), however less <strong>de</strong>termined (Aerts 1999; Silvertown 2004). In<strong>de</strong>ed,<br />

the use of niche concept represents an interesting approach to study the importance of<br />

competition along environmental gradients. The <strong>de</strong>viation of species distribution from their<br />

physiological range of tolerance in response to interspecific competition can assess<br />

competition importance at species distribution limits (Wel<strong>de</strong>n & Slauson 1986; Jung et al.<br />

2009). Thus, only a small effect of competition on niche positions is expected among high<br />

competitive species (Parrish & Bazzaz 1982). Moreover, a reduction of competition<br />

importance with the increase of stress is expected, i.e. at each si<strong>de</strong> of the physiological<br />

optimum (Jung et al. 2009), but competition importance may be different between each si<strong>de</strong><br />

inducing an asymmetric species’ distribution along the gradient (Austin 1990).<br />

Different fundamental niches of species may be <strong>de</strong>scribed along a flooding gradient<br />

(see part 1): as a consequence, the reduction of performances by competition <strong>de</strong>pends on<br />

species and on locations along the gradient (Jung et al. 2009). In addition, the knowledge of<br />

the fundamental niches is <strong>de</strong>cisive in the outcome of interspecific competition because this<br />

97


Chapitre III<br />

outcome may <strong>de</strong>pend of individual responses (Greiner la Peyre et al. 2001; Choler et al. 2001;<br />

Bruno et al. 2003; Corcket et al. 2003; Liancourt et al. 2005; Gaucherand et al. 2006): this<br />

information is essential to extrapolate the result of tra<strong>de</strong>-off from the species level to the<br />

community level aggregating responses of all species having various ranges of tolerance<br />

(Greiner la Peyre et al. 2001). In<strong>de</strong>ed, this tra<strong>de</strong>-off at the species level may not be observed<br />

at the community level (Violle et al. 2010). Consequently, these different ranges of tolerance<br />

imply different levels of competition importance which may explain the difficulty to evaluate<br />

the variation of competition importance along environmental gradients.<br />

The use of niche concept has to be paired with a precise characterization of the<br />

environmental factor acting strongly on species distribution, which is a complex logistic,<br />

especially when a resource gradient coinci<strong>de</strong>s with a stress gradient (Kotowski et al. 2006):<br />

disentangling the effects of competition from physical factors is essential to un<strong>de</strong>rstand<br />

communities’ distribution. In<strong>de</strong>ed, communities’ distribution may come from species<br />

segregation according to their competitive ability, from distinct physiological preference for<br />

habitats, or from an interaction between species responses to both physical and resource<br />

gradients (Emery et al. 2001; Kotowski et al. 2006).<br />

The aim of this present study is to <strong>de</strong>termine the relative importance of flooding<br />

conditions and competition in the distribution of species along a flooding gradient, through<br />

the measure of the in<strong>de</strong>x of competition importance (Brooker et al. 2005). We hypothesize,<br />

first, the existence of a tra<strong>de</strong>-off between competitive abilities and tolerance to stress, driving<br />

species ecological success along the flooding gradient. Second, there should be an exclusion<br />

of less competitive species from the higher stressed parts of the flooding gradient, as the<br />

product of strong competitive interactions in the lesser stressed parts of the gradient.<br />

To respond to this objective, the concept of niche will be used through the comparison<br />

of fundamental and realized niches of species, informing then on the competitive ability of<br />

species measured as the <strong>de</strong>gree of niche <strong>de</strong>viation. Moreover, the <strong>de</strong>termination of the<br />

individual strain (Suding et al. 2003; Gross et al. 2010) representing an interesting way to<br />

quantify stress at the species level (see part 1), will allow comparing the importance of<br />

competition between species according to the intensity of strain experimented (Brooker et al.<br />

2005; Gaucherand et al. 2006).<br />

98


Material and methods<br />

Study site<br />

Chapitre III<br />

The study site is a flood meadow located in the Marais Poitevin on the French Atlantic<br />

coast (46°28’N; 1°13’W) characterized by a mild Atlantic climate. The study area is<br />

characterized by a micro-topographical gradient, with an elevation range of approximately 40<br />

cm, with floo<strong>de</strong>d <strong>de</strong>pressions, intermediate slopes and higher level flats resulting in a range of<br />

flooding intensity, frequency and duration (Violle et al. 2011). Three plant communities have<br />

been distinguished on these three topographical compartments (Tourna<strong>de</strong> & Bouzillé 1995).<br />

The higher level flats are never floo<strong>de</strong>d: the mesophilous community is dominated by Lolium<br />

perenne, Agrostis stolonifera, Elymus repens, Cynosurus cristatus and Hor<strong>de</strong>um secalinum.<br />

The hygrophilous community dominated by Agrostis stolonifera, Glyceria fluitans, Oenanthe<br />

fistulosa and Eleocharis palustris in the floo<strong>de</strong>d <strong>de</strong>pressions are submitted to 6 months long<br />

flooding. The intermediate slopes experience a variable duration of flooding <strong>de</strong>pending on<br />

climate during some weeks per year, dominated by Juncus gerardii, Hor<strong>de</strong>um marinum,<br />

Alopecurus bulbosus.<br />

The productivity of the study site is stable along the gradient: flooding intensity did<br />

not affect biomass production, so nutrients availability and flooding factors are not<br />

confoun<strong>de</strong>d (Violle et al. 2011).<br />

Floristic samplings<br />

To <strong>de</strong>termine species distribution according to hydrological conditions in the presence<br />

of interspecific competition, 266 floristic samplings were realized randomly in paddocks<br />

managed by mo<strong>de</strong>rate grazing. Within each quadrate (25 cm*25 cm), the presence of species<br />

was noted. As for the part 1, measurements of the elevation of plots with a Trimble (Trimble<br />

M3, Ohio, USA) were realized in or<strong>de</strong>r to <strong>de</strong>termine the intensity of the constraint<br />

experienced by species following the proposition of Gowing et al. (1998) (See appendix).<br />

99


Data analysis<br />

Assessment of species competitive ability<br />

Chapitre III<br />

The <strong>de</strong>gree of displacement of the ecological (realized) optima from the fundamental<br />

optima sensu Austin (1990) was assessed to <strong>de</strong>termine the existence of a tra<strong>de</strong>-off between<br />

competitive ability and tolerance to stress at the species level. Following this hypothesis of<br />

competitive hierarchy, a coinci<strong>de</strong>nce of fundamental and realized optima is expected for high<br />

competitive species, whereas a displacement of optima is expected for low competitive<br />

species (competitive exclusion) (Fig. 1).<br />

Fig 1: A. The response of a high competitive species to the presence of neighbors, i.e. a<br />

reduction of performance without <strong>de</strong>viation of optima; B. The response of a less competitive<br />

species to the presence of neighbors, i.e. a competitive exclusion observable through the<br />

displacement of the ecological optima (adapted from Austin 1990).<br />

Measure of competition importance<br />

Following the prediction of fundamental niches realized with GLM in the part 1 of the<br />

paper, the probability of occurrence of each species in presence of competitors along the<br />

gradient was predicted from binary data. The Akaike Information Criterion (AIC) was used to<br />

select the better fitting mo<strong>de</strong>l between quadratic GLM and GAM (Thuiller et al. 2003; Austin<br />

et al. 2009). In<strong>de</strong>ed a <strong>de</strong>bate exists concerning the prediction of response curves with GLM<br />

100


Chapitre III<br />

and with GAM: better fitting of asymmetric response curves were expected with GAM<br />

(Guisan et al. 1999; Austin 2009; Heikkinen & Makipaa 2010). Bootstrap procedure though a<br />

re-sampling of residuals was realized for each regression to <strong>de</strong>termine whether parameters of<br />

mo<strong>de</strong>ls were well predicted because the number of observations per species was low. Chi-<br />

square tests from analysis of variance were realized to test significance of regressions.<br />

The in<strong>de</strong>x of competition importance Cimp (Brooker et al. 2005) was used to quantify<br />

the importance of competition relative to the impact of all others factors in the environment.<br />

This in<strong>de</strong>x was calculated with the probability of occurence data for each value of SEV for<br />

aeration <strong>de</strong>scribed at the 6 locations along the gradient, following this formula:<br />

Cimp= (Occcomp-Occtarget)/(MaxOcc-y)<br />

Occcomp is the probability of occurence in presence of competitors, Occtarget is the probability<br />

of occurence without competitors, MaxOcc is the maximum value of the probability of<br />

occurence without competitors along the gradient and y is the smaller of either Occcomp or<br />

Occtarget. The more Cimp tends to -1, the more important is the contribution of competition<br />

against the hydrological conditions; whereas the more Cimp tends to +1, the more important is<br />

the contribution of abiotic factor relative to competition.<br />

Linear regressions were performed to test whether the importance of competition was<br />

related to the intensity of strain.<br />

Characterization of niche shape<br />

To <strong>de</strong>termine whether an interaction between hydrological conditions and competition<br />

occurs and was more important for one part of the gradient, the skewness ( ) was calculated<br />

to characterize niche shape of realized niches (Heikkinen & Makipaa, 2010). It was calculated<br />

for a threshold equal to e -1/2 of the ecological optimum of niche, threshold <strong>de</strong>termined by<br />

previous mo<strong>de</strong>ling works (Heegaard 2002; Heikkinen & Makipaa 2010). However, this<br />

calculation can be applied only for non-truncated responses curves. Skewness gives a value<br />

between +1 and -1: a negative value indicates that niche is larger for low values of the<br />

hydrological conditions whereas a positive value indicates that niche is larger for high values<br />

of the hydrological conditions.<br />

Statistical tests were computed with R software (R Development Core Team 2008),<br />

with the boot and GAM packages.<br />

101


Results<br />

Niche shifts and <strong>de</strong>viation of optima in response to diffuse competition<br />

Chapitre III<br />

The range of hydrological conditions experienced by transplants was only 1/4 long<br />

(from 0 to 8.27 cm.day -1 ) (see part 1) compared with the range along the various sites where<br />

the floristic samplings were realized (from 0 to 36.92cm.day -1 ). The quantification of the<br />

importance of competition was possible only on this 1/4 range of hydrological conditions.<br />

Ecological responses curves were better predicted by GAM than by quadratic GLM<br />

because of low AIC values (see Appendix): the bootstrap procedure indicated reliable<br />

predictions because estimators were reasonably unbiased and the bootstrap standard errors<br />

smaller (Table 2).<br />

Table 2: Results of bootstrapping realized on residuals of simple GLM for fundamental niches<br />

and on residuals of GAM for realized niches, indicating unbiased estimators of regressions<br />

paired with small standard errors.<br />

Fundamental niches Realized niches<br />

Parameters Intercept ~SEVa Intercept ~SEVa<br />

Bias Std. error Bias Std. error Bias Std. error Bias Std. error<br />

A. stolonifera -0.063 0.084 0.024 0.032 -0.004 0.105 0.0002 0.006<br />

B.perennis -0.010 0.068 0.004 0.026 0.149 0.071 -0.009 0.004<br />

C. cristatus -0.078 0.102 0.030 0.039 0.059 0.083 -0.003 0.005<br />

C. divisa -0.090 0.162 0.034 0.062 -0.263 0.118 0.015 0.007<br />

G. fluitans -0.017 0.134 0.007 0.051 0.022 0.108 -0.001 0.006<br />

H. secalinum -0.046 0.144 0.018 0.055 -0.011 0.108 0.0006 0.006<br />

J. articulatus -0.039 0.125 0.015 0.048 0.219 0.140 -0.013 0.008<br />

J. gerardii -0.073 0.150 0.028 0.057 -0.068 0.127 0.004 0.007<br />

L. autumnalis -0.109 0.111 0.041 0.042 0.120 0.055 -0.007 0.003<br />

L. perenne -0.067 0.151 0.025 0.058 -0.201 0.112 0.011 0.006<br />

M. pulegium 0.013 0.156 -0.005 0.059 -0.063 0.108 0.004 0.006<br />

T. fragiferum -0.071 0.160 0.027 0.061 -0.070 0.126 0.004 0.007<br />

For overall species, interspecific competition induced niche shifts with some<br />

displacements of niche optima. A succession of ecological optima was observed along the<br />

overall range of hydrological conditions, and four response groups were <strong>de</strong>scribed (Fig 3).<br />

The first group was characterized by species presenting a probability of occurrence<br />

strongly reduced for Juncus gerardii, Bellis perennis, Leontodon autumnalis, Mentha<br />

pulegium and Trifolium fragiferum. On the range of 0 to 8.27cm.day -1 , the ecological optima<br />

of these species were displaced to lower elevations of the gradient for the hydrological<br />

102


Chapitre III<br />

conditions superior to 8.27cm.day -1 : ecological optima were equal to 17, 13, 19, 24 and<br />

20cm.day -1 respectively. Among these species, the characterization of the skewness was<br />

possible only for the non-truncated ecological niches of Bellis perennis, Leontodon<br />

autumnalis, Mentha pulegium and Trifolium fragiferum. For the two first species, the<br />

ecological niche was larger in the higher elevations of the gradient ( =-0.11 and =-0.45<br />

respectively) whereas the ecological niche of the two last species was larger for the lower<br />

elevations of the gradient ( =0.14 and =0.27 respectively).<br />

Fig. 3: Realized niches of species predicted through the probability of occurrence using GAM<br />

in response to the variation of hydrological conditions and to the presence of interspecific<br />

competition with the result of the chi-square test.<br />

The second group was composed of Carex divisa, Lolium perenne, Hor<strong>de</strong>um<br />

secalinum and Cynosurus cristatus where the probability of occurrence was reduced in<br />

response to interspecific competition except for Carex divisa on the range of SEV aeration<br />

from 0 to 8.27cm.day -1 . Moreover, along the range of hydrological conditions, the success of<br />

species was greater for a SEV for aeration superior to 8.27cm.day -1 in presence of<br />

interspecific competition.<br />

103


Chapitre III<br />

The third group was composed of Glyceria fluitans and Juncus articulatus for which<br />

the ecological optima were <strong>de</strong>scribed for high values of SEV for aeration. The ecological<br />

success of these species was reduced for low values of SEV for aeration. Agrostis stolonifera<br />

composed the last group: this ecological success was high along the flooding gradient and<br />

appeared bimodal.<br />

Contribution of competition against hydrological conditions: the quantification on a short<br />

range of hydrological conditions.<br />

Through the use of the probability of occurrence from fundamental and realized niches<br />

for the 6 values of SEV for aeration corresponding to the 6 locations of the field experiment,<br />

the contribution of competition against hydrological conditions seemed to be higher (Cimp<br />

closed to -1) for Mentha pulegium, Trifolium fragiferum, Leontodon autumnalis, Bellis<br />

perennis and Juncus articulatus for which the contribution increased slowly with SEV for<br />

aeration. Cimp was closed to -1 for Bellis perennis, Leontodon autumnalis and Juncus gerardii<br />

for higher locations of the gradient and increased with SEV for aeration, especially strongly<br />

for Bellis perennis. The contribution of competition against hydrological conditions was the<br />

lowest (Cimp closed to 1) for Carex divisa.<br />

Figure 4: Pattern of competition importance, Cimp, for the twelve species in each location (1<br />

for the driest to 6 for the wettest) of the in situ experiment, calculated from the probability of<br />

occurrence of species in presence and in absence of interspecific competition. We observed<br />

that the importance of competition is generally higher for the hygrophilous and mesohygrophilous<br />

species (Cimp closed to -1).<br />

104


Chapitre III<br />

Significant regressions were <strong>de</strong>scribed between the intensity of strain and the<br />

importance of competition, except for Cynosurus cristatus, Carex divisa, Hor<strong>de</strong>um secalinum,<br />

Leontodon autumnalis and Juncus articulatus (Table 5). For Agrostis stolonifera, the<br />

competition importance was high where the intensity of strain was high. On the other hand,<br />

for Bellis perennis, Glyceria fluitans, Lolium perenne, Mentha pulegium, Trifolium<br />

fragiferum, the importance of competition was high where the strain intensity was low; a<br />

ten<strong>de</strong>ncy was <strong>de</strong>scribed for Juncus gerardii.<br />

Table 5: Significant regressions between competition importance Cimp and strain intensity for<br />

A. stolonifera, B. perennis, G. fluitans, L. perenne, M. pulegium and T. fragiferum.<br />

A. stolonifera B. perennis C. cristatus C. divisa<br />

t 3.74 -8.21 -1.44 -0.74<br />

df 4 3 2 4<br />

P-value 0.02 0.0038 0.2869 0.498<br />

cor 0.882 -0.978 -0.713 -0.348<br />

G. fluitans H. secalinum J. articulatus J. gerardii<br />

t -12.69 0.79 0.72 -2.34<br />

df 4 4 4 4<br />

P-value 0.0002 0.4759 0.512 0.079<br />

cor -0.988 0.366 0.338 -0.761<br />

L. autumnalis L. perenne M. pulegium T. fragiferum<br />

t -0.56 -3.23 -3.46 -32.07<br />

df 4 4 4 4<br />

P-value 0.607 0.032 0.026 5.63e-6<br />

cor -0.268 -0.850 -0.875 -0.998<br />

Discussion<br />

Competitive ability-strain tolerance tra<strong>de</strong>-off driving species success along the flooding<br />

gradient<br />

Interspecific competition induces a spatial segregation of species niches along the<br />

flooding gradient, particularly according to the <strong>de</strong>gree of anoxia which is the variable strongly<br />

explaining niches segregation (Silvertown et al. 1999; Silvertown 2004; Araya et al. 2010).<br />

This segregation confirms our hypothesis of a tra<strong>de</strong>-off between species competitive ability<br />

and their tolerance to strain driving species distribution along the flooding gradient with a<br />

displacement of species (Lenssen et al. 1999; Brose & Tielbörger 2005; Pennings et al. 2005;<br />

Jung et al. 2009). However all species are not driven by this tra<strong>de</strong>-off.<br />

105


Chapitre III<br />

Even if ranges of hydrological conditions were not equivalent between the<br />

topographical sequence where the field experiment was implanted and the various sites where<br />

floristic samplings were realized, it is possible to formulate hypothesis and extrapolate results<br />

on the importance of competition quantified over the short range of the hydrological<br />

conditions to the overall gradient.<br />

Importance of competition <strong>de</strong>pending on the strain intensity<br />

A niche shift is <strong>de</strong>fined as any change in the position, in the mo<strong>de</strong> and/or a reduction<br />

in the variance of the distribution of a species along a niche axis (Silvertown 2004). In our<br />

study, interspecific competition is responsible for the niche shifts and for the optima<br />

displacements for Juncus gerardii, Bellis perennis, Leontodon autumnalis, Mentha pulegium<br />

and Trifolium fragiferum. For these species, competition importance <strong>de</strong>creases with strain<br />

intensity which is in accordance with others studies (Sammul et al. 2000; Greiner La Peyre et<br />

al. 2001; Gaucherand et al. 2006; Jung et al. 2009). In other words, competition is important<br />

when environmental conditions are close to the optimal hydrological conditions of the<br />

species. Thus, these species are exclu<strong>de</strong>d from floo<strong>de</strong>d elevations of the gradient due to a low<br />

competitive ability. Moreover, at the limits of species’ distribution, i.e. at each si<strong>de</strong> of the<br />

ecological niche, the contribution of competition is not equal (Austin 1990). In<strong>de</strong>ed,<br />

interspecific competition strongly limits the species’ upper distribution of the ecological<br />

niches except for Juncus gerardii, Bellis perennis and Leontodon autumnalis whereas<br />

flooding strongly limits their distribution in the lower elevations. The distribution of previous<br />

species is thus driven by the tra<strong>de</strong>-off between competitive abilities and tolerance to stress at<br />

species level, resulting in the prediction of competition importance, <strong>de</strong>pending on the<br />

individual tolerance to the absence of flooding or to the presence of long durations of<br />

flooding.<br />

Our study only allows hypotheses to be formulated for Glyceria fluitans and Juncus<br />

articulatus. In<strong>de</strong>ed, no available data on species performances for high <strong>de</strong>gree of anoxia<br />

exists. However the absence of flooding represents the constraint (part 1). This may<br />

confirmed the relationship between competition importance and the strain intensity: in<strong>de</strong>ed,<br />

the contribution of competition in species success is high in the higher parts of the gradient<br />

for Glyceria fluitans. This relationship is not significant for Juncus articulatus: the low<br />

number of replicates may influence the result of this relationship. In opposition to previous<br />

species, no displacement of niche optima could occur in response to interspecific competition<br />

106


Chapitre III<br />

because of their preference for floo<strong>de</strong>d habitats. However, the range of environmental<br />

conditions is strongly reduced by the presence of neighbors, otably in low floo<strong>de</strong>d parts of the<br />

gradient.<br />

Enlargement of mesophilous species’ realized niches in presence of competitors<br />

No displacement of niche optima are observed for Cynosurus cristatus, Carex divisa,<br />

Hor<strong>de</strong>um secalinum and Lolium perenne in response to interspecific competition but only a<br />

reduction of the ecological success. This indicates a high competitive ability for these species<br />

(Parrish & Bazzaz 1982) which are certainly responsible for the exclusion of the previous<br />

meso-hygrophilous species and hygrophilous species (Mentha pulegium and Trifolium<br />

fragiferum).<br />

Moreover, an expected result is observed: the unrelated variations of competition<br />

importance and strain intensity, except for Lolium perenne. This may be explained by the<br />

increase realized niche width of mesophilous species compared to fundamental niches. This<br />

may suggest a facilitative effect of neighbors then increasing their range of tolerance to<br />

flooding. In<strong>de</strong>ed, following Bruno et al. (2003), in the context of niche theory, facilitation<br />

may extent the realized niche of species until this extent can be larger than the fundamental<br />

niche. Recent studies <strong>de</strong>monstrated that facilitation did not only occur in harsh environments<br />

but <strong>de</strong>pends on target species and on the kind of resources (Choler et al. 2001; Liancourt et al.<br />

2005; Maestre et al. 2009; Bowker et al. 2010). In<strong>de</strong>ed facilitation <strong>de</strong>pends on the low<br />

tolerance of species to the constraint and at the same time on their high competitive ability<br />

(Liancourt et al. 2005). In stressful environments, an amelioration of habitats by neighbors<br />

may be observed, thus facilitating less adapted species (Choler et al. 2001) but this was<br />

<strong>de</strong>monstrated for resource gradient where the availability of water was low. Nevertheless,<br />

neighbors may improve the availability of soil oxygen (Keddy 1994): plants may compete for<br />

this soil oxygen, and this proposition to consi<strong>de</strong>r the aeration as a resource stress seems to be<br />

a<strong>de</strong>quate (Bartelheimer et al. 2010), contrary to the proposition of Austin (1990, 2002)<br />

consi<strong>de</strong>ring this abiotic variable as a non-resource gradient. However, we cannot exclu<strong>de</strong> the<br />

effect of the mo<strong>de</strong>rate grazing present in this study, although grazing effect was <strong>de</strong>monstrated<br />

to restrict realized niches (Lau et al. 2008), not enlarge them. Moreover, an experimental bias<br />

cannot be exclu<strong>de</strong>d because the removal of vegetation may locally change the environmental<br />

conditions.<br />

107


The particular case of Agrostis stolonifera<br />

Chapitre III<br />

One species presents a high success along the gradient with the presence of<br />

competitors, Agrostis stolonifera: its ecological success does not seem to be driven by the<br />

competitive ability-stress tolerance tra<strong>de</strong>-off. This may be explained by a high plasticity of<br />

this species (Callaway et al. 2003), not <strong>de</strong>monstrated in this study. In<strong>de</strong>ed, plastic<br />

morphological responses of this species may reduce the effect of interspecific competition<br />

(Silvertown & Gordon, 1989). Moreover, this species <strong>de</strong>monstrated a high competitive ability<br />

after flooding events thanks to an efficient clonal growth (Lenssen et al. 2004).<br />

A possible extrapolation at the community scale<br />

To extrapolate results of the tra<strong>de</strong>-off from the species level to the community level,<br />

our results support the i<strong>de</strong>a suggesting an aggregation of species responses, which could have<br />

various tolerances to the abiotic variables (Greiner la Peyre et al. 2001). The exclusion of<br />

Juncus gerardii, Bellis perennis and Leontodon autumnalis, <strong>de</strong>scribing the meso-<br />

hygrophilous community, from stressful habitats (slopes) by high competitive species<br />

characterizing the mesophilous community (Carex divisa, Lolium perenne, Hor<strong>de</strong>um<br />

secalinum and Cynosurus cristatus) allows the extrapolation of this tra<strong>de</strong>-off at these two<br />

communities’ scale. In<strong>de</strong>ed the meso-hygrophilous community is certainly composed of a<br />

large number of subordinate species, resulting in a <strong>de</strong>crease of competition importance with<br />

the increase of the constraint. Moreover, the predicted ecological success seems to be<br />

concordant with the competitive hierarchy with subordinate presenting a weak relative<br />

abundance, especially for Leontodon autumnalis and Bellis perennis. However relative<br />

abundances of Cynosurus cristatus and Hor<strong>de</strong>um secalinum were low compared to the<br />

predicted ecological success. This may be explained by the fact that these species are more<br />

sensitive to interspecific competition than Carex divisa and Lolium perenne. In<strong>de</strong>ed, the study<br />

of competitive hierarchy among mesophilous community <strong>de</strong>monstrated that the low<br />

competitive abilities of Cynosurus cristatus and Hor<strong>de</strong>um secalinum and the high competitive<br />

abilities of other species especially Lolium perenne (Marion et al. in prep.).<br />

However, the extrapolation at the community level of the reduction of the contribution<br />

of competition with the increase of the constraint is not possible because of floo<strong>de</strong>d habitats’<br />

preference of two species: Juncus articulatus and Glyceria fluitans, explaining then their<br />

abundance patterns. Then, because of these habitats preferences, the assessment of<br />

competition importance along environmental gradients is not easy: in our study, at the<br />

108


Chapitre III<br />

community scale, the competitive ability-tolerance of stress tra<strong>de</strong>-off cannot be <strong>de</strong>monstrated<br />

(Violle et al. 2010).<br />

Conclusion<br />

The competitive ability-stress tolerance tra<strong>de</strong>-off drives species distribution along the<br />

flooding gradient resulting in a competitive hierarchy with a ranked or<strong>de</strong>r from competitive<br />

dominant species to subordinate species. The predictability of the constraint, i.e. a regular<br />

occurrence as it is the case in our study site (occurrence at each autumn and finish at the end<br />

of spring) is a prerequisite to observe the competitive ability-stress tolerance tra<strong>de</strong>-off (Keddy<br />

1989; Costa et al. 2003). In<strong>de</strong>ed, in irregular environment characterized by large variations<br />

inducing a low persistent gradient of stress across time, competitive hierarchy does not appear<br />

as a major process structuring plant communities (Costa et al. 2003). However the non<br />

persistence of stress over time is not the only reason explaining the non-observation of the<br />

tra<strong>de</strong>-off. In<strong>de</strong>ed, the high plasticity of species allowing the adaption to local abiotic and<br />

biotic conditions may reduce the effects of interspecific competition, as it the case for<br />

Agrostis stolonifera.<br />

This study <strong>de</strong>monstrated the essential role of the <strong>de</strong>termination of fundamental niches<br />

in the assessment of competition importance against environmental conditions and in the<br />

extrapolation of results from the species scale to the community scale, thus allowing to<br />

un<strong>de</strong>rstand abundance patterns and species coexistence. Responses are species specific for<br />

both the <strong>de</strong>termination of range of tolerance to the abiotic variable and for the outcome of<br />

competition. Our results could <strong>de</strong>monstrate that the variation of competition importance could<br />

not be applied generally at the community scale because of the presence of potential<br />

competitive species in each part of the flooding gradient.<br />

In spite of a short range of hydrological conditions experimented by transplants, this<br />

study helps to un<strong>de</strong>rstand community’ distribution due to a running field experiment along a<br />

realistic range of environmental conditions and to the characterization and the quantification<br />

of the abiotic variable, expressed here by the <strong>de</strong>gree of aeration, which is less realized in<br />

studies although essential (Aerts 1999).<br />

109


Literature cited<br />

Chapitre III<br />

Aerts, R. (1999) Interspecific competition in natural plant communities: mechanisms, tra<strong>de</strong>offs<br />

and plant-soil feedbacks. Journal of Experimental Botany, 50, 29-37.<br />

Austin, M.P. (1990) Community theory and competition in vegetation In: Grace, J.B. &<br />

Tilman, D. (eds.) Perspectives on plant competition. pp. 27–49. Aca<strong>de</strong>mic Press, San Diego.<br />

Austin, M.P. (2002) Spatial prediction of species distribution: an interface between ecological<br />

theory and statistical mo<strong>de</strong>lling. Ecological Mo<strong>de</strong>lling, 157, 101–118.<br />

Austin, M.P., Smith, T.M., Van Niel, K.P. & Wellington, A.B. (2009) Physiological<br />

responses and statistical mo<strong>de</strong>ls of the environmental niche: a comparative study of two cooccurring<br />

Eucalyptus species. Journal of Ecology, 97, 496–507.<br />

Barber, K.R., Leeds-Harrison, P.B., Lawson, C.S. & Gowing, D.J.G. (2004) Soil aeration<br />

status in a lowland wet grassland. Hydrological Processes, 18, 329–341.<br />

Bartelheimer, M., Gowing, D. & Silvertown, J. (2010) Explaining hydrological niches: the<br />

<strong>de</strong>cisive role of below-ground competition in two closely related Senecio species. Journal of<br />

Ecology, 98, 126–136.<br />

Blom, C. & Voesenek, L. (1996) Flooding: the survival strategies of plants. Trends in<br />

Ecology & Evolution, 11, 290–295.<br />

Bowker, M.A., Soliveres, S. & Maestre, F.T. (2010) Competition increases with abiotic stress<br />

and regulates the diversity of biological soil crusts. Journal of Ecology, 98, 551–560.<br />

Brose, U. & Tielbörger, K. (2005) Subtle differences in environmental stress along a flooding<br />

gradient affect the importance of inter-specific competition in an annual plant community.<br />

Plant Ecology, 178, 51–59.<br />

Bruno, J.F., Stachowicz, J.J. & Bertness, M.D. (2003) Inclusion of facilitation into ecological<br />

theory. Trends in Ecology & Evolution, 18, 119–125.<br />

Casanova, M.T. & Brock, M.A. (2000) How do <strong>de</strong>pth, duration and frequency of flooding<br />

influence the establishment of wetland plant communities? Plant Ecology, 147, 237–250.<br />

Choler, P., Michalet, R. & Callaway, R.M. (2001) Facilitation and competition on gradients in<br />

alpine plant communities. Ecology, 82, 3295–3308.<br />

Costa, C.S., Marangoni, J.C. & Azevedo, A.M. (2003) Plant zonation in irregularly floo<strong>de</strong>d<br />

salt marshes: relative importance of stress tolerance and biological interactions. Journal of<br />

Ecology, 91, 951–965.<br />

Damgaard, C. & Fayolle, A. (2010) Measuring the importance of competition: a new<br />

formulation of the problem. Journal of Ecology, 98, 1–6.<br />

Désilets, P. & Houle, G. (2005) Effects of resource availability and heterogeneity on the slope<br />

of the species-area curve along a floodplain-upland gradient. Journal of Vegetation Science,<br />

16, 487–496.<br />

Emery, N.C., Ewanchuk, P.J. & Bertness, M.D. (2001) Competition and salt-marsh plant<br />

zonation: stress tolerators may be dominant competitors. Ecology, 82, 2471–2485.<br />

Gau<strong>de</strong>t, C.L. & Keddy, P.A. (1995) Competitive performance and species distribution in<br />

shortline plant communities: a comparative approach. Ecology, 76, 280–291.<br />

Gilbert, J.C., Gowing, D.J. & Bullock, R.J. (2003) Influence of seed mixture and hydrological<br />

regime on the establishment of a diverse grassland sward at a site with high phosphorus<br />

availability. Restoration Ecology, 11, 424–435.<br />

110


Chapitre III<br />

Goldberg, D.E. 1990. Components of resource competition in plant communities. In: Grace,<br />

J.B. & Tilman, D. (eds.) Perspectives on plant competition. pp. 27–49. Aca<strong>de</strong>mic Press, San<br />

Diego.<br />

Goldberg, D. & Novoplansky, A. (1997) On the relative importance of competition in<br />

unproductive environments. Journal of Ecology, 85, 409–418.<br />

Gowing, D.J.G., Youngs, E.G., Gilbert, J.C. & Spoor, G. (1998). Predicting the effect of<br />

change in water regime on plant communities. In Hydrology in a Changing Environment, vol.<br />

I, Wheater H, Kirby C (eds). John Wiley: Chichester; 473–483.<br />

Graff, P., Aguiar, M.R. & Chaneton, E.J. (2007) Shifts in positive and negative plant<br />

interactions along a grazing intensity gradient. Ecology, 88, 188–199.<br />

Greiner La Peyre, M.K., Grace, J.B., Hahn, E. & Men<strong>de</strong>lssohn, I.A. (2001) The importance of<br />

competition in regulating plant species abundance along a salinity gradient. Ecology, 82, 62–<br />

69.<br />

Grime, J.P. (1977) Evi<strong>de</strong>nce for the existence of three primary strategies in plants and its<br />

relevance to ecological and evolutionary theory. American naturalist, 111, 1169–1194.<br />

Grime, J.P. (1979) Plant strategies and Vegetation processes. Wiley, Chichester, England.<br />

Gross, N., Liancourt, P., Choler, P., Suding, K.N. & Lavorel, S. (2010) Strain and vegetation<br />

effects on local limiting resources explain the outcomes of biotic interactions. Perspectives in<br />

Plant Ecology, Evolution and Systematics, 12, 9–19.<br />

Guisan, A., Weiss, S.B. & Weiss, A.D. (1999) GLM versus CCA spatial mo<strong>de</strong>ling of plant<br />

species distribution. Plant Ecology, 143, 107–122.<br />

Heegaard, E. (2002) The outer bor<strong>de</strong>r and central bor<strong>de</strong>r for species-environmental<br />

relationships estimated by non-parametric generalised additive mo<strong>de</strong>ls. Ecological Mo<strong>de</strong>lling,<br />

157, 131–139.<br />

Heikkinen, J. & Mäkipää, R. (2010) Testing hypotheses on shape and distribution of<br />

ecological response curves. Ecological Mo<strong>de</strong>lling, 221, 388–399.<br />

Hirzel, A.H. & Le Lay, G. (2008) Habitat suitability mo<strong>de</strong>lling and niche theory. Journal of<br />

Applied Ecology, 45, 1372–1381.<br />

Keddy, P.A. (1992) Assembly and response rules: two goals for predictive community<br />

ecology. Journal of Vegetation Science, 3, 157–164.<br />

Keddy, P.A., Twolan-Strutt, L. & Wisheu, I.C. (1994) Competitive effect and response<br />

rankings in 20 wetland plants: are they consistent across three environments? Journal of<br />

Ecology, 82, 635–643.<br />

Körner, C. (2004) Individuals have limitations, not communities—A response to Marrs,<br />

Weiher and Lortie et al. Journal of Vegetation Science, 15, 581–582.<br />

Körner, C. (2003) Limitation and stress-always or never? Journal of Vegetation Science, 14,<br />

141–143.<br />

Kotowski, W., Thörig, W., Diggelen, R. & Wassen, M.J. (2006) Competition as a factor<br />

structuring species zonation in riparian fens-a transplantation experiment. Applied Vegetation<br />

Science, 9, 231–240.<br />

Lau, J.A., McCall, A.C., Davies, K.F., McKay, J.K. & Wright, J.W. (2008) Herbivores and<br />

edaphic factors constrain the realized niche of a native plant. Ecology, 89, 754–762.<br />

Liancourt, P., Callaway, R.M. & Michalet, R. (2005) Stress tolerance and competitiveresponse<br />

ability <strong>de</strong>termine the outcome of biotic interactions. Ecology, 86, 1611–1618.<br />

111


Chapitre III<br />

Lenssen, J., Menting, F., van <strong>de</strong>r Putten, W. & Blom, K. (1999) Control of plant species<br />

richness and zonation of functional groups along a freshwater flooding gradient. Oikos, 86,<br />

523–534.<br />

Lortie, C.J., Brooker, R.W., Kikvidze, Z. & Callaway, R.M. (2004) The value of stress and<br />

limitation in an imperfect world: A reply to K\örner. Journal of Vegetation Science, 15, 577–<br />

580.<br />

Maestre, F.T., Callaway, R.M., Valladares, F. & Lortie, C.J. (2009) Refining the stressgradient<br />

hypothesis for competition and facilitation in plant communities. Journal of Ecology,<br />

97, 199–205.<br />

McGill, B.J., Enquist, B.J., Weiher, E. & Westoby, M. (2006) Rebuilding community ecology<br />

from functional traits. Trends in Ecology & Evolution, 21, 178–185.<br />

Mommer, L., Pe<strong>de</strong>rsen, O. & Visser, E.J.W. (2004) Acclimation of a terrestrial plant to<br />

submergence facilitates gas exchange un<strong>de</strong>r water. Plant, Cell & Environment, 27, 1281–<br />

1287.<br />

Parrish, J. A. D. & Bazzaz, F. A. (1982) Competitive interactions in plant communities.<br />

Ecology ,63, 314–320.<br />

Pennings, S.C. & Callaway, R.M. (1992) Salt marsh plant zonation: the relative importance of<br />

competition and physical factors. Ecology, 73, 681–690.<br />

Pennings, S.C., Grant, M.B. & Bertness, M.D. (2005) Plant zonation in low-latitu<strong>de</strong> salt<br />

marshes: disentangling the roles of flooding, salinity and competition. Journal of Ecology, 93,<br />

159–167.<br />

Pollock, M.M., Naiman, R.J. & Hanley, T.A. (2008) Plant species richness in riparian<br />

wetlands—a test of biodiversity theory.<br />

Shipley, B., Keddy, P.A. & Lefkovitch, L.P. (1991) Mechanisms producing plant zonation<br />

along a water <strong>de</strong>pth gradient: a comparison with the exposure gradient. Canadian Journal of<br />

Botany, 69, 1420–1424.<br />

Silvertown, J., Dodd, M.E., Gowing, D.J. & Mountford, J.O. (1999) Hydrologically <strong>de</strong>fined<br />

niches reveal a basis for species richness in plant communities. Nature, 400, 61–63.<br />

Silvertown, J. (2004) Plant coexistence and the niche. Trends in Ecology & Evolution, 19,<br />

605–611.<br />

Sommer, U. & Worm, B. (2002) Competition and coexistence. Springer-Verlag Berlin<br />

Hei<strong>de</strong>lberg New York. ISBN 3-540-43311-2.<br />

Suding, N. K., Goldberg, D.E. & Hartman, K.M. (2003) Relationships among species traits:<br />

separating levels of response and i<strong>de</strong>ntifying linkages to abundance. Ecology, 84, 1–16.<br />

Swetnam, R.D., Mountford, J.O., Armstrong, A.C., Gowing, D.J.G., Brown, N.J.,<br />

Manchester, S.J. & Treweek, J.R. (1998) Spatial relationships between site hydrology and the<br />

occurrence of grassland of conservation importance: a risk assessment with GIS. Journal of<br />

Environmental Management, 54, 189–203.<br />

Tilman, D. 1988. Plant strategies and the dynamics and structure of plant communities.<br />

Princeton University Press, Princeton, New Jersey, USA.<br />

Thuiller, W., Araújo, M.B. & Lavorel, S. (2003) Generalized mo<strong>de</strong>ls vs. classification tree<br />

analysis: predicting spatial distributions of plant species at different scales. Journal of<br />

Vegetation Science, 14, 669–680.<br />

Tourna<strong>de</strong>, F. & Bouzillé, J.B. (1995) Déterminisme pédologique <strong>de</strong> la diversité végétale<br />

d’écosystèmes prairiaux du Marais Poitevin. Etu<strong>de</strong> et Gestion <strong>de</strong>s Sols 2, 57–72.<br />

112


Chapitre III<br />

Violle, C. & Jiang, L. (2009) Towards a trait-based quantification of species niche. Journal of<br />

Plant Ecology, 2, 87.<br />

Violle, C., Bonis, A., Plantegenest, M., Cu<strong>de</strong>nnec, C., Damgaard, C., Marion, B., Le Coeur,<br />

D. & Bouzillé, J.B. (2011) Plant functional traits capture species richness variations along a<br />

flooding gradient. Oikos, 120(3), 389-398.<br />

Violle, C., Pu, Z. & Jiang, L. (2010) Experimental <strong>de</strong>monstration of the importance of<br />

competition un<strong>de</strong>r disturbance. Proceedings of the National Aca<strong>de</strong>my of Sciences, 107,<br />

12925.<br />

Warwick, N.W. & Brock, M.A. (2003) Plant reproduction in temporary wetlands: the effects<br />

of seasonal timing, <strong>de</strong>pth, and duration of flooding. Aquatic Botany, 77, 153–167.<br />

Weiher, E. & Keddy, P.A. (1995) The assembly of experimental wetland plant communities.<br />

Oikos, 73, 323–335.<br />

Weiher, E., Werf, A., Thompson, K., Ro<strong>de</strong>rick, M., Garnier, E. & Eriksson, O. (1999)<br />

Challenging Theophrastus: a common core list of plant traits for functional ecology. Journal<br />

of Vegetation Science, 10, 609–620.<br />

Wel<strong>de</strong>n, C.W. & Slauson, W.L. (1986) The intensity of competition versus its importance: an<br />

overlooked distinction and some implications. Quarterly Review of Biology, 61, 23–44.<br />

113


Appendix<br />

Material and Methods<br />

The characterization of the elevation gradient<br />

Chapitre III<br />

A precise characterization of these waterlogging and drought constraints experienced<br />

by plants was neglected before the work of Gowing et al. (1998). They proposed the<br />

characterization of these aeration and drought stresses by the <strong>de</strong>velopment of two indices: the<br />

aeration SEV (Sum Exceedance Value) and the drought SEV expressed in cm.day -1 or in<br />

m.week -1 . Each in<strong>de</strong>x consisted of the <strong>de</strong>termination of the timing and the extent of the<br />

constraint by the <strong>de</strong>termination of two thresholds from which stress occurs from March to<br />

September (Gowing et al., 1998; Barber et al., 2004; Gilbert et al., 2003). The aeration SEV<br />

reflects changes in microtopography: aeration mainly drove species success and community<br />

distribution whereas drought does not (Gowing et al., 1998).<br />

For that, dipwells were placed along the topographical gradient in the aim to survey water-<br />

table movements across time with automatic probes (Fig. 1). The SEV aeration was calculated<br />

as the difference between the water table <strong>de</strong>pth and a reference level following this equation<br />

(Swetnam et al., 1998):<br />

where<br />

where WTt is the water-table <strong>de</strong>pth at time t from March to September and RV the reference<br />

water-table <strong>de</strong>pth above which the plants are expected to be aeration-stressed. The reference<br />

water-table <strong>de</strong>pth was expressed as a threshold calculated from a soil moisture release curve<br />

as the <strong>de</strong>pth that gives 10% air-filled porosity inducing an insufficient oxygen diffusion to<br />

supply the respiratory <strong>de</strong>mands of roots during growth period (Gowing et al., 1998; Swetnam<br />

et al., 1998). The aeration threshold thus correspon<strong>de</strong>d to a water table <strong>de</strong>pth of -0.191 m:<br />

when water table level is above this threshold, aeration stress occurs.<br />

114


Chapitre III<br />

Fig 1: Hydrograph realized along the topographical sequence where field experiment ran. The<br />

black line corresponds to the threshold of aeration at -0.191m below the ground surface. In<br />

grey: the area of the hydrograph which rise over the threshold aeration value indicative of the<br />

extent and the time that plants are aeration-stressed.<br />

The SEV aeration was positively correlated with flooding duration, measured as the number<br />

of days when water is above the soil surface from March to September (Pearson correlation,<br />

r=0.862, t=27.71, df=265, P


Chapitre III<br />

Table 1: Corresponding values of duration of submergence and SEV for aeration <strong>de</strong>termined<br />

by Pearson correlation.<br />

Results<br />

Duration of submergence in days Range of SEV aeration (cm.days -1 )<br />

0-7 [0; 4]<br />

10-18 [4.3; 5]<br />

21-27 [5.8; 7.2]<br />

31 7.6<br />

43 8.3<br />

54-57 [8.8; 8.9]<br />

62-65 [9.1; 9.4]<br />

70-74 [9.7; 10.2]<br />

80-89 [10.9; 14.7]<br />

90 [14.9; 16.9]<br />

91 [17.1; 20.7]<br />

92 [20.9; 25.5]<br />

93 [25.6; 30.1]<br />

94 [34.1; 36.9]<br />

Results of AIC between GLM and GAM realized for predictions of realized niches (Table<br />

2) <strong>de</strong>monstrated that realized of the twelve species were better predicted with GAM than with<br />

quadratic GLM because of low AIC values .<br />

Table 2: AIC values of quadratic GLM and GAM predicting realized niches of species.<br />

Quadratic GLM GAM<br />

A. stolonifera 208.91 190.02<br />

B. perennis 103.34 83.88<br />

C. cristatus 115.65 109.39<br />

C. divisa 236.71 223.47<br />

G. fluitans 194.87 187.21<br />

H. secalinum 198.66 181.05<br />

J. articulatus 302.17 289.16<br />

J. gerardii 281.52 245.72<br />

L. autumnalis 67.71 58.08<br />

L. perenne 209.75 193.14<br />

M. pulegium 205.14 195.18<br />

T. fragiferum 288.81 243.95<br />

116


DONNEES COMPLEMENTAIRES<br />

La mesure <strong>de</strong> la fluorescence a le potentiel <strong>de</strong> quantifier la contrainte subie par les<br />

espèces, à travers la mesure <strong>de</strong>s capacités photosynthétiques <strong>de</strong>s plantes, car l’appareil<br />

photosynthétique peut être affecté par la contrainte (Saltmarsh et al. 2006). La fluorescence a<br />

été mesurée lors d’une expérimentation similaire que celle <strong>de</strong> l’article 1.<br />

En raison <strong>de</strong> résultats similaires, certains résultats <strong>de</strong> cette expérimentation<br />

(matériel et métho<strong>de</strong>s en Annexe 3) sont présentés uniquement si <strong>de</strong>s méthodologies<br />

différentes ont été utilisées.<br />

Les résultats <strong>de</strong> l’analyse du ren<strong>de</strong>ment photosynthétique confirment les précé<strong>de</strong>ntes<br />

tendances, c’est-à-dire que l’augmentation en hauteur et en durée <strong>de</strong>s inondations diminue les<br />

performances photosynthétiques <strong>de</strong> J. gerardii (one-way ANOVA, P=1.08e-7), celle <strong>de</strong> L.<br />

autumnalis (one-way ANOVA, P=0.015) excepté pour le <strong>de</strong>rnier traitement, et augmente pour<br />

J. articulatus (one-way ANOVA, P=0.021). En revanche, les traitements n’ont pas d’effet sur<br />

les performances photosynthétiques <strong>de</strong> M. pulegium (one-way ANOVA, P=0.36) et <strong>de</strong> L.<br />

perenne (one-way ANOVA, P=0.57) (Fig. 19). Ces résultats nous permettent <strong>de</strong> directement<br />

approcher la réponse <strong>de</strong>s espèces aux conditions d’inondation.<br />

La tolérance <strong>de</strong>s espèces à l’augmentation <strong>de</strong>s conditions d’inondation a pu être reliée<br />

à la longueur <strong>de</strong> la tige, permettant aux espèces <strong>de</strong> renouer contact avec l’atmosphère<br />

augmentant la disponibilité à la lumière. Pour Juncus articulatus, les tiges dépassaient la<br />

surface <strong>de</strong> l’eau présentant une hauteur supérieure à 50cm.<br />

La présence <strong>de</strong> compétiteurs peut être vue comme une contrainte pour les espèces en<br />

raison <strong>de</strong> la réduction <strong>de</strong> la disponibilité <strong>de</strong>s ressources par les compétiteurs. La présence <strong>de</strong><br />

compétiteurs réduit les performances photosynthétiques <strong>de</strong> L. perenne et J. articulatus pour<br />

les traitements 20cm à durée d’inondation courte (Fig. 19). Les performances <strong>de</strong> L.<br />

autumnalis sont améliorées en présence <strong>de</strong> voisins pour les traitements 20cm, et cette<br />

différence significative pour le traitement 20cm à durée longue d’inondation. Les<br />

performances photosynthétiques <strong>de</strong> M. pulegium et <strong>de</strong> J. gerardii mesurées en présence <strong>de</strong><br />

voisins ne sont pas significativement <strong>de</strong> ses performances en l’absence <strong>de</strong> voisins (two-way<br />

ANOVA, P>0.05).<br />

117


Fig. 19 : Ren<strong>de</strong>ment photosynthétique <strong>de</strong>s espèces soumises à 4 traitements expérimentaux<br />

croissant en monocultures (losanges blancs) et en présence d’un compétiteur A. stolonifera<br />

(losange noir) (moyenne ± écart-type). 10S : 10cm durée courte, 20s : 20cm durée courte,<br />

20L : 20cm durée longue ; 40L : 40cm durée longue. Les différences significatives entre les<br />

ren<strong>de</strong>ments photosynthétiques sans et avec compétiteur sont indiquées par <strong>de</strong>s lettres<br />

distinctes (résultat <strong>de</strong>s tests <strong>de</strong> Tukey).<br />

118


RESUME PARTIE 1<br />

Dans cette partie, nous avons mis en évi<strong>de</strong>nce l’intérêt <strong>de</strong> caractériser les niches<br />

fondamentales <strong>de</strong>s espèces pour comprendre le rôle <strong>de</strong> la compétition dans les patrons<br />

d’abondance <strong>de</strong>s espèces.<br />

Deux types <strong>de</strong> contraintes ont été mis en évi<strong>de</strong>nce le long du gradient d’inondation :<br />

l’augmentation <strong>de</strong> la durée d’inondation et l’absence d’inondation. L’augmentation <strong>de</strong> la<br />

durée d’inondation est une contrainte pour une gran<strong>de</strong> partie <strong>de</strong>s espèces étudiées (9 espèces<br />

sur 12). Les espèces présentent <strong>de</strong>s niches fondamentales contrastées avec différents <strong>de</strong>grés <strong>de</strong><br />

tolérance à l’augmentation <strong>de</strong> la durée d’inondation. Ces contrastes sont observés à la fois<br />

entre espèces appartenant à <strong>de</strong>s communautés végétales différentes, mais ils sont également<br />

observés entre espèces appartenant à une même communauté.<br />

Dans le premier chapitre, nous avons mis en évi<strong>de</strong>nce différents patrons <strong>de</strong> variation<br />

<strong>de</strong> la réponse à la compétition (une composante <strong>de</strong> l’intensité <strong>de</strong> la compétition) le long du<br />

gradient expérimental d’inondations. Certaines espèces (Hor<strong>de</strong>um secalinum, Juncus gerardii<br />

et Mentha pulegium) ont <strong>de</strong>s réponses à la compétition qui varient entre les traitements<br />

d’inondation ; alors que d’autres espèces (Lolium perenne, Carex divisa, Bellis perennis,<br />

Glyceria fluitans et Juncus articulatus) ont <strong>de</strong>s réponses à la compétition qui ne varient pas<br />

entre les traitements expérimentaux d’inondation. Dans ce chapitre, il a également été montré<br />

que l’importance <strong>de</strong> la compétition est négativement corrélée à l’intensité <strong>de</strong> la contrainte<br />

mesurée à l’échelle <strong>de</strong> l’espèce.<br />

La comparaison <strong>de</strong>s niches fondamentales et réalisées <strong>de</strong>s espèces met en évi<strong>de</strong>nce un<br />

déplacement <strong>de</strong>s optima <strong>de</strong>s niches pour les espèces méso-hygrophiles (Juncus gerardii,<br />

Bellis perennis et Leontodon autumnalis) et pour les espèces hygrophiles (Mentha pulegium et<br />

Trifolium fragiferum). Les espèces les moins compétitives seraient exclues vers les parties du<br />

gradient plus contraignantes pour leur développement, c’est-à-dire vers les parties inondées<br />

du gradient. Ces espèces seraient exclues par les espèces mésophiles plus compétitives. La<br />

mise en évi<strong>de</strong>nce <strong>de</strong> ce déplacement <strong>de</strong> niche, couplée à la quantification <strong>de</strong> l’importance <strong>de</strong><br />

la compétition, démontre le rôle structurant <strong>de</strong> la compétition le long du gradient<br />

d’inondation.<br />

119


120


Partie 2 :<br />

La démographie comme outil d’étu<strong>de</strong> <strong>de</strong>s<br />

filtres écologiques<br />

121


122


- Chapitre 4 -<br />

Effect of the environment on species competitive effect and<br />

importance of competition along a flooding gradient.<br />

En préparation<br />

123


124


Chapitre IV<br />

Effect of the environment on species competitive effect and importance of<br />

competition along a flooding gradient.<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1,2 , Christian Damgaard 3 , François Mesléard 2,4 , Anne Bonis 1<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, 35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc F-13200, France<br />

3 Department of Terrestrial Ecology, NERI, Aarhus University, Vejlsøvej 25, 8600 Silkeborg,<br />

Denmark<br />

4 <strong>Université</strong> d’Avignon - IMEP, IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09,<br />

France<br />

Abstract<br />

Competition is little measured in natural communities. In this study, we examined the<br />

effect of two constraints (flooding and soil drying) observed during two different periods<br />

(during flooding period and after the end of flooding) on the competitive effect of species and<br />

importance of competition, to <strong>de</strong>termine the shaping role of competition on species<br />

distribution using a competition mo<strong>de</strong>l.<br />

The mo<strong>de</strong>l measured the competitive effect of species and the importance of<br />

competition using two ecological variables estimated from pin point measurements realized in<br />

natural communities: species cover and species vertical <strong>de</strong>nsity. The two integrative variables<br />

<strong>de</strong>scribed the plant performances: survival, colonization and biomass growth. This<br />

competition mo<strong>de</strong>l was tested for five wetland species presenting different ecological ranges<br />

along the flooding gradient.<br />

Along the gradient, the increase in flooding duration and in water availability had a<br />

positive effect on survival, colonization, biomass growth and competitive effect of Agrostis<br />

stolonifera and Juncus gerardii. These environmental conditions had no significant effect on<br />

the competitive effect of Lolium perenne and G. fluitans. For Cynosurus cristatus, the<br />

competitive effect <strong>de</strong>creased only with the availability of water resources. Along the gradient,<br />

competition importance of overall studied species increased with the <strong>de</strong>crease in soil drying.<br />

During the floo<strong>de</strong>d period, the importance of competition for L. perenne, J. gerardii and G.<br />

fluitans increased with flooding duration whereas the patterns of competition importance<br />

varied greatly along the gradient for A. stolonifera and C. cristatus.<br />

This competition mo<strong>de</strong>l, which directly measures competition in natural communities,<br />

shows the importance of flooding as a filter influencing species performances such as the<br />

shaping role of competition along the gradient. This mo<strong>de</strong>l provi<strong>de</strong>s a powerful tool to<br />

<strong>de</strong>termine processes acting on communities’ composition and structure.<br />

Key-words: competition mo<strong>de</strong>l, survival, colonization, growth, pinpoint method<br />

125


Chapitre IV<br />

126


Introduction<br />

Chapitre IV<br />

Along environmental gradients, plant species are simultaneously submitted to the<br />

presence of a constraint as well as the presence of competitors interacting for similar<br />

resources affecting species performances. Measurements of the impact of competition relative<br />

to the impact of other environmental factors on species performances are essential for<br />

<strong>de</strong>termining population dynamics and then predict their consequences at the level of the<br />

whole community. However, few studies have measured competition integrating the different<br />

measures of population dynamics (Brooker & Kikvidze 2008; Frekleton et al. 2009).<br />

Along flooding gradients, frequency and duration of flooding represent constraints for<br />

weak tolerant species reducing plant survival (Van Eck et al. 2004; Fraser & Karnesis 2005),<br />

plant recruitment (Lenssen et al. 1998; Casanova & Brock 2000; Fraser & Karnesis 2005;<br />

Kotowski et al. 2010), plant reproduction (Warwick & Brock 2003) or plant growth (Visser et<br />

al. 2003; Fraser & Karnesis 2005). Competition between species also affects these different<br />

measures of plant performances (Howard & Goldberg 2001; Fayolle et al. 2009). Competition<br />

is <strong>de</strong>scribed by two components: the intensity of competition, i.e. the absolute reduction of the<br />

fitness of a species due to the presence of neighbors (Wel<strong>de</strong>n & Slauson 1986) and the<br />

importance of competition, i.e. the reduction of the fitness of a species by the presence of<br />

competitors and by any other process and condition (Wel<strong>de</strong>n & Slauson 1986). Competition<br />

does not only regulate species performances, competition can also represent a main<br />

structuring factor regulating the community structure along environmental gradients.<br />

Studies measuring the two components of competition have been mainly realized<br />

un<strong>de</strong>r controlled conditions. Experimentally, studies have highlighted an increase in the<br />

importance of competition with the <strong>de</strong>crease in the intensity of the constraint along<br />

productivity (Gaucherand et al. 2006), salinity (Greiner la Peyre et al. 2001) and flooding<br />

gradients (Jung et al. 2009). By contrast, generalization of results is not easy for the intensity<br />

of competition. In some cases, it has been <strong>de</strong>monstrated to be constant along salinity,<br />

productivity and disturbance gradients (e.g. Greiner la Peyre et al. 2001; Gaucherand et al.<br />

2006; Carlyle et al. 2010). In other cases, it has been shown to <strong>de</strong>crease with the increase in<br />

the intensity of the constraint as well as with gradients of productivity, salinity and flooding<br />

gradients (e.g. Fraser & Keddy 2005; Pennings et al. 2005). These contrasted results may be<br />

explained by different aspects of the competitive ability of species measured: the competitive<br />

response, measuring the ability of species to tolerate the presence of competitor (Goldberg<br />

1990), or the competitive effect, measuring the ability of species to <strong>de</strong>plete resources for<br />

127


Chapitre IV<br />

others, thus reducing neighbors performances (Goldberg 1990). Competitive response informs<br />

on the species strategies implemented to tolerate the reduction of resources level due to the<br />

presence of neighbors. Competition response <strong>de</strong>pends on the type of resources available in the<br />

environment. Competition response thus appears more variable along gradients than the<br />

competitive effect (Keddy et al. 1994). However, the methodologies used as well as the<br />

distinction between the competitive response and the competitive effect appear insufficient to<br />

predict the real effect of competition experienced by species in natural communities, because<br />

species are affected by the presence of competitors and induce a <strong>de</strong>pletion of resources<br />

simultaneously.<br />

Competition mo<strong>de</strong>ls are now required to measure competition in natural communities.<br />

The <strong>de</strong>velopment of mo<strong>de</strong>ls integrating <strong>de</strong>mographic data of individual plants have a great<br />

potential for measuring competition in natural communities (Goldberg 1999; Aarssen &<br />

Keogh 2002) and providing an interesting approach to dissect out the various processes<br />

influencing the dynamics of populations. Recently, Frekleton et al. (2009) proposed a<br />

dynamic mo<strong>de</strong>l for annual species measuring the two components of competition using<br />

different measures of annual ecological success such as seed production and survival,<br />

consi<strong>de</strong>ring species <strong>de</strong>nsity. Another promising mo<strong>de</strong>l quantifying the importance of<br />

competition and <strong>de</strong>riving from the <strong>de</strong>finition of Wel<strong>de</strong>n & Slauson (1986) (previously<br />

enounced) has been <strong>de</strong>veloped by Damgaard & Fayolle (2010). They proposed a general<br />

method to calculate the two components at any measure of population or individual ecological<br />

success, consi<strong>de</strong>ring the <strong>de</strong>nsity <strong>de</strong>pen<strong>de</strong>nce of interactions and the various levels of<br />

environmental gradients. They thus <strong>de</strong>monstrated that the importance of competition relative<br />

to the level of herbici<strong>de</strong>s appears species-specific.<br />

The aim of this study is to measure the two components of competition in natural<br />

communities. We particularly aimed to: i) test the effect of the environment on the<br />

competitive effect of different species presenting different ecological ranges along a flooding<br />

gradient, ii) predict competition importance relative to environmental conditions on species<br />

ecological success, and iii) compare these results for two periods characterized by two kinds<br />

of constraint: flooding occurring between growing seasons and the reduction of the<br />

availability of water resources occurring during the growing season.<br />

To respond to these objectives, the competition mo<strong>de</strong>l <strong>de</strong>veloped by Damgaard<br />

(submitted) has been used to test the effect of the environmental conditions on the competitive<br />

effect of species and to quantify the importance of competition at species level. The<br />

originality of this approach is the non-<strong>de</strong>structive character of this method measuring two<br />

128


Chapitre IV<br />

ecological variables relevant for the perennial species ecological success: the cover and the<br />

vertical <strong>de</strong>nsity (Damgaard et al. 2009). The competition mo<strong>de</strong>l has been <strong>de</strong>veloped to<br />

estimate competition coefficients between species, i.e. estimating the competitive effect of<br />

species on neighbors’ performances in natural communities (Damgaard et al. 2009). In this<br />

mo<strong>de</strong>l, the vertical <strong>de</strong>nsity is a function of the cover of species during the growing season, as<br />

the change in vertical <strong>de</strong>nsity is assimilated to the biomass growth during the growing season.<br />

Between growing seasons, the cover of species is function of their vertical <strong>de</strong>nsity at the end<br />

of the previous growing season, as cover can be assimilated to survival and colonization<br />

abilities of species in response to the environment. In this way, the effect of the environment<br />

on the processes controlling the biomass growth during the growing season and the processes<br />

controlling the translation of biomass in cover between growing seasons can be studied.<br />

Methods<br />

Study site and vegetation analysis<br />

The study site was the productive and wet grasslands of the Marais Poitevin on the<br />

French Atlantic coast (46°28’N; 1°13’W). The climate is atlantic with a mean annual rainfall<br />

of 655 mm (Amiaud & Touzard 2004). These wet grasslands are characterized by an elevation<br />

gradient with a range of 45 cm, leading to a flooding gradient varying in duration and water<br />

<strong>de</strong>pth (Violle et al. 2007). Annual cycles of flooding start in autumn <strong>de</strong>pending on rainfall<br />

amount. Within a year, two constraints succeed across time. During the floo<strong>de</strong>d/anoxic season<br />

(from end of October to beginning of June), the main constraint is anoxia paired with<br />

submergence, particularly observed in the lower parts of the gradient. During summer-<br />

beginning of autumn (from beginning of June to end of October), the main constraint is the<br />

low availability of water for plants.<br />

Along two different topographical sequences, 70 permanent plots (25*25 cm) were<br />

placed systematically along the gradient spaced by 20 cm from each other (35 plots per<br />

sequence). Along both diagonals of each plot, relevés using the pinpoint method, were<br />

realized every 4 cm, for a total of 17 points per plot. Samples were taken at the beginning<br />

(end of October 2008) and at the end of the floo<strong>de</strong>d/anoxic period (beginning of June 2009),<br />

as well as at the end of summer, right before the beginning of the next flooding period (end of<br />

October 2009). By sampling at these dates, we are able to study competition un<strong>de</strong>r different<br />

water availability constraints.<br />

129


Chapitre IV<br />

As the topographical sequences are located in two different floo<strong>de</strong>d <strong>de</strong>pressions, we<br />

have verified whether hydrological functioning were similar. For that, we have quantified the<br />

intensity of flooding conditions during the flooding period, as well as water availability (i.e.<br />

intensity of drought constraint) during summer, following the proposition of Gowing et al.<br />

(1998) (see Appendix). This method consists in the quantification of flooding intensity<br />

represented by the Sum Excee<strong>de</strong>nce Value (SEV) for aeration (aeration SEV) and the<br />

quantification of the intensity of drought represented by the SEV for drought (soil drying<br />

SEV). The plot of the aeration SEV in relation to the measure of elevation <strong>de</strong>monstrated that<br />

for a same measure of elevation correspon<strong>de</strong>d two different intensities of flooding conditions<br />

(Fig. 1). The hydrological functioning was different between the two topographical<br />

sequences. The measure of the elevation did not represent the best variable to study<br />

competition according to the flooding conditions: we chose to analyze data according to the<br />

intensity of flooding conditions and soil drying, known to better explain species distribution<br />

(Silvertown et al. 1999; Silvertown 2004). The duration of flooding was correlated with<br />

aeration SEV (Pearson correlation, r=0.862, P


Chapitre IV<br />

hygrophilous species recor<strong>de</strong>d in the part of the gradient characterized by long duration of<br />

flooding conditions, is constrained by the absence of flooding conditions. The competition of<br />

these four species was calculated in association with another one: Agrostis stolonifera. This<br />

last species was distributed all along the elevation gradient.<br />

Competition mo<strong>de</strong>l<br />

The species competitive effect was measured for two periods: the flooding season and<br />

the growing season. In or<strong>de</strong>r to evaluate the species competitive effect during the flooding<br />

season, the samples taken after and before flooding were compared as explained below.<br />

Similarly, to evaluate the species competitive effect during the growing season, the samples<br />

taken after and before.<br />

For a given year, the vertical <strong>de</strong>nsity of a species i, Yi, represents the total number of<br />

contact points between the species and any of the 17 pinpoints in the frame. The vertical<br />

<strong>de</strong>nsity appeared to be integrative of the biomass growth of a species during the growing<br />

season (Damgaard et al. 2009). The cover of species i, Xi, is the total number of occurrence of<br />

the species at any of the 17 pinpoints of the frame. It is assumed that the vertical <strong>de</strong>nsity of a<br />

species i at time t2 is an increasing function of the cover of species i, function of the cover of<br />

species j and k at time t1, and function of the environmental gradient zr. Competitive growth<br />

of species i was mo<strong>de</strong>led as:<br />

with r the pin-point frame; the residual process variation during the growing season of species<br />

i across different years and pin-point frames. ai(zr) corresponds to the growth of species i<br />

directly affected by the relation between the cover and the vertical <strong>de</strong>nsity of the species;<br />

cj(zr) and ck(zr) correspond to the competitive effects of species j and k affecting growth of<br />

species i. The parameters a and c are functions of the environmental gradient as linear<br />

functions a0+a1z and c0+c1z.<br />

(1)<br />

131


Chapitre IV<br />

Flooding occurred at the end of autumn and en<strong>de</strong>d in spring of the following year:<br />

colonization and survival abilities of species appeared integrative of fitness of species during<br />

this period (Damgaard et al. 2009). Among growing seasons, it is assumed that the cover of<br />

species i at year t+1 is an increasing function of the vertical <strong>de</strong>nsity of species at the year t (at<br />

the end of the growing season), function of the vertical <strong>de</strong>nsity of species j and k at year t and<br />

function of the environmental gradient zr. Survival and colonization of species i at year t+1<br />

was mo<strong>de</strong>led as:<br />

with the residual process variation from one season to the next of species i among years and<br />

pin-point frames.<br />

Parameters estimation<br />

Mo<strong>de</strong>l parameters (a, b, c and d) were estimating using the Bayesian method: the joint<br />

posterior distribution of mo<strong>de</strong>l parameters were calculated using the MCMC method<br />

(Metropolis-Hastings algorithm with 120000 iterations) and a multivariate distribution. The<br />

sampled chains of all parameters were inspected to check the properties of the sampling<br />

procedure. The effect of the environmental conditions on species performances (parameter a1)<br />

and on species competitive effect (parameter c1) were assessed by the 2.5%, 50% and 97.5%<br />

percentiles taken from the marginal posterior distribution of the parameters.<br />

Measurement of competition importance<br />

From equations (1) and (2), the importance of competition was quantified along the<br />

environmental gradient for the two studied periods, i.e. during the growing season (3) and<br />

during the flooding season (4) following the proposition of Damgaard & Fayolle (2010):<br />

(3)<br />

(4)<br />

Where and represent the absolute changes in the ecological<br />

success of the species i by changing respectively the cover of species j (Xj) and the vertical<br />

(2)<br />

132


Chapitre IV<br />

<strong>de</strong>nsity of species j (Yj). and represent the absolute changes in the<br />

ecological success of the species i by changing the level of the environment z. Absolute<br />

values indicate that the importance of competition is comprised between 0 and 1. All analyses<br />

were run with Mathematica (Wolfram 2003).<br />

Results<br />

Effect of environmental conditions on species performances<br />

The effect of the environmental conditions on species performances (survival,<br />

colonization and growth) was assessed by the percentiles of the parameter a1 (Table 1).<br />

During the growing season, the <strong>de</strong>crease in the intensity of soil drying along the gradient had<br />

a positive effect on growth for L. perenne, C. cristatus, J. gerardii, G. fluitans and A.<br />

stolonifera for the four groups of species tested. For the aggregated species, the effect of the<br />

<strong>de</strong>crease in the intensity of soil drying along the gradient varied between the groups of species<br />

studied: this effect was negative for the A. stolonifera-C. cristatus group and positive only for<br />

the A. stolonifera-J. gerardii group .<br />

During the flooding season, the increase in flooding duration along the gradient had a<br />

positive effect on the survival and colonization for J. gerardii. The increase in flooding<br />

duration had a positive effect on the survival and colonization for A. stolonifera only for the<br />

A. stolonifera-L. perenne group. For the aggregated species, increase in flooding duration had<br />

a positive effect for the four groups of species tested, except the A. stolonifera-L. perenne<br />

group.<br />

Effect of environmental conditions on species competitive effect<br />

The effect of the environmental conditions on species competitive effect was assessed<br />

by the percentiles of the parameter c1 (Table 1). During the growing season, the <strong>de</strong>crease in<br />

the intensity of soil drying along the gradient had a positive effect only on the competitive<br />

effect for J. gerardii and a negative effect only on the competitive effect for C. cristatus. For<br />

A. stolonifera and others aggregated species, the <strong>de</strong>crease in the intensity of soil drying had a<br />

positive effect on their competitive effects for the four groups of species tested.<br />

133


Chapitre IV<br />

During the flooding season, the increase in flooding duration along the gradient had a<br />

positive effect only on the competitive effect for other aggregated species for all groups of<br />

species tested, except for the A. stolonifera-C. cristatus group.<br />

Table 1: Percentiles provi<strong>de</strong>d from the marginal posterior distribution of the parameters a1<br />

and c1, measuring the effect of environment on species ecological success (a1) and species<br />

competitive effect (c1), for both studied periods and for the four groups of species tested. In<br />

bold: parameters significantly <strong>de</strong>viated from zero.<br />

Mo<strong>de</strong>l parameters a1 c1<br />

Confi<strong>de</strong>nce interval 2.5% 50% 97.5% 2.5% 50% 97.5%<br />

During A. stolonifera 10.115 14.415 31.355 0.167 0.208 0.262<br />

growing<br />

season<br />

L.perenne<br />

4.453 9.610 17.900 -0.355 -0.060 0.081<br />

Other aggregated<br />

species<br />

-4.219 -2.340 1.0152 0.048 0.074 0.115<br />

During A. stolonifera 0.0001 0.003 0.009 -0.006 0.001 0.007<br />

flooding L.perenne -0.004 0.002 0.018 -0.023 -0.001 0.013<br />

season Other aggregated<br />

species<br />

-0.0006 0.030 0.103 0.003 0.018 0.034<br />

During A. stolonifera 6.819 8.855 11.196 0.190 0.211 0.223<br />

growing C. cristatus 4.971 5.022 5.694 -0.940 -0.692 -0.497<br />

season Other aggregated<br />

species<br />

-5.760 -3.886 -2.012 0.063 0.067 0.088<br />

During A. stolonifera -0.004 0.0001 0.003 -0.006 0.005 0.022<br />

flooding C. cristatus -0.005 -0.002 0.004 -0.001 0.009 0.029<br />

season Other aggregated<br />

species<br />

0.560 0.630 0.723 -0.004 0.005 0.014<br />

During A. stolonifera 20.337 23.812 27.136 0.278 0.295 0.316<br />

growing J. gerardii 2.017 2.158 2.528 0.130 0.200 0.274<br />

season Other aggregated<br />

species<br />

17.176 22.662 31.033 0.073 0.090 0.109<br />

During A. stolonifera -0.003 0.0005 0.005 -0.003 0.006 0.016<br />

flooding J. gerardii 0.194 0.296 0.446 -0.057 -0.026 0.008<br />

season Other aggregated<br />

species<br />

0.935 1.109 1.412 0.004 0.017 0.030<br />

During A. stolonifera 10.916 18.438 31.817 0.225 0.250 0.283<br />

growing G. fluitans 47.157 67.449 100.842 -0.144 0.208 0.638<br />

season Other aggregated<br />

species<br />

-4.427 1.274 8.262 0.066 0.086 0.117<br />

During A. stolonifera -0.002 0.0005 0.005 -0.018 -0.006 0.0002<br />

flooding G. fluitans -0.001 0.002 0.004 -0.037 -0.007 0.006<br />

season Other aggregated<br />

species<br />

0.906 1.047 1.149 0.002 0.581 0.973<br />

134


Importance of competition<br />

Chapitre IV<br />

The importance of competition was calculated for each group of species tested: the<br />

curves (Fig. 2 & 3) represented the proportion of change in ecological success caused by<br />

competition relative to environmental conditions.<br />

During the growing season, the patterns of competition importance along the gradient<br />

were similar between species (Fig. 2). The importance of competition increased with the<br />

<strong>de</strong>crease in soil drying for L. perenne, C. cristatus and G. fluitans: this importance of<br />

competition was lower in the higher parts of the gradient, followed by a strong increase in<br />

importance of competition and reached a plateau at high values of competition importance for<br />

the rest of the gradient (closed to 1 for each species). For J. gerardii, the importance of<br />

competition increased progressively with the <strong>de</strong>crease in soil drying. For A. stolonifera, the<br />

pattern of competition importance increased with the <strong>de</strong>crease in soil drying but not linearly.<br />

Fig. 2: Pattern of competition importance along the gradient calculated for each group of<br />

tested species during the growing season, with an initial cover of species, i, j and k: 1, 1, 0<br />

135


Chapitre IV<br />

During the flooding season, the patterns of competition importance for L. perenne, J.<br />

gerardii and G. fluitans increased with the flooding duration along the gradient (Fig. 3). The<br />

importance of competition was especially high for J. gerardii but less important for G.<br />

fluitans and L. perenne in floo<strong>de</strong>d parts of the gradient. The importance of competition for C.<br />

cristatus was the lowest for an aeration SEV value of 7cm.day -1 (Fig. 3). The importance of<br />

competition for A. stolonifera varied between the four groups of species tested (Fig. 3). The<br />

importance of competition for A. stolonifera increased with aeration SEV for the A.<br />

stolonifera-L. perenne group. The importance of competition for A. stolonifera <strong>de</strong>creased with<br />

the increase in aeration SEV for the A. stolonifera-G. fluitans group. For the A. stolonifera-J.<br />

gerardii and A. stolonifera-C. cristatus groups, the importance of competition for A.<br />

stolonifera was lowest for aeration SEV values equal to 4 and 8cm.day -1 .<br />

Fig. 3: Pattern of competition importance along the gradient calculated for each group of<br />

species tested during the flooding season with an initial vertical <strong>de</strong>nsity of species, i, j and k:<br />

1, 1, 0.<br />

136


Discussion<br />

Chapitre IV<br />

The analysis of species cover and vertical <strong>de</strong>nsity allows, on the one hand, the<br />

estimation of the effect of the environmental conditions on the competitive effect of species<br />

on neighbors, and on the other hand, the <strong>de</strong>termination whether competition represents a main<br />

factor structuring species assemblages relative to environmental conditions.<br />

Effect of the environmental conditions on species competitive effect<br />

Species performances and competitive effects varied along the gradient for the two<br />

studied periods, i.e. during the floo<strong>de</strong>d period and during the growing season. For J. gerardii<br />

and A. stolonifera, the increase in flooding duration along the gradient increases their survival<br />

and colonization abilities. However, the flooding duration has no effect on their ability to<br />

occupy space. Moreover, the increase in water availability along the gradient, i.e. during the<br />

growing season, positively impacted the biomass growth of these two species and also their<br />

competitive effect. Only performances of these two species increased along the gradient for<br />

the two studied periods. This result suggests a relationship between species performances<br />

between the two studied periods. We can hypothesize that the ability of species to respond to<br />

flooding is an advantage to be competitive afterwards. This hypothesis could be supported by<br />

the growth strategies of these two species. In<strong>de</strong>ed, in response to flooding conditions, A.<br />

stolonifera <strong>de</strong>veloped long stems along which adventitious roots grow, i.e. long above-ground<br />

runners (obs. pers.). This growth strategy is recognized to be a <strong>de</strong>terminant response to<br />

flooding conditions representing an efficient strategy to occupy space and to re-growth after<br />

the end of the flooding period (Soukupova 1994; Lenssen et al. 2004; Benot et al. 2011).<br />

In<strong>de</strong>ed, the adventitious roots allow the rooting of stems after the end of flooding events<br />

(Santamaria 2002; Lenssen et al. 2004). The ability to store resources in rhizomes during a<br />

constrained period such as flooding, known to represent resources storage organs (Dong & <strong>de</strong><br />

Kroon 1994), may explained the increase in plant survival and colonization with flooding<br />

duration highlighted for J. gerardii. The stored resources may be efficiently used when the<br />

<strong>de</strong>crease in flooding conditions starts as it was observed for rhizomatous species such as<br />

Mentha aquatica (Lenssen et al. 2000) or as Phragmites australis (Clevering 1998).<br />

In literature, a relationship between plant biomass and species competitive effect<br />

(Fraser & Miletti 2008), as well as plant size and species competitive effect (Peltzer & Kochy<br />

2001; Wang et al. 2010) has been highlighted. Few plant strategies exist to suppress<br />

neighbors: in high productive environments as ours, the increase in size and consequently in<br />

137


Chapitre IV<br />

biomass represents a strong advantage in capturing light (Grime 1977). An increase in<br />

biomass growth responsible of the increase of species competitive effect with the availability<br />

of water resource during the growing season could be expected. However, this biomass<br />

growth does not appear sufficient to allow species to be competitive during the growing<br />

season in our study. In<strong>de</strong>ed, for C. critatus, L. perenne and G. fluitans, <strong>de</strong>spite a biomass<br />

growth increased with water availability, neither colonization and survival abilities nor the<br />

ability to occupy space are impacted during the flooding period, resulting in an absence of<br />

their ability to suppress neighbors. A hypothesis concerning the growth strategy of these<br />

species may be formulated to un<strong>de</strong>rstand these results. Cynosurus critatus, L. perenne and G.<br />

fluitans present caespitose growths form, highly efficient in semi-arid environments (Derner<br />

& Briske 2001), but less efficient to store resources than the rhizomatous organs in<br />

waterlogged and floo<strong>de</strong>d environment. Thus, the lower ability of species to store resources<br />

during the constrained period may limit the ability of species to re-growth at the end of<br />

flooding event and to be strong competitive species. This may suggest the importance of the<br />

strategies implemented by species, such as the vegetative reproduction and the storage<br />

resources, to respond to flooding (Lenssen et al. 2000) and to explain the dynamic of the<br />

population over the studied periods.<br />

Importance of competition<br />

The importance of competition relative to environmental conditions varied between<br />

the studied periods. In<strong>de</strong>ed the competition has a greater impact than environmental<br />

conditions on species ecological success during the growing season for all studied species and<br />

all along the gradient. While the importance of competition explains 45% of survival and<br />

colonization abilities of L. perenne and C. cristatus during the floo<strong>de</strong>d season compared to<br />

70% in floo<strong>de</strong>d parts of the gradient for J. gerardii and G. fluitans. During the floo<strong>de</strong>d season,<br />

competition does not represent the main filter acting on species ecological success except for<br />

J. gerardii and G. fluitans in floo<strong>de</strong>d parts of the gradient, whereas competition appears to be<br />

the main process acting on species ecological success during the growing season. This then<br />

suggests an important role of flooding conditions in regulating plant populations and species<br />

distribution along the gradient and also a shaping role of competition all along the gradient<br />

after the end of flooding event.<br />

For a given period, patterns of competition importance along the gradient are similar<br />

among species. In<strong>de</strong>ed, the importance of competition increases along the gradient with the<br />

138


Chapitre IV<br />

availability of water resources for all species studied; in general the importance of<br />

competition along the gradient increases with flooding duration. However, the importance of<br />

competition does not <strong>de</strong>crease with the intensity of environmental conditions known to<br />

constrain species performances – such as the increase in flooding duration for C. cristatus, L.<br />

perenne and J. gerardii – as it could be expected (Brooker et al. 2005). Only the importance<br />

of competition increased linearly with flooding duration for G. fluitans, arguing for the<br />

expected negative relationship between the importance of competition and the constraint at<br />

species level, <strong>de</strong>monstrated using experimental data (<strong>Merlin</strong> et al. in prep). Such a linear<br />

relationship is not observed for the studied species affected by the increase in flooding<br />

duration (C. cristatus, L. perenne and J. gerardii) and for A. stolonifera. These results suggest<br />

that the importance of competition cannot always be predicted by the intensity of the<br />

environmental conditions known to reduce species performances. The results may be<br />

explained by the stages of plant life-cycle used as measures of plant performances: these<br />

stages may be differently affected by competition and abiotic factors. Moreover the species<br />

studied have been submitted to several annual cycles of flooding conditions: their response to<br />

abiotic and biotic factors and their distribution along the gradient probably represent the<br />

outcome of past and present processes (Wel<strong>de</strong>n & Slauson 1986; Violle et al. 2010) resulting<br />

in variable patterns of the competition importance.<br />

Despite a similarity in the pattern of competition importance between the species<br />

studied for a given period, the values of competition importance appear different along the<br />

gradient, especially during the floo<strong>de</strong>d period. This suggests that the patterns of competition<br />

importance along the flooding gradient are species-specific.<br />

Conclusion<br />

This study, quantifying both components of competition in natural communities using<br />

a competition mo<strong>de</strong>l, emphasizes the interest to integer the different measures of species life-<br />

cycle such as survival, colonization and biomass growth, as well as other factors such as<br />

environmental conditions, in or<strong>de</strong>r to un<strong>de</strong>rstand the role of competition on species ecological<br />

success (Brooker & Kikvidze 2008; Frekleton et al. 2009). This study thus represents a step<br />

forward the un<strong>de</strong>rstanding of the shaping role of competition on species and populations in a<br />

dynamics context. Further investigations are required for studying competition and predicting<br />

the importance of competition to explain species distribution and communities’ organization<br />

139


Chapitre IV<br />

consi<strong>de</strong>ring long term monitoring. In<strong>de</strong>ed we lack a step back to really un<strong>de</strong>rstand the patterns<br />

of competition importance.<br />

Literature cited<br />

Aarssen, L.W. & Keogh, T. (2002) Conundrums of competitive ability in plants: what to<br />

measure? Oikos, 96, 531-542.<br />

Benot, M.L., Mony, C., <strong>Merlin</strong>, A., Marion, B., Bouzillé, J.B. & Bonis, A. (2011) Clonal<br />

growth strategies along flooding and grazing gradients in Atlantic coastal meadows. Folia<br />

Geobotanica, 46, 219-235.<br />

Brooker, R. & Kikividze, Z. (2008) Importance: an overlooked concept in plant interaction<br />

research. Journal of Ecology, 96, 703-708.<br />

Carlyle, C. N., Fraser, L.H. & Turkington, R. 2010. Using three pairs of competitive indices<br />

to test for changes in plant competition un<strong>de</strong>r different resource and disturbance levels.<br />

Journal of Vegetation Science, 21: 1–10.<br />

Casanova, M.T. & Brock, M.A. (2000) How do <strong>de</strong>pth, duration and frequency of flooding<br />

influence the establishment of wetland plant communities? Plant Ecology, 147, 237–250.<br />

Clevering, O.A. (1998) Effects of litter accumulation and water table on morphology and<br />

productivity of Phragmites australis. Wetlands Ecology and Management 5, 275–287.<br />

Damgaard, C. & Fayolle, A. (2010) Measuring the importance of competition: a new<br />

formulation of the problem. Journal of Ecology, 98, 1–6.<br />

Damgaard, C., Riis-Nielsen, T. & Schmidt, I.G. (2009) Estimating plant competition<br />

coefficients and predicting community dynamics from non-<strong>de</strong>structive pin-point data: a case<br />

study with Calluna vulgaris and Deschampsia flexuosa. Plant Ecol. 201, 687–697.<br />

Derner J. D. & Briske, D. D. (2001) Below-ground carbon and nitrogen accumulation in<br />

perennial grasses: a comparison of caespitose and rhizomatous growth forms. Plant and Soil<br />

237, 117-127.<br />

Dong, M., & <strong>de</strong> Kroon, H. (1994) Plasticity in morphology and biomass allocation in<br />

Cynodon dactylon, a grass species forming stolons and rhizomes. Oikos, 70, 99-106.<br />

Fayolle, A., Violle, C. & Navas, M.L. (2009) Differential impacts of plant interactions on<br />

herbaceous species recruitment disentangling factors controlling emergence, survival and<br />

growth seedlings. Oecologia, 159(4), 817-825.<br />

Fraser, L.H. & Karnesis, J.P. (2005) A comparative assessment of seedling survival and<br />

biomass accumulation for fourteen wetland plant species grown un<strong>de</strong>r minor water-<strong>de</strong>pth<br />

differences. Wetlands, 25(3), 520–530.<br />

Fraser, L.H. & Keddy, P.A. (2005) Can competitive ability predict structure in experimental<br />

plant communities? Journal of Vegetation Science. 16, 571-578<br />

Fraser, L.H. & Miletti, T.E. (2008) Effect of minor water <strong>de</strong>pth treatments on competitive<br />

effect and response of eight wetland plants. Plant Ecology, 195, 33–43.<br />

Frekleton, R.P., Watkinson, A.R. & Rees, M. (2009) Measuring the importance of<br />

competition in plant communities. Journal of Ecology, 97, 379–384.<br />

140


Chapitre IV<br />

Fréville, H. & Silvertown, J. (2005) Analysis of interspecific competition in perennial plants<br />

using Life Table Response Experiments. Plant Ecology, 176, 69-78.<br />

Gaucherand, S., Liancourt, P. & Lavorel, S. (2006) Importance and intensity of competition<br />

along a fertility gradient and across species. Journal of Vegetation Science, 17, 455-464.<br />

Goldberg, D.E. (1990) Components of resource competition in plant communities. In: Grace,<br />

J.B. & Tilman, D. (eds.) Perspectives on plant competition. pp. 27–49. Aca<strong>de</strong>mic Press, San<br />

Diego.<br />

Goldberg, D.E., Rajaniemi, T., Gurevitch, J. & Stewart-Oaten, A. (1999) Empirical<br />

Approaches to Quantifying Interaction Intensity: Competition and Facilitation along<br />

Productivity Gradients. Ecology, 80(4), 1118-1131.<br />

Gowing D.J.G., Youngs, E.G., Gilbert, J.C. & Spoor, G. (1998). Predicting the effect of<br />

change in water regime on plant communities. In Hydrology in a Changing Environment, vol.<br />

I, Wheater H, Kirby C (eds). John Wiley: Chichester; 473–483. Published by: British<br />

Ecological Society<br />

Greiner La Peyre, M.K., Grace, J.B., Hahn, E. & Men<strong>de</strong>lssohn, I.A. (2001) The importance of<br />

competition in regulating plant species abundance along a salinity gradient. Ecology, 82, 62–<br />

69.<br />

Grime, J.P. (1977) Evi<strong>de</strong>nce for the existence of three primary strategies in plants and its<br />

relevance to ecological and evolutionary theory. American naturalist, 111, 1169–1194.<br />

Howard, T.G. & Goldberg, D.E. (2001) Competitive response hierarchies for germination,<br />

growth, and survival and their influence on abundance. Ecology, 82(4), 979–990.<br />

Jung, V., Mony, C., Hoffmann, L. & Muller, S. (2009) Impact of competition on plant<br />

performances along a flooding gradient: a multi-species experiment. Journal of Vegetation<br />

Science, 20, 433–441.<br />

Keddy, P.A., Twolan-Strutt, L. & Wisheu, I.C. (1994) Competitive Effect and Response<br />

Rankings in 20 Wetland Plants: Are They Consistent Across Three Environments? Journal of<br />

Ecology, 82(3), 635-643.<br />

Kotowski, W., Beauchard, O., Op<strong>de</strong>kamp, W., Meire, P. & van Diggelen, R. (2010)<br />

Waterlogging and canopy interact to control species recruitment in floodplains. Functional<br />

Ecology, doi: 10.1111/j.1365-2435.2009.01682.x.<br />

Lenssen, J.P.M., Ten Dolle, G.E. & Blom, C. (1998) The effect of flooding on the recruitment<br />

of reed marsh and tall forb plant species. Plant Ecology, 139, 13–23.<br />

Lenssen, J.P., van <strong>de</strong> Steeg, H.M. & <strong>de</strong> Kroon, H. (2004) Does disturbance favour weak<br />

competitors? Mechanisms of changing plant abundance after flooding. Journal of Vegetation<br />

Science, 15, 305–314.<br />

Peltzer, D.A. & Kochy, M. (2001) Competitive Effects of Grasses and Woody Plants in<br />

Mixed-Grass Prairie. Journal of Ecology, 89(4), 519-527.<br />

Pennings, S.C., Grant, M.B. & Bertness, M.D. (2005) Plant zonation in low-latitu<strong>de</strong> salt<br />

marshes: disentangling the roles of flooding, salinity and competition. Journal of Ecology, 93,<br />

159–167.<br />

141


Chapitre IV<br />

Santamaria, L., (2002) Why are most aquatic plants wi<strong>de</strong>ly distributed? Dispersal, clonal<br />

growth and small-scale heterogeneity in a stressful environment. Acta Oecologica, 23,137-<br />

154<br />

Silvertown, J., Dodd, M.E., Gowing, D.J. & Mountford, J.O. (1999) Hydrologically <strong>de</strong>fined<br />

niches reveal a basis for species richness in plant communities. Nature, 400, 61–63.<br />

Silvertown, J. (2004) Plant coexistence and the niche. Trends in Ecology & Evolution, 19,<br />

605–611.<br />

Soukupova, L. (1994) Allocation plasticity and modular structure in clonal graminoids in<br />

response to waterlogging. Folia Geobotanica Phytotaxa, 29, 227-236.<br />

Van Eck, W. H. J. M., van <strong>de</strong> Steeg, H.M., Blomand, C. W. P. M. & <strong>de</strong> Kroon, H. (2004) Is<br />

tolerance to summer flooding correlated with distribution patterns in river floodplains? A<br />

comparative study of 20 terrestrial grassland species. Oikos 107, 393-405.<br />

Violle, C., Cu<strong>de</strong>nnec, C., Plantgenest, M., Damgaard, C., Le Cœur, D., Bouzillé, J.B. &<br />

Bonis, A. 2007. Indirect assessment of flooding duration as a driving factor of plant diversity<br />

in wet grasslands. Proceedings of Symposium S7 held during the Seventh IAHS Scientific<br />

Assembly at Foz do Iguacu, Brazil,. IAHS Publication 303, 2006, pp. 334<br />

Violle, C., Pu, Z. & Jiang, L. (2010) Experimental <strong>de</strong>monstration of the importance of<br />

competition un<strong>de</strong>r disturbance. Proceedings of the National Aca<strong>de</strong>my of Sciences, 107,<br />

12925.<br />

Visser, E.J.W., Voesenek, L.A.C.J., Vartapetian, B.B. & Jackson, M.B. (2003) Flooding and<br />

plant growth. Annals of Botany, 91, 107-109.<br />

Wang, P., Stieglitz, T., Zhou, D.W. & Cahill, J.F. (2010). Are competitive effect and response<br />

two si<strong>de</strong>s of the same coin, or fundamentally different? Functional Ecology, 24, 196-207.<br />

Warwick, N.W. & Brock, M.A. (2003) Plant reproduction in temporary wetlands: the effects<br />

of seasonal timing, <strong>de</strong>pth, and duration of flooding. Aquatic Botany, 77, 153–167.<br />

Wel<strong>de</strong>n, C.W. & Slauson, W.L. (1986) The intensity of competition versus its importance: an<br />

overlooked distinction and some implications. Quarterly Review of Biology, 61, 23–44.<br />

Wolfram S (2003) Mathematica. Wolfram Research, Inc, Champaign, USA<br />

142


Appendix<br />

Material and Methods<br />

The characterization of the gradient<br />

Chapitre IV<br />

A precise characterization of these waterlogging and drought constraints experienced<br />

by plants was neglected before the work of Gowing et al. (1998). They proposed the<br />

characterization of these aeration and drought stresses by the <strong>de</strong>velopment of two indices: the<br />

aeration SEV (Sum Exceedance Value) and the drought SEV expressed in cm.day -1 or in<br />

m.week -1 . Each in<strong>de</strong>x consisted of the <strong>de</strong>termination of the timing and the extent of the<br />

constraint by the <strong>de</strong>termination of two thresholds from which stress occurs from March to<br />

September (Gowing et al., 1998; Barber et al., 2004; Gilbert et al., 2003). The aeration SEV<br />

reflects changes in microtopography: aeration mainly drove species success and community<br />

distribution whereas drought does not (Gowing et al., 1998).<br />

For that, dipwells were placed along the topographical gradient in the aim to survey water-<br />

table movements across time with automatic probes (Fig. 1). The SEV aeration was calculated<br />

as the difference between the water table <strong>de</strong>pth and a reference level following this equation<br />

(Swetnam et al., 1998):<br />

where<br />

where WTt is the water-table <strong>de</strong>pth at time t from March to September and RV the reference<br />

water-table <strong>de</strong>pth above which the plants are expected to be aeration-stressed. The reference<br />

water-table <strong>de</strong>pth was expressed as a threshold calculated from a soil moisture release curve<br />

as the <strong>de</strong>pth that gives 10% air-filled porosity inducing an insufficient oxygen diffusion to<br />

supply the respiratory <strong>de</strong>mands of roots during growth period (Gowing et al., 1998; Swetnam<br />

et al., 1998). The aeration threshold thus correspon<strong>de</strong>d to a water table <strong>de</strong>pth of -0.191 m:<br />

when water table level is above this threshold, aeration stress occurs. Similarly, the soil<br />

drying threshold corresponds to a water table <strong>de</strong>pth of -0.42 m: when water table level is<br />

below this threshold, drought stress occurs.<br />

143


Chapitre IV<br />

144


- Chapitre 5 -<br />

The <strong>de</strong>mography of space occupancy: measuring plant<br />

colonization and survival probabilities using repeated pin-point<br />

measurements<br />

Accepté dans Methods in Ecology and Evolution<br />

MEE (2011) 2: 100-115<br />

145


146


Chapitre V<br />

The <strong>de</strong>mography of space occupancy: measuring plant colonisation and<br />

survival probabilities using repeated pin-point measurements<br />

Christian Damgaard 1 , <strong>Amandine</strong> <strong>Merlin</strong> 2,3 , François Mesléard 3,4 , Anne Bonis 2<br />

1 Department of Terrestrial Ecology, NERI, Aarhus University, Vejlsøvej 25, 8600 Silkeborg,<br />

Denmark<br />

2 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, 35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

3 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc F-13200, France<br />

4 UMR CNRS-IRD 6116 IMEP, <strong>Université</strong> d’Avignon - IUT site Agroparc BP 1207, F-84911<br />

Avignon Ce<strong>de</strong>x 09, France<br />

Summary<br />

1. The study of plant <strong>de</strong>mography has long been an important component of plant ecological<br />

studies. However in plant communities, e.g. grasslands, where individual plants are not<br />

easily distinguished and often vary in size, a convenient method of <strong>de</strong>scribing the<br />

<strong>de</strong>mography and its ecological consequences has been lacking.<br />

2. The aim of the present paper is to discuss the potential for using the change in the<br />

probability of space occupancy as a measure of ecological success in plant population<br />

biology. This will be done by <strong>de</strong>monstrating how the change in the probability of space<br />

occupancy <strong>de</strong>pends on the processes of colonisation and survival and <strong>de</strong>monstrating how<br />

colonisation and survival probabilities may be estimated from repeated pin-point<br />

recordings at the same pin-position. Furthermore, it will be <strong>de</strong>monstrated how to calculate<br />

the sensitivity and elasticity of the change in the probability of space occupancy to<br />

colonisation and survival probabilities.<br />

3. The method is applied to a case of repeated pin-point data of the perennial grass<br />

Brachypodium phoenicoï<strong>de</strong>s that un<strong>de</strong>rwent different grazing regimes from 2001 to 2008<br />

on xero-halophytic grasslands.<br />

4. It was <strong>de</strong>monstrated that grazing affected both plant survival and colonisation probabilities<br />

so that the estimated colonisation and survival probabilities of B. phoenicoï<strong>de</strong>s were<br />

highest in the non-grazed regime. Furthermore, using the calculated elasticity of the change<br />

in space occupancy, we found that survival was more important than colonization events in<br />

<strong>de</strong>termining the ecological success of B. phoenicoï<strong>de</strong>s in the non-grazed regime, whereas<br />

colonization events were more important than survival in the grazed regime.<br />

5. Synthesis and applications. We expect that the relatively simple method may be wi<strong>de</strong>ly<br />

used on existing and future repeated pin-point recordings from the same pin-position and,<br />

consequently, that plant <strong>de</strong>mographic questions may be addressed with larger precision in<br />

plant communities where individual plants are not easily distinguished.<br />

Keywords: colonisation, <strong>de</strong>mography, elasticity, plant cover, sensitivity, survival<br />

147


Chapitre V<br />

148


Introduction<br />

Chapitre V<br />

Knowledge of recruitment and survival probabilities at different environmental<br />

conditions is important for un<strong>de</strong>rstanding fundamental plant ecological processes, which often<br />

play a significant role in the dynamics of the entire ecosystem (Harper, 1977). Recruitment<br />

and survival probabilities are, typically, estimated from <strong>de</strong>mographic data of individual<br />

plants. For example, Batista et al. (1998) studied the effect of a hurricane on the <strong>de</strong>mography<br />

of Fagus gradifolia trees by censusing individual trees. However, in many natural plant<br />

communities dominated by perennial plants, e.g. grasslands, it is often difficult to distinguish<br />

individual plants due to their vegetative growth pattern and, consequently, to obtain reliable<br />

<strong>de</strong>mographic data at the level of the individual. Furthermore, once individual plants can be<br />

distinguished, they often vary markedly in size so that the number of individuals is<br />

insufficient for <strong>de</strong>scribing the plant population. Instead, plant abundance may be <strong>de</strong>scribed by<br />

the cover, i.e. the relative area covered by different plant species, rather than using the number<br />

of individuals (Kent and Coker, 1992). Plant cover data may be used to classify the studied<br />

plant community into a vegetation type, to test different ecological hypotheses on plant<br />

abundance, and in gradient studies, where the effects of different environmental gradients on<br />

the abundance of specific plant species are investigated (Damgaard, 2008, Damgaard, 2009,<br />

Damgaard and Ejrnæs, 2009).<br />

A popular and objective method of measuring plant cover is the pin-point method, also<br />

known as the point-intercept method (Kent and Coker, 1992, Levy and Mad<strong>de</strong>n, 1933). In a<br />

pin-point analysis, a frame with a number of fixed pin positions is placed above the vegetation<br />

and a pin is inserted vertically into the vegetation at each pin position. The cover of plant<br />

species A is measured by , where is the number of pins that hit species A out of a total<br />

of n pins. Since a single pin often will hit more than a single species, the sum of the plant<br />

cover of the different species may be larger than unity when estimated by the pin-point<br />

method. The sum of the estimated plant cover is expected to increase with the number of plant<br />

species in a plot and with increasing 3-dimensional structuring of the plants in the<br />

community.<br />

Instead of calculating plant cover by estimating the mean probability of being hit by<br />

one of the n pins, it is possible to consi<strong>de</strong>r each pin-position separately and <strong>de</strong>fine the<br />

probability of space occupancy by plant species A at the specific pin-position. If the location<br />

of the pin-point frame is marked on the ground so that the pin-point frame may be placed at<br />

exactly the same location the next time the plot is visited, then it is possible to track the fate of<br />

149


Chapitre V<br />

a species at a specific pin-position through time. If the species was present at time t but absent<br />

at time t +1, we may loosely speak of a mortality event, and if the species is absent from a<br />

specific pin-position at time t and present at time t +1, we may loosely speak of a colonisation<br />

event. We have chosen to use the term “colonisation” rather than “recruitment” since the<br />

event is <strong>de</strong>fined by a novel occurrence in space (Adler et al., 2006, Adler et al., 2009).<br />

However, in the interpretation of the results, it is important to remember that the concepts of<br />

colonisation and mortality have a different meaning than usual in studies where individuals<br />

are consi<strong>de</strong>red. For perennial plant species that spread clonally by forming well <strong>de</strong>fined<br />

ramets, the concepts of colonisation and mortality at a certain pin position make apparent<br />

biological sense whereas, for species characterised by more continuous plant growth, the<br />

concepts are ina<strong>de</strong>quate <strong>de</strong>scriptions of the un<strong>de</strong>rlying biological causes of a change in plant<br />

abundance. Bearing this issue of terminology in mind, we may, generally, assert that the<br />

probability of space occupancy of a plant species will increase with colonisation and <strong>de</strong>crease<br />

with mortality.<br />

The quantitative effect of birth and <strong>de</strong>ath processes on population growth of<br />

individuals is most effectively summarised using matrix population mo<strong>de</strong>ls (Caswell, 2001),<br />

and the calculation of sensitivity and elasticity of different birth and <strong>de</strong>ath processes on<br />

population growth rates has particularly been shown to be a powerful tool to investigate the<br />

importance of different <strong>de</strong>mographic variables in <strong>de</strong>termining population growth. For<br />

example, Dalgleish et al. (2010) used elasticity calculations to conclu<strong>de</strong> that climate change<br />

was more likely to influence population growth through effects on recruitment than through<br />

survival for 10 short-lived forb species. Likewise, when studying the effects of a variable<br />

environment, which may affect both colonisation and survival probabilities, it will be<br />

important to be able to calculate the sensitivity and elasticity of the change in the probability<br />

of space occupancy as influenced by variable colonisation and survival. Such sensitivity and<br />

elasticity calculations have not been discussed before.<br />

The aim of the present paper is, therefore, to discuss the potential of using the change<br />

in the probability of space occupancy as a measure of ecological success in plant population<br />

biology. This will be done by i) showing how the change in the probability of space<br />

occupancy <strong>de</strong>pends on colonisation and survival, ii) <strong>de</strong>monstrating how colonisation and<br />

survival probabilities may be estimated from repeated recordings at the same pin-position,<br />

and iii) calculating the sensitivity and elasticity of the change in the probability of space<br />

occupancy to colonisation and survival probabilities.<br />

150


Mo<strong>de</strong>l<br />

Chapitre V<br />

If absence-presence data of species A from two successive recordings from the same<br />

pin-position are consi<strong>de</strong>red, there are four possible transition events and corresponding<br />

probabilities (Table 1). These transition probabilities <strong>de</strong>pend on i) the probability (p) that a<br />

plant of species A is present at time t, ii) survival; the probability (s) that a plant of species A<br />

is present at the pin-position at both time t and time t+1, and iii) colonisation; the probability<br />

(c) that a plant of species A is present at the pin-position at time t+1 but was absent at time t.<br />

Table 1. The four possible events of two successive recordings of presence absence data and<br />

their corresponding probabilities, p: current probability, c: colonisation probability, s: survival<br />

probability.<br />

Event Description Probability<br />

X1: At , At+1<br />

A plant of species A was present in year t and was also<br />

present in year t+1<br />

X2: nAt , nAt+1 A plant of species A was not present in year t and was<br />

X3: At , nAt+1<br />

X4: nAt , At+1<br />

also not present in year t+1<br />

A plant of species A was present in year t but was not<br />

present in year t+1<br />

A plant of species A was not present in year t but was<br />

present in year t+1<br />

The change in the probability that species A is present in year t to t +1 is <strong>de</strong>noted by<br />

and may be <strong>de</strong>fined, analogous to the population growth rate λ of individuals (Caswell, 2001),<br />

as the ratio between the probability that species A is present at time t + 1 and the<br />

probability p that species A is present at time t :<br />

where are <strong>de</strong>fined in Table 1.<br />

(1),<br />

151


Chapitre V<br />

From the above <strong>de</strong>finition, it is apparent that if , then the probability that species<br />

A is present <strong>de</strong>creases, and if , then the probability that species A is present increases.<br />

Furthermore, the change in the probability that species A is present is a function of the current<br />

probability (p), the colonisation probability and the survival probability probabilities (Fig 1).<br />

Note also, that the plant cover of species A at time t is estimated by taking the mean of p<br />

across space.<br />

Fig. 1. The change in the probability that species A is present ( ) as a function of the<br />

colonisation (c) and survival (s) probabilities. The current probability that species A is present<br />

(p) is set to 0.5.<br />

The change in the probability of being present, as <strong>de</strong>fined above in equation (1), is<br />

always positive, which assures that it is possible to calculate both the sensitivity and elasticity<br />

of the change (Caswell, 2001). The sensitivity of the change in the probability that species A<br />

is present is a function of the colonisation and survival probabilities which are <strong>de</strong>fined as:<br />

Likewise, the elasticity of the change in the probability that species A is present is a<br />

function of the colonisation and survival probabilities which are <strong>de</strong>fined as:<br />

(2).<br />

152


Estimation and statistical inferences<br />

(3).<br />

Chapitre V<br />

The colonisation and survival probabilities may be estimated from data of the four<br />

possible events of two successive recordings of presence-absence data ,<br />

which are <strong>de</strong>fined in Table 1, by maximising the likelihood function of the multinomial<br />

distribution using the transition probabilities specified in Table 1, and assuming that<br />

the plants are distributed homogenously across the site of the investigation, such that the<br />

probability that species A is present at time t may be estimated<br />

as :<br />

or by simulating the Bayesian joint posterior distribution of the colonisation and survival<br />

probabilities using MCMC methods (Carlin and Louis, 1996), where the prior distribution of<br />

the colonisation and survival probabilities may be assumed to be beta-distributed (Chew,<br />

1971).<br />

Different hypotheses on the colonisation and survival probabilities may be tested by<br />

standard hierarchical likelihood ratio procedures or by comparing different joint posterior<br />

distributions of the parameters. Furthermore, the credibility interval of the sensitivity and<br />

elasticity may be calculated from the joint posterior distribution of the colonisation and<br />

survival probabilities (Ghosh et al., 2006).<br />

Example<br />

In or<strong>de</strong>r to <strong>de</strong>monstrate the potential of the method, repeated pin-point data of the<br />

dominant perennial grass Brachypodium phoenicoï<strong>de</strong>s on xero-halophytic grasslands at the<br />

Tour du Valat estate, Rhône <strong>de</strong>lta, Southern France (43°29’ N, 4°40’ E) were analysed. The<br />

xero-halophytic grasslands, which are a priority habitat (co<strong>de</strong> 6220) in the European Union<br />

(4),<br />

153


Chapitre V<br />

Habitats Directive (1992) and the richest habitat of the Camargue (Molinier and Tallon,<br />

1968), are managed by extensive livestock grazing using traditional methods (Otero and<br />

Bailey, 2003). The xero-halophytic grasslands are distributed along an elevation gradient with<br />

variable salinity, and they experience large variation in annual rain fall (Fig 2).<br />

Fig 2 : Histogram reprensenting the mean probability of space occupancy (= plant cover) of B.<br />

phoenicoï<strong>de</strong>s in non-grazed (dark grey) and grazed regimes (light gray) from 2001 to 2008<br />

plotted together with the annual rainfall (black line) at Tour du Valat Station, Camargue,<br />

France.<br />

In autumn 2001, an experimental <strong>de</strong>sign of paddocks with three different grazing<br />

regimes was set up in or<strong>de</strong>r to study the effect of domestic grazing on space occupancy and<br />

population persistence. Two of those grazing regimes are consi<strong>de</strong>red in this example: (i) an<br />

extensive grazing pressure that has been applied continuously since 1970 and (ii) a treatment<br />

where grazing was prevented by fencing at the start of the experiment in 2001. Twenty-one<br />

permanent quadrates of 0.16 m² were placed along the elevation gradient in each grazing<br />

regime. In these permanents quadrates, a set of crossing lines was used to position 36 pins so<br />

that the distance between the pins along the crossing line was three centimetres. Pin-point<br />

cover data were collected in April over an eight year period, from 2001 to 2008.<br />

Two different plant community types may be discriminated along the elevation<br />

gradient (Mesléard et al., 1991), which loosely may be categorised as either grassland or salt<br />

marches, and it is the grassland plant community that is dominated by B. phoenicoï<strong>de</strong>s, which<br />

is investigated in this example with five quadrates in the grazed regime and sixteen quadrates<br />

in the non-grazed regime.<br />

154


Chapitre V<br />

For both grazing regimes the colonisation and survival probabilities were estimated<br />

from two successive yearly recordings of B. phoenicoï<strong>de</strong>s and the sensitivities and elasticities<br />

of the change in space occupancy were calculated (Table 2).<br />

Table 2. The estimated colonisation and survival probabilities of B. phoenicoï<strong>de</strong>s, the change<br />

in the probability of space occupancy ( ), and the calculated sensitivities and elasticities of<br />

the change in the probability of space occupancy to colonisation and survival in the two<br />

grazing regimes.<br />

Grazing<br />

regime<br />

Non-<br />

grazed<br />

Grazed<br />

Year Colonisation Survival Colonisation<br />

sensitivity<br />

Colonisation<br />

elasticity<br />

Survival<br />

sensitivity<br />

Survival<br />

elasticity<br />

2001/2002 0.35 0.66 1.51 2.42 0.56 0.65 0.28<br />

2002/2003 0.65 0.85 1.62 1.19 0.47 0.36 0.19<br />

2003/2004 0.20 0.69 0.80 0.57 0.14 0.80 0.68<br />

2004/2005 0.33 0.71 1.00 0.86 0.29 0.67 0.47<br />

2005/2006 0.57 0.91 1.29 0.66 0.29 0.43 0.30<br />

2006/2007 0.22 0.84 0.93 0.37 0.09 0.78 0.71<br />

2007/2008 0.27 0.73 0.89 0.58 0.18 0.73 0.60<br />

mean 0.37 0.77 1.15 0.95 0.29 0.63 0.46<br />

CV 46.43 12.36 28.15 73.41 60.24 27.45 45.18<br />

2001/2002 0.20 0.44 0.86 2.06 0.49 0.80 0.41<br />

2002/2003 0.38 0.77 1.58 2.14 0.52 0.62 0.30<br />

2003/2004 0.20 0.45 0.72 1.38 0.37 0.81 0.51<br />

2004/2005 0.20 0.51 0.92 2.03 0.45 0.80 0.44<br />

2005/2006 0.23 0.56 1.08 2.21 0.48 0.77 0.40<br />

2006/2007 0.20 0.29 0.74 2.29 0.62 0.80 0.31<br />

2007/2008 0.27 0.44 1.27 3.02 0.65 0.73 0.25<br />

mean 0.24 0.49 1.02 2.16 0.51 0.76 0.37<br />

CV 28.05 29.58 30.35 22.26 18.77 8.93 23.89<br />

Grazing affected plant survival and colonisation probabilities, and the estimated<br />

colonisation and survival probabilities of B. phoenicoï<strong>de</strong>s were highest in the non-grazed<br />

regime. Moreover, the estimated colonisation probabilities were more variable across years<br />

than the estimated survival probabilities, particularly in the non-grazed regime, and it was<br />

hypothesised that these variations in the colonisation probabilities may be linked to variation<br />

in annual rainfall (Fig. 2). It was expected that high water availability would increase<br />

colonisation probabilities, but there was no significant correlation between annual rainfall and<br />

155


Chapitre V<br />

colonisation probability (Spearman correlations; non-grazed regime: P = 0.302 and grazed<br />

regime: P=0.435). This unexpected result may be due to <strong>de</strong>nsity-<strong>de</strong>pen<strong>de</strong>nce which may<br />

induce a temporal variation in the estimated probabilities (Grant and Benton, 2000) due to<br />

overlapping generations.<br />

Based on the calculated elasticity of the change in space occupancy (Table 2), it was<br />

found that survival was more important than colonization events in <strong>de</strong>termining the ecological<br />

success of B. phoenicoï<strong>de</strong>s in the non-grazed regime, i.e. low mortality played a major role in<br />

the increase of the cover of B. phoenicoï<strong>de</strong>s from 2001 to 2008 (Fig. 2). Oppositely,<br />

colonization events were found to be more important than survival in the grazed regime.<br />

In the non-grazed regime, sensitivity and elasticity of estimated probabilities were<br />

inconsistent, i.e. colonisation seemed to be more important than mortality according to the<br />

sensitivity calculations in the first years after cessation of grazing, whereas the contrary was<br />

observed when elasticity was used.<br />

Discussion<br />

Space occupancy in <strong>de</strong>nse vegetation is a sign of successful colonisation and survival<br />

of individual plants, and the probability of space occupancy, or plant cover, is, therefore, a<br />

relevant measure of the ecological success of plant species (Damgaard et al., 2009).<br />

Furthermore, an increased emphasis on the empirical study of the spatial dynamics of plant<br />

populations is in excellent agreement with the recent <strong>de</strong>velopment in the theoretical<br />

un<strong>de</strong>rstanding on the spatial dynamics of plant communities (e.g. Law et al., 1997, Bolker and<br />

Pacala, 1999, Dieckmann et al., 2000). Here, theoretical and empirical plant ecology are<br />

linked by showing how a change in the probability of space occupancy <strong>de</strong>pends on the<br />

ecological processes of colonisation and survival, and by introducing methods for calculating<br />

the sensitivity and elasticity of the change in the probability of space occupancy on the<br />

ecological processes. The two measures to compare the effect of colonisation and survival on<br />

space occupancy complement each other, but since sensitivity is an absolute measure and is<br />

less robust than elasticity; elasticity is often the favoured measure when comparing the effect<br />

of <strong>de</strong>mographic parameters between and within species (<strong>de</strong> Kroon et al., 1986, van<br />

Groenendael et al., 1994).<br />

As stated in the introduction, the concepts of mortality and colonisation are only<br />

loosely connected to the usual biological <strong>de</strong>finitions of the two concepts that are tightly linked<br />

156


Chapitre V<br />

to the <strong>de</strong>scription of the fate of individual plants. For example, the estimated survival<br />

probability inclu<strong>de</strong>s both true survival as well as mortality followed by colonisation, and the<br />

colonisation probability aggregates both vegetative and reproductive growth. For some<br />

applications and for some species, such an aggregation of the <strong>de</strong>mographic processes may<br />

obscure a more mechanistic un<strong>de</strong>rstanding of the ecological processes occurring at the level<br />

of individual plants.<br />

In the presented methodology, it is <strong>de</strong>monstrated how colonisation and survival<br />

probabilities may be estimated from repeated recordings at the same pin-position, and it is<br />

implicitly assumed that the plant is either present or absent at the pin-position. However,<br />

consi<strong>de</strong>r a plant that is rooted some distance away for the pin-position; in favourable years,<br />

the leaves of such a plant may grow sufficiently to be hit by the pin, whereas in unfavourable<br />

years, growth may be insufficient to be hit by the pin. Such a plant will be misclassified in the<br />

pin-point analysis and this misclassification will lead to an upwards biased estimate of<br />

colonisation probability and a downwards biased estimate of survival probability. A similar<br />

effect will arise if the pin-point frame is not situated exactly the same place each year. The<br />

possible effect of such misclassification is expected to <strong>de</strong>pend on the phenology and growth<br />

pattern of the specific plant species and the consequences for the ecological interpretation will<br />

have to be evaluated for each species and may be estimated with confi<strong>de</strong>nce intervals of space<br />

occupancy calculation using re-sampling methods as proposed by Valver<strong>de</strong> and Silvertown<br />

(1998).<br />

The method may be generalised so that the survival probability varies with plant age,<br />

i.e. the survival probability is assumed to <strong>de</strong>pend on the uninterrupted number of years the<br />

species is present at the position. Such a generalisation of the method may be especially<br />

important when the effect of e.g. extreme climatic events or other forms of disturbances on<br />

plant mortality and recruitment <strong>de</strong>pend on plant age or size (e.g. O'Connor, 1993).<br />

In the present study, it is assumed that the plants are distributed homogenously across<br />

the site of investigation, but if this assumption is known to not be approximately true, then it<br />

may be appropriate to partition the site of investigation into areas with approximately equal<br />

local <strong>de</strong>nsities and estimate the colonisation and survival probabilities for each area.<br />

Additionally, it may be beneficial to characterise the spatial variation using methods of spatial<br />

point pattern analysis (Diggle, 2003, Wiegand et al., 2007, Law et al., 2009). Furthermore, in<br />

the present analysis, the possible competitive effects among neighbouring plants are inclu<strong>de</strong>d<br />

in the estimated probabilities of survival and colonisation for each species, instead of<br />

examining the spatial intra- and inter-specific interactions among the studied plant species<br />

157


Chapitre V<br />

(e.g. Rees et al., 1996, Law et al., 1997, Adler et al., 2009). The mo<strong>de</strong>l choice whether to<br />

inclu<strong>de</strong> the competitive effects among neighbouring plants into “non-spatial” probabilities of<br />

mortality and colonisation, or to examine the competitive interaction in a spatial competition<br />

mo<strong>de</strong>l, hinges on i) the spatial <strong>de</strong>sign, for example, if the pin-point data arise from linear<br />

transects, as in the present case, or if the distance between the pins is relatively large, then the<br />

spatial information is not sufficiently <strong>de</strong>tailed to allow a spatial analysis of the effect of<br />

neighbouring plants; ii) statistical power, since the spatial mo<strong>de</strong>ls need more parameters than<br />

the non-spatial mo<strong>de</strong>l, the size of the <strong>de</strong>sign may be insufficient to mo<strong>de</strong>l the spatial inter-<br />

specific interactions; iii) the number of species, if the number of species in the plant<br />

community is relatively high and the species-specific competition kernels are expected to vary<br />

among species, then it may be unmanageable to mo<strong>de</strong>l the spatial inter-specific interaction<br />

and instead it may be useful, as in the present study, to consi<strong>de</strong>r the neighbouring vegetation<br />

as a variable matrix that influences mortality and colonisation; iv) the scientific question, for<br />

some scientific question species, specific probabilities of survival and colonisation may be the<br />

most relevant measure of the ecological success.<br />

In this manuscript, space occupancy data are assumed to be sampled by the pin-point<br />

method, which is especially relevant for plant species with a relatively high cover. However,<br />

for a long time, it has been technically possible to measure plant cover in continuous two-<br />

dimensional space by using a pantograph (Adler et al., 2006, Adler et al., 2009), but since the<br />

pantograph method is cumbersome, such plant cover data are relatively rare compared to pin-<br />

point data. In the near future, the use of digital imagery for measuring the change in plant<br />

cover through time is expected to become much more important (e.g. Seefeldt and Booth,<br />

2006) and the forthcoming of such plant cover data will allow a more sophisticated mo<strong>de</strong>lling<br />

of plant <strong>de</strong>mography, e.g. by including the effect of competitive interactions (Adler et al.,<br />

2006, Adler et al., 2009).<br />

Acknowledgements<br />

This work is a publication from the DIVHERB Project (French national program ECOGER<br />

fun<strong>de</strong>d by the Institut National <strong>de</strong> la Recherche Agronomique) and a contribution to GDR<br />

2574 “TRAITS”. Also thanks to College Doctoral International <strong>de</strong> Bretagne for providing a<br />

mobility grant, and two anonymous reviewers and Charlotte Elisabeth Kler for comments on a<br />

previous version of the manuscript.<br />

158


References<br />

Chapitre V<br />

Adler, P. B., HilleRisLambers, J., Kyriakidis, P. C., Guan, Q. & Levine, J. M. (2006) Climate<br />

variability has a stabilizing effect on the coexistence of prairie grasses. Proceedings of the<br />

National Aca<strong>de</strong>my of Sciences (USA), 103, 12793-12798.<br />

Adler, P. B., HilleRisLambers, J. & Levine, J. M. (2009) Weak effect of climate variability on<br />

coexistince in a sagebrush steppe community. Ecology, 90, 3303-3312.<br />

Batista, W. B., Platt, W. J. & Macchiavelli, R. E. (1998) Demography of a sha<strong>de</strong>-tolerant tree<br />

(Fagus grandifolia) in a hurricane-disturbed forest. Ecology, 79, 38–53.<br />

Bolker, B. M. & Pacala, S. W. (1999) Spatial moment equations for plant competition:<br />

Un<strong>de</strong>rstanding spatial strategies and the advantages of short dispersal. American Naturalist,<br />

153, 575-602.<br />

Carlin, B. P. & Louis, T. A. (1996) Bayes and empirical Bayes methods for data analysis.<br />

Chapman & Hall, London.<br />

Caswell, H. (2001) Matrix population mo<strong>de</strong>ls: Construction, analysis, and interpretation.<br />

Sinauer, Sun<strong>de</strong>rland.<br />

Chew, V. (1971) Point estimation of the parameter of the binomial distribution. American<br />

Statistician, 25, 47-50.<br />

Dalgleish, H. J., Koons, D. N. & Adler, P. B. (2010) Can life-history traits predict the<br />

response of forb populations to changes in climate variability? Journal of Ecology, 98 209–<br />

217.<br />

Damgaard, C. (2008) Mo<strong>de</strong>lling pin-point plant cover data along an environmental gradient.<br />

Ecological Mo<strong>de</strong>lling, 214, 404-410.<br />

Damgaard, C. (2009) On the distribution of plant abundance data. Ecological Informatics, 4,<br />

76–82.<br />

Damgaard, C. & Ejrnæs, R. (2009) Quantification of the intra-plot correlation in plant<br />

abundance data: A possible test of the neutral theory. Ecological Complexity, 6, 64-69.<br />

Damgaard, C., Riis-Nielsen, T. & Schmidt, I. K. (2009) Estimating plant competition<br />

coefficients and predicting community dynamics from non-<strong>de</strong>structive pin-point data: a case<br />

study with Calluna vulgaris and Deschampsia flexuosa. Plant Ecology, 201, 687–697.<br />

<strong>de</strong> Kroon, H., Plaisier, A., van Groenendael, J. & Caswell, H. (1986) Elasticity: the relative<br />

contribution of <strong>de</strong>mographic parameters to population growth rate. Ecology, 67, 1427-1431.<br />

Dieckmann, U., Law, R. & Metz, J. A. J. (2000) The geometry of ecological interactions:<br />

Simplifying spatial complexity. Cambridge University Press, Cambridge.<br />

Diggle, P. J. (2003) Statistical analysis of spatial point patterns. Arnold, London.<br />

EU (1992) Council Directive 92/43/EEC of 21 May 1992 on the conservation of natural<br />

habitats and of wild fauna and flora. (ed E. Commission).<br />

Ghosh, J. K., Delampady, M. & Samanta, T. (2006) An introduction to Bayesian analysis.<br />

Theory and methods. Springer, New York.<br />

159


Chapitre V<br />

Grant, A. & Benton, T. G. (2000) Elasticity analysis for <strong>de</strong>nsity-<strong>de</strong>pen<strong>de</strong>nt populations in<br />

stochastic environments. Ecology, 81, 680-693.<br />

Harper, J. L. (1977) Population biology of plants. Aca<strong>de</strong>mic Press, London.<br />

Kent, A. & Coker, P. (1992) Vegetation <strong>de</strong>scription and analysis - A practical approach. John<br />

Wiley & Sons, New York, USA.<br />

Law, R., Herben, T. & Dieckmann, U. (1997) Non-manipulative estimates of competition<br />

coefficients in a montane grassland community. Journal of Ecology, 85, 505-517.<br />

Law, R., Illian, J., Burslem, D. F. R. P., Georg Gratzer, G., Gunatilleke, C. V. S. &<br />

Gunatilleke, I. A. U. N. (2009) Ecological information from spatial patterns of plants: insights<br />

from point process theory. Journal of Ecology, 97 616–628.<br />

Levy, E. B. & Mad<strong>de</strong>n, E. A. (1933) The point method of pasture analyses. New Zealand<br />

Journal of Agriculture, 267-279.<br />

Mesléard, F., Grillas, P. & Lepart, J. (1991) Plant community succession in a coastal wetland<br />

after abandonment of cultivation: the example of the Rhône <strong>de</strong>lta. Vegetatio, 94, 35-45.<br />

Molinier, R. & Tallon, G. (1968) Etu<strong>de</strong>s botaniques en Camargue. II vers la forêt en<br />

Camargue. Revue d’Ecologie (Terre Vie), 19, 3-197.<br />

O'Connor, T. G. (1993) The influence of rainfall and grazing on the <strong>de</strong>mography of some<br />

African savanna garsses: a matrix mo<strong>de</strong>lling approach. Journal of Applied Ecology, 30, 119-<br />

132.<br />

Otero, C. & Bailey, T. (2003) La Tour du Valat, France, European natural and cultural<br />

heritage. Friends of the countrysi<strong>de</strong>pp. 463-467. The European Estate Friends of the Country<br />

si<strong>de</strong>, Brussel.<br />

Rees, M., Grubb, P. J. & Kelly, D. (1996) Quantifying the impact of competition and spatial<br />

heterogeneity on the structure and dynamics of a four-species guild of winter annuals.<br />

American Naturalist, 147, 1-32.<br />

Seefeldt, S. S. & Booth, D. T. (2006) Measuring plant cover in sagebrush steppe rangelands:<br />

A comparison of methods. Environmental Management, 37, 703-711.<br />

Valver<strong>de</strong>, T. & Silvertown, J. (1998) Variation in the <strong>de</strong>mography of a woodland un<strong>de</strong>rstorey<br />

herb (Primula vulgaris) along the forest regeneration cycle: projection matrix analysis.<br />

Journal of Ecology, 86, 545-562.<br />

van Groenendael, J., <strong>de</strong> Kroon, H., Kalisz, S. & Tuljapurkar, S. (1994) Loop analysis:<br />

evaluating life history pathways in population projection matrices. . Ecology, 75, 2410-2415.<br />

Wiegand, T., Gunatilleke, S., Gunatilleke, N. & Okuda, T. (2007) Analysing the spatial<br />

structure of a Sri Lankan tree species with multiple scales of clustering. Ecology, 88, 3088–<br />

3102.<br />

160


RESUME PARTIE 2<br />

Dans ces <strong>de</strong>ux chapitres, la métho<strong>de</strong> <strong>de</strong>s points-contact a été utilisée pour déterminer<br />

les probabilités <strong>de</strong> survie et <strong>de</strong> colonisation <strong>de</strong>s espèces, ainsi que la croissance en biomasse, à<br />

partir du recouvrement d’une espèce et du volume occupé par l’espèce.<br />

Dans les prairies humi<strong>de</strong>s du Marais Poitevin (Chapitre 4), un modèle a été construit<br />

afin d’estimer <strong>de</strong>ux composantes <strong>de</strong> la compétition : l’effet compétitif <strong>de</strong>s espèces et<br />

l’importance <strong>de</strong> la compétition. Ces <strong>de</strong>ux composantes ont été estimées sur <strong>de</strong>ux pério<strong>de</strong>s<br />

d’étu<strong>de</strong> : la pério<strong>de</strong> d’inondation et la saison estivale faisant suite à la pério<strong>de</strong> d’inondation.<br />

Cette étu<strong>de</strong> montre l’intérêt <strong>de</strong> tenir compte <strong>de</strong>s paramètres démographiques <strong>de</strong>s espèces pour<br />

comprendre la capacité <strong>de</strong>s espèces à être compétitive face aux voisins. En effet l’importance<br />

<strong>de</strong> la survie <strong>de</strong>s espèces aux inondations et à leur reprise via la colonisation du milieu est<br />

montrée pour que l’espèce puisse être compétitive notamment après la fin <strong>de</strong>s inondations.<br />

Cette étu<strong>de</strong> met en évi<strong>de</strong>nce le rôle structurant <strong>de</strong>s inondations qui affectent la survie et la<br />

colonisation <strong>de</strong>s espèces, ainsi que le rôle structurant <strong>de</strong> la compétition qui augmente avec<br />

l’arrêt <strong>de</strong>s inondations.<br />

Le chapitre 5 propose une analyse d’élasticité <strong>de</strong>s paramètres démographiques dans le<br />

but <strong>de</strong> comprendre quel paramètre démographique influence le plus fortement la dynamique<br />

<strong>de</strong>s populations, et plus particulièrement la probabilité <strong>de</strong>s espèces pérennes à occuper<br />

l’espace. Cette analyse d’élasticité réalisée sur une espèce pérenne méditerranéenne,<br />

Brachypodium phoenicoï<strong>de</strong>s, a mis en évi<strong>de</strong>nce un effet <strong>de</strong> la gestion <strong>de</strong> pâturage sur la<br />

probabilité <strong>de</strong> cette espèce à occuper l’espace. En effet, en milieu non pâturé, la survie <strong>de</strong><br />

cette espèce expliquerait son succès à occuper l’espace et par conséquent son augmentation en<br />

abondance au cours <strong>de</strong>s années. En milieu pâturé, la probabilité <strong>de</strong> colonisation expliquerait<br />

plus la dynamique <strong>de</strong> cette espèce. Ces résultats suggèrent un effet négatif <strong>de</strong> la présence <strong>de</strong>s<br />

herbivores sur la survie <strong>de</strong> cette espèce.<br />

161


162


Partie 3 :<br />

La part <strong>de</strong>s filtres abiotiques et biotiques dans<br />

la structure <strong>de</strong>s communautés végétales<br />

163


164


- Chapitre 6 -<br />

Detecting community assembly patterns in floo<strong>de</strong>d and productive<br />

grasslands<br />

En préparation<br />

165


166


Chapitre VI<br />

Detecting community assembly patterns in floo<strong>de</strong>d and productive<br />

grasslands.<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1, 2 , Jan-Bernard Bouzillé 1 , François Mesléard 2, 3 , Anne Bonis 1<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, F-35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc, F-13200, France<br />

3 UMR CNRS-IRD 6116 Institut Méditerranéen d’Ecologie et <strong>de</strong> Paléoécologie, <strong>Université</strong><br />

d’Avignon IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09, France<br />

Abstract<br />

Trait-based approaches are currently <strong>de</strong>veloped to study mechanisms un<strong>de</strong>rlying<br />

species coexistence. Few studies have investigated the distribution of functional traits with<br />

respect to theoretical predictions formulated according to a given environmental gradient. In<br />

wet and productive grasslands located on the French Atlantic coast, we investigated whether<br />

plant species assembly show non-random organization with respect to Specific Leaf Area<br />

(SLA) and plant height, two functional traits associated with tolerance to flooding and<br />

resource acquisition. We used a null mo<strong>de</strong>l approach to test: (i) convergent trait values<br />

resulting from habitat filtering process in response to flooding conditions, (ii) divergent trait<br />

values resulting from niche differentiation process, and (iii) convergent trait values attributed<br />

to competition in highly competitive environments.<br />

Significant patterns of mean traits values were <strong>de</strong>tected for SLA and plant height;<br />

traits values increased with flooding conditions. Significant patterns of divergent trait values<br />

between nearest neighbors were <strong>de</strong>tected for SLA and plant height. This divergence in plant<br />

height was higher in flooding environments than in non-floo<strong>de</strong>d environments.<br />

Simultaneously, significant pattern of traits values convergence between nearest neighbors<br />

was found along the gradient for plant height in non-floo<strong>de</strong>d environment.<br />

Our results may be interpreted as evi<strong>de</strong>nce for habitat filtering selecting species<br />

according to their ability to tolerate flooding conditions. Competition between coexisting<br />

species may have led to a functional niche differentiation along the gradient, but coexisting<br />

species require a certain <strong>de</strong>gree of similarity to coexist in highly competitive environments.<br />

Our findings support hypotheses tested concerning traits distribution and reveal<br />

mechanisms responsible for species assembly along the flooding gradient.<br />

Keywords: Community assembly, habitat filtering, niche differentiation, functional<br />

convergence, plant height, Specific Leaf Area<br />

167


Chapitre VI<br />

168


Introduction<br />

Chapitre VI<br />

Un<strong>de</strong>rstanding processes un<strong>de</strong>rlying community assembly along environmental<br />

gradients and predicting which subset of the regional species pool occurs in a specified<br />

habitat, <strong>de</strong>fying the assembly rules, are one of the main aims of community ecology<br />

(Diamond 1975; Keddy 1992; Weiher & Keddy 1995; McGill et al. 2006). The <strong>de</strong>velopment<br />

of trait-based approach, studying the traits organization within and among communities, has a<br />

great potential to reveal general processes (Weiher et al. 1998; Mc Gill et al. 2006; Ackerly &<br />

Cornwell 2007; Violle & Jiang 2009). The hypothesis of the trait-based approach is that<br />

functional traits reflect species responses to environmental factors, i.e. response traits<br />

(Lavorel & Garnier 2002; Wright et al. 2004; McGill et al. 2006; Ackerly & Cornwell 2007;<br />

Diaz et al. 2007). Despite the recent and strong <strong>de</strong>velopment of trait-based approaches to test<br />

assembly rules, few studies investigated the distribution of functional traits with respect to<br />

theoretical predictions formulated according to a given environmental gradient (Pillar et al.<br />

2009; Schamp et al. 2010; Pakeman et al. 2011).<br />

Along environmental gradients, two main processes un<strong>de</strong>rline assembly rules; habitat<br />

filtering and niche differentiation. However, they do not have the same importance according<br />

to the location along the gradient (Weiher & Keddy 1995). In<strong>de</strong>ed, habitat filtering is<br />

predicted in the constrained parts of the gradient which consist in a convergence of species<br />

traits associated with the tolerance to abiotic factors among coexisting species (Weiher &<br />

Keddy 1995). Accordingly, species are selected according to their trait values allowing them<br />

to persist un<strong>de</strong>r given environmental conditions. This was notably <strong>de</strong>monstrated for Specific<br />

Leaf Area (SLA) and plant height in wetlands (Weiher et al. 1998; Jung et al. 2010), and for<br />

SLA and canopy height in coastal forest (Cornwell & Ackerly 2009). In the less constrained<br />

parts of the gradient, the distribution of traits values is expected to be mainly driven by<br />

competition for limiting resources (Weiher & Keddy 1995). Accordingly, divergence in traits<br />

associated with competition is interpreted as evi<strong>de</strong>nce of niche differentiation among co-<br />

occurring species, thus reflecting the concept of limiting similarity (Mac Arthur & Levins<br />

1967). This process exclu<strong>de</strong>s the most similar species (Connell 1980), thereafter reducing<br />

competition and permitting species coexistence. Such a niche differentiation has been<br />

<strong>de</strong>monstrated, for instance, consi<strong>de</strong>ring water-use strategies in sand dunes community (Stubbs<br />

& Wilson 2004), growth forms in tropical forests (Kraft et al. 2008) and plant height and SLA<br />

in wetlands (Weiher et al. 1998; Jung et al. 2010).<br />

169


Chapitre VI<br />

In contrast to Weiher & Keddy’s point of view, Grime (2006) predicts a convergence<br />

in traits within productive and undisturbed communities, illustrating by the similarities in<br />

resource dynamics. To persist in high productive environment, i.e. in a competitive<br />

environment, an efficient capture of resources is nee<strong>de</strong>d: coexisting species consequently<br />

present similar strategies to capture resources. Accordingly, convergence in traits associated<br />

with competition is expected in these habitats due to the increasing effect of competition as a<br />

filtering process (Pakeman et al. 2011). This i<strong>de</strong>a of the effect of competition on trait<br />

convergence was supported by a recent study in tropical forest (Swenson & Enquist 2009).<br />

However, the two former points of view, suggesting different patterns of traits distribution<br />

associated with competition are not necessarily opposite. In<strong>de</strong>ed coevolution led to the<br />

divergence between species allowing coexistence through niche differentiation, but coexisting<br />

species may be similar in or<strong>de</strong>r to compete relatively equally (Scheffer & van Nes 2006).<br />

The aim of this study is to <strong>de</strong>termine assembly rules acting on community assembly<br />

along a flooding gradient in wet and productive grasslands. We aimed to test Weiher &<br />

Keddy’s and Grime’s points of view according to the effect of competition on traits dispersion<br />

within three plant communities distributed along a flooding gradient. As the selection of<br />

functional traits is fundamental to address the assembly mechanisms (Götzenberger et al.<br />

2011), two relevant traits have been chosen: Specific Leaf Area (SLA) and plant height.<br />

These two functional traits reflect species tolerance to flooding conditions: they increased<br />

with flooding conditions to improve gas diffusion un<strong>de</strong>rwater (Blom & Voesenek 1996;<br />

Clevering et al. 1998; Lenssen et al. 1998; Mommer et al. 2004). These traits are also related<br />

to species strategies to capture resources and species competitive ability (Westoby 1998;<br />

Weiher et al. 1999; Violle et al. 2011).<br />

To reach our goal, measurements of functional diversity, i.e. the range and value of<br />

species traits within a community, present a crucial importance to reveal community assembly<br />

functioning (Diaz & Cabido 2001; Tilman 2001; Petchey & Gaston 2002; Mouillot et al.<br />

2005). Three statistics, measuring trait distribution, have been used to test theoretical<br />

predictions of habitat filtering, niche differentiation and convergence due to competition<br />

previously enounced. These statistics allow the distinction between convergence in the trait<br />

distribution due to habitat filtering or due to competition. Habitat filtering restricts the mean<br />

of traits values (Kraft et al. 2008; Jung et al. 2010). Niche differentiation, through the concept<br />

of limiting similarity, affects traits spacing, increasing the mean distance between nearest<br />

neighbors and reducing the variance in distance between nearest neighbors (Schamp et al.<br />

2010). Convergence of traits values associated with competition reduces the mean distance in<br />

170


Chapitre VI<br />

the nearest neighbor (Schamp et al. 2010). Null mo<strong>de</strong>ls approach will be used to compare<br />

observed traits distribution in the field with randomly traits distribution (Gotelli & McCabe<br />

2002).<br />

Following Weiher & Keddy’s (1995) predictions, we hypothesized that in the floo<strong>de</strong>d<br />

parts of the gradient, a convergence in trait values will be observed resulting in habitat<br />

filtering, while in the non-floo<strong>de</strong>d parts the traits distribution will result in niche<br />

differentiation, signs of limiting similarity due to competition. In our study site, the<br />

productivity is high and constant all along the gradient (Violle et al. 2011): we cannot test the<br />

variation in the level of productivity on trait distribution. However, we hypothesized that the<br />

high productivity of the study site will result in convergence in traits associated with<br />

competition all along the gradient.<br />

Methods<br />

Study site<br />

The studied site is a floo<strong>de</strong>d grasslands of the Marais Poitevin on the French Atlantic<br />

coast (46°28’N; 1°13’W), characterized by an atlantic climate with an annual mean of the<br />

amount of rainfall equal to 655 mm (Amiaud & Touzard 2004). The floo<strong>de</strong>d grasslands are<br />

characterized by an elevation gradient with a mean range of 45 cm. The elevation was related<br />

to the duration of flooding (r²=0.776; P


Chapitre VI<br />

abundance of species was noted. Along the elevation, a total of 38 species were sampled.<br />

Measurements of the elevation of each plot were realized using a Trimble (Trimble M3, Ohio,<br />

USA) which is an integrative measure of flooding events that species experienced.<br />

Community level is consi<strong>de</strong>red here as a proxy to approach the environmental conditions that<br />

species experienced.<br />

Within each community and for each species, two functional traits were measured: the<br />

Specific Leaf Area (SLA) and the plant height. These two functional traits reflect species<br />

response to flooding conditions due to the facilitation of gas exchanges in oxygen-<strong>de</strong>ficient<br />

conditions (e.g. Blom & Voesenek 1996; Insausti et al. 2001; Mommer et al. 2004). They<br />

reflect also species strategies to capture resources (Weiher et al. 1999) and plant height can<br />

assess species competitive ability (Westoby 1998; Violle et al. 2011). Ten replicates of each<br />

trait were measured by species and by community when abundance of species was sufficient<br />

to sample it. These traits were measured following the standardized protocols (Weiher et al.<br />

1999; Cornelissen et al. 2001): SLA was measured on the youngest fully grown leaf in the<br />

light using a leaf area meter (Li-Cor, Lincoln, Nebraska, USA) and plant height as the<br />

difference between the elevation of the highest photosynthetic leaf and the base of the plant.<br />

Data analysis<br />

To test the hypotheses of habitat filtering and niche differentiation, we consi<strong>de</strong>red the<br />

distribution of functional traits within communities, taking into account intra-specific<br />

variability in trait values. In<strong>de</strong>ed, species may adjust their trait values to respond to changing<br />

environments and promoting species coexistence, highlighted in recent studies (Violle &<br />

Jiang 2009; Albert et al. 2010; Jung et al. 2010). Accordingly, 10 replicates were measured<br />

for each species within each community and the mean trait value was calculated per<br />

community. Then, species occurring in two different communities may then have different<br />

mean trait values.<br />

Three statistics were calculated in this study to test the dispersion of trait values within<br />

communities and test the hypotheses of habitat filtering and niche differentiation (Table 1): i)<br />

the mean of trait values <strong>de</strong>monstrating habitat filtering when observed values are higher than<br />

the expected values coming from null distribution (Kraft et al. 2008), ii) the mean nearest<br />

Eucli<strong>de</strong>an trait distance (meanNTD), measuring how traits are spaced within plots (Weiher et<br />

al. 1998; Schamp et al. 2010; Decaens et al. 2011), <strong>de</strong>monstrating competitive convergence<br />

when observed values are lower than the expected values, and <strong>de</strong>monstrating limiting<br />

172


Chapitre VI<br />

similarity when observed values are higher than the expected values coming the null<br />

distribution (Schamp et al. 2008; Schamp et al. 2010) and iii) the variance in nearest<br />

Eucli<strong>de</strong>an trait distance (varNTD) between each species measuring both how species are<br />

regularly spaced within plots (Kraft et al. 2008; Schamp et al. 2008; Schamp et al. 2010),<br />

<strong>de</strong>monstrating limiting similarity when observed values are lower than the expected values<br />

coming from the null distribution.<br />

Table 1: Theoretical prediction of processes un<strong>de</strong>rlying trait dispersion according to the<br />

direction of <strong>de</strong>viation for each statistic relatively to the null distribution.<br />

Test Statistic Habitat filtering Limiting similarity<br />

Convergence due to<br />

competition<br />

Mean Observed>Expected - -<br />

MeanNTD - Observed>Expected Observed


Chapitre VI<br />

where O is the observed value of trait, M and S are, respectively, the mean and the standard<br />

<strong>de</strong>viation for each test across the 1000 randomizations. A positive value indicates a <strong>de</strong>viation<br />

above the mean of the null distribution, a negative value a <strong>de</strong>viation below the mean of null<br />

distribution.<br />

To <strong>de</strong>termine whether variations in elevation within each community influence<br />

statistics values, regressions have been calculated between elevation and SES value of each<br />

metric in each community. To test changes of these patterns along the environmental gradient,<br />

regressions were realized between SES value of each trait and elevation all along the gradient.<br />

Results<br />

Community assembly within each community<br />

Coexisting species within the mesophilous community were more similar in SLA<br />

(MeanNTD; Table 2) and were more regularly distributed with respect to SLA than expected<br />

by chance (VarNTD; Table 2). Coexisting species showed higher SLA values on average<br />

(Mean; Table 3) but were less similar with respect to plant height than expected by chance<br />

(MeanNTD; Table 3). For both functional trait, the three statistics did not vary with the<br />

elevation range occurring within community (non significant regressions, P>0.05), except for<br />

the Mean SLA (Adjusted r²= 0.105, P=0.045).<br />

Coexisting species within the mesohygrophilous community were more similar in<br />

SLA than expected by chance (MeanNTD; Table 2). These coexisting species showed higher<br />

SLA values on average than expected by chance (Mean; Table 3). They were less similar with<br />

respect to plant height (MeanNTD; Table 3) but were more regularly distributed with respect<br />

to plant height than expected by chance (VarNTD; Table 3). For each functional trait, the<br />

three statistics did not vary with the elevation range occurring within community (non<br />

significant regressions, P>0.05), except for the VarNTD SLA (Adjusted r²= 0.525; P


Chapitre VI<br />

the elevation range occurring within community (non significant regressions, P>0.05), except<br />

for the VarNTD SLA (Adjusted r²= 0.106; P=0.05).<br />

Table 2: Results from the null distribution for the SLA trait for the three studied communities.<br />

Mesophilous<br />

community<br />

Mesohygrophilous<br />

community<br />

Hygrophilous<br />

community<br />

Test Observed>Expected ObservedExpected Observed


Chapitre VI<br />

Fig. 1: Regressions analysis showing changes in the Standardized Effect Size (SES) of each<br />

metric across the elevation gradient: the mean of plant height (a), the MeanNTD of plant<br />

height (b) and VarNTD in plant height (c), all negatively correlated with elevation gradient.<br />

Significant boundaries (-1.96; 1.96) are indicated by dashed horizontal lines.<br />

Mean SLA increased with the duration of flooding (<strong>de</strong>crease in elevation) (r²=0.409,<br />

P


Chapitre VI<br />

in accordance with predictions ma<strong>de</strong> by Weiher & Keddy (1995), i.e. coexisting species show<br />

convergence in their trait values. This convergence in trait values represents the species<br />

response to environmental conditions. In<strong>de</strong>ed in the hygrophilous community, the mean<br />

values for SLA and plant height are higher than expected by chance, providing an evi<strong>de</strong>nce of<br />

the habitat filtering. Coexisting species in the hygrophilous community are filtered with<br />

respect to traits allowing tolerance to flooding conditions, high plant height and high SLA<br />

facilitating gas exchanges among individuals un<strong>de</strong>rwater (Blom & Voesenek 1996; Clevering<br />

et al. 1998; Lenssen et al. 1998; Mommer et al. 2004; Violle et al. 2011). Such an increase in<br />

these two morphological traits has already been highlighted for other wetlands species<br />

(Weiher et al. 1998; Jung et al. 2010). At intermediate positions along the elevation gradient,<br />

on the slopes, habitat filtering also occurred but only for plant height. A high SLA together<br />

with tall plants occurs un<strong>de</strong>r long and floo<strong>de</strong>d conditions, while a high plant height appears<br />

sufficient to tolerate short and shallow flooding (Colmer & Voesenek 2009). The intensity of<br />

the constraint thus drive the range of traits involved in the filtering processes.<br />

Following the predictions of Weiher & Keddy (1995), a divergence in the trait values<br />

was expected for traits associated with species competitive abilities in the higher parts of the<br />

gradient. The metrics calculated indicated that competition process leads to an increase in the<br />

mean of the nearest neighbors among co-occuring species (MeanNTD) and a reduction in the<br />

variance of the nearest neighbors among co-occuring species (VarNTD). Our results<br />

<strong>de</strong>monstrate that in the higher elevations of the gradient, traits values for plant height and<br />

SLA are regularly spaced among mesophilous coexisting species. This provi<strong>de</strong>s an evi<strong>de</strong>nce<br />

of niche differentiation for these two functional traits, in accordance with the predictions of<br />

Weiher & Keddy (1995). The niche differentiation, following the concept of limiting<br />

similarity, allows a share of resources between coexisting species. This share of resources is<br />

realized according to both traits, which are related to resources acquisition (Weiher et al.<br />

1999; Wright et al. 2004). The process of niche differentiation has been <strong>de</strong>monstrated in other<br />

wetland communities (Weiher et al. 1998; Jung et al. 2010), in sand dunes (Stubbs & Wilson<br />

2004) and in tropical forests (Kraft et al. 2008).<br />

The divergence in traits values is also <strong>de</strong>monstrated in our communities submitted to<br />

flooding conditions. These morphological traits, and especially plant height, are filtered by<br />

environmental conditions to tolerate flooding conditions, and are also involved in resource<br />

partitioning between species. In these floo<strong>de</strong>d parts of the gradient, where the productivity is<br />

high and equal to the productivity in the non-floo<strong>de</strong>d parts (Violle et al. 2011), the niche<br />

differentiation is <strong>de</strong>monstrated. This suggests that competition between species impacts<br />

177


Chapitre VI<br />

species assemblage in floo<strong>de</strong>d habitats, probably because of the high productivity of the<br />

community leading to increased competition for light among species.<br />

Because of the high availability of soil resources, vegetation is <strong>de</strong>nse all along the<br />

gradient. Following Grime’s proposition (2006), we expect an increasing effect of<br />

competition as a filtering process in productive environments, thus leading to a convergence<br />

in traits values (Grime 2006; Pakeman et al. 2011). Our results indicate a <strong>de</strong>crease in the<br />

mean distance between the nearest neighbors (MeanNTD) involving SLA, a convergence<br />

accentuated by the increase of the elevation. Simultaneously, a <strong>de</strong>crease in the mean distance<br />

between the nearest neighbors (MeanNTD) for plant height is showed, and this convergence<br />

in traits values is higher in the non-floo<strong>de</strong>d parts of the gradient. Our results support Grime’s<br />

proposition (2006) about traits related to competition. To persist in high competitive<br />

environment, an efficient capture of resources is nee<strong>de</strong>d: coexisting species consequently<br />

present similar strategies to capture resources, and this could involve close SLA and plant<br />

height values. These similarities in resource dynamics within coexisting species have been<br />

highlighted in other vegetation types (Swenson & Enquist 2009; Pakeman et al. 2011).<br />

However, traits involved in convergence are different <strong>de</strong>pending on the location along<br />

the gradient. In the mesophilous community, niche differentiation of coexisting species was<br />

possible through the differentiation in SLA, and a convergence due to competition involving<br />

plant height was observed suggesting a competition for light. In the hygrophilous community,<br />

niche differentiation of coexisting species was possible through the differentiation in plant<br />

height and simultaneously a convergence in traits distribution due to competition for SLA are<br />

observed. As SLA is a proxy of vegetative growth (Westoby et al. 1998), species may coexist<br />

through different growth abilities during and after flooding. These results are consistent with<br />

the proposition of Scheffer & van Nes (2006) that for traits linked to competition, a<br />

convergence and divergence may be simultaneously observed, supporting the points of view<br />

of Weiher & Keddy (1995) and Grime (2006). The coevolution un<strong>de</strong>r competitive<br />

environment leads to species sufficiently divergent to coexist through niche differentiation,<br />

but species coexistence requires a <strong>de</strong>gree of similarity to persist in competitive environments.<br />

Conclusion<br />

Our results show that flooding and competition represent two filtering processes, and<br />

competition represents a filtering process throughout the flooding gradient. We show a<br />

178


Chapitre VI<br />

simultaneous effect of habitat filtering and niche differentiation, as it has been already<br />

highlighted in wetlands (Weiher et al. 1998; Jung et al. 2010), in Californian foothills<br />

(Cornwell & Ackerly 2009), and in tropical forests (Kraft et al. 2008). We also show a<br />

simultaneous effect of niche differentiation and convergence due to competition. In the plant<br />

communities studied here, the niche differentiation and the response to competition appear to<br />

involve different traits (at the community level) leading to the possibility of <strong>de</strong>monstrating it.<br />

These results are in contrast with Schamp et al. (2008) using the same metrics, but who<br />

supposed that the studied traits contribute simultaneously to niche differentiation and to the<br />

response to competition, leading to an indistinguishable pattern from pattern expected by<br />

chance. The choice of the functional traits to study thus appears crucial to test these<br />

mechanisms (Kraft et al. 2008; Götzenberger et al. 2011). In our study, SLA and plant height<br />

appear as relevant response traits to abiotic and biotic environments to <strong>de</strong>tect assembly rules<br />

in wet grasslands.<br />

Different limits may be observed in this study. First, the species abundance has not<br />

been consi<strong>de</strong>red, while its incorporation in analyses may address processes more precisely<br />

(Götzenberger et al. 2011). Moreover, the community level has been chosen to test assembly<br />

rules. However, some statistics appear sensitive to variation in elevation within a community<br />

as VarNTD. Taking into account the abiotic constraints at the plot level and adopting a null<br />

mo<strong>de</strong>l consi<strong>de</strong>ring the structure of each plot (richness, species abundance) could allow to test<br />

precisely assembly rules. We are currently testing these two elements.<br />

179


Cited literature<br />

Chapitre VI<br />

Ackerly D.D. & Cornwell W.K. (2007) A trait-based approach to community assembly:<br />

partitioning of species trait values into within- and among-community components. Ecology<br />

Letters, 10: 135-145.<br />

Albert, C.H., Thuiller, W., Yoccoz, N.G., Soudant, A., Boucher, F., Saccone, P. & Lavorel, S.<br />

(2010) Intraspecific functional variability: extent, structure and sources of variation. Journal<br />

of Ecology, 98: 604–613.<br />

Amiaud, B. & Touzard, B. (2004) The relationships between soil seed bank, aboveground<br />

vegetation and disturbances in old embanked marshlands of Western France. Flora, 199: 25-<br />

35.<br />

Blom, C. & Voesenek, L. (1996) Flooding: the survival strategies of plants. Trends in<br />

Ecology & Evolution, 11: 290–295.<br />

Bouzillé, J.B., 1992. Structure et dynamique <strong>de</strong>s paysages, <strong>de</strong>s communautés et <strong>de</strong>s<br />

populations végétales <strong>de</strong>s marais <strong>de</strong> l’ouest. Thèse d’état, spécialité écologie. <strong>Université</strong> <strong>de</strong><br />

<strong>Rennes</strong> 1, pp. 303.<br />

Clevering, O.A. (1998) Effects of litter accumulation and water table on morphology and<br />

productivity of Phragmites australis. Wetlands Ecology and Management, 5: 275–287.<br />

Colmer, T.D. & Voesenek, L. (2009) Flooding tolerance: suites of plant traits in variable<br />

environments. Functional Plant Biology, 36: 665–681.<br />

Connell, J.H. (1980) Diversity and the coevolution of competitors, or the ghost of competition<br />

past. Oikos, 35(2): 131-138.<br />

Cornelissen, J.H.C., Lavorel, S., Garnier, E., Diaz, S., Buchmann, N., Gurvich, D.E., Reich,<br />

P.B., Steege, H., Morgan, H.D., Van Der Heij<strong>de</strong>n, M.G.A. & others. (2001) A handbook of<br />

protocols for standardised and easy measurement of plant functional traits worldwi<strong>de</strong>.<br />

Australian Journal of Botany, 51: 335–380.<br />

Cornwell, W.K. & Ackerly, D.D. (2009) Community assembly and shifts in plant trait<br />

distributions across an environmental gradient in coastal California. Ecological Monographs,<br />

79(1): 109–126.<br />

Decaens, T., Margerie, P., Renault, J., Bureau, F., Aubert, M. & Hed<strong>de</strong>, M. (2011) Niche<br />

overlap and species assemblage dynamics in an ageing pasture gradient in north-western<br />

France. Acta Oecologica, in press.<br />

Diamond, J. M. (1975). Assembly of species communities. In Ecology and Evolution of<br />

Communities (eds M. L. Cody and J. M. Diamond), pp. 342–444. Harvard University Press,<br />

Harvard<br />

Diaz, S. & Cabido, M. (2001) Vive la différence: plant functional diversity matters to<br />

ecosystem processes. Trends in Ecology & Evolution, 16(11): 646-655.<br />

Díaz, S., Lavorel, S., McIntyre, S., Falczuk, V., Casanoves, F., Milchunas, D.G., Skarpe, C.,<br />

Rusch, G., Sternberg, M., Noy-Meir, I., Landsberg, J., Zhang, W., Clark, H. & Campbell,<br />

B.D. (2007) Plant-trait responses to grazing-a global synthesis. Global Change Biology,<br />

13(2): 313-341.<br />

Gotelli, N.J. & McCabe, D.J. (2002) Species co-occurence: a meta-analysis of J.M.<br />

Diamond’S assembly rules mo<strong>de</strong>l. Ecology, 83(8): 2091–2096.<br />

180


Chapitre VI<br />

Götzenberger, L., <strong>de</strong> Bello, F., Brathen, K.A., Davison, J., Dubuis, A., Guisan, A., Leps, J.,<br />

Lindborg, R, Moora, M., Pärtel, M., Pellissier, L., Pottier, J., Vittoz, P., Zobel, K. & Zobel,<br />

M. (2011) Ecological assembly rules in plant communities—approaches, patterns and<br />

prospects. Biological Reviews, doi: 10.1111/j.1469-185X.2011.00187.x<br />

Grime, J.P. (2006) Trait convergence and trait divergence in herbaceous plant communities:<br />

Mechanisms and consequences. Journal of Vegetation Science 17: 255-260.<br />

Insausti, P., Grimoldi, A.A., Chaneton, E.J. & Vasellati, V. (2001) Flooding induces a suite of<br />

adaptive plastic responses in the grass Paspalum dilatatum. New Phytologist, 152, 291–299.<br />

Jung, V., Violle, C., Mondy, C., Hoffmann, L. & Muller, S. (2010) Intraspecific variability<br />

and trait-based community assembly. Journal of Ecology, 98: 1134–1140.<br />

Keddy, P.A. (1992) Assembly and response rules: two goals for predictive community<br />

ecology. Journal of Vegetation Science, 3: 157–164.<br />

Kraft, N.J.B., Valencia, R. & Ackerly, D.D. (2008) Functional traits and niche-based tree<br />

community assembly in an Amazonian forest. Science, 322: 580-582.<br />

Lavorel S. & Garnier E. (2002) Predicting changes in community composition and ecosystem<br />

functioning from plant traits: revisiting the Holy Grail. Functional Ecology, 16: 545- 556<br />

Lenssen, J.P.M., Ten Dolle, G.E. & Blom, C. (1998) The effect of flooding on the recruitment<br />

of reed marsh and tall forb plant species. Plant Ecology, 139: 13–23.<br />

MacArthur R.H. & Levins R. (1967) Limiting similarity convergence and divergence of<br />

coexisting species. American Naturalist, 101: 377-387<br />

McGill, B.J., Enquist, B.J., Weiher, E. & Westoby, M. (2006) Rebuilding community ecology<br />

from functional traits. Trends in Ecology & Evolution, 21: 178–185.<br />

Mommer, L., Pe<strong>de</strong>rsen, O. & Visser, E.J.W. (2004) Acclimation of a terrestrial plant to<br />

submergence facilitates gas exchange un<strong>de</strong>r water. Plant, Cell & Environment, 27: 1281–<br />

1287.<br />

Mouillot, D., Stubbs, W., Faure, M., Dumay, O., Tomasini, J.A., Wilson, J.B. & Chi, T.D.<br />

(2005) Niche overlap estimates based on quantitative functional traits: a new family of nonparametric<br />

indices. Oecologia, 145: 345–353.<br />

Pakeman, R.J., Lennon, J.J. & Brooke, R.W. (2011) Trait assembly in plant assemblages and<br />

its modulation by productivity and disturbance. Oecologia, doi 10.1007/s00442-011-1980-6.<br />

Petchey, O. L. & Gaston, K. J. (2002). Functional diversity (FD), species richness and<br />

community composition. Ecology Letters, 5: 402–411.<br />

Pillar, V.D., Leandro, D., Enio, S. & Fernando, J. (2009) Discriminating trait-convergence<br />

and trait-divergence assembly patterns in ecological community gradients. Journal of<br />

Vegetation Science, 20 : 334–348.<br />

Schamp, B., Chau, J. & Aarssen, L.W. (2008) Dispersion of traits related to competitive<br />

ability in an old-field plant community. Journal of Ecology, 96: 204–212.<br />

Schamp, B., Horsak, M. & Hajek, M. (2010) Deterministic assembly of lan snail communities<br />

according to species size and diet. Journal of Animal Ecology, 79: 803-810.<br />

Scheffer, M. & van Ness, E.H. (2006) Self-organized similarity, the evolutionary emergence<br />

of groups of similar species. Proceedings of the National Aca<strong>de</strong>my of Science, 103: 6230–<br />

6235.<br />

181


Chapitre VI<br />

Stubbs, W.J. & Wilson, J.B. (2004) Evi<strong>de</strong>nce for limiting similarity in a sand dune<br />

community. Journal of Ecology, 92: 557–567.<br />

Swenson, N.G. & Enquist, B. J. (2009). Opposing assembly mechanisms in a Neotropical dry<br />

forest: implications for phylogenetic and functional community ecology. Ecology, 90: 2161–<br />

2170.<br />

Tilman D. (2001) Functional diversity. In: Encyclopedia of Biodiversity (ed. Levin S), pp.<br />

109-120. Aca<strong>de</strong>mic Press, San Diego<br />

Violle, C., Bonis, A., Plantegenest, M., Cu<strong>de</strong>nnec, C., Damgaard, C., Marion, B., Le Coeur,<br />

D. & Bouzillé, J.B. (2011) Plant functional traits capture species richness variations along a<br />

flooding gradient. Oikos, 120(3): 389-398.<br />

Violle, C. & Jiang, L. (2009) Towards a trait-based quantification of species niche. Journal of<br />

Plant Ecology, 2(2): 87–93.<br />

Weiher, E. & Keddy, P.A. (1995) The assembly of experimental wetland plant communities.<br />

Oikos, 73: 323–335.<br />

Weiher, E., Clarke, J.D.P. & Keddy, A. (1998) Community assembly rules, morphological<br />

dispersion and the coexistence of plant species. Oikos, 81: 309-322.<br />

Weiher, E., Werf, A., Thompson, K., Ro<strong>de</strong>rick, M., Garnier, E. & Eriksson, O. (1999)<br />

Challenging Theophrastus: a common core list of plant traits for functional ecology. Journal<br />

of Vegetation Science, 10: 609–620.<br />

Westoby, M. (1998). A Leaf-Height-Seed (LHS) plant ecology strategy scheme. - Plant and<br />

Soil 199: 213-227.<br />

Wright, I.J., Reich, P.B., Westoby, M., Ackerly, D.D., Baruch, Z, Bongers, F., Caven<strong>de</strong>r-<br />

Bares, J., Chapin, T., Cornelissen, J.H.C., Diemer, M., Flexas, J., Garnier, E., Groom, P.K.,<br />

Gulias, J., Hikosaka, K., Lamont, B.B., Lee, T., Lee, W., Lusk, C., Midgley, J.J., Navas, M-<br />

L., Niinemets, U., Oleksyn, J., Osada, N., Poorter, H., Poot, P., Prior, L., Pyankov, V.I.,<br />

Roumet, C., Thomas, S.C., Tjoelker, M.G., Veneklass, E.J. & Villar, R. (2004) The<br />

worldwi<strong>de</strong> leaf economics spectrum. Nature, 428: 821–827.<br />

182


183


184


- Chapitre 7 -<br />

Stress level influences the effect of grazing regime and of<br />

variability of rainfall patterns on xero-halophytic communities<br />

Soumis à Plant Ecology<br />

185


186


Chapitre VII<br />

Stress level influences the effect of grazing regime and of variability of<br />

rainfall patterns on xero-halophytic communities<br />

<strong>Amandine</strong> <strong>Merlin</strong> 1,2 , Anne Bonis 1 & François Mesléard 2,3<br />

1 UMR CNRS 6553 EcoBio, <strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1, Campus Beaulieu, F-35042 <strong>Rennes</strong><br />

Ce<strong>de</strong>x, France<br />

2 Centre <strong>de</strong> recherche La Tour du Valat, Le sambuc, F-13200, France<br />

3 UMR CNRS-IRD 6116 Institut Méditerranéen d’Ecologie et <strong>de</strong> Paléoécologie, <strong>Université</strong><br />

d’Avignon IUT site Agroparc BP 1207, F-84911 Avignon Ce<strong>de</strong>x 09, France<br />

Abstract<br />

Questions: What are the respective roles of grazing and rainfall pattern in <strong>de</strong>termining plant<br />

community structure over 8 years? Does the abiotic stress level mediate the effects of grazing<br />

and of variability in rainfall patterns?<br />

Location: Mediterranean xero-halophytic grasslands of the Camargue (Southern France).<br />

Methods: Yearly monitoring was conducted for eight years on two vegetation communities:<br />

(1) xeric meadows, (2) Mediterranean salt-marshes. The two communities were subjected to<br />

three grazing treatments: (1) Status Quo treatment (control); (2) Short Grazing treatment and<br />

(3) the Exclusion of grazing.<br />

Results: Rainfall patterns controlled the structure of the two communities in the controls. The<br />

cessation of grazing led to a loss of diversity in xeric meadows through the increase of the<br />

dominant species Brachypodium phoenicoï<strong>de</strong>s. This cessation of grazing had no effect on the<br />

structure of the salt-marshes. The change in the modality of grazing had a weak effect mainly<br />

noticeable in xeric meadows: it increased the contribution of Bromus madritensis. The<br />

consequence is an increase in the similarity of the two communities.<br />

Conclusions: The level of abiotic stress buffers the impact of changes in grazing<br />

management. In the most highly environmentally constrained community, the structure<br />

(species diversity) remains unchanged over the 8 years studied <strong>de</strong>spite changes in grazing<br />

modality. While in the least environmentally constrained community, grazing represents the<br />

main structuring factor. The level of stress experienced by community should be consi<strong>de</strong>red<br />

before using grazing as a management tool.<br />

Keywords: Diversity; Dominant species, Evenness; NMDS; Response pathways<br />

Nomenclature: Tutin, T.G. et al. 1964-1980, Flora Europea, Cambridge University Press,<br />

Cambridge.<br />

187


Chapitre VII<br />

188


Introduction<br />

Chapitre VII<br />

Most of the time, grazing plays a <strong>de</strong>terminant role in the organization of herbaceous<br />

communities. Grazing management represents a biotic filter selecting species through<br />

<strong>de</strong>foliation and trampling actions, in relation with the respective value of their traits.<br />

According to the intensity and the frequency of grazing, different plant growth forms and<br />

functional groups are selected (Diaz et al. 1998; Díaz et al. 2001; Díaz et al. 2007; Golo<strong>de</strong>ts et<br />

al. 2010; Milchunas and Noy-Meir 2002; Noy-Meir et al. 1989; Sternberg et al. 2000)<br />

modifying community structure (Peco et al. 1983; Peco et al. 2006).<br />

Grazing represents a disturbance because it induces a partial or total loss of plant<br />

biomass (Grime 1979; White and Pickett 1985). Following the Intermediate-Disturbance<br />

Hypothesis <strong>de</strong>scribing a hump-shaped relationship (Grime 1973), a high level of diversity is<br />

to be expected for intermediate levels of disturbance due to the reduction of competitive<br />

interactions between species and due to the creation of gaps of bare soil representing vacant<br />

niches for species (Johnstone 1986), favoring the spread of a large number of species.<br />

However, at high levels of disturbance, selected species present morphological and<br />

anatomical traits allowing them to persist un<strong>de</strong>r these conditions (Golo<strong>de</strong>ts et al. 2010; Noy-<br />

Meir et al. 1989; Noy-Meir 1995; Olff and Ritchie 1998; Peco et al. 2005), while competitive<br />

exclusion by competitive species operates in the absence of disturbance (Milchunas et al.<br />

1988; Noy-Meir et al. 1989; Peco et al. 2005; Sternberg et al. 2000): both extreme levels of<br />

grazing pressure lead to a loss of diversity.<br />

Grazing can modify the environmental conditions in three ways (Cingolani et al.<br />

2003). Firstly, grazing can lead to an amplification of environmental conditions that<br />

communities are subjected to, increasing the existing divergence between communities.<br />

Secondly, grazing can reduce contrasts in environmental conditions between communities,<br />

leading to a convergence between them. Finally, grazing can preserve the contrasting<br />

environmental conditions between communities, but can lead to species replacement<br />

preserving the difference between communities, but not amplifying this difference.<br />

Alternatively, these three effects of grazing may be mediated by abiotic conditions, such as<br />

the level of stress. Highly environmentally constrained environments have been shown to<br />

reduce the magnitu<strong>de</strong> of floristic changes by grazing (Cingolani et al. 2003) especially in salt<br />

constrained environments (Allison 1992; Callaway et al. 1994; Milchunas et al. 1988). Soil<br />

salinity appears to be a <strong>de</strong>termining environmental filter (Callaway et al. 1990; Callaway and<br />

Sabraw 1994), which makes plants thus selected tolerant to other stresses. In<strong>de</strong>ed, in systems<br />

189


Chapitre VII<br />

characterized by a low availability of water resources, species’ strategies implemented to<br />

tolerate the presence of salt may also help species to tolerate other stresses or disturbances<br />

such as grazing (Milchunas et al. 1988).<br />

Despite the evi<strong>de</strong>nce of interactions between grazing management and environmental<br />

conditions, such as the variability in rainfall patterns and the soil salinity reducing community<br />

productivity, few studies have attempted to evaluate their respective importance with regard<br />

to community structure (Alday et al. 2010). The aim of this study is first to evaluate the<br />

respective roles of grazing and rainfall pattern and secondly to <strong>de</strong>termine whether the eventual<br />

effects are mediated by the abiotic stress level affecting the structure of Mediterranean<br />

grasslands. In particular, we aimed to <strong>de</strong>termine the impact of changes in grazing<br />

management on the composition and structure of the xero-halophytic grasslands over 8 years,<br />

exposed to different levels of abiotic stress: the xeric-meadows and the Mediterranean salt<br />

marshes.<br />

We hypothesized that (i) community composition responds strongly to the inter-annual<br />

variability in rainfall patterns due to the presence of a large proportion of annual species, (ii)<br />

a cessation of grazing leads to a strong modification of community structure and composition,<br />

with strong <strong>de</strong>velopment of perennial species reducing diversity, (iii) a change in the modality<br />

of grazing, from grazing by herbivores during six months to grazing during three days,<br />

representing a disturbance, leads to an increase in species diversity (iv) in the constrained<br />

community (i.e. the Mediterranean salt marshes), the impact of inter-annual rainfall<br />

variations and grazing changes on the community are weak due to the previous selection of<br />

species tolerant to various stresses.<br />

Methods<br />

Study site<br />

The study was carried out on the xero-halophytic grasslands of the Tour du Valat<br />

estate (Rhône <strong>de</strong>lta, Southern France, 43°29’ N, 4°40’ E). The xero-halophitic grasslands are<br />

characterized by a wi<strong>de</strong> diversity in annual species (Molinier and Tallon 1968) and represent<br />

the most diverse habitat of the Camargue. They are a priority habitat (co<strong>de</strong> 6220) un<strong>de</strong>r the<br />

terms of the European Union Habitats Directive (1992). They have been drastically reduced<br />

over the last <strong>de</strong>ca<strong>de</strong>s due to intensive agriculture (Lemaire et al. 1987).<br />

190


Chapitre VII<br />

The Camargue is characterized by a Mediterranean climate with cold winters and<br />

warm dry summers (Heurteaux 1970). Precipitation occurs mainly during spring and autumn.<br />

Average annual precipitation (1988-2008) is 570 mm, with strong inter-annual variations:<br />

from 252 mm (1989) up to 1049 mm (1996). Over the eight years studied, P-PET<br />

(Precipitation minus Potential Evapotranspiration), quantifying the water resource availability<br />

for plants (Scarnati et al. 2009), was variable (Fig. 1). We found three years which can be<br />

qualified as dry: 2001, 2006 and 2007. Spring was always dry, particularly in 2006, whereas<br />

the amount of autumn rainfall was variable but low for two years out of the eight years (2007<br />

and 2008) (see data in supplementary material).<br />

Fig 1 Annual P minus ETP (P-ETP) values for the 8 years of the study. Continuous line<br />

represents the mean of annual P-ETP values of 20 years (1989-2008) and dotted lines<br />

represent the limits of 95% confi<strong>de</strong>nce interval<br />

The xero-halophitic grasslands extend along a weak elevation gradient with the soil<br />

salinity <strong>de</strong>creasing with the increase in elevation, due to the influence of a superficial salty<br />

water table (Heurteaux 1970). Floristic composition varies along the gradient in response to<br />

the variation in soil salinity. Two different communities were discriminated according to a<br />

Correspon<strong>de</strong>nce analysis (CA) (Fig. 2) run on the first year dataset: the community occurring<br />

at higher elevations is dominated by Brachypodium phoenicoï<strong>de</strong>s L., Crepis vesicaria subsp.<br />

taraxacifolia Thuill., Scorpiorus muricatus L., Crepis sancta L., Brachypodium distachyon<br />

L., Psilurus aristatus L., Evax pygmaea L., Galium murale L. and other characteristic species<br />

of xeric meadows (habitat directive co<strong>de</strong> 6220-2). The second community occurring at lower<br />

elevations is composed of Halimione portulacoï<strong>de</strong>s Aellen, Limonium narbonense Mill,<br />

191


Chapitre VII<br />

Bromus hor<strong>de</strong>aceus L., Hymenolobus procumbens L., Parapholis filiformis Roth, Bupleurum<br />

semicompositum L. that can be assimilated to a Mediterranean salt-marsh community (habitat<br />

directive co<strong>de</strong> 1310-4).<br />

Fig. 2 Correspon<strong>de</strong>nce analysis (CA) run on vegetation records for 2001 discriminated two<br />

variations of composition of the xero-halophytic grasslands. Two communities are<br />

discriminated along axis 1, corresponding to a salinity gradient. In the left hand part of the<br />

plot are grouped samples presenting the strong contribution of Brachypodium phoenicoï<strong>de</strong>s,<br />

Crepis vesicaria, Vulpia sp. characterizing xeric meadows. In the right hand part of the plot,<br />

samples are grouped according to the contribution of Halimione portulacoï<strong>de</strong>s, Limonium<br />

narbonense, Bromus madritensis characterizing salt-marshes<br />

Community productivity was found to be significantly different between the two<br />

communities, with productivity significantly higher in the xeric meadows with a mean<br />

productivity in 2007 equal to 3.97 t/ha compared to 1.74 t/ha in Mediterranean salt-marshes.<br />

According to the <strong>de</strong>finition of stress sensu Grime (1979), the Mediterranean salt-marshes<br />

seemed to be more highly constrained than xeric meadows, because are less productive.<br />

In situ grazing treatment<br />

In autumn 2001, an experimental <strong>de</strong>sign was set up which consisted of three grazing<br />

treatments. The first treatment corresponds to extensive grazing pressure with a continuous<br />

annual stocking rate of 0.07 ha -1 .year -1 applied during six months. This treatment was applied<br />

from 1970 and it can be consi<strong>de</strong>red as the control: the Status Quo treatment. The second<br />

192


Chapitre VII<br />

grazing treatment consisted of the same annual stocking rate as the previous treatment but was<br />

applied during only two days at the beginning of February, representing a strong<br />

instantaneous disturbance, hereafter called the Short Grazing treatment. The third treatment<br />

consisted of a cessation of domestic grazing through fencing: the Exclusion treatment. Three<br />

replicates (plots) were set up for each treatment (4ha), except for the Status Quo treatment,<br />

applied over a 350 ha area. Because of the large size of the study area, the replicates of each<br />

grazing treatment could be well spaced out in or<strong>de</strong>r to avoid too much spatial <strong>de</strong>pen<strong>de</strong>nce<br />

between them.<br />

Twenty-one permanent 40 cm × 40 cm quadrats were established randomly in each of<br />

the three grazing treatment plots. Before any analysis, the plot effect was tested to verify<br />

whether there was any confusion between the plot and grazing treatment effects. First year<br />

floristic samplings were un<strong>de</strong>rtaken at random without taking into account the slight elevation<br />

gradient. CA of our first year floristic surveys was used to classify each quadrat as one of the<br />

two communities. Intermediate quadrats were exclu<strong>de</strong>d from further analyses. This results in<br />

an unbalanced number of quadrats between grazing treatments and between the two<br />

communities (Table 1).<br />

Table 1 Number of quadrats classified by CA realized on floristic samplings of 2001;<br />

intermediate quadrats were exclu<strong>de</strong>d for Exclusion treatment and for Short Grazing treatment<br />

Exclusion Status Quo Short Grazing<br />

treatment treatment treatment<br />

Xeric meadows 15 5 9<br />

Salt-marshes 5 16 10<br />

Data were collected using the pinpoint method: in each permanent quadrat, a set of<br />

two crossing lines was established materializing 36 pins, three cm apart from each other.<br />

Species presence was recor<strong>de</strong>d at each of the 36 pins per quadrat resulting in species<br />

frequencies which were obtained per quadrat for each grazing treatment and for each<br />

community. Data were collected every year in the two last weeks of April over an eight year<br />

period, from 2001 to 2008.<br />

193


Characterization of community structure and composition<br />

Chapitre VII<br />

Community structure was <strong>de</strong>scribed by the Shannon diversity in<strong>de</strong>x (H’) (Shannon<br />

and Weaver 1949) and species evenness (J’) (Pielou 1975) with:<br />

These indices were calculated firstly at quadrat scale. Secondly a mean of these indices was<br />

calculated for each year and for each grazing treatment within each community. But the wi<strong>de</strong><br />

difference in the number of quadrats sampled within each grazing treatment affects the<br />

number of species within each quadrat. Therefore, to calculate the mean, a bootstrap<br />

procedure was performed to select randomly a balanced number of quadrats for each<br />

community and for each grazing treatment (n = 5).<br />

Data analyses<br />

Characterization of the communities’ pathways<br />

Distance matrices ordination was used to <strong>de</strong>tect species composition changes in each<br />

community in response to grazing treatment and to climatic variability, because this<br />

ordination is useful to quantify vegetation changes across time and to express the<br />

community’s displacement after disturbance (Halpern 1988). Bray-Curtis similarity matrix<br />

was calculated as the sum of contacts by species in each year and in a given grazing treatment<br />

plot relative to the total number of contacts: then dissimilarities among communities were<br />

calculated on the basis of species composition (Halpern 1988; Myers and Harms 2011). The<br />

ordination of these similarity matrices was performed using a Non-Metric Dimensional<br />

Scaling ordination (NMDS, isoNMDS function) using the vegan package in R software (R<br />

version 2.7.2., 2008, R Foundation for Statistical Computing, Vienna, Austria). Using this<br />

technique, slow compositional changes in response to variation of environmental factors are<br />

translated into annual samples plotted close together. The robustness of the NMDS was<br />

assessed by measuring the discrepancy between the inter-point distances on the NMDS plot<br />

versus distances in the original distance matrix called “stress” (Legendre and An<strong>de</strong>rson 1999;<br />

Lock et al. 2010). The NMDS analysis was run simultaneously for the two communities,<br />

allowing comparison of their response pattern to grazing and climatic pattern over the years.<br />

194


Chapitre VII<br />

To test changes in community composition over time for each pathway plotted on the NMDS,<br />

distance-based permutational multivariate analyses of variance (PERMANOVA, version 1.6,<br />

2005, Department of Statistics, University of Auckland, New Zealand) were performed for<br />

each community and grazing treatment.<br />

Dynamics of plant life forms<br />

Changes in plant life form cover (i.e. annual and perennial forbs, annual and perennial<br />

grasses and annual legumes) in each community with grazing treatment and years were<br />

addressed by mixed effect mo<strong>de</strong>ls, with community, grazing treatment and time as fixed<br />

factors, and replicates (quadrats) as random factor.<br />

Dynamics of dominant species<br />

Dominant species, <strong>de</strong>termined as species with abundance higher than 10% (Grime<br />

1998), are likely to drive community dynamics over time (Del Pozo et al. 2006). Changes in<br />

cover of dominant species were addressed by mixed effect mo<strong>de</strong>ls, with community, grazing<br />

treatment and time as fixed factors, and replicates (quadrats) as random factor.<br />

Effect of dominant species on community structure<br />

Changes in species diversity and evenness in each community relative to the cover of<br />

dominant species were addressed by mixed effect mo<strong>de</strong>ls. In these mo<strong>de</strong>ls, grazing treatment,<br />

cover of dominant species and time were consi<strong>de</strong>red as fixed factors and replicates (quadrats)<br />

as random factor.<br />

Relationship between community structure and rainfall patterns<br />

To investigate whether species diversity and evenness were influenced by the<br />

variability and the seasonality of rainfall, three regression analyses were run between<br />

community diversity, evenness and P-PET values for both communities and for each grazing<br />

treatment: first consi<strong>de</strong>ring the spring period only (from February until the end of April),<br />

secondly consi<strong>de</strong>ring the autumn period only (from September until the end of November),<br />

while the third regression consi<strong>de</strong>red P-PET during the period between autumn yearn-1 and<br />

spring yearn.<br />

The use of mixed effects mo<strong>de</strong>ls consi<strong>de</strong>ring quadrats as random factor allows the<br />

highlighting of the effects of tested factors on the studied variables (e.g. indices of diversity,<br />

195


Chapitre VII<br />

evenness, cover of dominant species and of plant life-forms) relative to the structure of the<br />

<strong>de</strong>sign (Bennington and Thayne 1994).<br />

Statistical tests and multivariate analyses were performed using the a<strong>de</strong>4, vegan and<br />

MASS packages of the R software (R version 2.7.2., 2008, R Foundation for Statistical<br />

Computing, Vienna, Austria).<br />

Results<br />

Directional response pathways<br />

The NMDS gave a good representation of the un<strong>de</strong>rlying similarity matrix with a<br />

stress factor of 8.9% (0.089) for the first two axes. The first axis of the NMDS biplot (Fig. 3)<br />

represented the elevation gradient, i.e. the salinity gradient with the higher salt concentrations<br />

in the right part of the biplot, distinguishing the two communities. The second axis was<br />

related to changes in the proportion of perennials, <strong>de</strong>creasing along this axis. The analysis<br />

showed that significant changes in community composition occurred between years for both<br />

communities, except with the exclusion treatment in the Mediterranean salt-marshes<br />

(PERMANOVA, xeric meadows: exclusion treatment: F=1.51, P=0.016; Short Grazing<br />

treatment: F=1.64, P=0.005; Status Quo treatment: F=1.96, P=0.002; salt-marshes: Exclusion<br />

treatment: F=1.33, P=0.111; Short Grazing treatment: F=2.80, P=0.001; Status Quo<br />

treatment: F=3.30, P=0.001). We recor<strong>de</strong>d weak changes in xeric meadows with the Short<br />

Grazing treatment and these changes followed axis one of the NMDS biplot. Changes in<br />

composition in the Status Quo treatment did not appear to follow a strictly linear trajectory,<br />

contrary to the Cessation of grazing. In Mediterranean salt-marshes, the range of composition<br />

changes was similar between grazing treatments, except for the Short Grazing treatments<br />

where the changes were weak.<br />

196


Chapitre VII<br />

Fig. 3 Non-Metric Dimensional 10Scaling ordination (NMDS) on the basis of Eucli<strong>de</strong>an<br />

distance measures for centroids of years. Xeric meadows are plotted in the left part of the<br />

biplot, and salt-marshes in the right part of the biplot<br />

Changes in plant life forms in response to changes in grazing pressure<br />

Mixed effect mo<strong>de</strong>ls showed a significant impact of grazing, years and their<br />

interaction on the relative abundance of the various groups of plant life forms (Table 2) and<br />

on dominant species except for Halimoine portulacoï<strong>de</strong>s for which grazing showed no effect<br />

(Table 3). The perennial grasses (Brachypodium phoenicoï<strong>de</strong>s, Dactylis hispanica Roth)<br />

increased un<strong>de</strong>r the three grazing treatments over years, but especially un<strong>de</strong>r the exclusion<br />

treatment (from around 20% in 2001 to around 70% in 2008; Fig. 4a). This increase in the<br />

perennial grasses abundance mainly resulted from the increase in the dominant Brachypodium<br />

phoenicoï<strong>de</strong>s (Fig. 4c). Perennial forbs (Halimione portulacoï<strong>de</strong>s, Aetheorhiza bulbosa L.,<br />

Crepis vesicaria, Limonium narbonense, Limonium virgatum Willd), abundant in salt-<br />

marshes, also increased un<strong>de</strong>r the three grazing treatments (Fig. 4b), in particular the<br />

dominant species Halimione portulacoï<strong>de</strong>s (Fig 4d). Annual grasses showed clear responses<br />

to changes in grazing regime, <strong>de</strong>creasing in cover un<strong>de</strong>r grazing cessation treatment in both<br />

communities and un<strong>de</strong>r Status Quo treatment in salt-marshes, while increasing un<strong>de</strong>r the<br />

197


Chapitre VII<br />

Short Grazing treatment in both communities. However, the annual grass Bromus madritensis,<br />

the second most dominant species of salt-marshes, was negatively affected by Short Grazing<br />

and Exclusion treatments in salt-marshes (Fig. 4e).<br />

Table 2 Mixed effect mo<strong>de</strong>ls for differences in plant life forms between communities and<br />

grazing treatments across time, taking into account interactions between factors<br />

Source of variation<br />

Annual Perennial<br />

df MS F P df MS F P<br />

Community 1 828 9.12 0.003 1 3045 56.82 < 0.001<br />

Grazing 2 1213 13.37 < 0.001 2 182 3.4 0.03<br />

Year 7 1164 12.82 < 0.001 7 401.4 7.49 < 0.001<br />

Community×grazing 2 410 4.51 0.01 2 3.9 0.073 0.93<br />

Community×year 7 822 9.05 < 0.001 7 101.4 1.89 0.07<br />

Grazing×year 14 59 0.65 0.82 14 72.2 1.35 0.18<br />

Community×grazing×year 14 92 1.02 0.43 14 40.2 0.75 0.72<br />

Residuals 431 91 431 53.6<br />

Community 1 57369 456.57 < 0.001 1 46293 1229.4 < 0.001<br />

Grazing 2 1242 9.88 < 0.001 2 2222 59.01 < 0.001<br />

Year 7 1632 12.99 < 0.001 7 174 4.62 < 0.001<br />

Community×grazing 2 2848 22.67 < 0.001 2 1250 33.2 < 0.001<br />

Community×year 7 185 1.47 0.17 7 204 5.41 < 0.001<br />

Grazing×year 14 245 1.95 0.02 14 101 2.68 < 0.001<br />

Community×grazing×year 14 73 0.58 0.88 14 73 1.94 0.02<br />

Residuals 431 126 431 38<br />

Community 1 1503.9 71.52 < 0.001<br />

Grazing 2 126.3 6.01 0.03<br />

Year 7 1683.2 80.04 < 0.001<br />

Community×grazing 2 354.3 16.85 < 0.001<br />

Community×year 7 605.1 28.77 < 0.001<br />

Grazing×year 14 0.8 0.04 0.96<br />

Community×grazing×year 14 14.4 0.68 0.51<br />

Residuals 431 21<br />

df: <strong>de</strong>gree of freedom, MS: Mean-square, F: F-value, P: P-value<br />

Forbs<br />

Grasses<br />

Legumes<br />

198


Chapitre VII<br />

Table 3 Mixed effect mo<strong>de</strong>ls for differences in cover of Brachypodium phoenicoï<strong>de</strong>s<br />

(dominant species of xeric meadows), Halimoine portulacoï<strong>de</strong>s and Bromus madritensis<br />

(dominant species of salt-marshes) between communities and grazing treatments across time,<br />

taking into account interactions between factors<br />

Source of variation df MS F P<br />

Brachypodium<br />

phoenicoï<strong>de</strong>s<br />

Community<br />

1 43539<br />

1312.<br />

< 0.001<br />

9<br />

Grazing 2 2520 76.0 < 0.001<br />

Year 7 200 6.0 < 0.001<br />

Community×grazing 2 1409 42.5 < 0.001<br />

Community×year 7 219 6.6 < 0.001<br />

Grazing×year 14 113 3.4 < 0.001<br />

Community×grazing×year 14 72 2.2 < 0.01<br />

Residuals 43<br />

1<br />

33<br />

Halimoine Community 1 19022 509.9 < 0.001<br />

portulacoï<strong>de</strong>s Grazing 2 45 1.2 0.30<br />

Year 7 108.9 2.9 < 0.01<br />

Community×grazing 2 27.5 0.7 0.48<br />

Community×year 7 84 2.3 0.03<br />

Grazing×year 14 13.1 0.4 0.99<br />

Community×grazing×year 14 11.9 0.3 0.99<br />

Residuals 43<br />

1<br />

37.3<br />

Bromus<br />

madritensis<br />

Community<br />

1<br />

9962.<br />

4<br />

193.4 < 0.001<br />

Grazing<br />

2<br />

1780.<br />

8<br />

34.6 < 0.001<br />

Year 7 342.6 6.7 < 0.001<br />

Community×grazing 2 1022 19.8 < 0.001<br />

Community×year 7 184.4 3.6 < 0.001<br />

Grazing×year 14 69.4 1.3 0.18<br />

Community×grazing×year 14 28.7 0.6 0.90<br />

Residuals 43<br />

1<br />

51.5<br />

df: <strong>de</strong>gree of freedom, MS: Mean-square, F: F-value, P: P-value<br />

199


Chapitre VII<br />

Fig. 4 Dynamics of cover across the 8 years for a) perennial grasses, b) perennial forbs, c)<br />

Brachypodium phoenicoï<strong>de</strong>s, d) Halimoine portulacoï<strong>de</strong>s and e) Bromus madritensis in the<br />

two communities<br />

Effect of grazing on community structure<br />

Mixed effect mo<strong>de</strong>ls showed a significant impact of grazing on diversity in both<br />

communities and on evenness in xeric meadows (Table 4). In xeric meadows, diversity and<br />

evenness <strong>de</strong>creased in the Exclusion treatment (Fig. 5). Across the eight years, the diversity<br />

did not vary significantly per grazing treatment in salt-marshes, but a ten<strong>de</strong>ncy of a <strong>de</strong>crease<br />

of diversity was observed with the Short Grazing and the Status Quo treatments.<br />

200


Chapitre VII<br />

Table 4 Mixed effect mo<strong>de</strong>ls for differences in diversity and evenness between grazing<br />

treatments and in relation with cover of dominant species across time, taking into account<br />

interactions between factors<br />

Source of variation Shannon diversity Evenness<br />

df MS F P df MS F P<br />

Grazing 2 7.7 80.9 < 0.001 2 0.2 95.0 < 0.001<br />

Dominant cover 1 56.8 598.7 < 0.001 1 1.3 723.8 < 0.001<br />

Year 7 1.3 13.5 < 0.001 7 0.01 6.0 < 0.001<br />

Grazing ×dominant 2 1.4 15.1 < 0.001 2 0.04 21.3 < 0.001<br />

Grazing ×year 14 0.2 1.9 0.03 14 0.001 0.8 0.71<br />

Dominant ×year 7 0.2 1.8 0.10 7 0.004 2.0 0.06<br />

Grazing × dominant<br />

×year<br />

14 0.1 1.4 0.16 14 0.004 2.3 0.007<br />

Residuals 183 183<br />

Grazing 2 2.4 14.6 < 0.001 2 0.01 2.2 0.11<br />

Dominant cover 1 41.6 257.8 < 0.001 1 0.7 218.2 < 0.001<br />

Year 7 1.1 6.6 < 0.001 7 0.01 2.3 0.03<br />

Grazing ×dominant 2 0.2 1.4 0.24 2 0.01 3.6 0.03<br />

Grazing ×year 14 0.1 0.7 0.74 14 0.003 0.8 0.67<br />

Dominant ×year 7 0.3 2 0.06 7 0.02 4.5 < 0.001<br />

Grazing × dominant<br />

×year<br />

14 0.1 0.8 0.65 14 0.004 1.2 0.25<br />

Residuals 199 199 0.003<br />

df: <strong>de</strong>gree of freedom, MS: Mean-square, F: F-value, P: P-value<br />

Xeric meadows<br />

Salt-marshes<br />

201


Chapitre VII<br />

Fig. 5 Dynamics of Shannon diversity and evenness across the 8 years in the two<br />

communities and in each grazing treatments<br />

Relationships between community structure and cover of dominant species<br />

Mixed effect mo<strong>de</strong>ls showed a significant impact of the abundance of dominant<br />

species on diversity and evenness in xeric meadows and in Mediterranean salt-marshes (Table<br />

4). The interaction term Grazing treatment*Dominant cover*time was significant only in<br />

xeric meadows for evenness.<br />

202


Relationships between rainfall patterns and community structure<br />

Chapitre VII<br />

The diversity and evenness indices varied only slightly between years, both in the<br />

xeric meadows and in the Mediterranean salt-marshes. In xeric meadows, species diversity or<br />

evenness were not correlated with any of the seasonal variations of P-ETP consi<strong>de</strong>red. In the<br />

Mediterranean salt-marshes, diversity and evenness were positively correlated with the<br />

autumn P-PET value (F=5.679, P=0.055; F=8.169, P=0.029 respectively) for the Status Quo<br />

treatment. Plant diversity was positively correlated with both spring and autumn P-PET<br />

values for the Short Grazing and Status Quo treatments (F=6.986, P=0.036; F=79.44,<br />

P=0.0002 respectively).<br />

Discussion<br />

A response to variability of rainfall patterns stronger in the community the most exposed to<br />

stress.<br />

At the time of our eight year experiment, the grazing pressure in the control<br />

treatments had been constant and unchanged for over 40 years. The changes observed<br />

occurring in vegetation were therefore attributable to effects of the variability of rainfall<br />

patterns. During the experiment, we observed an effect of variability in rainfall patterns on<br />

structure in the two communities studied, confirming our hypothesis of the role of variability<br />

of rainfall patterns. This effect concerns species composition and functional groups, but does<br />

not alter species diversity and evenness. This effect is more marked in xeric meadows but<br />

follows a more strictly linear pathway in salt-marshes, especially for the last years of the<br />

study. This strong influence of the amount of rainfall characterizing the whole of the period<br />

studied on species composition is in accordance with other studies conducted in<br />

Mediterranean and semi-arid environments (e.g. Allen et al. 1995; Clarke et al. 2005; Kutiel<br />

et al. 2000; Sternberg et al. 2000).<br />

The strong increase in perennial species observed during the spring of 2008 seems to<br />

be largely attributable to the high amount of rainfall, which probably increased the water<br />

resource availability leading to the increase of competition between species. Such an increase<br />

in the contribution of perennials was especially observed in our salt-marshes, for Halimione<br />

portulacoï<strong>de</strong>s which is known to be a strong competitive species (Bouchard et al. 2003; Keihl<br />

et al. 1996). Despite the non-significant inter-annual variation of species diversity, we<br />

203


Chapitre VII<br />

observed a correlation between autumn rainfall and species diversity, with lower species<br />

diversity obtained for the autumns of 2006 and 2007, the least rainy during the experiment.<br />

A contrasting effect of the cessation of grazing between the two communities<br />

The cessation of grazing strongly altered the structure of xeric meadows, leading to a<br />

linear pathway, thus validating our hypothesis. Over years, a strong increase in perennial<br />

species is observed, in particular in Brachypodium phoenicoï<strong>de</strong>s clearly reducing the<br />

contribution of annual species by competition, due to the growth in height. This effect<br />

negatively affects the species diversity and evenness of xeric meadows, a result reported in<br />

many others studies (Allen et al. 1995; Díaz et al. 2001; Firincioglu et al. 2007; McIntyre and<br />

Lavorel 2001; Milchunas et al. 1988; Noy-Meir et al. 1989; Noy-Meir 1995; Peco et al. 2005;<br />

Sternberg et al. 2000).<br />

In our Mediterranean salt-marshes, we observed weak changes in species composition,<br />

and no change in diversity and evenness, showing a low impact of the cessation of grazing,<br />

and validating our hypothesis of a lesser role of grazing in communities subjected to abiotic<br />

stress. This result fits with observations ma<strong>de</strong> on Atlantic salt marshes where the contribution<br />

of Halimione portulacoï<strong>de</strong>s is reduced by cattle grazing (Tessier et al. 2003), as we were able<br />

to test the effect of the presence of grazing with the grazing treatment.<br />

The cessation of grazing in the two communities maintains their structural divergence,<br />

but the divergence only concerns a species replacement in the xeric meadows (a replacement<br />

of annuals by perennials).<br />

Response to change in extensive grazing modality is stronger in the community the most<br />

exposed to stress.<br />

The application of the same annual pressure but over a short period (i.e. from 6<br />

months to 3 days), unexpectedly leads to mo<strong>de</strong>rate changes in community structure in the two<br />

communities. Our hypothesis of a strong increase in species diversity in the xeric meadows<br />

was not confirmed. Changes in species composition are weak and the proportion of annual<br />

species is only maintained mainly by forbs and grasses, known to be grazing tolerant<br />

(Golo<strong>de</strong>ts et al. 2010; Illius and O’Connor 1999; Noy-Meir et al. 1989; Noy-Meir 1995; Peco<br />

et al. 2005). Although our two days un<strong>de</strong>r grazing application can be assimilated to<br />

disturbance (Grime 1979; White and Pickett 1985) by creating large gaps of bare soil (pers.<br />

obs.) increasing the vacant niches for species, it did not lead to an increase in species<br />

204


Chapitre VII<br />

diversity. Our hypothesis that a strong instantaneous grazing pressure will lead to an increase<br />

of species diversity is clearly ruled out; but the other hypothesis that the impact of change in<br />

grazing modality will be less perceptible for the community subjected to the greatest stress is<br />

confirmed. The tolerance of the water shortage generated by stress, like salt, minimizes the<br />

effect of the addition of a new stress or disturbance, such as grazing in our study, particularly<br />

for succulent species (Milchunas et al. 1988). In dry environments, succulent plants show<br />

high tolerance of stress such as salt (Hale and Orcutt 1987). This is the case for Halimione<br />

portulacoï<strong>de</strong>s, a dominant species in our salt-marshes.<br />

Compositional changes over the 8 years were less marked un<strong>de</strong>r this Short Grazing<br />

treatment than with the cessation of grazing, which was to be expected as the species pool<br />

was already filtered by the occurrence of long lasting grazing in the system studied (Metzger<br />

et al. 2005; Ward et al. 1998). However a change in floristic cortege in xeric meadows occurs<br />

along with the increase of the contribution of salt-tolerant species such as Bromus<br />

madritensis, a dominant species of the Mediterranean salt-marshes, leading to an increase in<br />

similarity between the two communities. The augmentation of such salt-tolerant species may<br />

result in an increase in soil salt content due to an increase of instantaneous pressure,<br />

particularly of trampling, as suggested by Amiaud et al. (1998). It highlights one of the three<br />

types of grazing impact <strong>de</strong>scribed by Cingolani et al. (2003) which consists in the<br />

amplification of the similarity between grazed communities due to the alteration of local<br />

abiotic conditions.<br />

Conclusion<br />

Our experiment was carried out over 8 years: this was a necessary precaution since<br />

the communities we studied may have shown slow responses. Probably, we need to un<strong>de</strong>rtake<br />

long term monitoring to be sure of characterizing the communities’ response to rainfall and<br />

changes in grazing management.<br />

Despite that, the annual variation of rainfall patterns induces an immediate response to<br />

this variability although species have been selected by these environmental conditions, due to<br />

the high proportion of annual species. Our results suggest that abiotic conditions, such as soil<br />

salinity resulting in salt-tolerant vegetation, buffer the effects of grazing. In consequence, the<br />

impact of grazing, shown to be <strong>de</strong>terminant in Mediterranean pastures, is mediated in<br />

environmentally constrained communities, even if subjected to a long history of grazing. The<br />

205


Chapitre VII<br />

effect of high instantaneous grazing pressure is not as strong as expected, although we know<br />

that this instantaneous pressure is an efficient management approach to control colonization<br />

by shrubs in xeric meadows (Mesléard et al. 2004; Mesléard et al. 2010).<br />

Acknowledgements<br />

We acknowledge the help of N. Yavercovski and L. Willm for technical assistance. This work<br />

is a publication from the DIVHERB Project (French national program ECOGER fun<strong>de</strong>d by<br />

the Institut National <strong>de</strong> la Recherche Agronomique) and a contribution to GDR 2574<br />

‘TRAITS’. This study was supported by a doctoral research grant from the Ministère <strong>de</strong><br />

l’Enseignement Supérieur et <strong>de</strong> la Recherche, the Fondation PRO-VALAT and the Fondation<br />

MAVA.<br />

206


Literature cited:<br />

Chapitre VII<br />

Alday JG, Marrs RH, Martinez-Ruiz C (2010) The importance of topography and climate on<br />

short-term revegetation of coal wastes in Spain. Ecological Engineering, 36(4):579-585.<br />

doi:10.1016/j.ecoleng.2009.12.005<br />

Allen RB, Bastow WJ, Mason CR (1995) Vegetation change following exclusion of grazing<br />

animals in <strong>de</strong>pleted grasslands, Central Otago, New Zealand. Journal of Vegetation Science,<br />

6(5):615-626<br />

Allison KS (1992) The influence of rainfall variability on the species composition of a<br />

northern California salt marsh plant assemblage. Vegetatio, 101:145-160.<br />

doi:10.1007/BF00033198<br />

Amiaud B, Bouzillé JB, Tourna<strong>de</strong> F, Bonis A (1998) Patterns of soil salinities in old<br />

embanked marshlands in Western France. Wetlands, 18(3),482-494.<br />

doi:10.1007/BF03161540<br />

Bennington CC, Thayne WV (1994) Use and misuse of mixed mo<strong>de</strong>l analysis of variance in<br />

ecological studies. Ecology, 75(3):717-722. doi:10.2307/1941729<br />

Bouchard V, Tessier M, Digaire F, Vivier JP, Valery L, Gloaguen JC, Lefeuvre JC (2003)<br />

Sheep grazing as management tool in Western European salt marshes. Comptes Rendus <strong>de</strong><br />

Biologies, 326(1):148-157. doi:10.1016/S1631-0691(03)00052-0<br />

Callaway, R.M., Jones, S., Ferren, W.R., Parikh, A. (1990) Ecology of a Mediterraneanclimate<br />

estuarine wetland at Carpinteria, California: plant distributions and soil salinity in the<br />

upper marsh. Canadian Journal of Botany, 68(5):1139-11146. doi: 10.1139/b90-144<br />

Callaway RM, Sabraw CS (1994) Effects of variable precipitation on the structure and<br />

diversity of a California salt marsh community. Journal of Vegetation Science, 5(3):433-438.<br />

doi:10.2307/3235867<br />

Clarke P, Latz PK, Albrecht DE (2005) Long-term changes in semi-arid vegetation: Invasion<br />

of an exotic perennial grass has larger effects than rainfall variability. Journal of Vegetation<br />

Science, 16(2):237-248. doi:10.1111/j.1654-1103.2005.tb02361.x<br />

Cingolani, A.M., Cabido, M.R., Renison, D., Solis Neffa, Viviana (2003) Combined effects of<br />

environment and grazing on vegetation structure in Argentine granite grasslands. Journal of<br />

Vegetation Science, 14: 223-232.<br />

Del Pozo A, Ovalle C, Casado MA, Acosta B, <strong>de</strong> Miguel JM (2006) Effects of grazing<br />

intensity in grasslands of the Espinal of central Chile. Journal of Vegetation Science, 17(6):<br />

791-798. doi:10.1111/j.1654-1103.2006.tb02502.x<br />

Díaz S, Noy-Meir I, Cabido M. (2001) Can grazing response of herbaceous plants be<br />

predicted from simple vegetative traits? Journal of Applied Ecology, 38(3):497-508.<br />

doi:10.1046/j.1365-2664.2001.00635.x<br />

Díaz S, Lavorel S, McIntyre S, Falczuk V, Casanoves F, Milchunas DG, Skarpe C, Rusch G,<br />

Sternberg M, Noy-Meir I, Landsberg J, Zhang W, Clark H, Campbell BD (2007) Plant-trait<br />

responses to grazing-a global synthesis. Global Change Biology, 13(2):313-341.<br />

doi:10.1111/j.1365-2486.2006.01288.x<br />

Firincioglu HK, Seefeldt SS, Sahin B (2007) The effects of long-term grazing exclosures on<br />

range plants in the central anatolian region of Tukey. Environmental Management, 39(3):<br />

326-337. doi:10.1007/s00267-005-0392-y<br />

207


Chapitre VII<br />

Golo<strong>de</strong>ts C, Kigel J, Sternberg M (2010) Recovery of plant species composition and<br />

ecosystem function after cessation of grazing in a Mediterranean grasslands. Plant Soil, 329<br />

(1-2):365-378. doi:10.1111/j.1365-2664.2011.02031.x<br />

Grime JP (1973) Competition exclusion in herbaceous vegetation. Nature 242:344-347.<br />

doi:10.1038/242344a0<br />

Grime JP (1979) Plant strategies and Vegetation processes. Wiley, Chichester, England.<br />

doi:10.1007/BF02895358<br />

Grime JP (1998) Benefits of plant diversity to ecosystems: immediate, filter and foun<strong>de</strong>r<br />

effects. Journal of Ecology, 86(6):902-910. doi:10.1046/j.1365-2745.1998.00306.x<br />

Hale MG, Orcutt DM (1987) The physiology of plants un<strong>de</strong>r stress. John Wiley and Sons,<br />

New York.<br />

Halpern CB (1988) Early successional pathways and the resistance and resilience of forest<br />

communities. Ecology, 69(6):1703-1715. doi:10.2307/1941148<br />

Heurteaux P (1970) Rapports <strong>de</strong>s eaux souterraines avec les sols halomorphes et la végétation<br />

en Camargue. Revue d’Ecologie (Terre et Vie), 24:467-510.<br />

doi:10.1017/S0376892900033683<br />

Illius AW, O’Connor TG (1999) On the relevance of nonequilibrium concepts to arid and<br />

semiarid grazing systems. Ecological Applications, 9(3):798–813.<br />

doi:10.1016/j.ecolmo<strong>de</strong>l.2005.02.001<br />

Johnstone IM (1986) Plant invasion windows: a time-based classification of invasion<br />

potential. Biological Reviews, 61(4):369-394. doi:10.1111/j.1469-185X.1986.tb00659.x<br />

Kiehl K, Eischeid I, Gettner S, Walter J (1996) Impact of different sheep grazing intensities<br />

on salt marsh vegetation in northern Germany. Journal of Vegetation Science, 7(1):99-106.<br />

doi:10.2307/3236421<br />

Kutiel P, Kutiel H, Lavee H (2000) Vegetation response to possible scenarios of rainfall<br />

variations along a Mediterranean-extreme arid climatic transect. Journal of Arid<br />

Environments, 44(3):277–290<br />

Legendre P, An<strong>de</strong>rson MJ (1999) Distance-based redundancy analysis: testing multispecies<br />

responses in multifactorial ecological experiments. Ecological Monographs, 69(1):1-24.<br />

doi:10.1890/0012-9615(1999)069[0001:DBRATM]2.0.CO;2<br />

Lemaire S, Tamisier A, Gagnier F (1987) Surfaces, distribution et diversité <strong>de</strong>s principaux<br />

milieux <strong>de</strong> Camargue (France). Evolution par analyse <strong>de</strong>s photos aériennes (1942-1984).<br />

Revue d’Ecologie (Terre Vie), 4:47-56<br />

Lock K, Adriaens T, Stevens M (2010) Distribution and ecology of the Belgian Campo<strong>de</strong>a<br />

species (Diplura: Campo<strong>de</strong>idae). European Journal of Soil Biology, 46(1):62–65.<br />

doi:10.1016/j.ejsobi.2009.09.003<br />

McIntyre S, Lavorel S (2001) Livestock grazing in subtropical pastures: steps in the analysis<br />

of attribute response and functional types. Journal of Ecology, 89(2):209-226.<br />

doi:10.1046/j.1365-2745.2001.00535.x<br />

Mesléard F, Desnouhes L, Pineau OM (2004) Herbivores domestiques ou sauvages ? leur role<br />

respectif dans la gestion conservatoire <strong>de</strong>s mileiux. Espaces Naturels 8:8-9<br />

Mesléard F, Mauchamp A, Pineau O, Dutoit T (2011) Rabbits are More Effective than Cattle<br />

for Limiting Shrub Colonization in Mediterranean Xero-Halophytic Meadows. Ecoscience<br />

18(1):37-41. doi:10.2980/18-1-3383<br />

208


Chapitre VII<br />

Metzger KL, Coughenour MB, Reich RM, Boone RB (2005) Effects of seasonal grazing on<br />

plant species diversity and vegetation structure in a semi-arid ecosystem. Journal of Arid<br />

Environments, 61(1):147-160. doi:10.1016/j.jari<strong>de</strong>nv.2004.07.019<br />

Milchunas DG, Sala OE, Lauenroth WK (1988) A generalized mo<strong>de</strong>l of the effects of grazing<br />

by large herbivores on grasslands community structure. The American Naturalist, 132(1):87-<br />

106. doi:10.1007/s00442-005-0019-2<br />

Milchunas DG, Noy-Meir I (2002) Grazing refuges, external avoidance of herbivory and plant<br />

diversity. Oikos, 99(1):113-130. doi:10.1034/j.1600-0706.2002.990112.x<br />

Molinier R, Tallon G (1968) Etu<strong>de</strong>s botaniques en Camargue. II vers la forêt en Camargue.<br />

Revue d’Ecologie. Terre Vie, 19:3-197<br />

Myers JA, Harms KE (1989) Seed arrival and ecological filters interact to assemble highdiversity<br />

plant communities. Ecology, 92(3):676–686. doi: 10.1111/j.1461-<br />

0248.2009.01373.x<br />

Noy-Meir I, Gutman M, Kaplan Y (1989) Responses of Mediterranean Grasslands Plants to<br />

Grazing and Protection. Journal of Ecology, 77(1):290-310<br />

Noy-Meir I (1995) Interactive effects of fire and grazing on structure and diversity of<br />

Mediterranean grasslands. Journal of Vegetation Science, 6(5):701-710. doi:10.2307/3236441<br />

Olff H, Ritchie ME (1998) Effects of herbivores on grasslands plant diversity. Trends in<br />

Ecology and Evolution, 13(7):261-265<br />

Peco B, Levassor C, Pineda FD (1983) Diversité et structure spatiale <strong>de</strong> pâturages<br />

méditérranéens en cours <strong>de</strong> succession. Ecologia Mediterranea, 9:223-234<br />

Peco B, <strong>de</strong> Pablos I, Traba J, Levassor C (2005) The effect of grazing abandonment on<br />

species composition and functional traits: the case of <strong>de</strong>hesa grasslands. Basic and Applied<br />

Ecology, 6(2):175-183. doi:10.1016/j.baae.2005.01.002<br />

Peco B, Sanchez AM, Azcarate FM (2006) Abandonment in grazing systems: Consequences<br />

for vegetation and soil. Agriculture, Ecosystems and Environment, 113(1-4):284–294.<br />

doi:10.1016/j.agee.2005.09.017<br />

Pielou EC (1975) Ecological diversity. John Wiley and Sons, New York.<br />

Scarnati L, Attore F, De Sanctis M, Farcomeni A, Francesconi F, Mancini M, Bruno F (2009)<br />

A multiple approach for the evaluation of the spatial distribution and dynamics of a forest<br />

with Taxus baccata and Ilex aquifolium. Biodiversity and Conservation, 18(12):3099-3113.<br />

doi:10.1007/s10531-009-9629-z<br />

Sternberg M, Gutman M, Perevolotsky A, Ungar ED, Kigel J (2000) Vegetation response to<br />

grazing management in a Mediterranean herbaceous community: a functional group approach.<br />

Journal of Applied Ecology, 37(2):224-237. doi:10.1046/j.1365-2664.2000.00491.x<br />

Tessier M, Vivier JP, Ouin A, Gloaguen JC, Lefeuvre JC (2003) Vegetation dynamics and<br />

plant species interactions un<strong>de</strong>r grazed and ungrazed conditions in a western European salt<br />

marsh. Acta Oecologica, 24(2):103-111. doi:10.1016/S1146-609X(03)00049-3<br />

Ward D, Ngairorue BT, Kathena J, Samuels R, Ofran Y (1998) Land <strong>de</strong>gradation is not a<br />

necessary outcome of communal pastoralism in arid Namibia. Journal of Arid Environments<br />

40(4):357–371. doi:10.1006/jare.1998.0458<br />

209


Chapitre VII<br />

210


RESUME PARTIE 3<br />

Dans les prairies du Marais Poitevin, l’utilisation <strong>de</strong>s modèles nuls nous a permis <strong>de</strong><br />

montrer que l’assemblage <strong>de</strong>s espèces le long du gradient résulterait <strong>de</strong> processus<br />

déterministes. En effet, une convergence <strong>de</strong>s valeurs <strong>de</strong> traits (SLA et hauteur) est mise en<br />

évi<strong>de</strong>nce : elle résulterait <strong>de</strong>s conditions abiotiques <strong>de</strong> l’habitat et plus particulièrement <strong>de</strong>s<br />

conditions d’inondation. De plus, il est à la fois observé une divergence dans les écarts <strong>de</strong>s<br />

valeurs <strong>de</strong> traits entre espèces qui coexistent et une convergence dans ces écarts au sein <strong>de</strong><br />

chaque communauté. Cependant cette divergence et cette convergence ne concernent pas les<br />

mêmes traits suivant la communauté. Dans la communauté mésophile par exemple, une<br />

divergence est plutôt observée pour la distribution <strong>de</strong>s valeurs <strong>de</strong> SLA et une convergence<br />

pour les valeurs <strong>de</strong> hauteur ; alors que l’inverse est observé dans la communauté hygrophile.<br />

Ces résultats nous indiquent que la compétition agirait comme un filtre écologique suivant les<br />

capacités compétitives <strong>de</strong>s espèces leur permettant <strong>de</strong> persister dans un milieu compétitif.<br />

Ainsi, les conditions d’inondations et la compétition entre plantes seraient responsables <strong>de</strong><br />

l’assemblage <strong>de</strong>s espèces tout le long du gradient.<br />

Les pelouses strictes méditerranéennes, c’est-à-dire la communauté la moins<br />

contrainte, apparaissent être contrôlées par les patrons <strong>de</strong> pluviosité dans le traitement témoin<br />

correspondant à 40 années d’une charge extensive <strong>de</strong> pâturage. En revanche, un changement<br />

dans la modalité <strong>de</strong> pâturage apparaît être un filtre écologique important, contrôlant la<br />

structure <strong>de</strong> la communauté. En effet, un arrêt du pâturage conduit à une perte <strong>de</strong> diversité, à<br />

travers l’augmentation en abondance d’une graminée pérenne Brachypodium phoenicoï<strong>de</strong>s.<br />

En revanche, le niveau <strong>de</strong> stress plus élevé dans les prés salés apparaît être un facteur qui<br />

module la réponse <strong>de</strong> cette communauté aux patrons <strong>de</strong> pluviosité et aux effets du<br />

changement dans les modalités <strong>de</strong> pâturage. L’espèce dominante <strong>de</strong>s prés salés, Halimione<br />

portulacoï<strong>de</strong>s, apparaît contrôler la structure <strong>de</strong> cette communauté.<br />

211


212


DISCUSSION GENERALE<br />

L<br />

’objectif <strong>de</strong> ce travail <strong>de</strong> recherche était d’i<strong>de</strong>ntifier les mécanismes structurant les<br />

communautés végétales le long <strong>de</strong> gradients environnementaux et d’évaluer leur<br />

importance. Dans cet objectif, nous avons privilégié <strong>de</strong>s méthodologies différentes,<br />

car aucune méthodologie ne peut, à elle seule, décrire les patrons <strong>de</strong> variation <strong>de</strong> composition<br />

et <strong>de</strong> structure <strong>de</strong>s communautés naturelles et expliquer les mécanismes sous-jacents (Noy-<br />

Meir & Whittaker 1977).<br />

Dans un premier temps, nous nous sommes intéressés à quantifier la part <strong>de</strong> la<br />

compétition relativement aux conditions d’inondation pour expliquer la structuration <strong>de</strong>s<br />

communautés végétales <strong>de</strong>s prairies humi<strong>de</strong>s du Marais Poitevin, en cherchant à répondre aux<br />

questions suivantes :<br />

- Les espèces sont-elles toutes contraintes par les mêmes conditions d’inondation ?<br />

- La compétition est-elle structurante le long du gradient d’inondation ? Comment<br />

varient les <strong>de</strong>ux composantes <strong>de</strong> la compétition (intensité et importance) le long du<br />

gradient d’inondation ?<br />

Pour cela, <strong>de</strong>s approches expérimentales en conditions contrôlées et en conditions<br />

naturelles ont été utilisées pour établir précisément si un déplacement <strong>de</strong>s niches est observé<br />

en réponse à la compétition (Silvertown 2004) (Chapitres 1 à 3). Ces approches ont été<br />

conduites au niveau <strong>de</strong> l’espèce. Ces approches, où la présence <strong>de</strong> voisins est contrôlée,<br />

limitent cependant les généralisations <strong>de</strong>s résultats <strong>de</strong>s interactions plante-plante testées: ces<br />

limites sont liées à la pério<strong>de</strong> d’étu<strong>de</strong> ainsi qu’à l’i<strong>de</strong>ntité et l’abondance <strong>de</strong>s voisins<br />

(Goldberg 1996). En complément, une approche fondée sur l’analyse <strong>de</strong> données<br />

démographiques a été utilisée. Les <strong>de</strong>ux composantes <strong>de</strong> la compétition (intensité et<br />

importance) ont été estimées à travers l’estimation <strong>de</strong>s probabilités <strong>de</strong> survie et <strong>de</strong><br />

colonisation <strong>de</strong>s espèces ainsi que l’estimation <strong>de</strong> la croissance en biomasse, le long du<br />

gradient d’inondation naturel (Chapitre 4). En effet, considérer une perspective dynamique<br />

<strong>de</strong>s populations ou <strong>de</strong>s communautés est nécessaire pour initier une meilleure compréhension<br />

du rôle <strong>de</strong> la compétition dans l’organisation <strong>de</strong> celles-ci (Aarssen & Keogh 2002 ; Frekleton<br />

et al. 2009). Enfin, une approche fonctionnelle a été utilisée pour évaluer l’effet <strong>de</strong>s filtres<br />

abiotique et biotique sur l’assemblage <strong>de</strong>s communautés, et ce, à l’échelle <strong>de</strong>s communautés<br />

(Chapitre 6).<br />

213


Dans un second temps, nous nous sommes intéressés aux mécanismes responsables <strong>de</strong><br />

la structure <strong>de</strong>s communautés végétales méditerranéennes <strong>de</strong> Camargue soumises à <strong>de</strong>ux<br />

niveaux <strong>de</strong> contraintes abiotiques (salinité du sol), en cherchant à répondre aux questions<br />

suivantes :<br />

- Quels sont les rôles respectifs <strong>de</strong> la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité (hétérogénéité<br />

<strong>de</strong> la ressource en eau) et <strong>de</strong> la gestion par pâturage sur la structure <strong>de</strong>s<br />

communautés ?<br />

- Le niveau <strong>de</strong> productivité <strong>de</strong> la communauté, proxy du niveau <strong>de</strong> stress à l’échelle <strong>de</strong><br />

la communauté sensu Grime (1979), influence-t-il les effets <strong>de</strong> la variabilité <strong>de</strong>s<br />

patrons <strong>de</strong> pluviosité et du pâturage?<br />

La dynamique <strong>de</strong> ces <strong>de</strong>ux communautés a été étudiée sur une pério<strong>de</strong> <strong>de</strong> 8 années par<br />

une approche multivariée (Chapitre 7). Elle a été couplée à une approche démographique axée<br />

sur les analyses d’élasticité <strong>de</strong>s probabilités <strong>de</strong> survie et <strong>de</strong> colonisation, pour apprécier<br />

l’importance du changement <strong>de</strong> modalité <strong>de</strong> pâturage sur le succès écologique d’une espèce<br />

dominante (Chapitre 5).<br />

1. La caractérisation <strong>de</strong>s contraintes<br />

La première étape pour comprendre les mécanismes qui structurent les communautés<br />

végétales consiste en la caractérisation et la quantification <strong>de</strong> la contrainte environnementale<br />

majeure. Il est en effet nécessaire <strong>de</strong> savoir comment les conditions environnementales<br />

peuvent influencer la performance <strong>de</strong>s espèces, afin <strong>de</strong> pouvoir replacer le rôle <strong>de</strong> la<br />

compétition parmi d’autres filtres écologiques affectant eux-mêmes la performance <strong>de</strong>s<br />

espèces.<br />

La notion <strong>de</strong> stress diffère si on considère le stress à l’échelle <strong>de</strong> la communauté ou à<br />

l’échelle <strong>de</strong> l’individu. A l’échelle <strong>de</strong> la communauté, le stress correspond à tout facteur qui<br />

réduit la productivité <strong>de</strong> la communauté (Grime 1979); alors qu’à l’échelle <strong>de</strong> l’individu, il<br />

correspond à la réponse <strong>de</strong> l’individu aux facteurs abiotiques ou « strain » (Wel<strong>de</strong>n & Slauson<br />

1986). Ces <strong>de</strong>ux niveaux d’organisation ont été considérés pour caractériser la contrainte dans<br />

les <strong>de</strong>ux milieux étudiés.<br />

Nous nous sommes placés au niveau <strong>de</strong> la communauté dans les milieux<br />

méditerranéens en suivant la définition <strong>de</strong> Grime (1979) énoncée précé<strong>de</strong>mment, afin <strong>de</strong><br />

214


comparer les effets <strong>de</strong> la pluviosité et du pâturage sur la structure <strong>de</strong> <strong>de</strong>ux communautés<br />

végétales soumises à <strong>de</strong>s niveaux <strong>de</strong> productivité différents en raison d’une salinité du sol<br />

différente. En revanche, afin <strong>de</strong> déterminer la part réelle <strong>de</strong>s conditions d’inondation et <strong>de</strong> la<br />

compétition dans le succès écologique <strong>de</strong>s espèces <strong>de</strong>s prairies humi<strong>de</strong>s du Marais Poitevin,<br />

nous avons choisi l’échelle <strong>de</strong> l’individu pour mesurer le « strain » ; ce travail fait l’objet <strong>de</strong>s<br />

sections 1.1 et 1.2 suivantes.<br />

1.1. Deux contraintes rencontrées le long du gradient d’inondation<br />

Dans les prairies humi<strong>de</strong>s du Marais Poitevin, <strong>de</strong>ux types <strong>de</strong> contraintes sont mises en<br />

évi<strong>de</strong>nce le long du gradient <strong>de</strong> durée d’inondation. Pour une gran<strong>de</strong> majorité <strong>de</strong>s espèces, et<br />

plus particulièrement pour les espèces décrites dans les communautés mésophiles et méso-<br />

hygrophiles, l’augmentation <strong>de</strong> la durée d’inondation, couplée à une augmentation <strong>de</strong> la<br />

hauteur d’eau, représentent la contrainte majeure limitant la performance <strong>de</strong>s individus<br />

(Chapitres 1 et 2). Cet effet négatif <strong>de</strong>s inondations est mis en évi<strong>de</strong>nce sur différentes<br />

mesures <strong>de</strong> performances <strong>de</strong>s espèces : leur survie, leur croissance en biomasse, et également<br />

sur leurs capacités photosynthétiques, résultats fréquemment rapportés dans la littérature<br />

(Baruch 1994 ; Grimoldi 1999 ; van Eck et al. 2004 ; Fraser & Karnesis 2005 ; Jung et al.<br />

2009). En revanche, l’absence d’inondation représente une contrainte majeure pour <strong>de</strong>ux<br />

espèces : Glyceria fluitans et Juncus articulatus. Les niches fondamentales <strong>de</strong> ces 12 espèces<br />

(β-niche, Wilson 1999) le long du gradient <strong>de</strong> durée d’inondation par conséquent se<br />

chevauchent, avec <strong>de</strong>s optima <strong>de</strong> niches fondamentales observés majoritairement dans les<br />

parties non inondées du gradient (Chapitre 2).<br />

Les <strong>de</strong>grés <strong>de</strong> tolérance aux inondations sont différents entre les 12 espèces étudiées, à<br />

la fois entre espèces appartenant à <strong>de</strong>ux communautés différents mais également entre espèces<br />

d’une même communauté (Chapitres 1 et 2). Par exemple, parmi les espèces <strong>de</strong> la<br />

communauté mésophile, Cynosurus cristatus apparaît moins tolérant à l’inondation que<br />

Lolium perenne. Un autre exemple est celui <strong>de</strong>s espèces <strong>de</strong> la communauté hygrophile avec<br />

Trifolium fragiferum, qui est contraint par l’augmentation <strong>de</strong> la durée d’inondation alors que<br />

Juncus articulatus est favorisé par les inondations.<br />

Les conclusions quant à l’impact <strong>de</strong>s conditions d’inondation sur les performances <strong>de</strong>s<br />

espèces en monocultures sont similaires entre l’expérimentation réalisée en conditions<br />

contrôlées (Chapitre 1) et celle réalisée in situ (Chapitre 2). Néanmoins, les mesures <strong>de</strong><br />

« strain » indiquent qu’en conditions naturelles, les individus transplantés sont plus contraints<br />

215


qu’en conditions contrôlées. Il semblerait donc qu’in situ d’autres facteurs environnementaux<br />

que les conditions d’inondation limitent également les performances <strong>de</strong>s individus. Pour<br />

l’expérimentation en conditions naturelles, les individus ont été transplantés en Octobre, avant<br />

l’hiver. Il est possible que les conditions hivernales, avec les gelées observées au cours <strong>de</strong>s<br />

mois <strong>de</strong> janvier et février, aient affecté les performances <strong>de</strong>s espèces.<br />

Notons qu’une seule espèce répond différemment aux conditions d’inondation<br />

naturelles et aux conditions expérimentales : il s’agit <strong>de</strong> Mentha pulegium. En conditions<br />

contrôlées, cette espèce est plutôt contrainte par l’absence d’inondation (Chapitre 1) alors<br />

qu’en conditions naturelles les durées longues d’inondation apparaissent être un facteur <strong>de</strong><br />

contrainte important (Chapitre 2). Il est possible que les fluctuations du niveau d’eau au cours<br />

d’une saison d’inondation, qui interviennent in natura (voir partie Modèles biologiques,<br />

section 2.1), représentent une contrainte supplémentaire pour cette espèce. Les fluctuations du<br />

niveau d’eau sont reconnues pour affecter négativement les performances <strong>de</strong>s espèces<br />

(Casanova & Brock 2000), comme cela pourrait être le cas pour Mentha pulegium.<br />

1.2. Les traits reliés à la tolérance aux inondations<br />

Avec l’augmentation <strong>de</strong> la durée d’inondation est observée une élongation <strong>de</strong>s tiges<br />

(Chapitres 1 et 6) ainsi qu’une augmentation <strong>de</strong> la surface spécifique foliaire ou SLA<br />

(Chapitre 6). Ces <strong>de</strong>ux traits fonctionnels facilitent les échanges gazeux au sein <strong>de</strong>s plantes<br />

sous <strong>de</strong>s conditions inondées, comme cela a été observé pour Juncus articulatus, Glyceria<br />

fluitans et Agrostis stolonifera. Ces résultats sont communément observés pour une gran<strong>de</strong><br />

diversité d’espèces (Blom & Voesenek 1996 ; Clevering et al. 1998 ; Lenssen et al. 1998 ;<br />

Insausti et al. 2001 ; Mommer et al. 2004 ; Voesenek et al. 2004 ; Mommer et al. 2006 ;<br />

Colmer & Voesenek 2009). Nous avons également observés l’existence d’aérenchyme dans<br />

les tiges pour Glyceria fluitans et Juncus articulatus, permettant également <strong>de</strong> faciliter ces<br />

échanges gazeux (Blom & Voesenek 1996 ; Mommer et al. 2004 ; Voesenek 2006 ; Colmer &<br />

Voesenek 2009) ; cet aérenchyme est probablement présent dans leurs racines. De plus, un<br />

développement <strong>de</strong> racines adventices a été observé en conditions inondées pour Mentha<br />

pulegium et Agrostis stolonifera.<br />

216


2. Variation <strong>de</strong> la compétition le long du gradient d’inondation<br />

Cet axe <strong>de</strong> travail montre clairement l’intérêt <strong>de</strong> différencier l’intensité et l’importance<br />

<strong>de</strong> la compétition, qui sont <strong>de</strong>ux composantes différentes <strong>de</strong> la compétition (Chapitre 1). Ce<br />

résultat est en accord avec la démarche scientifique actuelle proposant la différenciation <strong>de</strong><br />

ces <strong>de</strong>ux composantes pour une meilleure compréhension <strong>de</strong> la variation <strong>de</strong> la compétition le<br />

long <strong>de</strong> gradients environnementaux (Brooker et al. 2005 ; Brooker & Kikvidze 2008 ;<br />

Kikvidze et al. 2011). En effet, l’intensité <strong>de</strong> la compétition mesure l’effet <strong>de</strong> la présence <strong>de</strong><br />

compétiteurs per se sur les performances <strong>de</strong> l’espèce cible alors que l’importance <strong>de</strong> la<br />

compétition mesure l’effet <strong>de</strong> la présence <strong>de</strong> compétiteurs sur les performances <strong>de</strong>s espèces<br />

cibles relativement aux autres facteurs environnementaux qui impactent également les<br />

performances <strong>de</strong>s espèces. Jusqu’à récemment, l’importance <strong>de</strong> la compétition était évaluée à<br />

travers <strong>de</strong>s mesures d’intensité <strong>de</strong> la compétition permettant <strong>de</strong> conclure si oui ou non la<br />

compétition jouait un rôle important dans les patrons d’espèces. Actuellement et grâce au<br />

développement d’indices permettant <strong>de</strong> mesurer l’importance <strong>de</strong> la compétition, il est possible<br />

<strong>de</strong> déterminer si la compétition est structurante par rapport à d’autres facteurs<br />

environnementaux. Nous passons alors d’une approche binaire à une approche quantitative<br />

(Kikvidze et al. 2011), comme cela a été entrepris dans cette thèse.<br />

2. 1. Intensité <strong>de</strong> la compétition le long du gradient d’inondation<br />

2.1.1. Ségrégation spatiale <strong>de</strong>s niches le long du gradient d’inondation en réponse à la<br />

présence <strong>de</strong> compétiteurs<br />

En réponse à la présence <strong>de</strong> compétiteurs, un déplacement <strong>de</strong> niche, et plus<br />

particulièrement <strong>de</strong> l’optimum <strong>de</strong> développement, est observé pour certaines espèces (Juncus<br />

gerardii, Bellis perennis, Leontodon autumnalis, Trifolium fragiferum et Mentha pulegium)<br />

validant l’hypothèse d’un compromis entre les capacités compétitives <strong>de</strong>s espèces et la<br />

tolérance à la contrainte inondation (Chapitre 3). Ce déplacement <strong>de</strong> niche est mis en évi<strong>de</strong>nce<br />

par un changement dans la hiérarchie compétitive entre espèces le long du gradient<br />

expérimental d’inondation (Chapitre 1 ; Lenssen et al. 2004).<br />

Suivant l’hypothèse du compromis entre capacités compétitives <strong>de</strong>s espèces et la<br />

tolérance à la contrainte, les espèces exclues <strong>de</strong>s niveaux les moins contraints (non inondés)<br />

vers <strong>de</strong>s niveaux plus contraints (vers les parties inondées), seraient <strong>de</strong>s espèces faiblement<br />

217


compétitives par rapport aux espèces mésophiles ; ce serait le cas <strong>de</strong> Juncus gerardii ou Bellis<br />

perennis par exemple. En effet, les espèces mésophiles caractéristiques <strong>de</strong>s niveaux non-<br />

inondés du gradient (Lolium perenne, Hor<strong>de</strong>um secalinum, Cynosurus cristatus), ne subissent<br />

pas <strong>de</strong> déplacement d’optimum <strong>de</strong> développement. Ces espèces présentent <strong>de</strong>s stratégies<br />

d’acquisition rapi<strong>de</strong> <strong>de</strong>s ressources, c’est-à-dire <strong>de</strong> fortes valeurs <strong>de</strong> surfaces spécifiques<br />

foliaires ainsi qu’une croissance en hauteur importante (Chapitres 2 et 6). Les valeurs <strong>de</strong> la<br />

surface spécifique foliaire sont faibles pour Juncus gerardii (moyenne <strong>de</strong> 10m²/kg dans les<br />

niveaux les plus hauts du gradient) contrairement à celles <strong>de</strong> Lolium perenne (moyenne <strong>de</strong><br />

31.04 m²/kg dans les niveaux les plus hauts du gradient). Le port en rosette <strong>de</strong> Leontodon<br />

autumnalis et Bellis perennis ne leur permet pas croître en hauteur et serait par conséquent un<br />

facteur limitant pour leur développement et leur aptitu<strong>de</strong> compétitive dans les parties non<br />

inondées où la végétation est <strong>de</strong>nse et haute. Ces traits pourraient expliquer leur exclusion<br />

compétitive <strong>de</strong> la communauté mésophile.<br />

Cependant, ces <strong>de</strong>ux mêmes traits ne peuvent pas expliquer l’exclusion <strong>de</strong> toutes les<br />

espèces. Par exemple, l’espèce hygrophile Trifolium fragiferum présente <strong>de</strong>s capacités<br />

d’acquisition rapi<strong>de</strong> <strong>de</strong>s ressources (SLA en moyenne <strong>de</strong> 37.87 m²/kg dans les parties non<br />

inondées) et une croissance en hauteur équivalente aux espèces mésophiles. Globalement, il<br />

apparaît difficile <strong>de</strong> relier expérimentalement le <strong>de</strong>gré <strong>de</strong> similitu<strong>de</strong> entre le compétiteur et<br />

l’espèce cible pour ces <strong>de</strong>ux traits et l’intensité <strong>de</strong> la compétition (Chapitre 1). D’autres traits<br />

fonctionnels sont sans doute impliqués dans la réponse à la compétition, c’est-à-dire pour être<br />

capable <strong>de</strong> tolérer la réduction <strong>de</strong> la disponibilité <strong>de</strong>s ressources. Il existe en effet différentes<br />

stratégies <strong>de</strong> réponse à la compétition, et chacune implique différents ensembles <strong>de</strong> traits<br />

fonctionnels (Keddy et al. 1998 ; Wang et al. 2010). Les espèces peuvent soit présenter <strong>de</strong>s<br />

stratégies <strong>de</strong> conservation <strong>de</strong>s ressources, soit prélever rapi<strong>de</strong>ment ces ressources ou décaler<br />

ce prélèvement dans le temps (Keddy et al. 1998 ; Carlyle & Fraser 2006).<br />

La stratégie mise en place par une espèce pour répondre à la présence <strong>de</strong> compétiteurs<br />

peut varier suivant les conditions environnementales (Keddy et al. 1994 ; Wang et al. 2010).<br />

En conséquence, on peut s’attendre à <strong>de</strong>s variations dans la réponse à la compétition <strong>de</strong><br />

l’espèce selon les conditions environnementales. L’existence <strong>de</strong> compromis entre traits<br />

fonctionnels pourrait expliquer les variations <strong>de</strong> stratégie mise en place par l’espèce, car ces<br />

compromis entre traits peuvent varier suivant les conditions environnementales induisant <strong>de</strong>s<br />

capacités <strong>de</strong> réponse à la compétition différentes, modifiant alors les capacités d’une espèce à<br />

être capable <strong>de</strong> persister dans un environnement donné (Lortie et al. 2004). Cette sensibilité à<br />

l’environnement <strong>de</strong> la réponse à la compétition pourrait expliquer les variations <strong>de</strong> réponse à<br />

218


la compétition observées pour trois espèces le long du gradient expérimental d’inondation<br />

(Chapitre 1, Hor<strong>de</strong>um secalinum, Juncus gerardii et Mentha pulegium). Parallèlement, cette<br />

étu<strong>de</strong> a également montré que d’autres espèces (Lolium perenne, Carex divisa, Bellis<br />

perennis, Glyceria fluitans et Juncus articulatus) ont une réponse à la compétition qui ne<br />

varie pas suivant les conditions d’inondation. Ce résultat suggère une stabilité <strong>de</strong> la<br />

performance <strong>de</strong>s espèces, qui peut néanmoins s’accompagner d’une variation dans l’allocation<br />

<strong>de</strong>s ressources vers les racines ou les parties aériennes suivant les conditions (Tilman 1988),<br />

ce qui n’a pas été mesurée dans ce travail. On peut émettre l’hypothèse que les patrons <strong>de</strong><br />

réponse à la compétition observés (Chapitre 1) sont la résultante d’une stratégie mise en place<br />

par l’espèce dans un milieu donné suivant la disponibilité <strong>de</strong>s ressources, stratégie qui serait<br />

dépendante <strong>de</strong>s conditions locales où l’espèce se trouve, et également du besoin <strong>de</strong>s plantes en<br />

ressources dans un environnement donné comme le propose Taylor et al. (1990). De plus,<br />

l’i<strong>de</strong>ntité du voisin serait également une variable qui influencerait la réponse à la compétition<br />

<strong>de</strong>s espèces (Gau<strong>de</strong>t & Keddy1995 ; Fraser & Keddy 2005 ; Engel & Weltzin 2008). Tout<br />

cela pourrait alors expliquer la diversité <strong>de</strong> résultats concernant la variation <strong>de</strong> la réponse à la<br />

compétition d’espèces le long <strong>de</strong> gradients environnementaux observée dans ce travail et dans<br />

la littérature.<br />

Les méthodologies utilisées pour mesurer la réponse à la compétition dans les<br />

chapitres 1 et 3, c’est-à-dire le contrôle <strong>de</strong> voisins en conditions contrôlées et naturelles, sont<br />

complémentaires. La comparaison précise <strong>de</strong>s niches fondamentales et réalisées in situ<br />

(Chapitre 3) par une caractérisation précise du gradient d’inondation, couplée à une<br />

expérimentation en conditions contrôlées, ont permis <strong>de</strong> réellement établir une analyse <strong>de</strong> la<br />

ségrégation spatiale <strong>de</strong>s niches <strong>de</strong>s espèces en réponse à la compétition (α-niche, Wilson<br />

1999) le long du gradient d’inondation (Silvertown et al. 1999 ; Silvertown 2004).<br />

Il faut tout <strong>de</strong> même noter les limites <strong>de</strong> ce travail <strong>de</strong> comparaison <strong>de</strong>s niches<br />

fondamentales et réalisées <strong>de</strong>s espèces à partir <strong>de</strong>s données récoltées en conditions naturelles.<br />

La comparaison entre les niches fondamentales et réalisées ne s’est faite que sur une part<br />

restreinte du gradient naturel d’inondation. Nous n’avions tout simplement pas toutes les<br />

données <strong>de</strong> caractérisation <strong>de</strong>s inondations disponibles au moment <strong>de</strong> la mise en place <strong>de</strong><br />

l’expérimentation, pour transplanter idéalement les plantes sur une plus large gamme<br />

d’inondation. Il apparaît maintenant nécessaire d’avoir une gamme plus large pour déterminer<br />

si le déplacement <strong>de</strong> la niche en réponse à la compétition reste dans la gamme <strong>de</strong>s conditions<br />

que peuvent tolérer physiologiquement les espèces ; auquel cas les résultats pourraient<br />

219


indiquer <strong>de</strong> la facilitation car la largeur <strong>de</strong> la niche réalisée serait plus gran<strong>de</strong> que celle <strong>de</strong> la<br />

niche fondamentale (Bruno et al. 2003). Néanmoins, l’extrapolation <strong>de</strong>s niches fondamentales<br />

<strong>de</strong>s espèces méso-hygrophiles et <strong>de</strong> certaines hygrophiles tend à conclure que le déplacement<br />

<strong>de</strong> l’optimum <strong>de</strong> la niche reste dans la gamme <strong>de</strong> conditions environnementales que tolère<br />

physiologiquement l’espèce.<br />

2.1.2. Importance <strong>de</strong> la survie et <strong>de</strong> la colonisation pendant les inondations pour être<br />

compétitif<br />

A travers l’approche démographique, les probabilités <strong>de</strong> survie et <strong>de</strong> colonisation ont<br />

été estimées en réponse à l’inondation ainsi que la croissance en biomasse après la fin <strong>de</strong>s<br />

inondations (Chapitre 4). La modélisation <strong>de</strong> ces données nous a permis <strong>de</strong> déterminer le rôle<br />

essentiel <strong>de</strong>s espèces à occuper l’espace durant la phase inondée, via une augmentation <strong>de</strong> la<br />

survie et <strong>de</strong> la colonisation, pour être compétitif et réduire la performance <strong>de</strong>s autres espèces,<br />

notamment après les inondations. Nous avons pu ainsi relier les probabilités <strong>de</strong> survie et <strong>de</strong><br />

colonisation <strong>de</strong> <strong>de</strong>ux espèces dans les parties inondées du gradient à leurs stratégies <strong>de</strong><br />

croissance clonale. L’effet <strong>de</strong>s conditions d’inondation comme filtre écologique sélectionnant<br />

les espèces suivant leur stratégie clonale a d’ailleurs été mise en évi<strong>de</strong>nce dans les prairies<br />

humi<strong>de</strong>s du Marais Poitevin (Benot et al. 2011, Annexe 4).<br />

Tout d’abord, Agrostis stolonifera a un fort effet compétitif sur ces voisins, en raison<br />

<strong>de</strong> la survie et <strong>de</strong> la colonisation améliorées pendant les inondations (Chapitre 4). Ce succès<br />

peut s’expliquer par sa forte croissance clonale lui permettant à la fois <strong>de</strong> tolérer les<br />

inondations et <strong>de</strong> coloniser l’espace. En effet, sa croissance clonale, par le développement <strong>de</strong><br />

stolons photosynthétiques flottants, est reconnue pour être une stratégie importante pour<br />

répondre aux inondations (Soukupova 1994). Après la fin <strong>de</strong>s inondations ces stolons sont<br />

capables <strong>de</strong> s’implanter grâce aux racines adventives, représentant une stratégie d’occupation<br />

<strong>de</strong> l’espace et ainsi permettant d’être compétitif après la fin <strong>de</strong>s inondations (Santamaria<br />

2002 ; Lenssen et al. 2004). Ces stolons peuvent être assez longs : expérimentalement sous<br />

une profon<strong>de</strong>ur d’eau <strong>de</strong> 40 cm, nous avons mesuré <strong>de</strong>s longueurs <strong>de</strong> stolons supérieurs à 1m<br />

avec <strong>de</strong>s racines adventives espacées régulièrement le long <strong>de</strong> ces stolons.<br />

Cette approche permet également <strong>de</strong> comprendre comment une espèce qui est exclut<br />

en réponse à la compétition vers <strong>de</strong>s zones qui lui sont physiologiquement plus contraignantes<br />

(Chapitre 3), est tolérance aux inondations et compétitif dans ces parties inondées du gradient<br />

(Chapitre 4). C’est le cas <strong>de</strong> Juncus gerardii, qui voit sa survie et ses capacités <strong>de</strong> colonisation<br />

220


améliorées sous <strong>de</strong>s conditions d’inondation. Cette espèce est caractérisée par la présence <strong>de</strong><br />

rhizomes, connus pour être <strong>de</strong>s organes <strong>de</strong> réserve (Dong & <strong>de</strong> Kroon 1994). Cet organe <strong>de</strong><br />

stockage permet une utilisation rapi<strong>de</strong> <strong>de</strong>s ressources lorsqu’intervient la fin <strong>de</strong>s inondations,<br />

lui donnant un avantage pour la reprise <strong>de</strong> croissance, augmentant par conséquent sa<br />

colonisation, et ainsi être compétitif.<br />

2.2. Importance <strong>de</strong> la compétition le long du gradient d’inondation<br />

2.2.1. Importance <strong>de</strong> la compétition prédite par le « strain »<br />

La compétition a une gran<strong>de</strong> importance pour expliquer le succès écologique <strong>de</strong>s<br />

espèces car elle est responsable <strong>de</strong> la ségrégation spatiale <strong>de</strong>s niches <strong>de</strong>s espèces, à travers un<br />

compromis entre leurs capacités compétitives <strong>de</strong>s espèces et leur tolérance au « strain »<br />

(Chapitre 3).<br />

Par une approche quantitative, nous avons mis en évi<strong>de</strong>nce expérimentalement que<br />

l’importance <strong>de</strong> la compétition, mesurée à l’échelle <strong>de</strong> l’espèce, est inversement<br />

proportionnelle à l’intensité <strong>de</strong> la contrainte subie par cette espèce, c’est-à-dire l’intensité du<br />

« strain » (Chapitres 1 et 3). Cela signifie que la part <strong>de</strong> la compétition dans le succès<br />

écologique d’une espèce augmente quand cette espèce se trouve dans <strong>de</strong>s conditions<br />

environnementales proches <strong>de</strong> son optimum physiologique <strong>de</strong> développement, comme attendu<br />

par Wel<strong>de</strong>n & Slauson (1986) et Brooker et al. (2005).<br />

Dans les prairies humi<strong>de</strong>s du Marais poitevin, il est ainsi possible <strong>de</strong> conclure que les<br />

limites inférieures <strong>de</strong>s niches <strong>de</strong>s espèces mésophiles, <strong>de</strong>s espèces méso-hygrophiles et <strong>de</strong><br />

Trifolium fragiferum sont déterminées principalement par les inondations (Shipley et al.<br />

1991). Les limites supérieures <strong>de</strong>s niches <strong>de</strong>s espèces méso-hygrophiles et <strong>de</strong> Trifolium<br />

fragiferum sont déterminées principalement par la compétition. Parallèlement, les limites<br />

supérieures <strong>de</strong>s niches <strong>de</strong> Glyceria fluitans et <strong>de</strong> Juncus articulatus seraient d’abord<br />

déterminées par l’absence d’inondation. De plus, il a été mis en évi<strong>de</strong>nce l’importance <strong>de</strong> la<br />

compétition dans les parties faiblement inondées du gradient pour expliquer leur succès<br />

(Chapitre 3). En effet, la présence <strong>de</strong> compétiteurs réduit la gamme <strong>de</strong>s niches <strong>de</strong> ces <strong>de</strong>ux<br />

espèces vers <strong>de</strong>s zones longuement inondées. Ainsi les limites supérieures <strong>de</strong>s niches <strong>de</strong> ces<br />

espèces sont déterminées par l’absence d’inondations et par la compétition, exliquant ainsi<br />

leur distribution dans les parties longuement inondées du gradient.<br />

221


Nous remarquons que moitié moins d’espèces suivent la relation entre l’importance <strong>de</strong><br />

la compétition et l’intensité <strong>de</strong> la contrainte à l’échelle <strong>de</strong> l’espèce (« strain ») en conditions<br />

naturelles (Chapitre 3) qu’en conditions contrôlées (Chapitre 1), suggérant que d’autres<br />

facteurs interviennent in natura dans le succès écologiques <strong>de</strong>s espèces.<br />

2.2.2. Variation <strong>de</strong> l’importance <strong>de</strong> la compétition - apport <strong>de</strong> l’approche<br />

démographique<br />

La mesure <strong>de</strong> l’importance <strong>de</strong> la compétition in natura confirme la part importante <strong>de</strong>s<br />

inondations dans le succès écologique <strong>de</strong>s espèces mais également celle <strong>de</strong> la compétition<br />

(Chapitre 4). La mesure d’importance <strong>de</strong> la compétition a été réalisée à partir <strong>de</strong> la<br />

modélisation <strong>de</strong> différentes mesures <strong>de</strong> performances <strong>de</strong>s espèces : les probabilités <strong>de</strong> survie<br />

et <strong>de</strong> colonisation <strong>de</strong>s espèces durant la saison d’inondation, ainsi que la croissance en<br />

biomasse <strong>de</strong>s espèces estimée après la fin <strong>de</strong>s inondations. La prise en compte <strong>de</strong>s différentes<br />

mesures <strong>de</strong> performances <strong>de</strong>s espèces montrent que les probabilités <strong>de</strong> survie et <strong>de</strong><br />

colonisation sont largement impactées par les inondations que par la compétition pendant la<br />

pério<strong>de</strong> d’inondation. Les inondations ont une influence forte dans le succès écologique <strong>de</strong>s<br />

espèces et plus particulièrement dans leur capacité à occuper l’espace après la fin <strong>de</strong>s<br />

inondations. En revanche, après la fin <strong>de</strong>s inondations, la croissance en biomasse <strong>de</strong>s espèces<br />

est plus largement impactée par la compétition que par tout autre facteur environnemental, y<br />

compris la faible disponibilité en eau en été. Ces résultats montrent ainsi une variation<br />

temporelle <strong>de</strong> l’importance <strong>de</strong> la compétition.<br />

A cette variation temporelle <strong>de</strong> l’importance <strong>de</strong> la compétition entre saisons s’ajoute<br />

une variation spatiale <strong>de</strong> l’importance <strong>de</strong> la compétition le long du gradient. Ces patrons <strong>de</strong><br />

variation <strong>de</strong> l’importance <strong>de</strong> la compétition le long du gradient sont différents selon les<br />

espèces étudiées, spécialement durant la saison d’inondation, reflétant les différentes<br />

capacités <strong>de</strong> survie et <strong>de</strong> colonisation <strong>de</strong>s espèces suivant les changements <strong>de</strong> voisins et <strong>de</strong><br />

conditions environnementales.<br />

Cette approche considère différents paramètres démographiques et place la<br />

quantification in natura <strong>de</strong> l’importance <strong>de</strong> la compétition dans un cadre <strong>de</strong> dynamique <strong>de</strong>s<br />

populations et communautés : cela représente un complément dans la compréhension du rôle<br />

<strong>de</strong> la compétition dans l’organisation <strong>de</strong>s communautés (Goldberg et al. 1999 ; Frekleton et<br />

al. 2009). Nos résultats vont dans le sens <strong>de</strong>s propositions récentes <strong>de</strong> Frekleton et al. (2009)<br />

222


suggérant <strong>de</strong> ne pas se contenter d’une seule mesure <strong>de</strong> performances <strong>de</strong>s espèces pour<br />

déterminer le rôle <strong>de</strong> la compétition. En effet, en considérant seulement la mesure <strong>de</strong><br />

biomasse, nous avons mis en évi<strong>de</strong>nce une relation linéaire entre importance <strong>de</strong> la<br />

compétition et intensité <strong>de</strong> la contrainte (Chapitre 1). Or en considérant d’autres mesures <strong>de</strong><br />

performances, cette relation linéaire <strong>de</strong>vient moins évi<strong>de</strong>nte (Chapitre 4). Cela suggère que la<br />

compétition peut affecter différemment les différents sta<strong>de</strong>s du cycle <strong>de</strong> vie <strong>de</strong>s plantes<br />

(Howard & Goldberg 2001), mesurés dans cette thèse par les probabilités <strong>de</strong> colonisation et<br />

<strong>de</strong> survie <strong>de</strong>s espèces, et qu’il est nécessaire d’en tenir compte pour mesurer l’importance <strong>de</strong><br />

la compétition.<br />

2.3. Intensité et importance <strong>de</strong> la compétition le long du gradient d’inondation<br />

Aucune relation linéaire significative n’a été mise en évi<strong>de</strong>nce entre l’intensité et<br />

l’importance <strong>de</strong> la compétition, qui varient indépendamment le long du gradient expérimental<br />

d’inondation (Chapitre 1). Néanmoins, les résultats ten<strong>de</strong>nt à montrer que ces <strong>de</strong>ux<br />

composantes <strong>de</strong> la compétition sont tout <strong>de</strong> même reliées. En effet, pour les espèces qui sont<br />

affectées par une forte intensité <strong>de</strong> la compétition par les voisins, comme par exemple pour<br />

Juncus gerardii, Mentha pulegium, ou encore Leontodon autumnalis dans les parties non<br />

inondées du gradient (Chapitre 3), la compétition semble avoir une importance décisive pour<br />

leur succès et déterminante <strong>de</strong> leur limite <strong>de</strong> niches supérieures (i.e. dans les zones non<br />

inondées). Une intensité forte <strong>de</strong> la compétition apparaît donc être un pré-requis à une<br />

importance <strong>de</strong> la compétition forte (Bartelheimer et al. 2010 ; Kikvidze et al. 2011).<br />

3. Les traits fonctionnels comme outils d’étu<strong>de</strong> <strong>de</strong>s règles d’assemblage<br />

La ségrégation spatiale <strong>de</strong>s (α)-niches <strong>de</strong>s espèces le long du gradient d’inondation<br />

permet <strong>de</strong> comprendre l’assemblage <strong>de</strong>s espèces et <strong>de</strong> la structure <strong>de</strong>s communautés dans les<br />

prairies humi<strong>de</strong>s du Marais Poitevin. Néanmoins, il existe un chevauchement <strong>de</strong>s niches<br />

réalisées entre espèces le long du gradient (Chapitre 3). L’approche fonctionnelle analysant la<br />

distribution <strong>de</strong>s traits fonctionnels mesurés in situ (Chapitre 6) permet <strong>de</strong> comprendre<br />

comment les espèces peuvent coexister.<br />

L’analyse <strong>de</strong>s distributions <strong>de</strong>s valeurs <strong>de</strong>s <strong>de</strong>ux traits fonctionnels mesurés, le SLA et<br />

la hauteur, montre qu’elles sont significativement différentes d’une distribution aléatoire.<br />

223


L’assemblage <strong>de</strong>s espèces se fait donc suivant <strong>de</strong>s processus déterministes. L’analyse <strong>de</strong>s<br />

distributions <strong>de</strong> ces <strong>de</strong>ux traits met en évi<strong>de</strong>nce un effet du filtre abiotique (l’inondation) ainsi<br />

qu’un effet du filtre biotique (la compétition) sur ces distributions. Ces <strong>de</strong>ux filtres agissent<br />

simultanément sur l’assemblage <strong>de</strong>s espèces, et ce, au sein <strong>de</strong>s trois communautés végétales<br />

existant dans ces prairies (mésophile, méso-hygrophile et hygrophile). L’utilisation <strong>de</strong><br />

différentes métriques mesurant la distribution <strong>de</strong>s traits permet <strong>de</strong> différencier l’effet <strong>de</strong>s<br />

inondations agissant sur la valeur moyenne du trait <strong>de</strong> celui <strong>de</strong> la compétition agissant sur les<br />

écarts <strong>de</strong>s valeurs <strong>de</strong> traits entre espèces qui coexistent.<br />

Tout d’abord, l’analyse <strong>de</strong>s distributions <strong>de</strong> ces traits confirme que <strong>de</strong>s individus<br />

possédant <strong>de</strong>s traits associés à la tolérance à l’inondation, tels une hauteur et un SLA élevés,<br />

sont sélectionnés dans les parties inondées du gradient, conformément aux attendus <strong>de</strong><br />

l’« habitat filtering » (Weiher & Keddy 1995). Ce résultat confirme ceux concernant<br />

l’importance <strong>de</strong> ces traits reliés à la tolérance aux inondations (cf. discussion partie 1.2 ;<br />

Weiher et al. 1998 ; Mommer et al. 2006 ; Jung et al. 2010 ; Violle et al. 2011).<br />

Les distributions <strong>de</strong> <strong>de</strong>ux traits observées confirment également les prédictions du<br />

concept <strong>de</strong> « limiting similarity » (Mac Arthur & Levins 1967). Une maximisation <strong>de</strong>s écarts<br />

<strong>de</strong>s valeurs <strong>de</strong> traits entre espèces coexistantes est mise en évi<strong>de</strong>nce, limitant la similitu<strong>de</strong><br />

fonctionnelle entre voisins à l’échelle locale (α). Ainsi les résultats indiquent qu’une<br />

différenciation <strong>de</strong>s niches <strong>de</strong>s espèces semble avoir lieu. Cette différenciation <strong>de</strong>s niches se<br />

produit pour ces <strong>de</strong>ux traits mesurés, suggérant <strong>de</strong>s stratégies d’acquisition et <strong>de</strong> conservation<br />

<strong>de</strong>s ressources différentes entre espèces (Weiher et al. 1998; Kraft et al. 2008 ; Jung et al.<br />

2010). Le SLA et la hauteur apparaissent être <strong>de</strong>s traits <strong>de</strong> réponse à la fois aux conditions<br />

abiotiques, mais également à l’environnement biotique (Weiher et al. 1999 ; Wright et al.<br />

2004 ; Violle et al. 2011).<br />

Suivant les propositions <strong>de</strong> Weiher & Keddy (1995), une maximisation <strong>de</strong>s écarts <strong>de</strong>s<br />

valeurs <strong>de</strong> traits entre les espèces qui coexistent serait principalement attendue dans les<br />

niveaux non contraints d’un gradient environnemental, c’est-à-dire dans les parties non-<br />

inondées du gradient d’inondation étudié. Les résultats d’expérimentations montrent<br />

cependant que toutes les espèces ne sont pas contraintes par les inondations, et que la<br />

compétition est impliquée dans le succès écologique <strong>de</strong>s espèces à différentes positions le<br />

long du gradient (Partie 1). La compétition influencerait par conséquent la distribution <strong>de</strong>s<br />

traits le long du gradient d’inondation. Ceci est d’autant plus vraisemblable que les prairies<br />

humi<strong>de</strong>s du Marais Poitevin sont <strong>de</strong>s milieux productifs, où le niveau <strong>de</strong> ressources fort est<br />

propice à un fort niveau d’interactions compétitives. Il est donc envisageable que ces<br />

224


interactions compétitives influencent les écarts <strong>de</strong>s valeurs <strong>de</strong> traits entre les espèces<br />

coexistantes le long du gradient, et pas seulement dans les parties non-inondées. Nous<br />

mettons, en effet, en évi<strong>de</strong>nce une augmentation <strong>de</strong> la moyenne <strong>de</strong>s écarts <strong>de</strong>s valeurs <strong>de</strong>s<br />

<strong>de</strong>ux traits entre espèces coexistantes (MeanNTD) et une réduction <strong>de</strong> leur variance<br />

(VarNTD), suggérant une différenciation <strong>de</strong>s niches opérant tout le long du gradient<br />

d’inondation.<br />

Les prairies étudiées étant productives, la proposition <strong>de</strong> Grime (2006) formulée dans<br />

les milieux productifs concernant la distribution <strong>de</strong>s traits doit être prise en compte, c’est-à-<br />

dire une convergence dans la distribution <strong>de</strong>s traits d’acquisition <strong>de</strong>s ressources entre les<br />

espèces qui coexistent. La convergence dans les distributions <strong>de</strong> ces traits résulte <strong>de</strong><br />

l’augmentation <strong>de</strong> l’effet <strong>de</strong> la compétition comme filtre sélectionnant les espèces sur la base<br />

<strong>de</strong> leur aptitu<strong>de</strong> compétitive (Pakeman et al. 2011). Dans notre cas d’étu<strong>de</strong>, cette convergence<br />

est attendue le long du gradient d’inondation. Cette convergence tend à réduire les écarts <strong>de</strong>s<br />

valeurs <strong>de</strong>s traits entre espèces coexistantes. Le long du gradient d’inondation, nous avons<br />

ainsi pu mettre en évi<strong>de</strong>nce une réduction <strong>de</strong>s écarts <strong>de</strong>s valeurs <strong>de</strong>s traits entre espèces qui<br />

coexistent, impliquant la hauteur dans la communauté mésophile et impliquant le SLA dans la<br />

communauté hygrophile, suggérant une convergence due à la compétition. Grâce à<br />

l’utilisation <strong>de</strong> la métrique MeanNTD (distance moyenne entre plus proches voisins), il a été<br />

possible <strong>de</strong> différencier la convergence due à l’« habitat filtering » <strong>de</strong> la convergence due à la<br />

compétition. Néanmoins, en raison <strong>de</strong> l’utilisation <strong>de</strong> métriques différentes pour détecter ces<br />

<strong>de</strong>ux sources <strong>de</strong> convergence fonctionnelle entre traits, il n’est pas possible <strong>de</strong> mesurer leur<br />

importance respective.<br />

Concernant les métriques mesurant les écarts <strong>de</strong>s valeurs <strong>de</strong> traits entre voisins, nos<br />

résultats indiquent que dans les parties non inondées, une faible similitu<strong>de</strong> entre espèces<br />

implique plutôt une différenciation <strong>de</strong>s valeurs <strong>de</strong> SLA et une plus forte similitu<strong>de</strong> <strong>de</strong>s valeurs<br />

<strong>de</strong> hauteur entre espèces coexistantes due à la compétition. A l’inverse dans les niveaux<br />

inondés, une faible similitu<strong>de</strong> implique plutôt une différence <strong>de</strong>s valeurs <strong>de</strong> hauteur et une<br />

plus forte similitu<strong>de</strong> <strong>de</strong>s valeurs <strong>de</strong> SLA entre espèces coexistantes due à la compétition. Ces<br />

résultats soutiennent une similitu<strong>de</strong> dans la dynamique <strong>de</strong>s ressources : pour coexister dans<br />

<strong>de</strong>s environnements compétitifs comme les milieux productifs, une différenciation <strong>de</strong> niche<br />

est certes nécessaire, mais elle ne doit cependant pas être trop importante pour être compétitif<br />

face à <strong>de</strong>s voisins également compétitifs pour persister (Scheffer & van Nes 2006).<br />

225


Ces résultats nous indiquent que les inondations représentent un filtre abiotique<br />

important sélectionnant les individus suivant leur capacité à tolérer les inondations, et que la<br />

compétition agit le long du gradient permettant la coexistence <strong>de</strong>s espèces.<br />

4. Le stress à l’échelle <strong>de</strong> la communauté module les effets du pâturage et <strong>de</strong><br />

la pluviosité dans les milieux méditerranéens<br />

L’étu<strong>de</strong> <strong>de</strong> l’effet <strong>de</strong> différents facteurs abiotiques (variabilité <strong>de</strong>s patrons <strong>de</strong><br />

pluviosité) et biotiques (pâturage) dans les milieux méditerranéens met en évi<strong>de</strong>nce qu’en<br />

fonction du niveau <strong>de</strong> productivité <strong>de</strong> la communauté – proxy du stress à l’échelle <strong>de</strong> la<br />

communauté (Grime 1977, 1979) – les réponses à la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité et au<br />

pâturage sont différentes (Chapitre 7).<br />

Les communautés végétales méditerranéennes sont <strong>de</strong>s communautés très riches en<br />

espèces. Les espèces annuelles sont attendues à répondre très fortement à la pluviosité,<br />

variable déterminante <strong>de</strong> leurs patrons <strong>de</strong> germination et <strong>de</strong> croissance (Pake & Venable<br />

1996). Dans la communauté la plus productive (les pelouses strictes), c’est-à-dire la moins<br />

contrainte, la composition en espèces varie fortement entre années. Ces changements en<br />

composition sont attribués à la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité observée sur la pério<strong>de</strong><br />

2001-2008 étudiée. Des espèces xériques s’expriment plutôt les années sèches : on observe<br />

qu’un remplacement d’espèces intervient probablement entre années humi<strong>de</strong>s et années plus<br />

sèches, expliquant la diversité en espèces <strong>de</strong> ces pelouses strictes inchangée au cours <strong>de</strong> la<br />

pério<strong>de</strong> d’étu<strong>de</strong>. L’effet <strong>de</strong> la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité est moins marqué dans les<br />

prés salés, c’est-à-dire la communauté la moins productive caractérisée par une salinité du sol<br />

forte: la dynamique <strong>de</strong> la composition est linéaire. Néanmoins, on remarque une forte<br />

augmentation <strong>de</strong> la proportion <strong>de</strong>s espèces pérennes lors <strong>de</strong>s années humi<strong>de</strong>s, et ce dans les<br />

<strong>de</strong>ux communautés.<br />

Un <strong>de</strong>sign expérimental in situ contrôlant la modalité <strong>de</strong> pâturage a été mise en place<br />

après 40 années <strong>de</strong> charge <strong>de</strong> pâturage extensive. En plus <strong>de</strong> la modalité témoin <strong>de</strong> pâturage<br />

correspondant à ces 40 années <strong>de</strong> pâturage, les <strong>de</strong>ux communautés ont été soumises à : (i)<br />

l’arrêt du pâturage et (ii) un changement dans la pério<strong>de</strong> où le bétail est présent sur le site<br />

(<strong>de</strong>ux jours au lieu <strong>de</strong> 6 mois) mais sans aucun changement <strong>de</strong> la charge annuelle, appelé le<br />

pâturage instantané et caractérisé par une forte charge instantanée <strong>de</strong> pâturage.<br />

226


Dans les pelouses strictes, un arrêt du pâturage conduit à la dominance d’une espèce<br />

pérenne, Brachypodium phoenicoï<strong>de</strong>s : une fermeture du milieu est observée par cette espèce<br />

qui serait compétitive, et confirme les résultats fréquemment rapportés dans la littérature<br />

(Peco et al. 1983; Noy-Meir et al. 1989; Noy-Meir 1995; Díaz et al. 2001; Peco et al. 2005).<br />

L’analyse fine <strong>de</strong>s probabilités <strong>de</strong> colonisation et <strong>de</strong> survie <strong>de</strong> cette espèce nous indique que la<br />

survie <strong>de</strong> cette espèce est affectée par la présence <strong>de</strong>s herbivores (Chapitre 5). Le pâturage<br />

limite les capacités <strong>de</strong> survie <strong>de</strong> cette espèce (Fig. 20a) alors qu’en l’absence <strong>de</strong> pâturage, ces<br />

capacités <strong>de</strong> survie expliquent son succès et l’augmentation <strong>de</strong> son abondance, conduisant à<br />

une perte <strong>de</strong> diversité en espèces dans les pelouses strictes (Fig. 20b). Cette étu<strong>de</strong><br />

démographique met en évi<strong>de</strong>nce l’importance <strong>de</strong> la survie dans le succès <strong>de</strong>s espèces pérennes<br />

dans les milieux méditerranéens et surtout la sensibilité <strong>de</strong> ce paramètre démographique à la<br />

gestion par le pâturage, résultat observé dans d’autres milieux (Ehrlen 1995 ; Oliva et al.<br />

2005). Quelle que soit la modalité <strong>de</strong> pâturage, la présence d’herbivores limite l’augmentation<br />

en abondance <strong>de</strong> cette espèce dominante. Néanmoins, le pâturage instantané n’affecte pas la<br />

structure <strong>de</strong>s pelouses, suggérant que le pool d’espèces a déjà été filtrée par les 40 années <strong>de</strong><br />

pâturage extensif (Ward et al. 1998 ; Metzger et al. 2005). Une augmentation <strong>de</strong> la<br />

contribution d’annuelles caractéristiques <strong>de</strong>s prés salés, suggère une modification <strong>de</strong>s<br />

conditions abiotiques <strong>de</strong> salinité du sol, augmentant la similitu<strong>de</strong> entre les pelouses strictes et<br />

les prés salés en termes <strong>de</strong> composition (Cingolani et al. 2003).<br />

Fig 20 : Dans les pelouses strictes : a) Perte <strong>de</strong> diversité en espèces en raison <strong>de</strong> l’arrêt du<br />

pâturage favorisant le succès écologique <strong>de</strong> Brachypodium phoenicoï<strong>de</strong>s b) Maintien <strong>de</strong> la<br />

diversité en espèces par le pâturage limitant le succès écologique <strong>de</strong> l’espèce.<br />

227


Dans les prés salés, un changement dans le régime <strong>de</strong> pâturage, c’est-à-dire<br />

l’application <strong>de</strong> la forte charge instantanée ou l’arrêt du pâturage, n’affecte pas la diversité en<br />

espèces. Au cours <strong>de</strong> la pério<strong>de</strong> étudiée 2001-2008, la dynamique <strong>de</strong> composition en espèces<br />

est similaire entre ces <strong>de</strong>ux traitements et le traitement témoin. Une espèce en particulier<br />

apparaît contrôler cette dynamique <strong>de</strong> composition : l’espèce dominante Halimione<br />

portulacoï<strong>de</strong>s. Il s’agit d’une espèce succulente possédant les caractéristiques physiologiques<br />

lui permettant <strong>de</strong> se développer sous <strong>de</strong>s conditions <strong>de</strong> salinité fortes. Cette capacité à tolérer<br />

le sel est notée comme permettant <strong>de</strong> répondre à d’autres <strong>de</strong> contraintes et perturbation<br />

(Milchunas et al. 1988). On peut ainsi supposer que la tolérance au sel permet à cette espèce<br />

<strong>de</strong> tolérer les différentes modalités <strong>de</strong> pâturage, en contrôlant la composition en espèces<br />

probablement en raison <strong>de</strong> fortes capacités compétitives.<br />

Le niveau <strong>de</strong> stress à l’échelle <strong>de</strong> la communauté limite les effets du pâturage mis en<br />

évi<strong>de</strong>nce dans les pelouses. Ce résultat peut s’expliquer par l’espèce dominante <strong>de</strong>s prés salés,<br />

qui elle ne serait pas contrainte par la salinité du sol. En effet, un niveau <strong>de</strong> contrainte à<br />

l’échelle <strong>de</strong> la communauté n’implique pas forcément que les espèces <strong>de</strong> cette communauté<br />

sont contraintes (Lortie et al. 2004). Ces résultats suggèrent que se contenter du niveau <strong>de</strong><br />

stress à l’échelle <strong>de</strong> la communauté ne permet pas, à lui seul, <strong>de</strong> comprendre précisément les<br />

mécanismes impliqués dans la structure et la dynamique <strong>de</strong>s communautés végétales.<br />

5. Conclusion & Perspectives<br />

La contrainte apparaît déterminante dans les patrons d’espèces et <strong>de</strong> structure <strong>de</strong>s<br />

communautés. Le niveau <strong>de</strong> contrainte module l’importance <strong>de</strong>s autres filtres écologiques,<br />

comme la compétition dans les prairies humi<strong>de</strong>s du Marais Poitevin, ou le pâturage dans les<br />

prés-salés méditerranéens. De plus, le choix du niveau d’organisation considéré pour<br />

caractériser la contrainte, c’est-à-dire à l’échelle <strong>de</strong> l’espèce ou à l’échelle <strong>de</strong> la communauté,<br />

est déterminant pour comprendre les effets <strong>de</strong>s différents filtres écologiques. En effet, une<br />

contrainte mesurée à l’échelle <strong>de</strong> la communauté végétale n’est pas forcément une contrainte<br />

pour les espèces qui la composent. Mais ces <strong>de</strong>ux niveaux d’organisation apparaissent<br />

complémentaires pour comprendre les assemblages d’espèces et la structure <strong>de</strong>s communautés<br />

végétales dans les <strong>de</strong>ux milieux étudiés.<br />

Dans les prairies humi<strong>de</strong>s du Marais Poitevin, les inondations représentent un filtre<br />

abiotique majeur qui sélectionne les espèces suivant leur capacité à tolérer ces conditions. A<br />

228


ce filtre s’ajoute la compétition entre plantes, est particulièrement importante pendant les<br />

pério<strong>de</strong>s où les prairies ne sont plus inondées. Dans les milieux méditerranéens, la contrainte<br />

à l’échelle <strong>de</strong> la communauté limite les effets <strong>de</strong>s autres filtres écologiques, que sont la<br />

variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité et du pâturage. La réponse <strong>de</strong>s communautés à ces<br />

différents filtres écologiques est due aux espèces dominantes qui contrôlent la dynamique <strong>de</strong>s<br />

communautés et qui ne sont pas contraintes par les conditions <strong>de</strong> leur habitat respectif.<br />

Afin <strong>de</strong> compléter ce travail <strong>de</strong> compréhension <strong>de</strong>s mécanismes structurant les <strong>de</strong>ux<br />

modèles d’étu<strong>de</strong>, différentes perspectives <strong>de</strong> recherche sont envisageables.<br />

5.1. La mesure <strong>de</strong> la compétition en conditions naturelles<br />

La mesure <strong>de</strong> la compétition en conditions naturelles est une approche prometteuse car<br />

l’effet compétitif <strong>de</strong>s espèces sur les voisins et l’importance <strong>de</strong> la compétition par rapport aux<br />

autres facteurs environnementaux pouvant également contrôler la dynamique <strong>de</strong>s populations,<br />

ont été quantifiés dans un cadre dynamique considérant différentes mesures du cycle <strong>de</strong> vie<br />

<strong>de</strong>s plantes. Cette analyse a cependant été réalisée sur une seule saison d’inondation. Or une<br />

précé<strong>de</strong>nte étu<strong>de</strong> (Violle et al. 2007) a mis en évi<strong>de</strong>nce la nécessité <strong>de</strong> tenir compte <strong>de</strong><br />

plusieurs années successives, caractérisées par une variabilité <strong>de</strong>s durées d’inondation, pour<br />

comprendre le succès <strong>de</strong>s espèces et leurs patrons d’abondance. Grâce la réplication dans le<br />

temps <strong>de</strong>s relevés point contact, l’effet <strong>de</strong> la variabilité <strong>de</strong>s conditions d’inondation sur la<br />

dynamique <strong>de</strong>s populations pourrait être prise en compte, en corrélant ces probabilités à la<br />

variable durée d’inondation par exemple (Torang et al. 2010). Les effets <strong>de</strong> la variation <strong>de</strong>s<br />

conditions d’inondation sur l’importance <strong>de</strong> la compétition pourront ainsi être approchés et on<br />

pourrait mieux comprendre la distribution <strong>de</strong>s espèces et la structure <strong>de</strong>s communautés.<br />

5.2. Vers le développement <strong>de</strong>s tests <strong>de</strong>s règles d’assemblage<br />

Avec le test <strong>de</strong>s règles d’assemblage, nous avons pu montrer que l’assemblage <strong>de</strong>s<br />

espèces dans les prairies humi<strong>de</strong>s du Marais Poitevin se faisait suivant <strong>de</strong>s processus<br />

déterministes. Or cette analyse ne nous permet pas <strong>de</strong> connaître l’importance relative <strong>de</strong><br />

chacun <strong>de</strong> ces processus, notamment celui <strong>de</strong> la compétition, en raison <strong>de</strong> l’utilisation <strong>de</strong><br />

métriques différentes pour tester ces différents processus, et <strong>de</strong> généraliser les résultats.<br />

Avec ce type d’analyses, différents patrons <strong>de</strong> distribution <strong>de</strong>s traits peuvent être<br />

observés pour un même facteur (Podani 2009). En effet, une divergence dans la distribution<br />

<strong>de</strong>s traits peut être attendue par l’effet <strong>de</strong> la compétition ainsi qu’une convergence dans cette<br />

229


distribution ; cela dépend <strong>de</strong>s conditions environnementales <strong>de</strong> l’habitat (milieu productif ou<br />

non). De la même manière, il peut-être attendu à la fois une convergence dans la distribution<br />

<strong>de</strong>s traits ainsi qu’une divergence en réponse au facteur abiotique ; cela dépend <strong>de</strong> la nature<br />

du facteur abiotique, c’est-à-dire s’il s’agit ou non d’une perturbation. Cela amène à <strong>de</strong>s<br />

difficultés d’interprétation <strong>de</strong>s résultats et surtout à une difficulté à généraliser ces résultats,<br />

apparaissant comme une limite à l’utilisation <strong>de</strong> cette approche fonctionnelle <strong>de</strong>s règles<br />

d’assemblage.<br />

Suivant la proposition <strong>de</strong> Navas & Violle (2009), cette généralisation pourrait être<br />

possible en travaillant le long <strong>de</strong> gradients d’importance <strong>de</strong> la compétition. Relier la diversité<br />

fonctionnelle à l’importance <strong>de</strong> la compétition permettrait <strong>de</strong> déterminer le poids <strong>de</strong> cette<br />

importance dans l’assemblage <strong>de</strong>s espèces. Calculer un indice à l’échelle <strong>de</strong> la communauté<br />

en agrégeant les valeurs d’importance <strong>de</strong> la compétition <strong>de</strong>s espèces composant la<br />

communauté (Greiner la Peyre et al. 2001) permettrait d’avoir une valeur d’importance <strong>de</strong> la<br />

compétition à l’échelle <strong>de</strong> la communauté. Il serait ainsi possible <strong>de</strong> relier les mesures<br />

d’importance <strong>de</strong> la compétition aux mesures <strong>de</strong> diversité fonctionnelle pour déterminer plus<br />

précisément son rôle dans l’assemblage <strong>de</strong>s espèces et ainsi avoir une méthodologie plus<br />

prédictive et généralisable.<br />

5.3. De la facilitation observée ?<br />

Certains résultats expérimentaux suggèrent <strong>de</strong> la facilitation entre espèces dans les<br />

parties inondées du gradient (Chapitre 1, valeurs <strong>de</strong> lnRR(interaction) positives pour<br />

Leontodon autumnalis et performances photosynthétiques améliorées en présence <strong>de</strong> voisins<br />

sous <strong>de</strong>s conditions d’inondation). Les étu<strong>de</strong>s récentes sur le sujet montrent que la facilitation<br />

n’intervient pas seulement dans les milieux fortement contraints mais serait dépendante du<br />

type <strong>de</strong> contraintes (cf introduction 2.1) et <strong>de</strong> la stratégie <strong>de</strong> l’espèce (Michalet et al. 2006 ;<br />

Maestre et al. 2009). Il serait intéressant <strong>de</strong> regar<strong>de</strong>r ces résultats <strong>de</strong> plus près pour déterminer<br />

si réellement <strong>de</strong>s interactions positives peuvent intervenir dans notre milieu et ainsi expliquer<br />

la coexistence <strong>de</strong> certaines espèces.<br />

Il convient tout <strong>de</strong> même <strong>de</strong> rester pru<strong>de</strong>nt quand la présence <strong>de</strong> facilitation dans notre<br />

milieu d’étu<strong>de</strong>. Le choix <strong>de</strong>s méthodologies utilisées pour étudier la compétition est en effet<br />

susceptible <strong>de</strong> modifier les observations (Goldberg & Barton, 1992).<br />

230


5.4. Une étu<strong>de</strong> plus précise <strong>de</strong> l’effet <strong>de</strong>s patrons <strong>de</strong> pluviosité dans les milieux<br />

méditerranéens<br />

Les milieux méditerranéens étudiés sont caractérisés par une hétérogénéité et une<br />

faible disponibilité <strong>de</strong> la ressource en eau (variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité). Avec les<br />

relevés floristiques réalisés sur la pério<strong>de</strong> 2001-2008, il est difficile d’étudier précisément<br />

l’effet <strong>de</strong> la variabilité <strong>de</strong>s patrons <strong>de</strong> pluviosité sur la structure <strong>de</strong>s communautés végétales,<br />

car nous n’avons tout simplement pas <strong>de</strong> point <strong>de</strong> comparaison où la disponibilité <strong>de</strong> la<br />

ressource en eau serait plus gran<strong>de</strong>. L’expérimentation, qui est actuellement toujours en cours<br />

(cf. Partie Modèles Biologiques, section 2.2), nous permettra <strong>de</strong> connaître plus précisément<br />

l’effet <strong>de</strong> l’hétérogénéité <strong>de</strong> la ressource en eau sur la structure <strong>de</strong>s communautés végétales.<br />

Elle nous permettra également d’étudier le succès <strong>de</strong>s espèces annuelles et pérennes, et leurs<br />

interactions compétitives, face à l’augmentation <strong>de</strong> la disponibilité <strong>de</strong> l’eau. Grâce à cette<br />

expérimentation et aux méthodologies <strong>de</strong> mesure <strong>de</strong> la compétition in natura, il serait possible<br />

d’étudier les interactions compétitives entre pérennes et annuelles, et ainsi déterminer si la<br />

compétition a un rôle structurant <strong>de</strong>s communautés végétales et si ce rôle structurant varie<br />

entre années.<br />

231


232


BIBLIOGRAPHIE<br />

Aarssen, L.W., & Keogh, T. (2002) Conundrums of competitive ability in plants: what to<br />

measure? Oikos, 96, 531-542<br />

Ackerly, D.D. & Cornwell, W.K. (2007) A trait-based approach to community assembly:<br />

partitioning of species trait values into within- and among-community components. Ecology<br />

Letters, 10, 135-145<br />

Amiaud, B., Bouzillé, J.B., Tourna<strong>de</strong>, F., Bonis, A. (1998) Patterns of soil salinities in old<br />

embanked marshlands in Western France. Wetlands, 18(3), 482-494.<br />

Austin, M. P. (1980) Searching for a mo<strong>de</strong>l for use in vegetation analysis. Vegetatio, 42, 11-<br />

21<br />

Austin, M.P. (2005) Vegetation and environment: discontinuities and continuities. In:<br />

Vegetation Ecology, Blackwell Publishing, USA, UK, Australia, ISBN 0-632-05761-0<br />

Austin, M.P. & Smith, T.M. (1989) A new mo<strong>de</strong>l for the continuum concept. Vegetatio, 83,<br />

35-47.<br />

Austin, M.P. & Gaywood, M.J. (1994) Current problems of environmental gradients and<br />

species response curves in relation to continuum theory. Journal of Vegetation Science, 5,<br />

473-482.<br />

Bartelheimer, M., Gowing, D. & Silvertown, J. (2010) Explaining hydrological niches: the<br />

<strong>de</strong>cisive role of below-ground competition in two closely related Senecio species. Journal of<br />

Ecology, 98, 126–136.<br />

Baruch, Z. (1994) Responses to drought and flooding in tropical forage grasses. Plant and<br />

Soil, 164, 87–96.<br />

Belyea, L.R. & Lancaster, J. (1999). Assembly rules within a contingent ecology. Oikos,<br />

86(3), 402-416.<br />

Benot, M.L., Mony, C., <strong>Merlin</strong>, A., Marion, B., Bouzillé, J.B. & Bonis, A. (2011) Clonal<br />

growth strategies along flooding and grazing gradients in Atlantic coastal meadows. Folia<br />

Geobotanica, 46, 219-235.<br />

Berendse, F., Elberse, W.T., & Geerts, R. (1992) Competition and nitrogen loss from plants in<br />

grassland ecosystems. Ecology, 73, 46-53.<br />

Blom, C. & Voesenek, L. (1996) Flooding: the survival strategies of plants. Trends in<br />

Ecology & Evolution, 11, 290–295.<br />

Bonis, A., Bouzillé, J.B., Amiaud, B. & Loucougary, G. (2005). Plant community patterns in<br />

old embanked grasslands and the survival of halophytic flora. Flora, 200, 74-87.<br />

Briones, O., Montana, C. & Ezcurra, E. (1998). Competition intensity as a function of<br />

resource availability in a semiarid ecosystem. Oecologia, 116, 365-372.<br />

Briske, D.D., Fuhlendorf, F.E. & Smeins, F.R. (2005) State-and-transition mo<strong>de</strong>ls, thresholds,<br />

and rangeland health: a synthesis of ecological concepts and perspectives. Rangeland Ecology<br />

& Management, 58(1), 1-10.<br />

Brooker, R. (2006) Plant-plant interactions and environmental change. New Phytologist, 171,<br />

271-284<br />

233


Brooker R. & Kikividze Z. (2008) Importance: an overlooked concept in plant interaction<br />

research. Journal of Ecology, 96, 703-708<br />

Brooker R., Kikvidze Z., Pugnaire F., Callaway R., Choler P., Lortie C. & Michalet R.<br />

(2005). The importance of importance. Oikos, 109, 63-70<br />

Bruno, J.F., Stachowicz, J.J., Bertness, M.D. (2003) Inclusion of facilitation into ecological<br />

theory. Trends in Ecology and Evolution, 18(3), 199-125.<br />

Cadotte, M.W. (2007). Concurrent niche and neutral processes in the competitioncolonization<br />

mo<strong>de</strong>l of species coexistence. Proceedings of the royal society B., 274, 2739-<br />

2744.<br />

Carlyle, C. N., Fraser, L.H. & Turkington, R. (2010) Using three pairs of competitive indices<br />

to test for changes in plant competition un<strong>de</strong>r different resource and disturbance levels.<br />

Journal of Vegetation Science, 21, 1–10.<br />

Carlyle, C.N. & Fraser, L.H. (2006) A test of three juvebile plant competitive response<br />

strategies. Journal of Vegetation Science, 17, 11-18.<br />

Casanova, M.T. & Brock, M.A. (2000) How do <strong>de</strong>pth, duration and frequency of flooding<br />

influence the establishment of wetland plant communities? Plant Ecology, 147, 237–250.<br />

Casper, B.B. & Jackson, R.B. (1997) Plant competition un<strong>de</strong>rground. Annual Review of<br />

Ecology and Systematics, 28, 545-570.<br />

Chapin, F., Zavaleta, E., Eviner, V., Naylor, R., Vitousek, P., Reynolds, H., Hooper, D.,<br />

Lavorel, S., Sala, O., Hobbie, S., Mack, M. & Diaz, S. (2000) Consequences of changing<br />

biodiversity. Nature, 405, 234-242<br />

Chase J.M. & Leibold M.A. (2003) Ecological niches: linking classical and contemporary<br />

approaches. University of Chicago Press, Chicago.<br />

Chave, J. (2004) Neutral theory and community ecology. Ecology Letters, 7, 241–253.<br />

Chesson, P. (2000) Mechanisms of maintenance of species diversity. Annual Review of<br />

Ecology and Systematics, 31, 343-366.<br />

Choler, P., Michalet, R. & Callaway, R.M. (2001) Facilitation and competition on gradients in<br />

alpine plant communities. Ecology, 82, 3295-3308<br />

Cingolani, A.M., Cabido, M., Gurvich, D.E., Renison, D. & Díaz, S. (2007) Filtering<br />

processes in the assembly of plant communities: Are species presence and adundance driven<br />

by the same traits? Journal of Vegetation Science, 18, 911-920<br />

Clements, F.E. (1916) Plant succession: an analysis of the <strong>de</strong>velopment of vegetation.<br />

Carnegie Institute Publication, Washington, DC.<br />

Clevering, O.A. (1998) Effects of litter accumulation and water table on morphology and<br />

productivity of Phragmites australis. Wetlands Ecology and Management, 5, 275–287.<br />

Collins, S.L., Glenn, S.M. & Roberts, D.W. (1993) The hierarchical continuum concept.<br />

Journal of Vegetation Science, 4(2), 149-156.<br />

Colmer, T.D. & Voesenek, L. (2009) Flooding tolerance: suites of plant traits in variable<br />

environments. Functional Plant Biology, 36, 665–681.<br />

Corcket, E., Liancourt, P., Callaway, R.M. & Michalet, R. (2003) The relative importance of<br />

competition for two dominant grass species as affected by environmental manipulations in the<br />

field. Ecoscience, 10, 186–194.<br />

234


Cornwell, W.K. & Ackerly, D.D. (2009) Community assembly and shifts in plant trait<br />

distributions across an environmental gradient in coastal California. Ecological Monographs,<br />

79(1), 109–126.<br />

Craine, J.M. (2007). Plant strategy theories: replies to Grime and Tilman. Journal of Ecology,<br />

95, 235-240.<br />

Craine, J.M. (2005) Reconciling plant strategy theories of Grime and Tilman. Journal of<br />

Ecology, 93, 1041-1052<br />

Curtis, J.T. (1959) The vegetation of Wisconsin: an ordination of plant communities.<br />

University of Wisconsin Press, Madison<br />

Damgaard, C. (2004) Dynamics in a discrete two-species competition mo<strong>de</strong>l: coexistence and<br />

over-compensation. Journal of Theoretical Biology, 227, 197–203.<br />

Damgaard, C., Riis-Nielsen, T. & Schmidt, I. K. (2009) Estimating plant competition<br />

coefficients and predicting community dynamics from non-<strong>de</strong>structive pin-point data: a case<br />

study with Calluna vulgaris and Deschampsia flexuosa. Plant Ecology, 201, 687–697.<br />

<strong>de</strong>Bello, F., Lavorel, S., Albert, C.H., Thuiller, W., Grigulis, K., Dolezal, J., Janecek, S. &<br />

Leps, J. (2010) Quantifying the relevance of intraspecific trait variability for functional<br />

diversity. Methods in Ecology & Evolution, 2(2), 163-174<br />

<strong>de</strong> Bello, F., Thuiler, W., Leps, J., Choler, P., Clément, J.C., Macek, P., Sebastia, M.T. &<br />

Lavorel, S. (2009) Partitioning of functional diversity reveals the scale and extent of trait<br />

convergence and divergence. Journal of Vegetation Science, 20, 475-486.<br />

Casanova, M.T. & Brock, M.A. (2000) How do <strong>de</strong>pth, duration and frequency of flooding<br />

influence the establishment of wetland plant communities? Plant Ecology, 147, 237–250.<br />

Cingolani, A.M., Cabido, M.R., Renison, D. & Solis Neffa, V. (2003) Combined effects of<br />

environment and grazing on vegetation structure in Argentine granite grasslands. Journal of<br />

Vegetation Science, 14, 223-232.<br />

Diaz S., Lavorel S., McIntyre S., Falczuk V., Casanoves F., Milchunas D.G., Skarpe C.,<br />

Rusch G., Sternberg M., Noy-Meir I., Landsberg J., Zhang W., Clark H. & Campbell B.D.<br />

(2007) Plant trait responses to grazing - a global synthesis. Global Change Ecology, 13, 313-<br />

341<br />

Diaz, S. & Cabido, M. (2001) Vive la différence: plant functional diversity matters to<br />

ecosystem processes. Trends in Ecology & Evolution, 16(11), 646-655.<br />

Díaz S., Cabido M. & Casanoves F. (1998) Plant functional traits and environmental filters at<br />

the regional scale. Journal of Vegetation Science, 9, 113-122<br />

Dolé<strong>de</strong>c, S., Chessel, D. & Gimaret-Carpentier, C. (2000). Niche separation in community<br />

analysis: a new method. Ecology, 81, 2914-2927.<br />

Dong, M. & <strong>de</strong> Kroon, H. (1994) Plasticity in morphology and biomass allocation in Cynodon<br />

dactylon, a grass species forming stolons and rhizomes. Oikos, 70, 99-106.<br />

Ehrlen, J. (1995) Demography of the perennial herb Lathyrus vernus. II. Herbivory and<br />

population dynamics. Journal of Ecology, 83, 297-308.<br />

Elton C.S. (1927) Animal Ecology, London.<br />

Engel, E.C. & Weltzin, J.F. (2008) Can community composition be predicted from pairwise<br />

species interactions? Plant Ecology, 195, 77–85.<br />

235


Enright, N.J., Franco, M. & Silvertown, J. (1995) Comparing plant life histories using<br />

elasticity analysis: the importance of life span and the number of life-cycle stages. Oecologia,<br />

104, 79-84.<br />

Fayolle, A., Violle, C., Navas, M.L. (2009) Differential impacts of plant interactions on<br />

herbaceous species recruitment disentangling factors controlling emergence, survival and<br />

growth seedlings. Oecologia, 159(4), 817-825.<br />

Fraser, L.H. & Miletti, T.E. (2008) Effect of minor water <strong>de</strong>pth treatments on competitive<br />

effect and response of eight wetland plants. Plant Ecology, 195, 33–43.<br />

Fraser, L.H. & Karnesis, J.P. (2005) A comparative assessment of seedling survival and<br />

biomass accumulation for fourteen wetland plant species grown un<strong>de</strong>r minor water-<strong>de</strong>pth<br />

differences. Wetlands, 25(3), 520–530.<br />

Fraser, L.H. & Keddy, P.A. (2005) Can competitive ability predict structure in experimental<br />

plant communities? Journal of Vegetation Science. 16, 571-578<br />

Frekleton, R.P., Watkinson, A.R. & Rees, M. (2009) Measuring the importance of<br />

competition in plant communities. Journal of Ecology, 97, 379–384.<br />

Fréville, H. & Silvertown, J. (2005) Analysis of interspecific competition in perennial plants<br />

using Life Table Response Experiments. Plant Ecology, 176, 69-78.<br />

Fukami, T., Bezemer, T.M., Mortimer, S.R. & van <strong>de</strong>r Putten, W.H. (2005). Species<br />

divergence and trait convergence in experimental plant community assembly. Ecology letters,<br />

8, 1283-1290.<br />

Gaucherand, S., Liancourt, P. & Lavorel, S. (2006) Importance and intensity of competition<br />

along a fertility gradient and across species. Journal of Vegetation Science, 17, 455-464.<br />

Gau<strong>de</strong>t, C.L. & Keddy, P.A. (1995) Competitive performance and species distribution in<br />

shortline plant communities: a comparative approach. Ecology, 76, 280–291.<br />

Gause, G.F. (1934) The struggle for existence. Haffner, New York.<br />

Gewin, V. (2006) Beyond Neutrality—Ecology Finds Its Niche. PLoS Biology, 4(8), 1306-<br />

1310.<br />

Goldberg, D.E. (1996) Competitive ability: <strong>de</strong>finitions, contingency and correlated traits.<br />

Philosophical Transactions: Biological Sciences, 351(1345), 1377-1385.<br />

Goldberg, D.E. (1994). Influence of competition at the community level: an experimental<br />

version of the null mo<strong>de</strong>ls approach. Ecology, 75, 1503-1506.<br />

Goldberg, D.E. 1990. Components of resource competition in plant communities. In: Grace,<br />

J.B. & Tilman, D. (eds.) Perspectives on plant competition. pp. 27–49. Aca<strong>de</strong>mic Press, San<br />

Diego.<br />

Goldberg, D.E., Rajaniemi, T., Gurevitch, J. & Stewart-Oaten, A. (1999) Empirical<br />

Approaches to Quantifying Interaction Intensity: Competition and Facilitation along<br />

Productivity Gradients. Ecology, 80(4), 1118-1131.<br />

Goldberg, D.E. & Novoplanski, A. (1997) On the relative importance of competition in<br />

unproductive environments. Journal of Ecology, 85(4), 409-418.<br />

Goldberg D.E. & Barton A.M. (1992) Patterns and consequences of interspecific competition<br />

in natural communities: a review of field experiments with plants. American Naturalist,<br />

139(4), 771-801.<br />

236


Goldberg, D.E. & Landa, K. (1991) Competitive effect and response: hierarchies and<br />

correlated traits in the early stages of competition. Journal of Ecology, 79(4), 1013-1030.<br />

Gotelli, N.J. & McCabe, D.J. (2002) Species co-occurence: a meta-analysis of J.M.<br />

Diamond’S assembly rules mo<strong>de</strong>l. Ecology, 83(8), 2091–2096.<br />

Gowing, D.J.G., Youngs, E.G., Gilbert, J.C. & Spoor, G. (1998). Predicting the effect of<br />

change in water regime on plant communities. In Hydrology in a Changing Environment, vol.<br />

I, Wheater H, Kirby C (eds). John Wiley: Chichester; 473–483.<br />

Grace, J.B. (1995) On the measurement of plant competition intensity. Ecology, 76(1), 305-<br />

308.<br />

Greiner La Peyre, M.K., Grace, J.B., Hahn, E. & Men<strong>de</strong>lssohn, I.A. (2001) The importance of<br />

competition in regulating plant species abundance along a salinity gradient. Ecology, 82, 62–<br />

69.<br />

Grime J.P. (2007) Plant strategy theories: a comment on Craine (2005). Journal of Ecology,<br />

95, 227-230<br />

Grime J.P. (2006) Trait convergence and trait divergence in herbaceous plant communities:<br />

Mechanisms and consequences. Journal of Vegetation Science, 17, 255-260.<br />

Grime J.P. (1979) Plant Strategies and Vegetation Processes. John Wiley and Sons,<br />

Chichester, UK.<br />

Grime, J.P. (1977) Evi<strong>de</strong>nce for the existence of three primary strategies in plants and its<br />

relevance to ecological and evolutionary theory. American naturalist, 111, 1169–1194.<br />

Grimoldi, A.A., Inausti, P., Roitman, G.G. & Soriano, A. (1999) Responses to flooding<br />

intensity in Leontodon taraxacoi<strong>de</strong>s. New Phytologist, 141, 119-128.<br />

Grubb, P. (1977) Maintenance of species-richness in plant communities–– importance of the<br />

regeneration niche. Biol Rev Camb Philos Soc, 52,107–145<br />

Gustafsson, C. & Ehrlen, J. (2003) Effects of intraspecific and interspecific <strong>de</strong>nsity on the<br />

<strong>de</strong>mography of a perennial herb, Sanicula europaea. Oikos, 100, 317–324.<br />

Harper, J. L. (1977) Population biology of plants. Aca<strong>de</strong>mic Press, London.<br />

Hedges L.V., Gurevitch J. & Curtis P.S. (1999) The meta-analysis of response ratios in<br />

experimental ecology. Ecology, 80, 1150-1156.<br />

Heil, G.W. (2004) Mo<strong>de</strong>ling of plant community assembly in relation to <strong>de</strong>terministic and<br />

stochastic processes. In: Assembly rules and Restoration Ecology. Bring the gap between<br />

theory and practice. Island Press Washington, Covelo, London. ISBN 1-55963-374-3.<br />

Howard T.G. & Goldberg D.E. (2001) Competitive response hierarchies for germination,<br />

growth, and survival and their influence on abundance. Ecology, 82, 979-990.<br />

Hubbell S.P. (2001) The unified neutral theory of biodiversity and biogeography. Princeton<br />

University Press, Princeton.<br />

Hutchinson G.E. (1957) Concluding remarks. In: Cold Spring Harbor Symposia on<br />

Quantitative Biology, pp. 415-427<br />

Insausti, P., Grimoldi, A.A., Chaneton, E.J. & Vasellati, V. (2001) Flooding induces a suite of<br />

adaptive plastic responses in the grass Paspalum dilatatum. New Phytologist, 152, 291–299.<br />

Jung, V., Violle, C., Mondy, C., Hoffmann, L. & Muller, S. (2010) Intraspecific variability<br />

and trait-based community assembly. Journal of Ecology, 98, 1134–1140.<br />

237


Jung, V., Hoffmann, L. & Muller, S. (2009) Ecophysiological responses of nine floodplain<br />

meadow species to changing hydrological conditions. Plant Ecology, 201, 589-598.<br />

Keddy, P.A. (1992) Assembly and response rules: two goals for predictive community<br />

ecology. Journal of Vegetation Science, 3, 157–164.<br />

Keddy, P.A. (1990) Competitive hierarchies and centrifugal organization in plant<br />

communities. pp. 265-89: In J. Grace and D. Tilman (eds.) Perspectives on Plant<br />

Competition, Aca<strong>de</strong>mic Press, New York.<br />

Keddy P., Nielsen, K., Weiher, E. & Lawson, R. (2002) Relative competitive performance of<br />

63 species of terrestrial herbaceous plants. Journal of Vegetation Science, 13, 5-16.<br />

Keddy P., Fraser L. & Wisheu I. (1998) A comparative approach to examine competitive<br />

response of 48 wetland plant species. Journal of Vegetation Science, 9, 777-786<br />

Keddy, P.A., Twolan-Strutt, L. & Wisheu, I.C. (1994) Competitive Effect and Response<br />

Rankings in 20 Wetland Plants: Are They Consistent Across Three Environments? Journal of<br />

Ecology, 82(3), 635-643.<br />

Keddy P. & Shipley B. (1989) Competitive hierarchies in herbaceous plant-communities.<br />

Oikos, 54, 234-241.<br />

Kent, A. & Coker, P. (1992) Vegetation <strong>de</strong>scription and analysis - A practical approach. John<br />

Wiley & Sons, New York, USA.<br />

Kikvidze, Z., Suzuki, M. & Brooker, R. (2011) Importance versus intensity of ecological<br />

effects: why context matters. Trends in Ecology and Evolution. In press.<br />

Kikvidze, Z. & Armas, C. (2010) Plant interaction indices based on experimental plant<br />

performance data. Positive Plant Interactions and Community Dynamics (ed. F.I. Pugnaire),<br />

pp. 17–37. CRC Press, London, UK.<br />

Kikvidze, Z. & Brooker, R. (2010) Towards a more exact <strong>de</strong>finition of the importance of<br />

competition – a reply to Freckleton et al. (2009). Journal of Ecology, 98, 719–724.<br />

Kotowski, V., Werner, T., van Diggelen, R. & Joseph, W.M. (2006) Competition as a factor<br />

structuring species zonation in riparian fens – a transplantation experiment. Applied<br />

Vegetation Science, 9, 231-240.<br />

Kraft, N.J.B., Valencia, R. & Ackerly, D.D. (2008) Functional traits and niche-based tree<br />

community assembly in an Amazonian forest. Science, 322, 580-582.<br />

Lamb, E.G. & Cahill, F.J. (2008) When Competition Does Not Matter: Grassland Diversity<br />

and Community Composition. The American Naturalist, 171(6): 777-787.<br />

Lavorel S. & Garnier E. (2002) Predicting changes in community composition and ecosystem<br />

functioning from plant traits: revisiting the Holy Grail. Functional Ecology, 16, 545- 556<br />

Lavorel S., McIntyre S., Landsberg J. & Forbes T.D.A. (1997) Plant functional<br />

classifications: from general groups to specific groups based on response to disturbance.<br />

Trends in Ecology & Evolution, 12, 474-478<br />

Lenssen, J.P.M., Ten Dolle, G.E. & Blom, C. (1998) The effect of flooding on the recruitment<br />

of reed marsh and tall forb plant species. Plant Ecology, 139, 13–23.<br />

Lenssen, J.P., van <strong>de</strong> Steeg, H.M. & <strong>de</strong> Kroon, H. (2004) Does disturbance favour weak<br />

competitors? Mechanisms of changing plant abundance after flooding. Journal of Vegetation<br />

Science, 15, 305–314.<br />

238


Liancourt P., Callaway R.M. & Michalet R. (2005) Stress tolerance and competitive-response<br />

ability <strong>de</strong>termine the outcome of biotic interactions. Ecology, 86, 1611-1618.<br />

Lortie C.J., Brooker R.W., Kikvidze Z., & Callaway R.M. (2004) The value of stress and<br />

limitation in an imperfect world: A reply to Korner. Journal of Vegetation Science, 15, 577-<br />

580.<br />

MacArthur R.H. & Levins R. (1967) Limiting similarity convergence and divergence of<br />

coexisting species. American Naturalist, 101, 377-387<br />

Maestre, F.T., Callaway, R.M., Valladares, F. & Lortie, C.J. (2009) Refining the stressgradient<br />

hypothesis for competition and facilitation in plant communities. Journal of Ecology,<br />

97, 199-205.<br />

Marion, B. (2010) Impact du pâturage sur la structure <strong>de</strong> la végétation : interactions biotiques,<br />

traits et conséquences fonctionnelles. Thèse <strong>de</strong> l’<strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1.<br />

Mason, N.W.H., MacGillivray, K., Steel, J.B. & Bastow, W.J. (2003) An in<strong>de</strong>x of functional<br />

diversity. Journal of Vegetation Science, 14, 571-578.<br />

McGill B.J., Enquist B.J., Weiher E. & Westoby M. (2006) Rebuilding community ecology<br />

from functional traits. Trends in Ecology & Evolution, 21, 178-185.<br />

Metzger, K.L., Coughenour, M.B., Reich, R.M. & Boone, R.B. (2005) Effects of seasonal<br />

grazing on plant species diversity and vegetation structure in a semi-arid ecosystem. Journal<br />

of Arid Environments, 61(1), 147-160.<br />

Michalet R. (2006) Is facilitation in arid environments the result of direct or complex<br />

interactions? New Phytologist, 169, 3-6.<br />

Milchunas, D.G., Sala, O.E. & Lauenroth, W.K. (1988) A generalized mo<strong>de</strong>l of the effects of<br />

grazing by large herbivores on grasslands community structure. The American Naturalist,<br />

132(1), 87-106.<br />

Mokany, K. & Roxburgh, S. H. (2010) The importance of spatial scale for trait-abundance<br />

relations. Oikos, 119, 1504–1514.<br />

Mommer, L., Lenssen, J.P., Huber, H., Visser, E.J. & D.E. Kroon, H. (2006)<br />

Ecophysiological <strong>de</strong>terminants of plant performance un<strong>de</strong>r flooding: a comparative study of<br />

seven plant families. Journal of Ecology, 94, 1117–1129.<br />

Mommer, L., Pe<strong>de</strong>rsen, O. & Visser, E.J.W. (2004) Acclimation of a terrestrial plant to<br />

submergence facilitates gas exchange un<strong>de</strong>r water. Plant, Cell and Environment, 27, 1281–<br />

1287.<br />

Mouillot, D., Mason, N.W.H. & Wilson, J.B. (2007) Is the abundance of species <strong>de</strong>termined<br />

by their functional traits? A new method with a test using plant communities. Oecologia, 152,<br />

729–737.<br />

Mouillot, D., Mason, W.H.N., Dumay, O. & Wilson, J.B. (2005) Functional regularity: a<br />

neglected aspect of functional diversity. Oecologia, 142, 353-359.<br />

Navas, M.L., Violle, C. (2009) Plant traits related to competition : how do they shape the<br />

functional diversity of communities ? Community Ecology, 10(1), 131-137.<br />

Novoplanski, A. & Goldberg, D.E. (2001) Effects of water pulsing on individual performance<br />

and competitive hierarchies in plants. Journal of Vegetation Science, 12, 199-208.<br />

Noy-Meir, I. (1995) Interactive effects of fire and grazing on structure and diversity of<br />

Mediterranean grasslands. Journal of Vegetation Science, 6(5):701-710.<br />

239


Noy-Meir, I., Gutman, M. & Kaplan, Y. (1989) Responses of Mediterranean Grasslands<br />

Plants to Grazing and Protection. Journal of Ecology, 77(1): 290-310.<br />

Noy-Meir, I. & Whittaker, R.H. (1977) Continuous multivariate methods in community<br />

analysis: some problems and <strong>de</strong>velopments. Plant Ecology, 33(2-3), 79-98.<br />

Oliva, G., Collantes, M. & Humano, G. (2005) Demography of grazed tussock grass<br />

populations in Patagonia. Rangeland Ecology and Management, 58(5), 466-473.<br />

Pake, C.E. & Venable, L. (1996) Seed banks in <strong>de</strong>sert annuals: implications for persistence<br />

and coexistence in variable environments. Ecology, 77(5), 1427-1435.<br />

Pakeman, R.J., Lennon, J.J. & Brooke, R.W. (2011) Trait assembly in plant assemblages and<br />

its modulation by productivity and disturbance. Oecologia, in press<br />

Peco, B., <strong>de</strong> Pablos, I., Traba, J. & Levassor, C. (2005) The effect of grazing abandonment on<br />

species composition and functional traits: the case of <strong>de</strong>hesa grasslands. Basic and Applied<br />

Ecology, 6(2):175-183.<br />

Peco, B., Levassor, C. & Pineda, F.D. (1983) Diversité et structure spatiale <strong>de</strong> pâturages<br />

méditérranéens en cours <strong>de</strong> succession. Ecologia Mediterranea, 9, 223-234.<br />

Pennings, S.C., Grant, M.B. & Bertness, M.D. (2005) Plant zonation in low-latitu<strong>de</strong> salt<br />

marshes: disentangling the roles of flooding, salinity and competition. Journal of Ecology, 93,<br />

159–167.<br />

Pennings, S.C. & Callaway, R.M. (1992) Salt marsh plant zonation: the relative importance of<br />

competition and physical factors. Ecology, 73, 681–690.<br />

Perkins, T.A., Holmes, W.R. & Weltzin, J.F. (2007) Multi-species interactions in competitive<br />

hierarchies: New methods and empirical test. Journal of Vegetation Science, 18, 685–692.<br />

Petchey, O.L. & Gaston, K.J. (2002) Functional diversity (FD), species richness and<br />

community composition. Ecology Letters, 5, 402-411.<br />

Podani J (2009) Convex hulls, habitat filtering and functional diversity: mathematical<br />

elegance versus ecological interpretability. Community Ecology, 10, 244–250.<br />

Potts, D.L., Huxma, T.E., Enquis, B.J., Weltzin, J.F. & Williams, D.G. (2005) Resilience and<br />

resistance of ecosystem functional response to a precipitation pulse in a semi-arid grassland.<br />

Journal of Ecology, 94(1), 23-30<br />

Rea<strong>de</strong>r, R. J., Wilson, S. D., Belcher, J. W., Wisheu, I., Keddy, P.A., Tilman, D., Morris,<br />

E.C., Grace, J.B., McGraw, J.B., Olff, H., Turkington, R., Klein, E., Leung, Y., Shipley, B.,<br />

van Hulst, R., Johansson, M.E., Nilsson, C., Gurevitch, J., Grigulis, K. & Beisner, B.E. (1994)<br />

Plant competition in relation to neighbor biomass: an intercontinental study with Poa<br />

pratensis. Ecology, 75, 1753-1760.<br />

Ricotta, C. (2005) A note on functional diversity measures. Basic and Applied Ecology, 6,<br />

479-486.<br />

Rossignol, N. (2006) Hétérogénéité <strong>de</strong> la végétation et du pâturage : conséquences<br />

fonctionnelles en prairie naturelle. Thèse <strong>de</strong> l’<strong>Université</strong> <strong>de</strong> <strong>Rennes</strong> 1.<br />

Salguero-Gomez, R. & <strong>de</strong> Kroon, H. (2010) Matrix projection mo<strong>de</strong>ls meet variation in the<br />

real world. Journal of Ecology, 98(2), 250-254.<br />

Santamaria, L. (2002) Why are most aquatic plants wi<strong>de</strong>ly distributed ? Dispersal, clonal<br />

growth and small-scale heterogeneity in a stressful environment. Acta Oecologica, 23, 137-<br />

154.<br />

240


Scarnati, L., Attore, F., De Sanctis, M., Farcomeni, A., Francesconi, F., Mancini, M. &<br />

Bruno, F. (2009) A multiple approach for the evaluation of the spatial distribution and<br />

dynamics of a forest with Taxus baccata and Ilex aquifolium. Biodiversity and Conservation,<br />

18(12), 3099-3113.<br />

Schamp B.S., Chau J. & Aarssen L.W. (2008) Dispersion of traits related to competitive<br />

ability in an old-field plant community. Journal of Ecology, 96, 204-212.<br />

Scheffer, M. & van Ness, E.H. (2006) Self-organized similarity, the evolutionary emergence<br />

of groups of similar species. Proceedings of the National Aca<strong>de</strong>my of Science, 103, 6230–<br />

6235.<br />

Seifan, M., Seifan, T. & Tielbörger, K. (2010) Facilitating an importance in<strong>de</strong>x. Journal of<br />

Ecology, 98, 356–361.<br />

Shipley, B., Keddy, P.A. & Lefkovitch, L.P. (1991) Mechanism producing plant zonation<br />

along a water <strong>de</strong>pth gradient: a comparison with the exposure gradient. Canadian Journal of<br />

Botany, 69, 1420-1424.<br />

Silvertown, J. (2004) Plant coexistence and the niche. Trends in Ecology & Evolution, 19,<br />

605–611.<br />

Silvertown, J., Dodd, M.E., Gowing, D.J. & Mountford, J.O. (1999) Hydrologically <strong>de</strong>fined<br />

niches reveal a basis for species richness in plant communities. Nature, 400, 61–63.<br />

Stubbs W.J. & Wilson J.B. (2004) Evi<strong>de</strong>nce for limiting similarity in a sand dune community.<br />

Journal of Ecology, 92, 557-567.<br />

Soukupova, L. (1994) Allocation plasticity and modular structure in clonal graminoids in<br />

response to waterlogging. Folia Geobotanica Phytotaxa, 29, 227-236.<br />

Swilck, D.W. & Ackerly, D.D (2005) Limiting similarity and functional diversity along<br />

environmental gradients. Ecology Letters, 8, 272–281.<br />

Taylor D.R., Aarssen L.W., & Loehle C. (1990) On the relationship between r/K selection and<br />

environmental carrying capacity: a new habitat templet for plant life-history strategies. Oikos,<br />

58, 239-250.<br />

Ter Braak, C.J.F. (1986) Canonical correspon<strong>de</strong>nce analysis: a new eigenvector technique for<br />

multivariate direct gradient analysis. Ecology, 67, 1167-1179.<br />

Thuiller W. (2004) Patterns and uncertainties of species' range shifts un<strong>de</strong>r climate change.<br />

Global Change Biology, 10, 2020-2027<br />

Tilman, D. (2004) Niche tra<strong>de</strong>offs, neutrality, and community structure: A stochastic theory<br />

of resource competition, invasion, and community assembly. PNAS, 101(30), 10854-10861.<br />

Tilman D. (2001) Functional diversity. In: Encyclopedia of Biodiversity (ed. Levin S), pp.<br />

109-120. Aca<strong>de</strong>mic Press, San Diego<br />

Tilman, D. (1988). Plant strategies and the dynamics and structure of plant communities.<br />

Princeton University Press, Princeton, New Jersey, USA.<br />

Tilman, G.D. (1984) Plant dominance along an experimental nutrient gradient. Ecology,<br />

65(5), 1445-1453.<br />

Tilman D. (1982) Resource competition and community structure. Princeton University Press.<br />

Torang, P., Ehrlen, J. & Agren, J. (2010) Linking environmental and <strong>de</strong>mographic data to<br />

predict future population viability of a perennial herb. Oecologia, 163(1), 99-109.<br />

241


Twolan-Strutt, L. & Keddy, P.A. (1996) Above- and belowground competition intensity in<br />

two contrasting wetland plant communities. Ecology, 77(1), 259-270.<br />

Van Eck, W. H. J. M., van <strong>de</strong> Steeg, H.M., Blomand, C. W. P. M. & <strong>de</strong> Kroon, H. (2004) Is<br />

tolerance to summer flooding correlated with distribution patterns in river floodplains? A<br />

comparative study of 20 terrestrial grassland species. Oikos, 107, 393-405.<br />

Vellend, M. (2010) Conceptual synthesis in community ecology. The Quarterly Review Of<br />

Biology, 85(2), 183-206.<br />

Violle, C., Bonis, A., Plantegenest, M., Cu<strong>de</strong>nnec, C., Damgaard, C., Marion, B., Le Coeur,<br />

D. & Bouzillé, J.B. (2011) Plant functional traits capture species richness variations along a<br />

flooding gradient. Oikos, 120(3), 389-398.<br />

Violle, C. & Jiang, L. (2009) Towards a trait-based quantification of species niche. Journal of<br />

Plant Ecology, 2, 87.<br />

Violle, C., Cu<strong>de</strong>nnec, C., Plantgenest, M., Damgaard, C., Le Cœur, D., Bouzillé, J.B. &<br />

Bonis, A. (2007) Indirect assessment of flooding duration as a driving factor of plant diversity<br />

in wet grasslands. Proceedings of Symposium S7 held during the Seventh IAHS Scientific<br />

Assembly at Foz do Iguacu, Brazil,. IAHS Publication 303, 2006, pp. 334<br />

Violle C., Navas M.-L., Vile D., Kazakou E., Fortunel C., Hummel I. & Garnier E. (2007).<br />

Let the concept of trait be functional! Oikos, 116, 882-892<br />

Violle C., Richarte J. & Navas M.-L. (2006) Effects of litter and standing biomass on growth<br />

and reproduction of two annual species in a Mediterranean old-field. Journal of Ecology, 94,<br />

196-205<br />

Voesenek, L.A.C.J., Colmer, T.D., Pierik, R., Millenaar, F.F. & Peeters, A.J.M. (2006) How<br />

plants cope with complete submergence. New Phytologist, 170(2), 213-226.<br />

Wang, P., Stieglitz, T., Zhou, D.W. & Cahill, J.F. (2010) Are competitive effect and response<br />

two si<strong>de</strong>s of the same coin, or fundamentally different? Functional Ecology, 24, 196–207.<br />

Ward, D., Ngairorue, B.T., Kathena, J., Samuels, R., Ofran, Y. (1998) Land <strong>de</strong>gradation is not<br />

a necessary outcome of communal pastoralism in arid Namibia. Journal of Arid Environments<br />

40(4): 357–371.<br />

Weigelt A. & Jolliffe P. (2003) Indices of plant competition. Journal of Ecology, 91, 707-720<br />

Weiher E. & Keddy P.A. (1995) Assembly rules, null mo<strong>de</strong>ls, and trait dispersion – new<br />

questions front old patterns. Oikos, 74, 159-164<br />

Weiher, E. & Keddy, P.A. (1995) The assembly of experimental wetland plant communities.<br />

Oikos, 73, 323–335.<br />

Weiher, E., Clarke, J.D.P. & Keddy, A. (1998) Community assembly rules, morphological<br />

dispersion and the coexistence of plant species. Oikos, 81, 309-322.<br />

Weiher E. & Keddy P.A. (1999) Ecological assembly rules - Perspectives, advances, retreats.<br />

Cambridge University Press, Cambridge.<br />

Wel<strong>de</strong>n, C.W. & Slauson, W.L. (1986) The intensity of competition versus its importance: an<br />

overlooked distinction and some implications. Quarterly Review of Biology, 61, 23–44.<br />

Westoby, M., Walker, B. & Noy-Meir, I. (1989) Opportunistic management for rangelands<br />

not at equilibrium. Journal of Range Management, 42(4), 266-274.<br />

242


Whittaker RH (1951) A criticism of the plant association and climatic climax concepts.<br />

Northwest Sci., 25, 17-31<br />

Wilson, S.D. (2007) Competition, resources, and vegetation during 10 years in native<br />

grassland. Ecology, 88, 2951–2958.<br />

Wilson, J.B. (1999) Guilds, Functional Types and Ecological Groups. Oikos, 86(3), 507-522.<br />

Wisheu, I.C. & Keddy P.A. (1992) Competition and centrifugal organization of plant<br />

communities: theory and tests. Journal of Vegetation Science, 3, 147-156.<br />

Wright, I.J., Reich, P.B., Westoby, M., Ackerly, D.D., Baruch, Z, Bongers, F., Caven<strong>de</strong>r-<br />

Bares, J., Chapin, T., Cornelissen, J.H.C., Diemer, M., Flexas, J., Garnier, E., Groom, P.K.,<br />

Gulias, J., Hikosaka, K., Lamont, B.B., Lee, T., Lee, W., Lusk, C., Midgley, J.J., Navas, M-<br />

L., Niinemets, U., Oleksyn, J., Osada, N., Poorter, H., Poot, P., Prior, L., Pyankov, V.I.,<br />

Roumet, C., Thomas, S.C., Tjoelker, M.G., Veneklass, E.J. & Villar, R. (2004) The<br />

worldwi<strong>de</strong> leaf economics spectrum. Nature, 428, 821–827.<br />

243


244


ANNEXE 1<br />

Liste <strong>de</strong>s abréviations <strong>de</strong>s espèces présentes dans les prairies humi<strong>de</strong>s du Marais Poitevin<br />

AGRSTO Agrostis stolonifera<br />

ALOBUB Alopecurus bulbosus<br />

ALOGEN Alopecuris geniculatus<br />

BALRAN Bal<strong>de</strong>llia ranunculoï<strong>de</strong>s<br />

BELPER Bellis perennes<br />

BROCOM Bromus commutatus<br />

CALLIT Callitriche<br />

CARDIV Carex divisa<br />

CARPRA Cardamine pratense<br />

CERGLO Cerastium glomeratum<br />

CIRARV Cirsium arvense<br />

CIRVUL Cirsium vulgare<br />

CYNCRI Cynosurus cristatus<br />

ELEPAL Eleocharis palustris<br />

ELYREP Elymus repens<br />

EPITET Eîlobium tetragonum<br />

FESARU Festuca arundinacea<br />

FESRUB Festuca rubra<br />

GALDEB Galium <strong>de</strong>bile<br />

GAUFRA Gaudinia fragilis<br />

GERDIS Geranium dissectum<br />

GLYFLU Glyceria fluitans<br />

HORMAR Hor<strong>de</strong>um marinum<br />

HORSEC Hor<strong>de</strong>um secalinum<br />

JUNART Juncus articulatus<br />

JUNGER Juncus gerardii<br />

LEOAUT Leontodon autumnalis<br />

LEOTAR Leontodon taraxacoï<strong>de</strong>s<br />

LOLPER Lolium perenne<br />

LOTTEN Lotus tenui<br />

MENPUL Mentha pulegium<br />

MOUSSE Mousse<br />

MYOLAX Myosotis laxa<br />

OENFIS Oenanthe fistulosa<br />

PARSTR Parapholis strigosa<br />

PLACOR Plantago coronopus<br />

PLAMAJ Plantage major<br />

POAANN Poa annua<br />

POATRI Poa trivialis<br />

POTANS Potentilla anserina<br />

PUCMAR Pucinnellia maritima<br />

RANAQU Ranunuculus aquaticus<br />

RANOPH Ranunculus ophioglossus<br />

RANREP Ranunculus repens<br />

RANSAR Ranunculus sardou<br />

RANSP Ranunculus sp.<br />

RANTRI Ranunculus<br />

RUMCON Rumex conglomerata<br />

SONASP Sonchus asper<br />

SPEMAR Spergularia maritima<br />

TAROFF Taraxacum officinale<br />

TRIFRA Trifolium fragiferum<br />

TRIMIC Trifolium michaelinum<br />

TRIORN Trifolium ornithopoï<strong>de</strong>s<br />

TRIRES Trifolium resupinatum<br />

TRISQU Trifolium squamosum<br />

VULBRO Vulpia bromoï<strong>de</strong>s<br />

245


ANNEXE 2<br />

Liste <strong>de</strong>s abréviations <strong>de</strong>s espèces présentes dans pelouses xéro-halophiles <strong>de</strong> Camargue<br />

Aetbul Aetheorhiza bulbosa<br />

Allvin Allium vineale<br />

Anaarv Anagallis arvensis<br />

Arelep Arenaria leptoclados<br />

Astste Asterolinum stellatum<br />

Avebar Avena barbata<br />

Belann Bellis annua<br />

Belper Bellis perennis<br />

Blasp. Blackstonia sp.<br />

Bradis Brachypodium distachyon<br />

Brapho Brachypodium phoenicoi<strong>de</strong>s<br />

Brohor Bromus hor<strong>de</strong>aceus<br />

Bromad Bromus madritensis<br />

Bupsem Bupleurum semicompositum<br />

Carsp Carduus sp.<br />

Carten Cardus tenuiflorus<br />

Catrig Catapodium rigidum<br />

Cenmel Centaurea melitensis<br />

Censp. Centaurium sp.<br />

Cerglo Cerastium glomeratum<br />

Cerpum Cerastium pumilum<br />

Cersemi Cerastium semi<strong>de</strong>candrum<br />

Cersic Cerastium siculum<br />

Cersp. Cerastium sp.<br />

Cirvul Cirsium vulgare<br />

Comsp. Composée sp.<br />

Crecap Crepis capillaris<br />

Cresan Crepis sancta<br />

Cresp. Crepis sp.<br />

Creves Crepis vesicaria<br />

Cyndac Cynodon dactylon<br />

Cynech Cynosurus echinatus<br />

Dachis Dactylis hispanica<br />

Elysp. Elytrigia sp.<br />

Erocic Erodium cicutarium<br />

Eupexi Euphorbia exigua<br />

Euppep Euhorbia peplus<br />

Eupsul Euphorbia sulcata<br />

Evapyg Evax pygmaea<br />

Filsp. Filago sp.<br />

Filvul Filago vulgaris<br />

Galmur Galium murale<br />

Galpar Galium parisiense<br />

Gerdis Geranium dissectum<br />

Germol Geranium molle<br />

Haicyl Hainardia cylindrica<br />

Halpor Halimione portulacoi<strong>de</strong>s<br />

Hedcre Hedypnois cretica<br />

Hipbif Hippocrepis biflora<br />

Hormar Hor<strong>de</strong>um marinum<br />

Hympro Hymenolobus procumbens<br />

Hypgla Hypochaeris glabra<br />

Hypsp. Hypochaeris sp.<br />

Leotub Leontodon tuberosus<br />

Limnar Limonium narbonense<br />

Limvir Limonium virgatum<br />

Linbie Linum bienne<br />

Linstr Linum strictum<br />

Lolper Lolium perenne<br />

Lolrig Lolium rigidum<br />

Lolsp. Lolium sp.<br />

Medlup Medicago lupulina<br />

Medmin Medicago minima<br />

Medpol Medicago polymorpha<br />

Medrig Medicago rigidula<br />

Medsp. Medicago sp.<br />

Medtru Medicago truncatula<br />

Minhyb Minuartia hybrida<br />

Minten Minuartia tenuifolia<br />

Myoarv Myosotis arvensis<br />

Myodis Myosotis discolor<br />

Myosp. Myosotis sp.<br />

Narmar Nardurus maritima<br />

Ophsph Ophrys sphego<strong>de</strong>s<br />

Parfil Parapholis filiformis<br />

Parinc Parapholis incurva<br />

Parlat Parentucellia latifolia<br />

Phiang Phillyrea angustifolia<br />

Placor Plantago coronopus<br />

Plalag Plantago lagopus<br />

246


Poaann Poa annua<br />

Poabul Poa bulbosa<br />

Polmar Polypogon maritimus<br />

Poltet Polycarpon tetraphyllum<br />

Psiari Psilurus aristatus<br />

Psiinc Psilurus aristatus<br />

Romram Romulea ramiflora<br />

Romsp. Romulea sp.<br />

Roscri Rostraria cristata<br />

Sagsp. Sagina sp.<br />

Salver Salvia verbenaca<br />

Scolac Scorzonera laciniata<br />

Scomur Scorpiurus muricatus<br />

Senvul Senecio vulgaris<br />

Shearv Sherardia arvensis<br />

Sidrom Si<strong>de</strong>ritis romana<br />

Sonasp Sonchus asper<br />

Sonole Sonchus oleraceus<br />

Sonsp. Sonchus sp.<br />

Tarery Taraxacum erythrospermum<br />

Tarobo Taraxacum obovatum<br />

Taroff Taraxacum officinale<br />

Tornod Torilis nodosa<br />

Triang Trifolium angustifolium<br />

Tricam Trifolium campestre<br />

Trilap Trifolium lappaceum<br />

Trimon Trigonella monspeliaca<br />

Trinig Trifolium nigrescens<br />

Trires Trifolium resupinatum<br />

Trisca Trifolium scabrum<br />

Trisp. Trifolium sp.<br />

Trisqu Trifolium squamosum<br />

Triste Trifolium stellatum<br />

Trisuf Trifolium suffocatum<br />

Tritom Trifolium tomentosum<br />

Valmic Valerianella microcarpa<br />

Valmur Valantia muralis<br />

Valmuri Valerianella muricata<br />

Verarv Veronica arvensis<br />

Vulsp. Vulpia sp.<br />

247


Protocole <strong>de</strong> l’étu<strong>de</strong> <strong>de</strong> la compétition en conditions contrôlées<br />

ANNEXE 3<br />

Une expérimentation en conditions contrôlées a été mise en place au printemps 2008<br />

pour étudier la variation <strong>de</strong> l’intensité <strong>de</strong> la compétition le long d’un gradient expérimental<br />

d’inondation variant en hauteur et en durée. Ci-<strong>de</strong>ssous la <strong>de</strong>scription <strong>de</strong> cette<br />

expérimentation, qui a duré <strong>de</strong>ux saisons <strong>de</strong> croissance.<br />

1) Les traitements expérimentaux :<br />

Les traitements expérimentaux consistaient en un couplage du facteur hauteur d’eau<br />

avec le facteur durée d’inondations, c’est-à-dire avec une durée d’inondation augmentant avec<br />

la hauteur, <strong>de</strong>ux variables corrélées (Casanova & Brock 2000). Ainsi différentes espèces ont<br />

été soumises à 4 traitements : 10 cm-courte durée ; 20cm-courte durée ; 20cm-longue durée et<br />

40cm-longue durée. La différence entre les traitements courts et les traitements longs étaient<br />

<strong>de</strong> un mois.<br />

Les espèces soumises à ces traitements étaient <strong>de</strong>s espèces représentatives <strong>de</strong>s<br />

différentes communautés végétales : Lolium perenne, Juncus gerardii, Leontodon autumnalis,<br />

Juncus articulatus et Mentha pulegium. Quatre transplants <strong>de</strong> chaque espèce ont été<br />

transplantés en monocultures et en mixture face à une espèce : Agrostis stolonifera, sous <strong>de</strong>ux<br />

<strong>de</strong>nsités <strong>de</strong> compétiteurs (1 faible, 1 forte).<br />

L’expérimentation a été conduite dans le jardin expérimental <strong>de</strong> l’université <strong>de</strong><br />

<strong>Rennes</strong> : les réplicas ont été assignés aléatoirement dans <strong>de</strong>s 4 bacs pour chaque traitement<br />

d’inondation.<br />

A la fin <strong>de</strong> l’expérimentation, les biomasses aériennes et racinaires ont été mesurées,<br />

après avoir été nettoyées et séchées à 70°C pendant 72h.<br />

2) La quantification <strong>de</strong> la contrainte :<br />

La fluorescence permet d’apprécier les capacités photosynthétiques <strong>de</strong>s plantes et<br />

représente un proxy <strong>de</strong> l’intensité <strong>de</strong> la contrainte subie par les espèces car l’appareil<br />

photosynthétique peut être affecté par le stress (Saltmarsh et al. 2006). Les mesures <strong>de</strong><br />

ren<strong>de</strong>ment photosynthétique ont été effectuées à l’ai<strong>de</strong> d’un diving-PAM (Pulse Amplitu<strong>de</strong><br />

Modulation). Ce fluorimètre permet <strong>de</strong> calculer la part du ren<strong>de</strong>ment associé aux processus<br />

photochimiques (e.g. photosynthèse principalement). Les plantes sont préalablement laissées<br />

248


30 min à l’obscurité, <strong>de</strong> sorte à initier les mesures sur <strong>de</strong>s échantillons en « dormance<br />

photosynthétique » (pas <strong>de</strong> transport d’électron). Puis, trois mesures vont être effectuées par<br />

plante cible (sur 3 tiges ou feuilles différentes) ainsi que trois mesures par feuille.<br />

L’équipement photosynthétique <strong>de</strong> l’échantillon est alors excité par 2 types <strong>de</strong> sources<br />

lumineuses :<br />

- un faible éclairement : MB <strong>de</strong> 0,15 µmole.photons.m -² .s -1 : cette première source<br />

induit une émission <strong>de</strong> fluorescence sans mise en œuvre <strong>de</strong> la photosynthèse et permet <strong>de</strong><br />

déterminer le paramètre F0 correspondant à la fluorescence minimum <strong>de</strong> l’échantillon.<br />

- un flash <strong>de</strong> lumière saturante (rouge vif, 655 nm) court et intense : SatPULSE <strong>de</strong><br />

8000 à 10.000 µmole.photons.m-2.s-1, <strong>de</strong> 0,1 à 1s : ce flash aboutit à la fermeture <strong>de</strong>s centres<br />

<strong>de</strong> réactions du PSII ce qui provoque une augmentation substantielle <strong>de</strong> la fluorescence émise<br />

et <strong>de</strong>s pertes thermiques, ce qui permet <strong>de</strong> déterminer le paramètre Fm correspondant à la<br />

fluorescence maximum équivalente à l’activité photosynthétique.<br />

Le ren<strong>de</strong>ment maximal <strong>de</strong> fluorescence du PSII est défini par :<br />

Yield = Fv/Fm= (Fm-F0)/Fm<br />

Ce ren<strong>de</strong>ment mesure la fraction maximale <strong>de</strong> photons absorbés qui peut être utilisée pour la<br />

mise en place <strong>de</strong>s processus d’échanges chimiques et <strong>de</strong> transport d’électrons pour <strong>de</strong>s<br />

échantillons préalablement adaptés au noir.<br />

3) Mesure <strong>de</strong>s traits fonctionnels:<br />

Afin <strong>de</strong> connaître la réponse <strong>de</strong>s espèces à l’augmentation <strong>de</strong>s conditions d’inondation,<br />

<strong>de</strong>ux traits foliaires ont été mesurés en fin d’expérimentation : le Specific Leaf Area (SLA) et<br />

la longueur <strong>de</strong> la tige, suivant les protocoles standardisés (Weiher et al. 1999; Cornelissen et<br />

al. 2001).<br />

Afin <strong>de</strong> déterminer si l’allocation <strong>de</strong>s ressources peut varier suivant les conditions<br />

d’inondations, pouvant sur plus long terme d’affecter la survie et la capacité compétitive <strong>de</strong>s<br />

espèces, le ratio racine/tige a été mesuré. En effet, une altération du métabolisme racinaire<br />

peut être observée en raison <strong>de</strong> l’anoxie <strong>de</strong>s sols, provoquant une réallocation <strong>de</strong>s ressources<br />

<strong>de</strong>s organes souterrains vers les aériens.<br />

249


ANNEXE 4<br />

Clonal growth strategies along flooding and grazing gradients in Atlantic coastal<br />

meadows.<br />

Marie-Lise Benot, Cendrine Mony, <strong>Amandine</strong> <strong>Merlin</strong>, Benoit Marion, Jan-Bernard Bouzillé<br />

& Anne Bonis<br />

Folia Geobotanica (2011) 46: 219-235<br />

250


251


252


253


254


255


256


257


258


259


260


261


262


263


264


265


266


267

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!