You are on page 1of 394

Iván 

Gómez
Pirjo Huovinen   Editors

Antarctic
Seaweeds
Diversity, Adaptation and Ecosystem
Services
Antarctic Seaweeds
Iván Gómez  •  Pirjo Huovinen
Editors

Antarctic Seaweeds
Diversity, Adaptation and Ecosystem Services
Editors
Iván Gómez Pirjo Huovinen
Instituto de Ciencias Marinas y Instituto de Ciencias Marinas y
Limnológicas, Facultad de Ciencias Limnológicas, Facultad de Ciencias
Universidad Austral de Chile Universidad Austral de Chile
Valdivia, Chile Valdivia, Chile

Research Center Dynamics of High Latitude Research Center Dynamics of High Latitude
Marine Ecosystems, (IDEAL) Marine Ecosystems, (IDEAL)
Valdivia, Chile Valdivia, Chile

ISBN 978-3-030-39447-9    ISBN 978-3-030-39448-6 (eBook)


https://doi.org/10.1007/978-3-030-39448-6

© Springer Nature Switzerland AG 2020


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

In 2002, Christian Wiencke and Margaret Clayton published the book Antarctic
Seaweeds in the series Synopsis of the Antarctic Benthos edited by J.W. Wägele. To
our knowledge, this volume represents the most recent and comprehensive guide
dedicated exclusively to Antarctic macroalgae, which has made it an obligate refer-
ence for further studies. Nearly two decades later, we believe that it is timely and
urgently needed to bring to light an update on the present state of the art of these
important organisms in a rapidly changing world. Therefore, we invited various
Antarctic researchers to contribute chapters covering recent advances in a variety of
related topics. The list of contributors reflects well the currently active role of South
American research groups in this field. Originally, the focus of the book was on the
contributions from the South America; however, the invitation was expanded to
some distinguished colleagues from other regions in order to improve the coverage
of the thematics. It is clearly not possible to include all the topics, but we believe
that a representative view of the current state of knowledge on the most relevant
aspects is given, providing useful information for both undergraduate and postgrad-
uate students as well as for scientific community. The book is organized in 5 parts
with a total of 18 chapters. Part 1 gives a brief overview of the individual chapters
and outlines the major gaps and challenges as well as the new directions in the study
of Antarctic seaweeds. The following parts summarize the recent advances in diver-
sity and biogeography (Part 2); physiology, productivity, and environmental
responses (Part 3); biological interactions and ecosystem processes (Part 4); and
chemical ecology of Antarctic seaweeds (Part 5). Many of the chapters discuss the
topics in the context of environmental threats, especially climate change that is
already affecting these ecosystems. Thus, unavoidably, there is some overlapping of
these themes in different chapters, however, from distinct points of view and in
other context.
We are grateful to all the colleagues who kindly accepted the invitation to con-
tribute a chapter. We would also like to warmly thank our colleagues Chuck Amsler,
Kai Bischof, Bernardo Broitman, Gabriela L.  Campana, Marie-Laure Guillemin,
Patrick Neal, Ellie Poulin, Martin Thiel, Nelson Valdivia, Christian Wiencke and
Katharina Zacher for dedicating their time and expertise for peer review and

v
vi Preface

improving the chapters of this book at their manuscript stages. We greatly acknowl-
edge Prof. Christian Wiencke for contributing the foreword as one of the leading
experts in polar seaweeds. Finally, we would like to thank the Universidad Austral
de Chile, the Comisión Nacional de Investigación Científica y Tecnológica
(CONICYT), and the Instituto Antártico Chileno (INACH) for permanent support
of our research activities in the Antarctic and to the publisher for giving us the
opportunity to make this volume. This publication is within the frame of the scien-
tific program of the Research Center Dynamics of High Latitude Marine Ecosystems
(IDEAL).

Valdivia, Chile Iván Gómez


 Pirjo Huovinen
Foreword

Seaweeds represent a group of photoautotrophic organisms of vital importance for


the function of coastal ecosystems. They provide diverse habitats and breeding
areas for uncountable numbers of organisms including crustaceans and fishes and
represent an important food source not only for herbivores but also for detritivores
feeding on degraded seaweed biomass. About 10% of the global oceanic production
is based on seaweeds, a value similarly high as that of tropical rain forests. We knew
for long that seaweed communities are well developed in the temperate regions of
the world, but recent research shows that this is also true for the submarine seaweed
forests of Antarctica: single species can attain biomasses of over 10 kg wet weight
per square meter.
The exploration of Antarctic seaweeds began in 1817 with the expeditions of
Gaudichaud, Bory, Montagne, Hooker, and Harvey. In a second phase of intensive
research during the first two decades of the twentieth century, Gain, Skottsberg, and
Kylin made important taxonomic studies on Antarctic seaweeds. In the 1960s, scuba
diving investigations started and so, by the early 1980s, basic knowledge was avail-
able on taxonomy, geographic distribution, and depth zonation of Antarctic sea-
weeds. During the expeditions of Clayton and Wiencke in the 1980s and 1990s,
numerous species of Antarctic seaweeds were isolated and taken in culture, allow-
ing the description of algal life histories and the performance of physiological
experiments in temperature-controlled rooms in the home laboratories. An impor-
tant side product was a monograph on Antarctic seaweeds containing the first iden-
tification key.
These studies were a booster for scientists especially from South America, but
also from Europe, Australia, and North America, to work further on the investiga-
tion of these interesting biotas growing in this remote, harsh, and unique environ-
ment. Purely descriptive surveys came to an end, and scientists applied the latest
available methods to study distribution and biodiversity, metabolic adaptations to
the extreme Antarctic environment characterized by low temperatures and long
periods of darkness, the importance of the seaweeds as primary producers for
Antarctic near-shore ecosystems, and the effects of global climate changes, in par-
ticular the increase of UV-B radiation due to stratospheric ozone depletion and the

vii
viii Foreword

increase of the water temperatures in the Antarctic Peninsula region and their influ-
ence on zonation patterns and geographic distribution. Besides traditional studies,
scientists used new approaches to study photosynthesis and carbon balance, genetic
diversity, transcriptomic responses, and trophic interactions by experimental ecol-
ogy and ecological network analysis, just to mention some.
To my knowledge, this is the first multi-authored book exclusively focused on
Antarctic seaweeds and their role in coastal ecosystems in Antarctica with respect
to their reaction to a changing environment from the metabolic, cellular, and organ-
ismic level to the level of communities. I am proud to say that I have supported
many studies and interacted with almost all authors of the book. The book repre-
sents the present state of the art in this research area and as such will serve as an
important baseline for future research.

Bremerhaven, Germany Christian Wiencke


Contents

Part I Introduction
1 Antarctic Seaweeds: Biogeography, Adaptation,
and Ecosystem Services ��������������������������������������������������������������������������    3
Iván Gómez and Pirjo Huovinen
1.1 Introduction: The Historical Context������������������������������������������������    4
1.2 Antarctic Seaweeds in the Wake of Climate Change������������������������    7
1.3 The Book������������������������������������������������������������������������������������������    8
1.4 Gaps, Emerging Challenges, and Future Directions ������������������������   15
References��������������������������������������������������������������������������������������������������   17

Part II Diversity and Biogeography


2 Diversity of Antarctic Seaweeds��������������������������������������������������������������   23
Mariana C. Oliveira, Franciane Pellizzari, Amanda S. Medeiros,
and Nair S. Yokoya
2.1 The Antarctic Environment��������������������������������������������������������������   24
2.2 Seaweeds in Antarctica: Definition and Importance ������������������������   25
2.3 Seaweed Taxonomic Studies in Antarctica: Toward a New Species
Compilation��������������������������������������������������������������������������������������   27
2.4 Molecular Taxonomy for the Study of Antarctic Seaweed Diversity  35
2.5 Seaweed Distribution in Antarctica��������������������������������������������������   37
2.6 Concluding Remarks: Gaps and Prospects for the Future����������������   38
References��������������������������������������������������������������������������������������������������   39
3 Biogeographic Processes Influencing Antarctic
and sub-Antarctic Seaweeds ������������������������������������������������������������������   43
Ceridwen I. Fraser, Adele Morrison, and Pamela Olmedo Rojas
3.1 Antarctica’s Place in the World: An Isolated Continent?������������������   44
3.2 Physical Oceanographic Processes Influencing Movement
of Seaweeds into or out of the Antarctic ������������������������������������������   46

ix
x Contents

3.3 Hitch-Hiking to the Antarctic: Passengers on Seaweed Rafts����������   50


3.4 Concluding Remarks������������������������������������������������������������������������   54
References��������������������������������������������������������������������������������������������������   54
4 Detached Seaweeds as Important Dispersal Agents Across
the Southern Ocean����������������������������������������������������������������������������������   59
Erasmo C. Macaya, Fadia Tala, Iván A. Hinojosa,
and Eva Rothäusler
4.1 Introduction��������������������������������������������������������������������������������������   60
4.2 Detached Seaweeds in Antarctica ����������������������������������������������������   62
4.3 Abiotic Factors Influencing Floating Seaweeds��������������������������������   69
4.4 Biotic Factors Affecting Floating Seaweeds ������������������������������������   71
4.5 Physiology of Floating and Drifting Seaweeds:
Traspassing Thermal Barriers ����������������������������������������������������������   72
References��������������������������������������������������������������������������������������������������   75
5 Biogeography of Antarctic Seaweeds Facing Climate Changes����������   83
Franciane Pellizzari, Luiz Henrique Rosa, and Nair S. Yokoya
5.1 The Abiotic Setting of the Southern Ocean��������������������������������������   84
5.2 Biogeographic Patterns ��������������������������������������������������������������������   85
5.3 Seaweed Assemblages: Are Antarctic Seaweed
Diversity and Richness Changing? ��������������������������������������������������   87
5.4 The Physiological Bases of Macroalgal Shifts ��������������������������������   89
5.5 Deception Island: A Case Study of Opportunistic,
Alien, Cryptic and Cryptogenic Species������������������������������������������   90
5.6 Reevaluating Eco-Regions, Isolation,
and Endemism in the Southern Ocean����������������������������������������������   93
5.7 Concluding Remarks: Prospects for the
Future Marine Flora of the Southern Ocean ������������������������������������   97
References��������������������������������������������������������������������������������������������������   98
6 Comparative Phylogeography of Antarctic Seaweeds:
Genetic Consequences of Historical Climatic Variations ��������������������  103
Marie-Laure Guillemin, Claudio González-Wevar, Leyla Cárdenas,
Hélène Dubrasquet, Ignacio Garrido, Alejandro Montecinos, Paula
Ocaranza-Barrera, and Kamilla Flores Robles
6.1 Historical Isolation of Antarctic Marine Macroalgae ����������������������  104
6.2 Antarctic Marine Macroalgae: Surviving Quaternary
Glacial Cycles in Situ������������������������������������������������������������������������  106
6.3 Persistence in Multiple Isolated Glacial Refugia Versus
a Single Antarctic Refugium������������������������������������������������������������  108
6.4 Antarctic Macroalgae Genetic Diversity:
COI and TufA Sequences Data Sets ������������������������������������������������  109
6.5 Brown, Red and Green Macroalgae: Sharing a Common Pattern
of Glacial Impact and Postglacial Populations Recovery? ��������������  111
6.6 Concluding Remarks������������������������������������������������������������������������  119
References��������������������������������������������������������������������������������������������������  121
Contents xi

Part III Physiology, Productivity and Environmental Reponses


7 Underwater Light Environment of Antarctic Seaweeds����������������������  131
Pirjo Huovinen and Iván Gómez
7.1 Introduction��������������������������������������������������������������������������������������  131
7.2 Optics of Antarctic Coastal Waters ��������������������������������������������������  132
7.3 Adaptations of Antarctic Seaweeds
to Extreme Light Conditions������������������������������������������������������������  137
7.4 Consequences for Light Field Under Current
and Future Threats����������������������������������������������������������������������������  140
7.5 Concluding Remarks������������������������������������������������������������������������  144
References��������������������������������������������������������������������������������������������������  145
8 Production and Biomass of Seaweeds in Newly Ice-Free Areas:
Implications for Coastal Processes in a Changing Antarctic
Environment ��������������������������������������������������������������������������������������������  155
María L. Quartino, Leonardo A. Saravia, Gabriela L. Campana,
Dolores Deregibus, Carolina V. Matula, Alicia L. Boraso,
and Fernando R. Momo
8.1 Introduction: Seaweeds in Coastal Marine Ecosystems ������������������  156
8.2 Macroalgae and Carbon Fluxes in Antarctic Coastal Areas��������������  157
8.3 Macroalgal Biomass Studies in Antarctica ��������������������������������������  159
8.4 The Ecosystem of Potter Cove: An Outstanding Case Study ����������  160
8.5 A Dynamic Growth Model for Antarctic Macroalgae
Under a Fast-­Changing Environment������������������������������������������������  162
8.6 Seaweed Production in Present and Future Warming Scenarios������  166
8.7 Future Prospects��������������������������������������������������������������������������������  167
References��������������������������������������������������������������������������������������������������  168
9 Carbon Balance Under a Changing Light Environment����������������������  173
Dolores Deregibus, Katharina Zacher, Inka Bartsch, Gabriela
L. Campana, Fernando R. Momo, Christian Wiencke, Iván Gómez,
and María L. Quartino
9.1 Introduction��������������������������������������������������������������������������������������  174
9.2 Carbon Balance: A Case Study in Potter Cove ��������������������������������  178
9.3 New Scenarios and Their Implications
for Algal Photosynthesis ������������������������������������������������������������������  183
9.4 Concluding Remarks and Future Prospects��������������������������������������  185
References��������������������������������������������������������������������������������������������������  186
10 Life History Strategies, Photosynthesis, and Stress Tolerance
in Propagules of Antarctic Seaweeds������������������������������������������������������  193
Nelso Navarro, Pirjo Huovinen, and Iván Gómez
10.1 Seasonal Strategies and Life History Cycles����������������������������������  194
10.2 Photosynthetic Light Requirements of Early Stages����������������������  199
10.3 Effects of Environmental Factors
on the Biology of Propagules����������������������������������������������������������  205
xii Contents

10.4 Concluding Remarks: Biology of Propagules under Climate


Change��������������������������������������������������������������������������������������������  211
References��������������������������������������������������������������������������������������������������  211
11 Form and Function in Antarctic Seaweeds:
Photobiological Adaptations, Zonation Patterns,
and Ecosystem Feedbacks ����������������������������������������������������������������������  217
Iván Gómez and Pirjo Huovinen
11.1 Brief Overview of Form and Function in Seaweeds ����������������������  218
11.2 Functional Groups of Seaweeds in the Antarctic����������������������������  219
11.3 The Vertical Zonation of Antarctic Seaweeds:
A Paradigm of Spatial Distribution
of Different Morpho-functional Traits��������������������������������������������  222
11.4 Light Use Characteristics as a Major Factor
Delineating Physiological Thallus Anatomy of Seaweeds ������������  225
11.5 Form and Function in the Context of Life Strategies
and Stress Tolerance������������������������������������������������������������������������  227
11.6 Functional Traits of Seaweeds and Properties of Benthic
Communities ����������������������������������������������������������������������������������  229
11.7 Concluding Remarks����������������������������������������������������������������������  232
References��������������������������������������������������������������������������������������������������  232

Part IV Biological Interactions and Ecosystem Processes


12 Successional Processes in Antarctic Benthic Algae ������������������������������  241
Gabriela L. Campana, Katharina Zacher, Fernando R. Momo,
Dolores Deregibus, Juan Ignacio Debandi, Gustavo A. Ferreyra,
Martha E. Ferrario, Christian Wiencke, and María L. Quartino
12.1 Introduction������������������������������������������������������������������������������������  242
12.2 Structural Patterns and Changes in Algal Community
Composition during Succession������������������������������������������������������  244
12.3 Ecological Factors Influencing Antarctic Algal Succession ����������  248
12.4 Experimental Approaches to Study In Situ Succession
of Antarctic Benthic Algae��������������������������������������������������������������  253
12.5 Concluding Remarks and Perspectives ������������������������������������������  254
References��������������������������������������������������������������������������������������������������  256
13 Seaweed-Herbivore Interactions: Grazing as Biotic
Filtering in Intertidal Antarctic Ecosystems ����������������������������������������  265
Nelson Valdivia
13.1 Biological Invasions and Their Impact on the Ecology
of Antarctic Coastal Systems����������������������������������������������������������  266
13.2 Recent Introductions of Exotic Macroalgae in Antarctica��������������  267
13.3 Can Grazers Control Alien Macroalgae in Antarctica?������������������  268
13.4 Ulva intestinalis as a Case Study in a Simple,
Two-­Species Assembly Model��������������������������������������������������������  270
Contents xiii

13.5 Concluding Remarks����������������������������������������������������������������������  274


References��������������������������������������������������������������������������������������������������  274
14 Diversity and Functioning of Antarctic Seaweed Microbiomes����������  279
Juan Diego Gaitan-Espitia and Matthias Schmid
14.1 Introduction: Environment and Antarctic Seaweed Host-­
Microbiome������������������������������������������������������������������������������������  280
14.2 Functional Interactions of Antarctic Seaweeds
and Their Associated Microbiota����������������������������������������������������  281
14.3 Deciphering the Structure and Diversity of Seaweed
Microbiomes ����������������������������������������������������������������������������������  282
14.4 Variation of Bacterial Community Diversity in Antarctic
Seaweeds����������������������������������������������������������������������������������������  284
14.5 Conclusions and Future Perspectives����������������������������������������������  287
References��������������������������������������������������������������������������������������������������  287
15 Seaweeds in the Antarctic Marine Coastal Food Web��������������������������  293
Fernando R. Momo, Georgina Cordone, Tomás I. Marina, Vanesa
Salinas, Gabriela L. Campana, Mariano A. Valli, Santiago R. Doyle,
and Leonardo A. Saravia
15.1 Introduction������������������������������������������������������������������������������������  294
15.2 Food Webs and Seaweeds ��������������������������������������������������������������  295
15.3 Network Dynamics and Robustness ����������������������������������������������  299
15.4 Non-Trophic Interactions����������������������������������������������������������������  302
15.5 Final Remarks ��������������������������������������������������������������������������������  304
References��������������������������������������������������������������������������������������������������  304
16 Trophic Networks and Ecosystem Functioning������������������������������������  309
Marco Ortiz, Brenda B. Hermosillo-Núñez, and Ferenc Jordán
16.1 Introduction������������������������������������������������������������������������������������  310
16.2 Macroscopic Ecosystem-Network Properties ��������������������������������  318
16.3 Keystone Species Complex (KSC) ������������������������������������������������  322
16.4 Contribution of Keystone Species Complex
to Macroscopic Network Properties������������������������������������������������  327
16.5 Constrains and Perspectives������������������������������������������������������������  328
Appendix 16.A����������������������������������������������������������������������������������������   330
References��������������������������������������������������������������������������������������������������  330

Part V Chemical Ecology


17 Chemical Mediation of Antarctic Macroalga-Grazer
Interactions����������������������������������������������������������������������������������������������  339
Charles D. Amsler, James B. McClintock, and Bill J. Baker
17.1 Introduction������������������������������������������������������������������������������������  340
17.2 Feeding Bioassay Methodology������������������������������������������������������  340
17.3 Antarctic Macroalgal Resistance to Herbivory ������������������������������  341
xiv Contents

17.4 Macroalga-Invertebrate Interactions on the Western Antarctic


Peninsula ����������������������������������������������������������������������������������������  352
17.5 Overview����������������������������������������������������������������������������������������  355
References��������������������������������������������������������������������������������������������������  356
18 Brown Algal Phlorotannins: An Overview
of Their Functional Roles������������������������������������������������������������������������  365
Iván Gómez and Pirjo Huovinen
18.1 Introduction������������������������������������������������������������������������������������  366
18.2 Synthesis and Cellular Localization
of Phlorotannins: Dual Functions as Secondary
Metabolites and Structural Compounds������������������������������������������  369
18.3 Phlorotannins as UV-Screening Substances�����������������������������������  370
18.4 Phlorotannins as Active Antioxidant Compounds��������������������������  372
18.5 Phlorotannins in Antarctic Seaweeds����������������������������������������������  374
18.6 Concluding Remarks����������������������������������������������������������������������  381
References��������������������������������������������������������������������������������������������������  382

Index������������������������������������������������������������������������������������������������������������������  389
Part I
Introduction
Chapter 1
Antarctic Seaweeds: Biogeography,
Adaptation, and Ecosystem Services

Iván Gómez and Pirjo Huovinen

Abstract  Seaweeds (macroalgae) represent the most striking benthic organisms in


the Antarctic near-shore ecosystems. Their abundance, relevant roles as primary
producers, and foundation organisms were recognized since the first Antarctic
explorations. Furthermore, especially since the 1960s, improvements in the sub-
aquatic survey techniques and laboratory facilities expanded considerably our
knowledge on ecology, reproduction, and environmental adaptation of seaweeds
whose biological processes determine much of the biogeochemical cycles in the
Antarctic coastal systems. In recent years, the imminence of the climate change and
the direct impact of human activities, which are affecting vast regions of the
Antarctica, have highlighted the importance of seaweeds as central components
shaping the structure, functions, and supporting services of benthic ecosystems
under changing polar environment. The present book is aimed to put together the
knowledge and experience gained in recent years by diverse research groups. Many
of these research efforts have long tradition, while others have brought more recently
important new approaches in the study of these organisms with benefits for the
whole polar science. We believe that this initiative is timely and urgently needed in
order to improve our scientific knowledge on these fascinating organisms. In this
chapter, we describe the book’s framework, summarizing the most important
advances in areas related with diversity, biogeography, ecophysiology, biological
interactions, and chemical ecology of Antarctic seaweeds. Finally, considerations
regarding the major gaps and challenges as well as the new directions in the study
of Antarctic seaweeds are outlined.

Keywords  Antarctic · Climate change · Antarctic marine flora · Ecosystem


functions

I. Gómez (*) · P. Huovinen


Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias,
Universidad Austral de Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems, (IDEAL), Valdivia, Chile
e-mail: igomezo@uach.cl; pirjo.huovinen@uach.cl

© Springer Nature Switzerland AG 2020 3


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_1
4 I. Gómez and P. Huovinen

1.1  Introduction: The Historical Context

In recent decades, important advances have been demonstrated in different areas of


knowledge on Antarctic seaweeds, from organisms to ecosystems. However, in
order to understand the roles and services of seaweed communities in Antarctica
currently marked by climate change, it is essential to go back to the history of this
endeavor. The first explorations of Antarctic seaweeds in the nineteenth century
(Gaudichaud 1826; Hooker 1847) had already documented the exuberant presence
of benthic seaweeds and recognized their importance for the coastal ecosystems in
the Antarctic, especially around the Antarctic Peninsula. In his book The Botany of
the Antarctic Voyage of H.M. discovery ships “Erebus” and “Terror” in the years
1839–1843, under the command of Captain Sir James Clark Ross (Fig. 1.1), one of
most complete records of marine and terrestrial flora of the Southern Ocean, the
British botanist Joseph D. Hooker disclosed much of the extraordinary conditions
that characterize the habitat of many Antarctic seaweeds. Later, another important
researcher, Karl Skottsberg, expanded this information from different Antarctic
expeditions in the early twentieth century (e.g., Skottsberg 1907). During the 1960s
and 1970s, descriptions based on scuba diving surveys carried out by Neushul
(1965), Delépine et  al. (1966), Zaneveld (1966), and Lamb and Zimmermann
(1977), among others, confirmed this, highlighting the dominance of large endemic
Desmarestiales at depths >10  m, where they occupy similar role as kelps as the
dominant seaweed group in the Northern Hemisphere and the Arctic. The unique
characteristics of the Antarctic marine flora reflect the complex biogeographic and
evolutionary processes that followed the formation of the Antarctic Circumpolar
Current (ACC) around 30–35 Ma and consequent full glaciation of the Antarctica
(Clayton 1994). Diverse surveys across different sites in the Antarctic, including
communities growing under ice shelves, expanded considerably our knowledge on
vertical distribution, biomass, and diversity of seaweeds (Zielinski 1981, 1990;
Amsler et al. 1995; Klöser et al. 1993, 1996; Brouwer et al. 1995).
Due to the harsh climatic conditions and logistic restrictions in Antarctica,
advances in our knowledge on reproduction, phenology, and acclimation to the
polar environment were only possible since the 1980s. Using cultured material,
Moe and Henry (1982) described for the first time various aspects of the develop-
ment of early phases of Ascoseira mirabilis. The first studies unraveling the sea-
sonal development, life history, and physiological performance of Antarctic
seaweeds were based on algae grown under cultivation conditions simulating the
Antarctic light regime (Wiencke 1990). Based on these findings, two main growth
strategies were defined: the season responders start growth and reproduction when
environmental conditions are optimal in spring and summer, while the season antic-
ipators develop during late winter and spring. Thereafter, the number of investiga-
tions focused on physiology of photosynthesis, growth, chemical ecology, etc.,
increased (revised in Wiencke 1996). A noticeable finding was that various endemic
Antarctic brown algae, such as Ascoseira mirabilis, Cystosphaera jacquinotii,
Desmarestia anceps, and Himantothallus grandifolius, exhibit thallus anatomical
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 5

Fig. 1.1  Cover page of Hooker’s publication describing the flora of the Southern Oceans

and functional  characteristics resembling those of large kelps from the Northern
Hemisphere (Drew and Hastings 1992; Gómez et al. 1995; Fig. 1.2). Here, the most
remarkable morpho-functional adaptations of large Antarctic brown algae are their
very low light demands for growth and photosynthesis and an efficient operation of
light-independent carbon fixation (LICF) at the meristematic zones, which allow
these organisms to display positive carbon balance at depth close to 30 m (Gómez
et  al. 1997). The knowledge on these structural and functional aspects of
6 I. Gómez and P. Huovinen

Fig. 1.2  Large endemic brown algae are the most representative components of the Antarctic
costal systems. (a) Himantothallus grandifolius, (b) Desmarestia anceps, (c) Cystosphaera jac-
quinotii. (Photos by Ignacio Garrido)

p­ hotosynthetic responses had important implications for understanding the biologi-


cal interactions between seaweeds and their associated biota (Zacher et al. 2007;
Iken et al. 2009; Amsler et al. 2011).
In the last decades, the Antarctic ozone depletion and associated increase in
UV-B radiation, as well as the environmental shifts driven by climate change, ori-
ented the research of Antarctic seaweeds. In this context, various studies have exam-
ined the effects of changing irradiance on different algal assemblages across
Antarctica (Schwarz et al. 2003; Zacher et al. 2007; Huovinen and Gómez 2013;
Clark et al. 2017; Deregibus et al. 2016). At an ecosystem level, seaweeds have been
commonly recognized as important sentinels of climatic change in the Antarctic,
highlighting the remarkable capacity of these organisms to adapt to new habitats
(Quartino et al. 2013), and also providing some key ecological ecosystem properties
that permit the maintenance of species richness and biomass (Valdivia et al. 2015).
Through their ecosystem engineering functions, especially large endemic brown
algae are able to minimize environmental variability enhancing the resilience of the
whole system (Ortiz et al. 2017).
Despite these advances, much of the predictions related with adaptation and fate
of Antarctic seaweeds are limited by scarce molecular evidence. From this perspec-
tive, the findings of an increasing number of cryptic species with ­Antarctic/sub-­
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 7

Antarctic or even more vast joint distribution (van Oppen et al. 1993; Hommersand
et al. 2009; Billard et al. 2015) challenge some traditional concepts related with the
evolution and biogeographic patterns of the Antarctic marine flora (Crame 1992;
Clayton 1994). According to current predictions, climatic anomalies, e.g., enhanced
temperature, increased storms, and winds, will be able to break the ecological isola-
tion of Antarctica and facilitate the arrival of temperate species (Fraser et al. 2018),
with impacts on diversity and genetic configuration of local communities yet not
well understood.

1.2  Antarctic Seaweeds in the Wake of Climate Change

The climate, oceanography, and related ecosystem processes in Antarctica and its
surrounding oceanic system have been changing rapidly in the last decades
(reviewed in Constable et  al. 2014). Accelerated regional warming was reported
especially in the WAP region almost 20 years ago (Vaughan et al. 2003). According
to the IPCC scenarios, the mean annual air temperature in this region was predicted
to increase by 1.4–5.8°C until 2100 (Clarke et al. 2007), although strong natural
variability seems characteristic in this region (Turner et  al. 2016). The surface
waters of the Bellingshausen Sea have warmed by 1°C in summer since the 1950s
(Meredith and King 2005), while Schloss et al. (2012) reported an increase of more
than 2°C in winter sea surface temperature between 1991 and 2006 in Potter Cove
(King George Island). This tendency and the possible effects on the polar system
were recently highlighted in the last IPCC report (IPCC 2019). As a synthesis the
report indicates that the Southern Ocean (area corresponding to 25% of world’s
oceans) has been warming at alarming rates, being responsible for 45–62% of the
global ocean warming during the period 2005–2017. Although no clear overall
trends in Antarctic sea ice cover were evident for the period 1979–2018, a strong
decline has been observed recently (2016–2018), which can pose threats to the pho-
tosynthetic organisms due to unpredictable changes in the light regime (see Chap. 7
by Huovinen and Gómez). In the Arctic, massive ice-sheet losses, exceeding the
rates of modeled estimations, have been observed (Bronselaer et al. 2018). Here, the
role of albedo-reducing light-absorbing impurities in ice and snow fields exacerbat-
ing ice loss has been emphasized (Benning et al. 2014; Tedesco et al. 2016; Tedstone
et  al. 2017). Dark snow phenomenon has recently also been associated with
decreased albedo in Maritime Antarctic (Huovinen et al. 2018). Recently, the active
role of ice sheets and icebergs in the global carbon cycle has been recognized
(reviewed by Barnes et al. 2018; Wadham et al. 2019) and can have important con-
sequences for the adjacent marine realm in areas like Maritime Antarctic (Hood
et al. 2015). Although various impacts of these changes are broadcasted for pelagic
realms, their implications for the processes occurring in the Antarctic shallow ben-
thos are much less known (Barnes and Conlan 2012; Constable et al. 2014).
The increasing number of volumes devoted to the present and projected impacts
of global climate changes on the Southern Ocean and their different ecosystems
8 I. Gómez and P. Huovinen

(e.g., Bargagli 2005; Bergstrom et  al. 2006; Rogers et  al. 2012; Tin et  al. 2014;
Kanao et  al. 2018) is a clear evidence of the importance of understanding their
global consequences. Antarctica can be regarded as a natural laboratory where its
physical environment brings the adaptation capacities of organisms to an extreme
limit. In this context, seaweeds, as fundamental components of the Antarctic coastal
systems, can give important insights into the structure and functioning of the biota
in the new scenarios driven by climate change.

1.3  The Book

Based on recent quantitative, observational, and experimental evidences, this book


updates the state of art about the diversity and geographic distribution of seaweeds
as well as their biological interactions and responses to the environment, which is
fundamental for understanding the coastal processes in a changing Antarctica. The
main themes and the overall scientific framework discussed in the book can be sum-
marized in Fig. 1.3.

1.3.1  Diversity and Biogeography

Compared to other biogeographical regions in the Southern Hemisphere, e.g.,


southern Australia, New Zealand, and the southern Chilean coast, the diversity of
seaweeds in the Antarctic has been traditionally considered low. Based on Wiencke
and Amsler (2014), the number of species is 124, showing high endemism (35%).
In their chapter (Chap. 2), Oliveira et al. indicate that the richness of Antarctic sea-
weeds has been underestimated. Based on previous information and recent molecu-
lar surveys, the authors report a diversity of 151 species of which 85 are Rhodophyta,
32 Chlorophyta, and 34 Ochrophyta (most of them brown algae). Likewise, this
update decreased the percentage of endemism to 24%. Overall, the increase in the
number of catalogued species can be explained by improvements in the identifica-
tion tools, e.g., the use of DNA barcoding, more complete gene databases, and more
efficient approaches to detect, e.g., cryptic species. However, a conclusive outcome
of this diversity is far from definitive: a lack of baseline datasets in order to accu-
rately detect local loss of native species, or their replacement by alien assemblages,
still persists. Thus, extending the geographical range and number of surveys, adjust-
ing better the inventories of phylogenetic markers, and deepening the examination
of less conspicuous algal groups, such as crustose and endophyte species, a hidden
diversity normally overlooked, are suggested.
The Antarctic Circumpolar Current has defined the structure, diversity, and func-
tioning of the biomes of the Southern Ocean. Fraser et al. (Chap. 3) make a compre-
hensive analysis of environmental and oceanographic conditions that characterize
Antarctic from the sub-Antarctic regions, the dual role of ACC acting as an efficient
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 9

Fig. 1.3  Schematic presentation of the major drivers, organismal processes, and biological inter-
actions of Antarctic seaweeds. The framework is based on the conclusions of the different chapters
in this book

barrier and also as a bridge connecting marine assemblages and the requisites of
organisms permitting their dispersal across these environmental gradients. Rafting
of floating seaweeds driven by prevailing winds across the different fronts in the
Southern Ocean appears as a central mechanism promoting transoceanic connec-
tions, not only of seaweeds but also invertebrates. The definitive establishment and
persistence of new taxa in these zones will depend on different environmental fil-
ters, e.g., physical and biological constraints, and also on various organismal fea-
tures related with reproductive viability, physiological capacities, etc.
Probably the extent of exchange of species and hence genetic fluxes between
sub-Antarctic and Antarctic regions lie in the diversity of taxa that can be trans-
ported across long distances and their ability to remain alive during their journey.
Macaya et  al. (Chap. 4) indicate that a total of 39 species (3 Chlorophyta, 14
Ochrophyta, and 22 Rhodophyta) have been reported drifting, stranded or floating
in Antarctica or crossing the Antarctic Polar Front (APF). Considering that many
cold and cold-temperate species at both sides of the ACC show remarkable physio-
logical adaptions to biotic and abiotic factors, e.g., grazing, UV radiation, and
10 I. Gómez and P. Huovinen

t­ emperature, they could be able to arrive and colonize different locations around the
Southern Ocean. An example is the floating large brown algae commonly used by
different hitchhiking biota (e.g., barnacles, amphipods, algae). Interestingly, the
authors suggest that various Antarctic seaweeds, some with floating or buoyancy
capacity, have the physiological potential to travel out of the Antarctic.
In their chapter (Chap. 5), Pellizari et al. indicate that the diversity and biogeo-
graphic patterns of Antarctic seaweeds have begun to change. Here, the changing
environmental scenarios in the Southern Ocean, related mostly with circulation and
warming, will determine the new seaweed diversity. Using the seaweed assemblages
of Deception Island in the South Shetlands as a case study, the authors describe an
important presence of species with broad geographical distribution, especially
Chlorophyceans, indicative of recent arrival. Apparently, areas like this character-
ized by peculiar physicochemical conditions could become key places to study the
new Antarctic biodiversity, its biogeographic divergences and connections.
The Antarctic continental margins or peri-Antarctic islands are zones that evi-
dence the long evolutionary history of seaweeds within the Southern Ocean.
Guillemin et al. (Chap. 6) analyzed the sequences of mitochondrial and chloroplast
markers in eight Antarctic species of green, brown, and red seaweeds in order to
determine the genetic patterns in the context of the quaternary climatic oscillations
(QCO). The haplotype network revealed that the studied Antarctic seaweeds show
very low genetic diversity, and significant signatures are indicative of a recent popu-
lation expansion after a massive constriction during the Last Glacial Maximum
(20 Ka). Thus, the authors agree with a theory that this marine flora survived in situ
in a unique refugium and subsequently recolonized the multiple postglacial open
areas using the ACC as a predominant driving force.
In all, Antarctica is not a physically isolated continent, and in a scenario of
increasing warming, the influx of marine organisms arriving, e.g., via rafting to its
coasts, can find new opportunities for colonization, which finally will modify the
local diversity (Fig. 1.3). Here, the examination of large-scale patterns of seaweeds
may provide clues to evaluate aspects of endemism, biological corridors, and expan-
sion of geographical distribution of various algal species. In this context, an account
of the genetic footprints of past diversity can help to understand not only the large-­
scale processes that occurred along the evolution of the Antarctic flora, but also its
future genetic structure.

1.3.2  Environment and Ecophysiology

Due to the harsh environmental conditions, the Antarctic has commonly been
regarded as an inhospitable place for living organisms. Antarctic biota has adapted
to these conditions and thrives in different types of habitats, some marked by
extreme physical variability. However, the new environmental features as a conse-
quence of regional warming and related phenomena occurring in the cryospheric
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 11

realm, as well as direct anthropogenic pressures, are challenging the adaptive strate-
gies of seaweeds in manners still not well understood.
Light is probably the most important environmental factor determining the phe-
nology, spatial distribution, and productivity of Antarctic seaweeds. In Chap. 7,
Huovinen and Gómez describe the underwater optics in the context of present and
future variability and its importance for seaweed photobiology. The optical proper-
ties of the coastal waters, including their light absorbing and scattering components,
define the underwater light environment at ecologically relevant depths (down to
40 m). Despite Antarctic seaweeds being regarded as shade-adapted organisms, they
also show a striking capacity to acclimate to sudden increases in solar radiation.
However, the natural variability in light regimes is being altered due to earlier sea
ice breakup, enhanced runoff from the terrestrial and glacial melting, enhanced
UV-B levels as a result of ozone depletion, etc. These new scenarios are accompa-
nied by emergent stressors (e.g., local freshening, acidification, increasing contami-
nant load) whose influence on the underwater light climate in the Antarctic up to
now is not well understood.
Probably one of the most striking signals of warming in the Antarctic is the
retreat of glaciers, which is creating new ice-free habitats for benthic organisms.
The question of how the future coastal scenarios driven by climate change will
affect the colonization and fate of seaweeds was addressed by Quartino et al. (Chap.
8). In fact, the increased seaweed biomass will enhance the carbon flux and hence
the organic matter towards the higher trophic levels. Due to some species attaining
biomass values close to 10 kg m−2 wet weight, a strong impact on the coastal pro-
ductivity can be expected. However, in these highly dynamic new habitats, reflected
in the model system of Potter Cove in King George Island, seaweed colonization
follows the sharp gradients set by the light penetration, which are strongly modified
by enhanced sedimentation. Considering their great abundance and functional role
as ecosystem engineers, benthic seaweeds can become important carbon sink in
these systems. For instance, it has been estimated that seaweeds can account for a
global net primary production of ca. 1.5 Tg C yr−1 (Krause-Jensen and Duarte 2016),
thus forming part of the “blue carbon” components.
Low water transparency in the new ice-free areas affects the physiological per-
formance of seaweeds in different ways. Deregibus et al. (Chap. 9), based on long-­
term records in areas nearby a retreating glacier at Potter Cove, describe the
photosynthetic carbon balance of seaweeds (the gain of C in photosynthesis versus
that lost in respiration) and its changes in relation with the light climate. Considering
light requirements and photosynthetic efficiency estimated from P-E curves, the
authors indicate that vertical distribution limits of some seaweed species changed as
a result of enhanced turbidity. Accuracy of the carbon balance estimations requires
a robust temporal set of solar irradiance data; thus, the importance of permanent in
situ monitoring accounting for variations at short (hours, days) and long (monthly,
inter-annual) timescale was highlighted.
The performance of seaweed populations under changing environmental regimes
depends on the survivorship of their early reproductive stages. However, life cycle
stages (e.g., spores, microscopic gametophytes, embryonic sporophytes, etc.) can
12 I. Gómez and P. Huovinen

be highly sensitive to environmental stressors. In Chap. 10, Navarro et al. make a


thorough review of the aspects of the physiology of propagules of Antarctic sea-
weeds and how they respond to major physical factors, e.g., solar  radiation, and
temperature, considering present and future settings. UV effects are in many cases
modified by temperature, showing interactions of factors. The response mechanisms
and degree of tolerance of early developmental stages mirror those observed in the
parental individuals. For example, differential responses to UV radiation deter-
mined in adult populations of congeneric and conspecific species from distinct
depth zones (e.g., subtidal versus intertidal) or geographical origin (e.g., Antarctic
versus sub-Antarctic) have also been observed in their propagules.
Antarctic seaweeds can be very abundant in terms of biomass and account by
more than 50% of the coastal primary productivity, especially around the Antarctic
Peninsula. Much of this ubiquity is strongly linked with efficient morpho-functional
adaptations that have permitted these organisms to occupy niches characterized by
sharp physical gradients. In Chap. 11, Gómez and Huovinen discuss the importance
of the form and function of seaweeds. In general, the functional forms are well dis-
tributed along the major groups of Antarctic seaweeds: coarsely branched and leath-
ery species, which can be regarded as the most robust and large-sized forms,
represent 49% of the total number of species. In this group, endemic brown and red
algae dominate, mainly growing at the subtidal zone. Filamentous, finely branched
and foliose species (41%) belong mostly to green and red algae, common at shallow
and intertidal sites, and are geographically widely distributed. Each of these morphs
are integrated in different life strategies and hence distinct ecosystem functions. For
example, perennial canopy-forming species show competitive abilities for light and
substrate, but in general prevail less in sites subject to strong physical perturbation.
Here, small colonizers and opportunistic species dominate in virtue of rapid meta-
bolic adjustments and turnover rates.
The different chapters reveal that the abiotic environment of Antarctica is chang-
ing in extent that is already affecting several aspects of the physiology of marine
biota in general and seaweeds in particular. The emergence of new habitats available
as the glaciers retreat is modifying the composition, structure, and trophic relations
of the benthic communities dominated by seaweeds. Apparently, a strategic factor
underlying these responses is the ability of adult plants and their propagules to
acclimate, via different functional traits, to the environmental shifts.

1.3.3  Ecological Functions

The ecological succession in Antarctic benthos determines the structure of the


mature community and its biological network. Different types of positive and nega-
tive interactions between algal assemblages, invertebrates, fish, and microorgan-
isms can be identified as the community develops. Based on in situ experiments,
Campana et al. (Chap. 12) describe the successional stages and their biotic interrela-
tions in a coastal site near Potter Cove. During the first three months, the incipient
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 13

community is dominated by microorganisms and benthic diatoms, which apparently


promote the development of small ephemeral filamentous green algae. The assem-
blage is successively enriched by the presence of various foliose and crustose forms
of red algae and during late algal succession (after 4 years) by some perennial spe-
cies of Desmarestia. The different components of the succession respond differently
to environmental factors such as UV radiation, grazing, glacier retreat, and sea ice,
and hence the structure and the biotic relations change dynamically within this early
community.
Grazing is probably the most important biotic factor controlling the structure and
composition of seaweed-dominated communities. In Antarctic benthic systems, the
early successional stages dominated by small-sized seaweeds and periphyton repre-
sent excellent models to study how grazing modifies different ecological properties
not only of native assemblages but also of alien species, whose arrival and establish-
ment will be stimulated by climate warming. In Chap. 13, Valdivia determined by
means of mathematical simulations the impact of mesograzers in sub-Antarctic and
Antarctic sites connected by dispersal. Ulva sp. was regarded an alien species, being
highly competitive in the Antarctic but not in the sub-Antarctic littoral. The results
indicated that Antarctic mesograzers have a deterministic and marked effect on the
biomass of the alien seaweeds; however, projected climate-change-driven shifts in
temperature or pH could decrease the potential of, for example, amphipod grazers
to control the development of invaders.
Antarctic seaweeds harbor complex and intricate microbiomes, which exert
important influence on different molecular and biochemical processes of the algal
host. Hitherto much of the coevolutionary processes of this association have been
little studied. However, it is reasonable to argue that microbiota plays important
functional roles in the ecology of Antarctic seaweeds. Gaitan-Spitia and Schmid
(Chap. 14) review various aspects of structure, diversity, and functioning of Antarctic
microbiomes and their implications for seaweeds. Members of phylum Actinobacteria
show high diversity and persistence among different seaweed species, while
Firmicutes are less represented. In general, the microbiomes associated to seaweeds
are different from those found in the surrounding environment, which suggest that
the bacterial composition is regulated by the seaweed host. Apparently, this feature
reflects adaptive strategies to respond to multiple environmental conditions, e.g.,
antioxidation, antimicrobial activity, photoprotection, etc.
Seaweeds and microphytobenthos represent the basis of the Antarctic coastal
food web. Because coastal areas can become highly perturbed, the dynamics and
stability of the interspecific interrelations have fundamental influence on the whole
benthic ecosystem at different spatial and temporal scales. Momo et al. (Chap. 15)
determined that the food web at Potter Cove is based on 24 seaweed species and
diverse other photosynthetic organisms, such as epiphytic and benthic diatoms and
phytoplankton as well as their detritus. The system is also hyperconnected indicat-
ing multiple energy pathways. Considering extinction thresholds, this network can
be regarded as relatively resilient to local losses of seaweed species. Similarly, using
as a model Fildes Bay, a coastal system geographically close to Potter Cove, Ortiz
et al. (Chap. 16) analyzed different keystone species complexes, which contribute
14 I. Gómez and P. Huovinen

importantly to the emergent network properties, such as growth, organization,


development, maturity, and health of the ecosystem. The theoretical framework
(based on network analysis, ascendency, and loop analysis) identified detritus, the
phyto-zooplankton complex, sea stars, sea urchins, and seaweeds as the major com-
ponents determining the overall structure and function of this system. Similar to
Potter Cove, Fildes Bay appears to be a less developed system compared to other
cold-temperate system, but being highly resilient to physical perturbations.
The described examples on the ecology of the coastal system in the Maritime
Antarctic reveal complex biological interactions between, e.g., algae, microbiota,
invertebrates, and fish, which are strongly regulated by the physical environment. In
this scenario, seaweeds are identified as key components from the early stages of
succession to the consolidate communities. Due to the strong influence of physical
factors from terrestrial, freshwater, cryospheric and atmospheric processes, the
structure, function, and trophic interrelations in these communities are in general
very resistant to disturbances (Fig. 1.3). Thus, it seems that there are internal mech-
anisms operating at individual (e.g., efficient growth strategies, multiple anti-stress
mechanisms) as well as at population and community (e.g., filters controlling native
and alien species, high biological complexity based on species and biomass rich-
ness) levels, providing the system with a high resilience.

1.3.4  Chemical Ecology

The trophic relations in the Antarctic benthic system show a balance between con-
sumption by herbivores and their deterrence. Amsler et al. (Chap. 17) review the
recent advances in relationship between seaweeds and, e.g., amphipods, gastropods,
and fish. Diverse halogenated monoterpenes and phlorotannins (phenolic com-
pounds found in brown algae), and probably various other compounds, confer many
species of Antarctic seaweeds unpalatability to different kinds of herbivores.
Interestingly, the relationship between some seaweeds and various species of
amphipods includes mutualism, in which chemically defended algae offer protec-
tion from, e.g., omnivorous fish, while amphipods reduce the biofouling and epi-
phytic load of the thalli.
Chemical defenses based on phlorotannins operate not only against grazing, but
also form part of a wide suite of constitutive anti-stress mechanisms. In Chap. 18,
Gómez and Huovinen summarize the different aspects that determine the synthesis
and accumulation of these substances, which in some Antarctic brown algae can
represent up to 12% of the dry weight. These compounds have different functions as
grazing deterrents, reactive oxygen species (ROS) scavenging agents, and metal
chelators and can be allocated in different thallus parts to optimize defense. Although
phlorotannins are regarded as UV screening substances, no evidence on UV induc-
tion in Antarctic seaweeds has been reported. However, the antioxidant capacity
increases substantially along with increasing phlorotannin concentrations in algal
extracts, even in algae not naturally exposed to UV radiation.
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 15

The high prevalence of chemical defenses observed in Antarctic seaweeds is


remarkable and suggests their central role in defining their dominance and biologi-
cal interactions in the Antarctic coastal ecosystems (Fig. 1.3). Particularly, the con-
stitutively high levels of phlorotannins measured in various dominant brown algae
(e.g., Desmarestia anceps, Himantothallus grandifolius) open interesting questions
about the activation of anti-stress mechanisms based on chemical substances with
multiple primary and secondary functions. For seaweed assemblages subjected to
climate-change-driven environmental shifts, such defenses could confer ecological
advantages.

1.4  Gaps, Emerging Challenges, and Future Directions

The different chapters throughout this book update the current knowledge and pro-
vide novel insight into various aspects on diversity, ecophysiology, and ecology of
Antarctic seaweeds, with particular emphasis on their responses to the changing
polar environment. However, several gaps still persist and new questions require
attention in the near future.
• Long-term assessment: Due to logistical constraints, research in Antarctica is
normally restricted to the spring-summer season. This time frame clearly does
not permit covering the entire environmental variability to which Antarctic
organisms, especially annual and perennial species, are exposed. For example,
many gaps exist on the metabolic performance of seaweeds (e.g., carbon and
nutrient metabolism, use and remobilization of photoassimilates, etc.) during the
long Antarctic winter. In fact, the few studies addressing photosynthesis in win-
ter or under ice cover suggest that seaweeds are at their physiological limit dur-
ing this period (Gutkowski and Maleszewski 1989; Drew and Hastings 1992;
Schwarz et  al. 2003). These studies should be complemented with long-term
monitoring of annual and inter-annual physical fluctuations in order to delimit
the ranges of acclimation and adaptation of organisms. Because most of the mon-
itoring platforms deployed around the Antarctica are designed to record changes
in the open ocean, long-term or real-time baseline information of near coastal
processes is still very limited. In this context, the long-term observations focused
on the impact of the retreating Fourcade Glacier in Potter Cove (King George
Island) represent an important effort in gaining insights into the responses of
benthos at ecological scales (Meredith et al. 2018; see Chap. 8 by Quartino et al.
and Chap. 9 by Deregibus et al. and references therein).
• Molecular ecology: Although remarkable improvements in biomolecular tools
have considerably expanded our capacities to record and elucidate the taxonomi-
cal status of Antarctic species (Held 2014), many seaweeds are still not well
classified, are cryptic or due to their life form (e.g., epiphytes, endophytes or
prostrates) remain undiscovered. Another important limitation challenging the
efforts to expand not only the genetic inventories, but also the general knowledge
16 I. Gómez and P. Huovinen

on Antarctic organisms, is that the surveyed areas are strongly biased towards
some regions, especially around the Antarctic Peninsula and in sites in direct
proximity to research stations, while other coasts, e.g., from the East Antarctic,
have been scarcely visited (Mormède et al. 2014). Thus, it is assumed that in the
near future, along with the advances in phylogeography and population genetics
as well as in geographic coverage, the number of Antarctic seaweed species, both
native and recently arrived, will increase (see Chap. 2 by Oliveira et al.).
There a considerable lags in our understanding of gene expression and regula-
tion. This is probably one of the weakest areas in the study of seaweeds in general
and Antarctic species in particular. Thus, use of molecular tools such as transcrip-
tomic analysis will help identify the metabolic pathways and adaptive strategies that
Antarctic seaweeds exhibit beyond their tolerance threshold. For example, recently
high and constitutive gene expression of various physiological reactions, including
photochemical and inorganic carbon utilization components, from RNA-Seq analy-
sis was reported for the first time for an Antarctic endemic species (the brown alga
Desmarestia anceps; Iñiguez et al. 2017). Clearly this type of techniques open new
avenues for the identification of transcripts that are differentially expressed under
different stress conditions. On the other hand, the new molecular tools together with
improved physiological methodologies are fundamental to predict whether key
Antarctic seaweeds exhibit the molecular machinery to respond to ongoing and
near-future impacts of climate change.
• Ontogenetic development and life cycle responses: Developmental phases (e.g.,
spores, gametes, and embryonic sporophytes) are highly sensitive to environ-
mental changes (reviewed in Chap. 10 by Navarro et al.). However, they are often
overlooked due to their small size or because the logistical constraints associated
with their isolation, culture, and experimentation in Antarctica (Wiencke 1988).
Considering that the fate of these cells determine the structure and dynamics of
further life phases, it is urgent to conduct research focused on the acquisition of
stress tolerance capacity at different developmental stages and how this resil-
ience is “transferred” over generations. Following important developments in the
identification and visualization techniques in microalgae, e.g., fluorescence cell-­
based sensing and “omics” approaches (metabolomics, proteomics, genomics), it
is now possible to quantify in real time the effects of different stressors on cel-
lular structures of early stages of seaweeds. Thus, it will be possible to track the
progressive expression of anti-stress mechanisms along the ontogeny or life
cycle phases, an essential approach to understand the adjustments in response to
environmental changes at an organismal level.
• Direct anthropogenic impacts and interaction of multiple stressors: Warming
and ozone depletion are not the only threats to Antarctic biota. Among other
concerns are ocean acidification and local decreases in salinity (freshening) due
to enhanced melting of glaciers. Furthermore, increase of pollution in the
Antarctic environment is generating new and not well-understood threats to
these ecosystems. As the identification of sources, concentrations, and persis-
tence of inorganic and organic pollutants poses considerable challenges (reviewed
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 17

in Caroli et al. 2001; Bargagli 2005), their effects on seaweeds and their com-
munities are hitherto widely unknown. Moreover, many contaminants are reac-
tive to other environmental factors (e.g., UV radiation), which may enhance their
detrimental impact on biota. Because all these different variables are changing
simultaneously, the research on the impact of their interactive effects (synergis-
tic, antagonistic, additive, etc.) is challenging  (see Chap. 7 by Huovinen and
Gómez).
Finally, the contents of this book are in agreement with the increasing awareness
of the importance of Antarctic and its biota in global processes and the urgency to
improve our understanding on the role and sentinel responses of seaweeds to global
climate change. We believe that a comprehensive account of the progress made in
the last decades is timely and urgent in order to put into perspective how diversity,
ecophysiological adaptations, and ecosystem relations of seaweeds will be molded
in the future Antarctica.

References

Amsler CD, Rowley RJ, Laur DR, Quetin LB, Ross RM (1995) Vertical distribution of Antarctic
Peninsular macroalgae: cover, biomass, and species composition. Phycologia 34:424–430
Amsler CD, Iken K, McClintock JB, Baker BJ (2011) Defenses of polar microalgae against her-
bivores and biofoulers. In: Wiencke C (ed) Biology of polar benthic algae. Walter de Gruyter
GmbH & Co, KG, Berlin, New York, pp 101–120
Bargagli R (ed) (2005) Antarctic ecosystems: environmental contamination, climate change, and
human impact. Springer, Berlin
Barnes DKA, Conlan KE (2012) The dynamic mosaic–disturbance and development of Antarctic
benthic communities. In: Rogers AD, Johnston NM, Murphy EJ, Clarke A (eds) Antarctic eco-
systems: an extreme environment in a changing world. Wiley, Chichester, UK
Barnes DKA, Fleming A, Sands CJ, Quartino ML, Deregibus D (2018) Icebergs, sea ice, blue
carbon and Antarctic climate feedbacks. Philos Trans R Soc A 376 (2122): 20170176
Benning LG, Anesio AM, Lutz S, Tranter M (2014) Biological impact on Greenland’s albedo. Nat
Geosci 7:691
Bergstrom DM, Convey P, Huiskes AHL (eds) (2006) Trends in Antarctic terrestrial and limnetic
ecosystems. Springer, Dordrecht, The Netherlands
Billard E, Reyes J, Mansilla A, Faugeron S, Guillemin M-L (2015) Deep genetic divergence
between austral populations of the red alga Gigartina skottsbergii reveals a cryptic species
endemic to the Antarctic continent. Polar Biol 38:2021–2034
Bronselaer B, Winton M, Griffies SM, Hurlin WJ, Rodgers KB, Sergienko OV, Stouffer RJ, Russell
JL (2018) Change in future climate due to Antarctic meltwater. Nature 564:53–58
Brouwer PEM, Geilen EFM, Gremmen NJM, van Lent F (1995) Biomass, cover and zonation
patterns of sublittoral macroalgae at Signy Island, South Orkney Islands, Antarctica. Bot Mar
38:259–270
Caroli S, Cescon P, Walton BT (eds) (2001) Environmental contamination in Antarctica, a chal-
lenge to analytical chemistry. Elsevier, Amsterdam, London
Clark GF, Stark JS, Palmer AS, Riddle MJ, Johnston EL (2017) The roles of sea-ice, light and sedi-
mentation in structuring shallow Antarctic benthic communities. PLoS One 12(1):e0168391.
https://doi.org/10.1371/journal.pone.0168391
18 I. Gómez and P. Huovinen

Clarke A, Murphy EJ, Meredith MP, King JC, Peck LS, Barnes DKA, Smith RC (2007) Climate
change and the marine ecosystem of the western Antarctic Peninsula. Philos Trans R Soc B
362:149–141
Clayton MN (1994) Evolution of the Antarctic benthic algal flora. J Phycol 30:897–904
Constable AJ, Melbourne -Thomas J, Corney SP, Arrigo KR, Barbraud C, Barnes DKA et al (2014)
Climate change and Southern Ocean ecosystems. I: how changes in physical habitats directly
affect marine biota. Glob Chang Biol 20:3004–3025. https://doi.org/10.1111/gcb.12623
Crame JA (1992) Evolutionary history of the polar regions. Hist Biol 6:37–60
Delépine R, Lamb JM, Zimmermann MH (1966) Preliminary report on the vegetation of the
Antarctic Peninsula. Proc Int Seaweed Symp 5:107–116
Deregibus D, Quartino ML, Campana GL, Momo FR, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in Potter
Cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166
Drew EA, Hastings RM (1992) A year-round ecophysiological study of Himantothallus grandifo-
lius (Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31:262–277
Fraser CI, Morrison AK, Hogg AM, Macaya EC, van Sebille E, Ryan PG, Padovan A, Jack C,
Valdivia N, Waters JM (2018) Antarctica’s ecological isolation will be broken by storm-­
driven dispersal and warming. Nat Clim Change 8:704–708. https://doi.org/10.1038/
s41558-018-0209-7
Gaudichaud C (1826) Botanique. In: Freycinet L (ed) Voyage autour du monde, entrepons par
ordre du Roi, execute sur le corvettes de S.M. l ‘Uranie’ et la ‘Physicienne’, pendant les années
1817, 1818, 1819 et 1820. Pillet Aîné, Paris
Gómez I, Thomas DN, Wiencke C (1995) Longitudinal profiles of growth, photosynthesis and
light independent carbon fixation in the Antarctic brown alga Ascoseira mirabilis. Bot Mar
38:157–164
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, daily carbon
balance and zonation of sublittoral macroalgae from King George Island (Antarctica). Mar
Ecol Prog Ser 148:281–293
Gutkowski R, Maleszewski S (1989) Seasonal changes of the photosynthetic capacity of the
Antarctic macroalga Adenocystis utricularis (Bory) Skottsberg. Pol Biol 10:145–148
Held C (2014) Chapter 10.5. Phylogeography and population genetics. In: De Broyer C, Koubbi P,
Griffiths HJ, Raymond B, d’ U d’A C, Van de Putte AP, Danis B, David B, Grant S, Gutt J et al
(eds) Biogeographic Atlas of the Southern Ocean. Scientific Committee on Antarctic Research,
Cambridge, pp 437–440
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and bio-
geographical relationships of Antarctic and subAntarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52:509–534
Hood E, Battin TJ, Fellman J, O’Neel S, Spencer RGM (2015) Storage and release of organic
carbon from glaciers and ice sheets. Nat Geosci 8:91–96
Hooker JD (1847) The botany of the Antarctic voyage of H. M. discovery ships Erebus and Terror
in the years 1839–1843, vol. 1, Flora Antarctica, Part 1 Botany of Lord Auklands Group and
Campbell’s Island, Part 2 Botany of Fuegia, The Falklands, Kerguelens Land. Reeve Brothers,
London
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
Huovinen P, Ramírez J, Gómez I (2018) Remote sensing of albedo-reducing snow algae and impu-
rities in the Maritime Antarctica. ISPRS J Photogram Rem Sens 146:507–517
Iken K, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009) Field studies on deterrent roles
of phlorotannins in Antarctic brown algae. Bot Mar 52:547–557
Iñiguez C, Heinrich S, Harms L, Gordillo FJL (2017) Increased temperature and CO2 alleviate
photoinhibition in Desmarestia anceps: from transcriptomics to carbon utilization. J Exp Bot
68(14):3971–3984
1  Antarctic Seaweeds: Biogeography, Adaptation, and Ecosystem Services 19

IPCC (2019) IPCC special report on the ocean and cryosphere in a changing climate. Pörtner H-O,
Roberts DC, Masson-Delmotte V, Zhai P, Tignor M, Poloczanska E, Mintenbeck K, Nicolai M,
Okem A, Petzold J, Rama B, Weyer N (eds)
Kanao M, Genti Toyokuni G, Yamamoto M-Y (eds) (2018) Antarctica–a key to global change.
IntechOpen. https://doi.org/10.5772/intechopen.82197
Klöser H, Ferreyra G, Schloss I, Mercuri G, Laturnus F, Curtosi A (1993) Seasonal variation of
algal growth conditions in sheltered Antarctic bays: the example of Potter Cove (King George
Island, South Shetlands). J Mar Syst 4:289–301
Klöser H, Quartino ML, Wiencke C (1996) Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiologia 333:1–17
Krause-Jensen D, Duarte CM (2016) Substantial role of macroalgae in marine carbon sequestra-
tion. Nat Geosci 9(10):737
Lamb IM, Zimmermann MH (1977) Benthic marine algae of the Antarctic Peninsula. Ant Res Ser
23:130–229
Meredith MP, King JC (2005) Rapid climate change in the ocean west of the Antarctic Peninsula
during the second half of the 20th century. Geophys Res Lett 32:L19604. https://doi.
org/10.1029/2005GL024042
Meredith MP, Falk U, Bers AV, Mackensen A, Schloss IR, Ruiz Barlett E, Jerosch K, Silva Busso
A, Abele D (2018) Anatomy of a glacial meltwater discharge event in an Antarctic cove. Phil
Trans R Soc A 376:20170163. https://doi.org/10.1098/rsta.2017.0163
Moe R, Henry EC (1982) Reproduction and early development of Ascoseira mirabilis Skottsberg
(Phaeophyta), with notes on Ascoseirales Petrov. Phycologia 21:55–56
Mormède S, Irisson JO, Raymond B (2014) Chapter 2.3. Distribution modelling. In: De Broyer C,
Koubbi P, Griffiths HJ, Raymond B, Udekem d’Acoz C d', Van de Putte AP, Danis B, David B,
Grant S, Gutt J et al (eds) Biogeographic Atlas of the Southern Ocean. Scientific Committee on
Antarctic Research, Cambridge, pp 27–29
Neushul M (1965) Diving observations of subtidal Antarctic marine vegetation. Bot Mar 8:234–243
Ortiz M, Hermosillo-Nuñez B, González J, Rodríguez-Zaragoza F, Gómez I, Jordán F (2017)
Quantifying keystone species complexes: Ecosystem-based conservation management in the
King George Island (Antarctic Peninsula). Ecol Ind 81: 453-460
Quartino ML, Deregibus D, Campana GL, Edgar G, Latorre J, Momo FR et al (2013) Evidence of
macroalgal colonization on newly ice-free areas following glacial retreat in Potter Cove (South
Shetland Islands), Antarctica. PLoS One 8:e58223
Rogers AD, Johnston NM, Murphy EJ, Clarke A (eds) (2012) Antarctic ecosystems: an extreme
environment in a changing world. Wiley, Chichester, UK
Schloss IR, Abele D, Moreau S, Demers S, Bers AV, Gonzalez O, Ferreyra GA (2012) Response of
phytoplankton dynamics to 19-year (1991–2009) climate trends in Carlini Station (Antarctica).
J Mar Syst 92:53–66. https://doi.org/10.1016/j.jmarsys.2011.10.006
Schwarz A, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photosyn-
thesis near the southern global limit for growth, Cape Evans, Ross Sea, Antarctica. Polar Biol
26:789–799
Skottsberg K (1907) Zur Kenntnis der subantarktischen und antarktischen Meeresalgen.
I.  Phaeophyceen. In: Nordenskjöld O (ed) Wissenschaftliche Ergebnisse der Schwedischen
Südpolar-Expedition 1901–1903, vol 4: 1. Lithographisches Institut des Generalstabs,
Stockholm, pp 1–172
Tedesco M, Doherty S, Fettweis X, Alexander P, Jeyaratnam J, Stroeve J (2016) The darkening of
the Greenland ice sheets: trends, drivers, and projections (1981–2100). Cryosphere 10:477–
496. https://doi.org/10.5194/tc-10-477-2016
Tedstone AJ, Bamber JL, Cook JM, Williamson CJ, Fettweis X, Hodson AJ, Tranter M (2017)
Dark ice dynamics of the south-west Greenland ice sheet. Cryosphere 11:2491–2506. https://
doi.org/10.5194/tc-11-2491-2017
20 I. Gómez and P. Huovinen

Tin T, Liggett D, Maher PT, Lamers M (2014) Antarctic futures: human engagement with the
Antarctic environment. Springer, The Netherlands
Turner J, Lu H, White I, King JC, Phillips T, Hosking JS, Bracegirdle TJ, Marshall GJ, Mulvaney
R, Deb P (2016) Absence of 21st century warming on Antarctic Peninsula consistent with
natural variability. Nature 535:411–415
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental stress gradients in a marine macroalgal-dominated subtidal community on the
Western Antarctic Peninsula. Plos One 10(9):e0138582
Van Oppen MJH, Olsen JL, Stam WT, Van der Hoek C, Wiencke C (1993) Arctic–Antarctic dis-
junction in the benthic seaweeds Acrosiphonia arcta (Chlorophyta) and Desmarestia viridis
(Phaeophyta) are of recent origin. Mar Biol 115:381–386
Vaughan DG, Marshall GJ, Connolley WM, Parkinson CL, Mulvaney R, Hodgson DA, King JC,
Pudsey CJ, Turner J (2003) Recent rapid regional climate warming on the Antarctic Peninsula.
Clim Change 60:243–274. https://doi.org/10.1023/A:1026021217991
Wadham JL, Hawkings JR, Tarasov L, Gregoire LJ, Spencer RGM, Gutjahr M, Ridgwell A,
Kohfeld KE (2019) Ice sheets matter for the global carbon cycle. Nat Comm 10 (1)
Wiencke C (1988) Notes on the development of some benthic marine macroalgae of King George
Island, Antarctica. Ser Cient INACh 37:23–47
Wiencke C (1990) Seasonality of brown macroalgae from Antarctica- a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600
Wiencke C (1996) Recent advances in the investigation of Antarctic macroalgae. Polar Biol
16:231–240
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1 Macroalgae. In: de Broyer C, Koubbi P,
Griffiths HJ, Raymond B, Udekem d’Acoz Cd’, Van de Putte PAP, Danis B, David B, Grant S,
Gutt J et al (eds) Biogeographic atlas of the Southern Ocean. Scientific Committee on Antarctic
Research, Cambridge, pp 66–73
Zacher K, Wulff A, Molis M, Hanelt D, Wiencke C (2007) UV radiation and consumer effects on
a field-grown intertidal macroalgal assemblages in Antarctica. Glob Chang Biol 13:1201–1215
Zaneveld JS (1966) Vertical zonation of Antarctic and subAntarctic benthic marine algae. Antarctic
J US 1(5):211–213
Zielinski K (1981) Benthic macroalgae of Admiralty Bay (King George Island, South Shetland
Islands) and circulation of algal matter between the water and the shore. Pol Polar Res 2:71–94
Zielinski K (1990) Bottom macroalgae of the Admiralty Bay (King George Island, South Shetlands,
Antarctica). Pol Polar Res 11:95–131
Part II
Diversity and Biogeography
Chapter 2
Diversity of Antarctic Seaweeds

Mariana C. Oliveira, Franciane Pellizzari, Amanda S. Medeiros,


and Nair S. Yokoya

Abstract  Antarctica is characterized by extremes of climate and biogeographic iso-


lation from other continents by distance, high depths, and the Antarctic Circumpolar
Current. Even under these harsh conditions, macroalgae thrive in different coastal
ecosystems contributing to primary production and serving as habitat and food for a
variety of species of marine fauna. However, it is known that the Antarctic marine
flora presents low species richness compared to other biogeographical regions: until
the past decade a number of 120 Antarctic seaweeds had been reported. On the other
hand, long geographical isolation and extreme climatic and oceanographic conditions
justify their high degree of endemism (ca. of 33–40%). A new compilation of the
Antarctic seaweed diversity is presented in this chapter, reporting a list of 151 species
cited to the entire Antarctica, comprising 85 Rhodophyta, 34 Ochrophyta
(Phaeophyceae and Chrysophyceae), and 32 Chlorophyta with an endemism degree
of 27%, lower than in previous reports. Molecular approaches based on different
markers (ITS, UPA, COI-5P) are being used to assist species identification. The col-
lection of marine specimens in Antarctica is expensive and still very difficult, and
therefore, the occurrence for many species can become inaccurate. The difficult
access to samples is another limitation, which could explain that most of the best
known species are concentrated around scientific stations. Consequently, the mac-
roalgal diversity in Antarctica and its distribution is probably underestimated. A bet-
ter knowledge on this diversity and its distribution is urgent, as the region is facing
significant climate changes that may drive shifts on the assemblages of macroalgae.

M. C. Oliveira (*) · A. S. Medeiros


Department of Botany, Institute of Biosciences, University of São Paulo,
São Paulo, SP, Brazil
e-mail: mcdolive@ib.usp.br
F. Pellizzari
Laboratório de Ficologia e Qualidade de Água Marinha, Universidade Estadual do Paraná,
Campus Paranaguá, Paranaguá, PR, Brazil
e-mail: franciane.pellizzari@unespar.edu.br
N. S. Yokoya
Institute of Botany, São Paulo, SP, Brazil
e-mail: nyokoya@pq.cnpq.br

© Springer Nature Switzerland AG 2020 23


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_2
24 M. C. Oliveira et al.

Keywords  Benthic algae · Diversity · Endemic species · Maritime Antarctica ·


Molecular taxonomy

2.1  The Antarctic Environment

Antarctica is characterized by extremes of climate, which makes its habitats and


biogeographic context quite unique. It is isolated from other continents by distance,
high depths, and oceanographic currents, mainly by the Antarctic Circumpolar
Current (Barker and Thomas 2003) (Fig. 2.1). Broad variations in temperature and
salinity can occur in the Antarctic habitat, with high incidence of alternating ultra-
violet radiation with long periods of absence of light, freezing, and thawing cycles
that determine highly variable physical mosaics (Wynn-Williams 1996; Vincent
2000; Clarke et al. 2005). However, few attempts have been made to provide sub-

Fig. 2.1  Schematic map showing Antarctica and surroundings, Africa (AF), Australia (AU), New
Zeeland (NZ), and South America (SA). King George Island (South Shetland Islands) is marked
by ∗ and Adelaide Island is marked by ★. The Antarctic Circumpolar Current is represented by a
black line and arrowheads. (https://freevectormaps.com/globes/antarctica/GLB-AN-01-0001)
2  Diversity of Antarctic Seaweeds 25

stantial insight into the implications of these gradients of abiotic factors for the
spatial distribution of biodiversity (Convey et al. 2014).
Low temperatures, high salinities, and the occurrence of long periods of ice
cover in coastal regions can characterize the Southern Ocean. The seasonality of
irradiance levels and photoperiod, as well as the ice cover, exert a strong effect on
supralittoral, eulittoral, and sublittoral communities (Hempel 1987; Drew and
Hastings 1992). The alternation of annual ice cycles impacts the physical and chem-
ical environment in many ways and, consequently, the local biota as a whole. The
effects of ice cover are most obvious in the intertidal zone and in the upper sublit-
toral due to the physical action of the displacement of ice blocks that remove sessile
organisms, including macroalgae. Ice formation, as well as summer melting, signifi-
cantly alters salinity and light penetration, exposing benthic organisms to extreme
values of these parameters.
Even so, the Southern Ocean is known to afford living space for high abundance
of benthic organisms compared to other regions of the world. Since the first studies
carried out in the region, the high densities and relatively high diversity of benthic
communities have been demonstrated (Clarke 1990).

2.2  Seaweeds in Antarctica: Definition and Importance

Similar to the terrestrial plants, seaweeds – or marine macroalgae – are photosyn-


thetic organisms that form underwater forests on consolidated or hard substrate. In
virtue of their diversity and biomass, seaweeds play a key role in sustaining the
primary production in coastal ecosystems and are important food sources, espe-
cially to the rockfish, Notothenia coriiceps; echinoderms, e.g., Odontaster validus
and Sterechinus neumayeri; and amphipods, e.g., Gondogeneia antarctica.
Moreover, macroalgae serve as shelter and substrate for the growth and reproduc-
tion of several marine organisms (Nedzarek and Rakusa-Suszcewski 2004, see also
Chap. 15 by Momo et al. and Chap. 16 by Ortiz et al.).
Antarctic seaweeds belong to the major divisions of algae: golden-brown and
brown (Ochrophyta: Chrysophyceae and Phaeophyceae, respectively), green
(Chlorophyta), and red algae (Rhodophyta). The morphology of Antarctic macroal-
gae varies from crustose forms or delicate filamentous forms of few centimeters to
large foliaceous, terete, and leathery forms that may reach several meters long
(Fig. 2.2). Some species deposit calcium carbonate in their cell walls, presenting a
rigid thallus. These calcareous species provide substrate for other marine species
(see also Chap. 11 by Gómez and Huovinen). Specific groups of algae are protected
against predation by synthesizing chemical compounds that make them unpalatable
(see Chap. 17 by Amsler et al.). The synthesis and accumulation of these bioactive
compounds is the result of a long and complex evolutionary and ecological process.
Some of these substances, especially for the species adapted to extreme polar eco-
systems, may have biotechnological and pharmaceutical applications. The repro-
ductive sexual and/or asexual cells of Antarctic seaweeds are well adapted to the
26 M. C. Oliveira et al.

Fig. 2.2 Macromorphology of Antarctic seaweeds. (a) Crustose calcareous red alga


Clathromorphum sp. (Rhodophyta). (Photo by F. Pellizari); (b) monostromatic saccate green alga
Monostroma hariotii (Chlorophyta); (c) terete red alga Plocamium sp. (Rhodophyta). (Photos by
A.S.  Medeiros); (d) leathery brown alga Himantothallus grandifolius (Phaeophyceae), with an
approximate length of 3.7 m. (Photo by E.C. Oliveira)
2  Diversity of Antarctic Seaweeds 27

environmental conditions marked by low temperature and limited solar radiation.


These cells, either an asexual stage, a gamete, or a zygote (after gametes fusion),
will float within the marine currents and settle and develop on a suitable substrate,
in a dynamic tuned with the seasonal Antarctic variability (see also Chap. 10 by
Navarro et al.).
In the Southern Ocean, macroalgae together with the phytoplankton are respon-
sible for 30–50% of the dissolved O2 released in the marine habitats surrounded by
the Southern Ocean (data retrieved from visibleearth.nasa.gov). In particular, the
robust and perennial endemic brown algae (Phaeophyceae), such as Himantothallus
grandifolius and Desmarestia spp., are well represented in terms of biomass and are
the main primary producers in these areas (Amsler et al. 1995; Wiencke et al. 2014).

2.3  S
 eaweed Taxonomic Studies in Antarctica: Toward
a New Species Compilation

Antarctic seaweed communities have been characterized by low diversity and high
levels of endemism (Lamb and Zimmermann 1977; Wiencke et al. 2014). The first
reports date back to the nineteenth century when seaweeds were collected during
cruises of the corvettes Uranie and Physicienne (Gaudichaud 1826). However, the
first compilations on diversity of Antarctic seaweeds along the Antarctic Peninsula
were published much later by Skottsberg (1906, 1941, 1953, 1964), who is consid-
ered the pioneer of the Antarctic phycology. The first scuba diving surveys were
conducted by Neushul (1959, 1961, 1963, 1965, 1968) and Skottsberg and Neushul
(1960). Skottsberg (1964) estimated the occurrence of 96 species of Antarctic sea-
weeds (16 Chlorophyta, 19 Phaeophyceae, and 61 Rhodophyta). In later surveys,
higher species richness was documented: 100 species (Papenfuss 1964; Moe 1985),
120 species (Clayton 1994), and 117–123 species (Wiencke and Clayton 2002). The
most recent review on seaweed diversity for the entire Antarctica reported a total of
124 taxa, comprising 80 species of Rhodophyta, 27 species of Phaeophyceae, and
17 Chlorophyta (Wiencke et al. 2014).
Different researchers have broadly studied seaweed communities from the South
Shetland Islands, mainly from the King George Island, including Zielinski (1990),
Quartino et al. (2005), Oliveira et al. (2009), Valdivia et al. (2014), and Gómez et al.
(2019). The macroalgal diversity from Deception Island and Livingston Island was
studied by Ramírez (1982), Clayton et  al. (1997), and Gallardo et  al. (1999).
Medeiros (2013) generated a set of sequences of macroalgal DNA barcodes and
phylogenetic markers from Admiralty Bay (King George Island), contributing to a
molecular database useful for future investigations on the diversity of Antarctic sea-
weeds. Molecular data were obtained for 8 species of Chlorophyta, 9 species of
Phaeophyceae, and 14 species of Rhodophyta. Prasiola sp., Protomonostroma rosu-
latum (Chlorophyta), Chordaria linearis (Phaeophyceae), Acanthococcus antarcti-
cus, and Plumariopsis peninsularis (Rhodophyta) are new records for Admiralty
Bay, and Callophyllis sp. is possibly a new species for science.
28 M. C. Oliveira et al.

Mystikou et  al. (2014) analyzed the seaweed diversity in the Southwestern
Antarctic Peninsula (Adelaide Island) over records of 35 years, reporting 41 spe-
cies. This part of the Antarctic Peninsula is a key region affected by contemporary
climate change, but has been rarely studied. Sanches et al. (2016) performed multi-
variate analyses of Antarctic and sub-Antarctic seaweed distribution patterns focus-
ing on a new evaluation of the role of the Antarctic Circumpolar Current. Regarding
species diversity, in this study the authors identified 129 and 145 macroalgal spe-
cies, respectively, for the Southern Antarctic Circumpolar Front (SACF) and for the
Polar Front (PF) and considering genera, 95 and 101, for the SACF and PF,
respectively.
Pellizzari et al. (2017) recorded a total of 104 species of benthic marine algae
along the South Shetland Islands (28 Phaeophyceae, 24 Chlorophyta, and 52
Rhodophyta), representing 82% of all seaweed taxa described in Antarctica. The
authors also reported nine new records, mainly previously recorded at other lati-
tudes (Protomonostroma rosulatum, Monostroma grevillei, Cladophora coelothrix,
Chaetomorpha irregularis, Dictyota decumbens, Asteronema ferruginea,
Microzonia velutina, Cladodonta lyalli, Rhodophyllis centrocarpa) and two puta-
tive new species of Prasiola sp. and Callophyllis sp. (see also Medeiros 2013).
Spatial variation in the species diversity was observed among the collecting sites:
Livingston Island and King George Island showed the highest diversity. Deception
Island, an area with geothermal activity and intense tourism, was dominated by
opportunistic and broadly distributed filamentous green algae (see also Chap. 5 by
Pellizzari et al.).
A new compilation of the Antarctic seaweed diversity is presented in this chap-
ter, based primarily on data reported by Wiencke and Clayton (2002), Oliveira et al.
(2009), Wiencke et al. (2014), Mystikou et al. (2014), and Pellizzari et al. (2017).
This survey reports a species richness of 151 species, comprising 85 Rhodophyta
(plus one uncertain species), 34 Phaeophyceae, and 32 Chlorophyta to the entire
Antarctic region (Tables 2.1, 2.2, and 2.3, based on Guiry and Guiry (2019) for
taxonomical nomenclature).
Antarctic marine flora can be characterized by a high number of endemic spe-
cies, approximately 33% (Wiencke and Clayton 2002), 35% (Neushul 1968; Wulff
et al. 2009; Wiencke and Amsler 2012), or 39.6% (Skottsberg 1964). However, this
percentage decreases to 27.1% of  species restricted to areas within the Antarctic
Circumpolar Current, i.e., Antarctica and sub-Antarctic islands (with latitude higher
than 55°S) in the compilation of the present chapter (Tables 2.1, 2.2, and 2.3).
Considering the degree of endemism close to 27.1% calculated from a total of 41
endemic species, Rhodophyta represents 61% of endemic taxa, Phaeophyceae
29.3%, and Chlorophyta 9.7%. However, if we consider separately and proportion-
ally (total group richness/group endemic richness), the algal group with the highest
endemism degree is Phaeophyceae (35.3%), followed by Rhodophyta (29.4%), and
Chlorophyta (12.5%).
Large brown seaweeds of the order Laminariales (kelps) are common structuring
species in cold-temperate coasts in the southern and northern hemisphere and also
in the Arctic, while in Antarctica this order is substituted by the Desmarestiales,
2  Diversity of Antarctic Seaweeds 29

Table 2.1  Compiled taxa list of Chlorophyta


South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Acrosiphonia arcta (Dillwyn) Gain 1912 X X X
Blidingia minima (Nägeli ex Kützing) Kylin X X
1947
Chaetomorpha irregularis (Zaneveld) X X
Cormaci, Furnari & Alongi 2014 (as Lola
irregularis) (E)
Chaetomorpha mawsonii Lucas 1919 X X
Cladophora coelothrix Kützing 1843 X
Cladophora repens Harvey 1849 X
Endophyton atroviride O’Kelly in Ricker X X
1987
Entocladia maculans (AD Cotton) X
Papenfuss 1964
Lambia antarctica (Skottsberg) Delépine X X
1967 (E)
Monostroma grevillei (Thuret) Wittrock X
1866
Monostroma hariotii Gain 1911 X X X X
Prasiola sp. (SP428305) X
Prasiola crispa (Lightfoot) Kützing 1843 X X X
Protomonostroma rosulatum X
Vinogradova (E)
Protomonostroma undulatum (Wittrock) X
Vinogradova 1969
Pseudothrix groenlandica (Agardh) Hanic X
& SC Lindstrom 2008 (as Capsosiphon
groenlandicus)
Rhizoclonium ambiguum (Hooker & X
Harvey) Kützing 1849
Rhizoclonium riparium (Roth) Harvey 1849 X
Spongomorpha pacifica (Montagne) X X
Kützing 1854 (as Acrosiphonia pacifica)
Ulothrix australis Gain 1911 (E) X X X
Ulothrix flacca (Dillwyn) Thuret in Le Jolis X X
1863
Ulothrix subflaccida Wille 1901 X
Ulothrix zonata (Weber & Mohr) Kützing X
1833
Ulva sp. (foliose) X
Ulva compressa Linnaeus 1753 X X
Ulva hookeriana (Kützing) HS Hayden, X X X
Blomster, Maggs, Silva, Stanhope &
Waaland 2003 (as Enteromorpha bulbosa)
(continued)
30 M. C. Oliveira et al.

Table 2.1 (continued)
South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Ulva intestinalis Linnaeus 1753 X X
Ulva lactuca Linnaeus 1753 X
Ulva rigida C.Agardh 1823 X
Ulvella leptochaete (Huber) Nielsen, X
O’Kelly & Wysor in Nielsen et al. 2013
Ulvella viridis (Reinke) Nielsen, O’Kelly & X X
Wysor in Nielsen et al. 2013 (as Entocladia
viridis)
Urospora penicilliformis (Roth) Areschoug X X X X
1866
Based on aWiencke and Clayton (2002) and Wiencke et al. (2014), bPellizzari et al. (2017), cOliveira
et al. (2009), and dMystikou et al. (2014)
E = endemic species (restricted to Antarctic and sub-Antarctica)

Table 2.2  Compiled species list of Ochrophyta (Phaeophyceae and Chrysophyceae)


South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Adenocystis utricularis (Bory) Skottsberg X X X X
1907
Antarctosaccion applanatum (Gain) X X X
Delépine (E)
Ascoseira mirabilis Skottsberg 1907 (E) X X X
Asteronema ferruginea (Harvey) X
Delépine & Asensi 1975
Australofilum incommodum (Skottsberg) X
AF Peters 2003
Chordaria linearis (Hooker & Harvey) X X
Cotton 1915
Cystosphaera jacquinotii (Montagne) X X X
Skottsberg 1907 (E)
Desmarestia anceps Montagne 1842 X X X
Desmarestia antarctica Moe & Silva X X X
1989 (E)
Desmarestia chordalis Hooker & Harvey X
1845
Desmarestia confervoides (Bory) X
Ramírez & Peters 1993
Desmarestia menziesii Agardh 1848 (E) X X X X
Dictyota decumbens (Ricker) Hörnig, X
Schnetter & Prud’homme van Reine
1992
(continued)
2  Diversity of Antarctic Seaweeds 31

Table 2.2 (continued)
South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Ectocarpus constanciae Hariot 1887 X
Ectocarpus siliculosus (Dillwyn) X
Lyngbye 1819
Elachista antarctica Skottsberg 1953 (E) X X X X
Geminocarpus austrogeorgiae Skottsberg X X X X
1907 (E)
Geminocarpus geminatus (Hooker & X X X X
Harvey) Skottsberg 1907
Halopteris corymbosa (Dickie) Draisma, X
Prud’homme & Kawai 2010
Halopteris obovata (Hooker & Harvey) X X
Sauvageau 1904
Haplogloia moniliformis RW Ricker X X
1987
Himantothallus grandifolius (A Gepp & X X X X
E Gepp) Zinova 1959 (E)
Leptonematella falklandica (Skottsberg) X
MJ Wynne 1969
Lithoderma antarcticum Skottsberg 1953 X X
(E)
Microzonia australe (Levring) Camacho X
& Fredericq 2018 (as Syringoderma
australe) (E)
Microzonia velutina (Harvey) Agardh X
1894
Petalonia fascia (Müller) Kuntze 1898 X X
Petroderma maculiforme (Wollny) X X
Kuckuck 1897
Phaeurus antarcticus Skottsberg 1907 X X X
(E)
Pylaiella littoralis (Linnaeus) Kjellman X X X X
1872
Ralfsia australis Skottsberg 1921 X X
Scytosiphon lomentaria (Lyngbye) Link X X
1833
Scytothamnus fasciculatus (Hooker & X X
Harvey) Cotton 1915
Utriculidium durvillei Skottsberg 1907 X X
a
Wiencke and Clayton (2002) and Wiencke et al. (2014), bPellizzari et al. (2017), cOliveira et al.
(2009), and dMystikou et al. (2014)
E = endemic species (restricted to Antarctic and sub-Antarctica)
32 M. C. Oliveira et al.

Table 2.3  Compiled taxa list of Rhodophyta


South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Acanthococcus antarcticus Hooker & X X
Harvey 1845 (E)
Ahnfeltia plicata (Hudson) Fries 1836 X
Antarcticothamnion polysporum Moe & X
Silva 1979 (E)
Antarctocolax lambii Skottsberg 1953 (E) X
Ballia callitricha (C Agardh) Kützing 1843 X X X X
Ballia sertularioides (Suhr) Papenfuss 1940 X
Bangia fuscopurpurea (Dillwyn) Lyngbye X X
1819
Callophyllis sp. 1 X
Callophyllis sp. 2 X
Callophyllis atrosanguinea (JD Hooker & X X X
Harvey) Hariot 1887
Callophyllis pinnata Setchell & Swezy Xe
1923
Callophyllis tenera Agardh 1849 X
Callophyllis variegata (Bory) Kützing 1843 X X
Carlskottsbergia antarctica (Hooker & X
Harvey) Athanasiadis 2018 (as
Lithophyllum antarcticum)
Ceramium involutum Kützing 1849 X
Cladodonta lyallii (Hooker & Harvey) X
Skottsberg 1923
Clathromorphum sp. X
Clathromorphum obtectulum (Foslie) Adey X X
1970
Curdiea racovitzae Hariot in De Wildemann X X X X
1900 (E)
Delisea pulchra (Greville) Montagne 1844 X X X
Erythrotrichia carnea (Dillwyn) Agardh X
1883
Falklandiella harveyi (Hooker) Kylin 1956 X
(as Dasyptilon harveyi)
Gainia mollis Moe 1985 (E) X X
Georgiella confluens (Reinsch) Kylin 1956 X X X
(E)
Gigartina skottsbergii Setchell & Gardner X X X
1936
Gracilariopsis longissima (Gmelin) X
Steentoft, Irvine & Farnham 1995 (as
Gracilaria verrucosa)
(continued)
2  Diversity of Antarctic Seaweeds 33

Table 2.3 (continued)
South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Griffithsia antarctica Hooker & Harvey in X
Hooker 1847
Gymnogongrus antarcticus Skottsberg 1953 X X X
(E)
Gymnogongrus turquetii Hariot 1907 (E) X X X
Hildenbrandia lecannellieri Hariot 1887 X X X
Hydrolithon sp. X
Hymenocladia sp. X
Hymenocladiopsis prolifera (Reinsch) X X X
Wynne 2004 (as H. crustigena) (E)
Iridaea cordata (Turner) Bory de Saint-­ X X X X
Vincent 1826
Iridaea mawsonii Lucas 1919∗ X X
Leptophytum coulmanicum (Foslie) Adey X
1970 (E)
Lithothamnion granuliferum Foslie 1905 X X
Meiodiscus concrescens (KM Drew) X
Gabrielson in Gabrielson et al. 2000 (as
Audouinella concrescens)
Mesophyllum sp. X
Microrhinus carnosus (Reinsch) Skottsberg X X
1923
Myriogramme livida (Hooker & Harvey) X
Kylin 1924
Myriogramme manginii (Gain) Skottsberg X X X
1953 (E)
Myriogramme smithii (Hooker & Harvey) X X X
Kylin 1924
Nereoginkgo adiantifolia Kylin in Kylin & X X
Skottsberg 1919 (E)
Neuroglossum delesseriae (Reinsch) Wynne X X X
1997 (as N. ligulatum) (E)
Notophycus fimbriatus Moe 1986 (E) X X X X
Pachymenia orbicularis (Zanardini) Setchell X X X
& Gardner 1934
Palmaria decipiens (Reinsch) Ricker 1987 X X X X
Palmaria georgica (Reinsch) Ricker 1987 X X
Pantoneura plocamioides Kylin 1919 (E) X X X
Paraglossum lancifolium (Agardh) Agardh X X X
(as Delesseria lancifolia)
Paraglossum salicifolium (Reinsch) Showe X X X
in Fredericq & Hommersand 2012 (as
Delesseria salicifolia)
(continued)
34 M. C. Oliveira et al.

Table 2.3 (continued)
South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Peyssonnelia harveyana P Crouan & H X
Crouan ex Agardh 1851
Phycodrys antarctica (Skottsberg) X X X
Skottsberg 1923 (E)
Phycodrys austrogeorgica Skottsberg 1923 X X X
(E)
Phycodrys quercifolia (Bory) Skottsberg X X
1922
Phyllophora abyssalis Skottsberg in Kylin X
& Skottsberg 1919 (E)
Phyllophora ahnfeltioides Skottsberg in X X
Kylin & Skottsberg 1919 (E)
Phyllophora antarctica A Gepp & ES Gepp X X
1905 (E)
Leptophytum foecundum (Kjellmann) Adey X
Phymatolithon lenormandii (Areschoug) X
Adey 1966
Picconiella plumosa (Kylin) De Toni 1936 X X X
Plocamium cartilagineum (Linnaeus) Dixon X X X X
1967
Plocamium secundatum (Kützing) Kützing X X X X
1866
Plocamium hookeri Harvey in Hooker & X X X
Harvey 1845 (E)
Plumariopsis eatonii (Dickie) De Toni 1903 X
Plumariopsis peninsularis Moe & Silva X X
1983 (E)
Polysiphonia abscissa Hooker & Harvey X
1845
Pseudolithophyllum sp. X
Pterothamnion antarcticum (Kylin) Moe & X
Silva 1980
Pterothamnion simile (Hooker & Harvey) X
Nägeli 1862
Ptilonia magellanica (Montagne) Agardh X
1852
Porphyra plocamiestris Ricker 1987 (E) X X X
Porphyra umbilicalis Kützing 1843 X
Porphyra woolhouseae Harvey 1863 X
Pyropia endiviifolia (A Gepp & E Gepp) X X X
Choi & Hwang in Sutherland et al. 2011 (as
Porphyra endiviifolia) (E)
(continued)
2  Diversity of Antarctic Seaweeds 35

Table 2.3 (continued)
South King
Shetland George Adelaide
Species Antarcticaa Islandsb Islandc Islandd
Rhodochorton purpureum (Lightfoot) X
Rosenvinge 1900 (as Audouinella purpurea)
Rhodophyllis centrocarpa (Montagne) X
Wynne
Rhodymenia coccocarpa (Montagne) X X X
Wynne 2007 (as Rhodymenia subantarctica)
Rubrointrusa membranacea (Magnus) X
Clayden & Saunders 2010 (as Audouinella
membranacea)
Sarcodia sp. X
Sarcodia montagneana (Hooker & Harvey) X X
Agardh 1852
Sarcothalia circumcincta (Agardh) X
Hommersand in Hommersand et al. 1993
Sarcothalia papillosa (Bory) Leister in X X
Hommersand, Guiry, Fredericq & Leister
1993
Adapted from aWiencke and Clayton (2002) and Wiencke et al. (2014), bPellizzari et al. (2017),
c
Oliveira et al. (2009), cMystikou et al. (2014), dMystikoy et al. (2014) and eYoneshigue-Valentin
et al. (2013)
E = endemic species (restricted to Antarctic and sub-Antarctica)
*This species is considered uncertain in Algaebase (https://www.algaebase.org/GuiryandGuiry
2019), and therefore was not considered in the calculations for total and endemic species.

e.g., Himantothallus grandifolius and various Desmarestia spp. (Moe and Silva
1977; Clayton 1994). The fact that Laminariales do not reproduce at temperatures
below 0°C and/or the competition with Desmarestiales could explain the absence of
kelps in Antarctica (Peters and Breeman 1993).

2.4  M
 olecular Taxonomy for the Study of Antarctic Seaweed
Diversity

Some Antarctic macroalgae exhibit phenotypic plasticity, which leads to identifica-


tion problems; therefore, it is important to use molecular markers (e.g., DNA bar-
coding) as an additional tool to gain accuracy in the classification (Medeiros 2013).
However, there are some difficulties in the applicability of this technique, due to the
occurrence of variability in the region of the primers, especially for cox1 DNA bar-
code, and the lack of effectiveness of this marker in the case of Chlorophyta. The
proposal of a quick, easy, and low-cost method for inventory of seaweed biodiver-
sity shows advantages over other techniques, since standardization of a DNA bar-
36 M. C. Oliveira et al.

code in the identification and report of new records or new species facilitates the
exchange of information among laboratories around the world (Le Gall and
Saunders 2010).
Medeiros (2013) has successfully used three DNA barcodes for the identification
of the macroalgal assemblages of Admiralty Bay, despite some limitations. The
three used DNA barcodes (UPA, cox1, and tufA), although presented different rates
of divergence, were consistent with other markers used in the analyses. The main
advantage found in the amplification of the UPA marker (Sherwood et al. 2010) was
the universality of the primers, since a single pair of primers was used for the three
seaweed groups, whereas different combinations of primers were required to
amplify the cox1. However, the low variability in UPA sequences may underesti-
mate the species diversity, while cox1 presents a higher level of divergence, being
more suitable as a specific marker. The absence of universal primers can limit the
amplification in some groups of algae. In other instance, tufA showed to be very
promising for green algae, since it was possible to amplify this marker with a single
set of primers and the levels of divergence found were relatively higher than for
UPA and comparable to those found for cox1 in red and brown algae. Finally, the
use of the UPA can be a fast and efficient tool for biodiversity monitoring, mainly
of cryptogenic and cryptic seaweed species from Antarctica. Medeiros (2013) con-
clude that due to the analysis of only a few specimens per species in Admiralty Bay,
it was not possible to establish a reliable limit between the values of intraspecific
and interspecific divergences (barcode gap) for the distinct genera, which according
to Meier et al. (2008) is necessary for successful species identification.
Thus, further studies are needed to establish these limits in order to use the DNA
barcoding technique extensively in the identification and surveys of macroalgae
diversity from Antarctica. Although larger molecular markers such as rbcL, ITS and
SSUrDNA, which have more sequences available for comparison in global ­database,
are relevant to obtain. In addition, the use of GenBank data for molecular species
identification purposes is not totally reliable, since accurate species identification is
not always guaranteed and identification based on morphological characters is still
indispensable.
Seaweed species from Admiralty Bay, King George Island, exhibit affinities with
seaweeds from South America and New Zealand, as well as with cold and polar
regions from the northern hemisphere. In Hommersand et  al. (2009), rbcL data
showed that the affinity between the Antarctic and South American Rhodophyta
species is complex, since the species groups are phylogenetically distant. Still,
according to Hommersand et al. (2009), there is a possibility that particularities in
the Antarctic climate caused a faster evolution of the rbcL gene, compared to tem-
perate waters from South America. This fact may justify the high divergence found
by Medeiros (2013) in sequences of rbcL for the species Iridaea cordata and
Plocamium aff. cartilagineum obtained in Antarctica, when compared to the same
species sampled in Chile. Finally, the author demonstrated that the DNA barcoding
tool, together with other markers, proved to be a very suitable approach for large-­
scale application in biodiversity and conservation studies, providing information for
2  Diversity of Antarctic Seaweeds 37

global database that would combine molecular, morphological, and distributional


data (see also Chap. 5 by Pellizzari et al.).

2.5  Seaweed Distribution in Antarctica

Seaweeds are conspicuous components of subtidal benthic communities; however,


their abundance and diversity vary strongly depending on the habitat characteristics
(Klöser et al. 1996). In many cases the hard-bottom substrata are often covered by
crustose red algae (e.g., Hildenbrandia lecannellieri and species of the order
Corallinales). Pebbles and boulders of granite, as well as volcanic outcrops, often
form the hard substrate. These formations are common on King George Island,
Penguin Island, Nelson Island, and Robert Island. Elephant Island, Livingston
Island, and Half Moon Island are formed of volcanic bedrocks and of pebbles in the
intertidal and subtidal zones. The slope and exposure to waves, ice melting, and
abrasion vary among Antarctic islands and the continental coast influencing directly
the recruitment of seaweeds (Pellizzari et al. 2017).
Large brown algae of the order Desmarestiales (e.g., Desmarestia anceps,
D. menziesii, and Himantothallus grandifolius) are dominant in the benthic com-
munities of Western Antarctica, growing in substrata below the zone affected by ice
scouring down to ≥30-m water depth (Wiencke and Clayton 2002). On the other
hand, red algae are dominant in the benthic communities at higher latitudes, as in
the Ross Sea. Many macroalgae may survive, after detached from the original sub-
strate, during long periods as free-floating thalli. Recent studies (synopsis in Fraser
et  al. 2018) report that these robust free-floating specimens may also transport
hitchhiker species by rafting, including smaller epiphyte seaweed species and sev-
eral marine invertebrates. These floating seaweed masses anchor other species,
being recognized as vectors of introduction of alien/cryptogenic taxa around
Antarctica (see also Chap. 3 by Fraser et al. and Chap. 4 by Macaya et al.).
Intertidal seaweeds are filamentous (e.g., the green algae Urospora penicillifor-
mis and Ulothrix spp.), foliose (e.g., Monostroma hariotii, Pyropia endiviifolia), or
saccate (e.g., Adenocystis utricularis), or possess perennial basal attachments (e.g.,
Palmaria decipiens). Intertidal species are well developed in the milder climates of
the Antarctic maritime islands, especially in the South Shetland and the South
Orkney archipelagos (Wiencke and Clayton 2002; Oliveira et  al. 2009). These
organisms display a suite of morpho-functional adaptations to cope with a high vari-
ability of physical factors, such as temperature, ice scour, light, salinity, etc. (Gómez
et al. 2019, see also Chap. 11 by Gómez and Huovinen).
In the Maritime Antarctica, which compress the Antarctic Peninsula and nearby
islands, the substrate availability, milder climate, and less marked seasonality per-
mit the colonization and persistence of a higher diversity of marine organisms
(Amsler et  al. 1995). Here, more than 90% of all Antarctic species have been
reported (Clayton 1994). Species richness decreases in the East Antarctic Peninsula
(EAP), and only 7 and 17 species occurred, respectively, on the coasts of the Ross
38 M. C. Oliveira et al.

Sea (Zaneveld 1966) and Terra Nova Bay (Cormaci et  al. 1992). Generally, an
inverse relationship between species diversity and latitude is observed in Antarctic
seaweeds (Wiencke and Clayton 2002). A total of 104 taxa was identified in South
Shetland Islands (ca. 60°S) by Pellizzari et al. 2017, a species number higher than
in Adelaide Island (67°S) with 41 taxa (Mystikou et al. 2014) and Terra Nova Bay
(Ross Sea, above latitude 70°S) with 17 taxa (Cormaci et al. 1992). Only few spe-
cies grow in latitudes above 76°S, such as the red algae Iridaea cordata, Phyllophora
antarctica, Phycodrys antarctica, and Hildenbrandia lecannellieri, the green alga
Monostroma hariotii, and the brown alga Desmarestia menziesii (Wiencke and
Clayton 2002).

2.6  Concluding Remarks: Gaps and Prospects for the Future

The knowledge on the seaweed diversity throughout Antarctica is essential, as they


are fundamental as primary producers and their composition affects the structure of
the ecosystems. Species of seaweeds are bioindicators of environmental changes,
including climate change, invasive species, and seawater pollution, among others.
Therefore, knowledge on seaweed diversity throughout Antarctica is urgently needed.
There are some clear gaps in the knowledge on seaweed diversity from the
Antarctic region, especially in deeper water. As described above, species richness
tends to increase with the number of collections in a wide range of locations and in
different environments, but collection in Antarctica is expensive and difficult.
Furthermore, as mentioned above, little is known on the diversity fluctuations of the
marine flora and how those are influenced by environmental factors or anthropo-
genic activities. Regular monitoring of sites should be done, especially of those that
are subjected to human activities such as those from scientific bases or touristic visi-
tation (Hughes and Ashton 2017).
In Antarctica, some of the strongest environmental gradients on the planet can be
found, therefore providing an ideal study ground to test hypotheses related with
environmental variability and its impact on biodiversity. The most important physi-
cal driver of Antarctic marine communities is the oceanographic boundary of the
Polar Front. At smaller spatial scales, ice cover, ice scour, and salinity gradients are
clearly important determinants of diversity at habitat and community level. However,
stochastic and extreme events remain an important driving force, particularly in the
context of local extinction, colonization, or recolonization of Antarctic biota
(Convey et al. 2014).
Biogeographic barriers are known to be fundamental in macroecological and
evolutionary processes. Ocean circulation, considering present and past patterns of
continental drift, can isolate or connect many groups of marine organisms, includ-
ing seaweeds (see also Chap. 3 by Fraser et al. and Chap. 5 by Pellizzari et al.).
These benthic organisms release spores and propagules as planktonic stages that
drift with currents and/or tides and have been a sensible indicator of changes in
biogeographic distribution patterns (Sanches et  al. 2016, see also Chap. 10 by
2  Diversity of Antarctic Seaweeds 39

Navarro et al.). According to Pellizzari et al. (2017) and Fraser et al. (2018), the
Antarctic Circumpolar Current, previously considered a biogeographic barrier, may
become a new pathway for biota interconnectivity and deserves further investiga-
tion. Mystikou et al. (2014) discuss that Antarctic seaweeds display plasticity and
adaptability in response to extreme environmental conditions such as low tempera-
tures and limited light availability (Wiencke and Amsler 2012). Thus, it is relevant
to examine how environmental alterations, such as those caused by climate change,
will be affecting algal seasonality, richness, depth zonation, and latitudinal
distribution.
The higher seaweed diversity observed in the South Shetland Islands (Pellizzari
et al. 2017), a transitional area, bring the imminent need of long-term biological and
abiotic monitoring in order to establish conservation guidelines across the Antarctic
and sub-Antarctic zones, especially upon increasing tourism, global climate, and
oceanographic changes.
Moreover, laboratory culture techniques associated with morphological and
molecular analyses could reveal the hidden diversity of Antarctic marine flora, espe-
cially with respect to small species (e.g., epiphytes and endophytes), including
potential alien/cryptogenic species. Integrative taxonomical studies are needed to
less known groups, e.g., crustose algae, and molecular studies are essential to clar-
ify the phylogenetic and biogeographic relationships of Antarctic seaweeds.
Finally, a reference baseline database of seaweed diversity is urgently necessary.
This should integrate different inventories based on morphology and life cycles with
molecular data from fast, high-throughput, and low-cost methods, such as DNA
barcodes and/or metagenomics (Oliveira et  al. 2018). Hence, researchers will be
enabled to effectively monitor environmental changes and help in the conservation
of this unique environment.

Acknowledgments We thank PROANTAR (Brazilian Antarctic Program 557030/2009-9,


407588/2013-2 and 442258/2018-2), INCT Criosfera 2, Brazilian Navy (Polar Ship Almirante
Maximiano–H41), Brazilian Air Force, MMA (Ministry of Environment), MCTIC (Ministry of
Science, Technology and Innovation), and CNPq (National Council of Research and Development;
for grant 310672/2016-3 to NSY and 301491/2013-5 to MCO). This study was financed in part by
the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior – Brasil (CAPES) – Finance
Code 001.

References

Amsler CD, Rowley RJ, Laur DR, Quetin LB, Ross RM (1995) Vertical distribution of Antarctic
peninsular macroalgae: cover, biomass and species composition. Phycologia 34:424–430
Barker PF, Thomas E (2003) Origin, signature and palaeoclimatic influence of the Antarctic
Circumpolar Current. Earth Sci Rev 66:143–162
Clarke DL (1990) Arctic Ocean ice-cover; geologic history and climatic significance. In: Grantz
A, Johnson L, Sweeney JL (eds) The Arctic Ocean region. The Geological Society of America,
Boulder, CO, pp 53–62
Clarke A, Barnes KA, Hodgson DA (2005) How isolated is Antarctica. Trends Ecol Evol 20:1–3
40 M. C. Oliveira et al.

Clayton MN (1994) Evolution of Antarctic benthic algal flora. J Phycol 30:897–904


Clayton MN, Wiencke C, Klöser H (1997) New records of temperate and sub-Antarctic marine
benthic macroalgae from Antarctica. Polar Biol 17:141–149. https://doi.org/10.1007/
s003000050116
Convey P, Chown SL, Clarke A, Barnes DKA, Schiaparelli S, Wall DH (2014) The spatial structure
of Antarctic biodiversity. Ecol Monogr 84:203–244
Cormaci M, Furnari G, Scammacca B (1992) The benthic algal flora of Terra Nova Bay (Ross Sea,
Antarctica). Bot Mar 35:541–552
Drew EA, Hastings RM (1992) A year-round ecophysiological study of Himantothallus grandifo-
lius (Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31:262–277
Fraser CI, Morrison AK, McC Hogg A, Macaya EC, Van Sebille E, Ryan PG, Padovan A, Jack
C, Valdivia N, Waters JM (2018) Antarctica’s ecological isolation will be broken by storm-­
driven dispersal and warming. Nat Clim Change Lett 8:704–708. https://doi.org/10.1038/
s41558-018-0209-7
Gallardo T, Pérez-Ruzafa IM, Flores-Moya A, Conde F (1999) New collections of benthic marine
algae from Livingston and Deception Islands (South Shetland Islands) and Trinity Island
(Bransfield Strait), Antarctica. Bot Mar 42:61–69. https://doi.org/10.1515/bot.1999.009
Gaudichaud C (1826) Botanique. In: Freycinet I (ed) Voyage au tur de monde…executé sur les
corvettes de SM l’Uranie et la Physicienne pendant les années 1817–1820. Pillet Ainé, Paris
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27
Guiry MD, Guiry GM (2019) AlgaeBase. World-wide electronic publication. National University
of Ireland, Galway. http://www.algaebase.org, searched on 24 August 2019
Hempel GV (1987) Die Polarmeere – ein biologischer Vergleich. Polarforschung 57:173–189
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52:509–534. https://doi.org/10.1515/bot.2009.081
Hughes KA, Ashton GV (2017) Breaking the ice: the introduction of biofouling organisms to
Antarctica on vessel hulls. Aquat Conserv Mar Freshw Ecosyst 27:158–164
Klöser H, Quartino ML, Wiencke C (1996) Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiologia 333:1–17
Lamb I, Zimmermann MH (1977) Benthic marine algae of the Antarctic Peninsula. In: Pawson DL
(ed) Biology of the Antarctic Seas V. American Geophysical Union, Antarctic Research Series.
23 (4), United States. 229 p. ISSN: 0066-4634
Le Gall L, Saunders GW (2010) DNA barcoding is a powerful tool to uncover algal diversity: a
case study of the Phyllophoraceae (Gigartinales, Rhodophyta) in the Canadian flora. J Phycol
46:374–389
Medeiros AS (2013) Macroalgae diversity of Admiralty Bay, King George Island, Antarctic
Peninsula based on DNA barcoding and other molecular markers. MSci Thesis, São Paulo
University, p. 164. http://www.teses.usp.br/teses/disponiveis/41/41132/tde-24032014-090801/
pt-br.php
Meier R, Zhang G, Ali F (2008) The use of mean instead of smallest interspecific distances exag-
gerates the size of the “Barcoding Gap” and leads to misidentification. Syst Biol 57:809–822
Moe RL (1985) Gainia and Gainiaceae: a new genus and family of crustose marine Rhodophyceae
from Antarctica. Phycologia 24:419–428
Moe RL, Silva PC (1977) Antarctic marine flora: uniquely devoid of kelps. Science 196:1206–1208
Mystikou A, Peters AF, Asensi AO, Fletcher KI, Brickle P, Van West P et al (2014) Seaweed bio-
diversity in the south-western Antarctic Peninsula: surveying macroalgal community com-
position in the Adelaide Island/Marguerite Bay region over a 35-year time span. Polar Biol
37:1607–1619. https://doi.org/10.1007/s00300-014-1547-1
2  Diversity of Antarctic Seaweeds 41

Nedzarek A, Rakusa-Suszcewski S (2004) Decomposition of macroalgae and the release of nutri-


ents in Admiralty Bay, King George Island, Antarctica. Polar Biosci 17:16–35
Neushul M (1959) Biological collecting in Antarctic waters. Veliger 2:15–17
Neushul M (1961) Diving in Antarctic water. Polar Rec 10:83–88
Neushul M (1963) Reproductive morphology of Antarctic kelps. Bot Mar 5:19–24
Neushul M (1965) Diving observation of sub-tidal Antarctic marine vegetation. Bot Mar 8:234–243
Neushul M (1968) Benthic marine algae. Antarctic Map Folio Ser 10:9–10
Oliveira EC, Absher TM, Pellizzari FM, Oliveira MC (2009) The seaweed flora of Admiralty
Bay, King George Island, Antarctic. Polar Biol 32:1639–1647. https://doi.org/10.1007/
s00300-009-0663-9
Oliveira MC, Repetti SI, Iha C, Jackson CJ, Díaz-Tapia P, Lubiana KMF, Cassano V, Costa JF,
Cremen MCM, Marcelino VR, Verbruggen H (2018) High-throughput sequencing for algal
systematics. Eur J Phycol 53:256–272. https://doi.org/10.1080/09670262.2018.1441446
Papenfuss GF (1964) Catalogue and bibliography of Antarctic and Subantarctic benthic marine
algae. Antarct Res Ser 1:1–76
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS et al (2017) Diversity
and spatial distribution of seaweeds in the South Shetland Islands, Antarctica: an updated data-
base for environmental monitoring under climate change scenarios. Polar Biol 40:1671–1685.
https://doi.org/10.1007/s00300-017-2092-5
Peters AF, Breeman AM (1993) Temperature tolerance and latitudinal range of brown algae from
temperate Pacific South America. Mar Biol 115:143–150
Quartino ML, Zaixso HE, Boraso de Zaixso AL (2005) Biological and environmental characteriza-
tion of marine macroalgal assemblages in Potter Cove, South Shetland Islands, Antarctica. Bot
Mar 48:187–197. https://doi.org/10.1515/BOT.2005.029
Ramírez ME (1982) Catálogo de las algas marinas del Territorio Chileno Antártico. Ser Cient
INACh 29:39–67
Sanches PF, Pellizzari FM, Horta PH (2016) Multivariate analyses of Antarctic and sub-Antarctic
seaweed distribution patterns: an evaluation of the role of the Antarctic Circumpolar Current. J
Sea Res 110:29–38. https://doi.org/10.1016/j.seares.2016.02.002
Sherwood AR, Kurihara A, Conklin KY, Sauvage T, Presting GG (2010) The Hawaiian
Rhodophyta biodiversity survey (2006–2010): a summary of principal findings. BMC Plant
Biol 10(258):1–29
Skottsberg C (1906) Observations on the vegetations of the Antarctic Sea. Botaniska Studier till-
agnade F.R. Kjelman, Uppsala, pp 245–264
Skottsberg C (1941) Communities of marine algae in subantarctic and Antarctic waters. K Sven
Velenslapsakad Handl Tredje Serien 19:1–92
Skottsberg C (1953) On two collections of Antarctic marine algae. Arkiv Botanik 2:531–566
Skottsberg C (1964) Antarctic phycology. In: Carrick R, Holgate M, Prevost J (eds) Biologie
Antarctique. 1st SCAR Symposium Hermann, Paris, pp 147–154
Skottsberg C, Neushul M (1960) Phyllogigas and Himantothallus. Antarctic Phaeophyceae Bot
Mar 2:164–173
Valdivia N, Diaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9(6):e100714. https://doi.org/10.1371/journal.
pone.0100714
Vincent WF (2000) Evolutionary origins of Antarctic microbiota: invasion, selection and ende-
mism. Antarct Sci 12:374–385
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Wiencke C,
Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and utilization.
Springer, Germany, pp 265–291
Wiencke C, Clayton MN (2002) Antarctic seaweeds. In: Wägele JW (ed) Synopses of the Antarctic
benthos, vol 9. A.R.G. Gantner Verlag, Rugell, Lichtenstein, p 239
42 M. C. Oliveira et al.

Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: DeBroyer C, Koubbi
P, Griffiths HJ, Raymond B, Udekem d’Acoz CD (eds) Biogeographic atlas of the Southern
Ocean. Scientific Committee on Antarctic Research, Cambridge, pp 66–73
Wulff A, Iken K, Quartino ML, Al-Handal A, Wiencke C, Clayton MN (2009) Biodiversity, bio-
geography and zonation of marine benthic micro–and macroalgae in the Arctic and Antarctic.
Bot Mar 52:491–507. https://doi.org/10.1515/bot.2009.072
Wynn-Williams DD (1996) Response of pioneer soil microalgal colonists to environmental change
in Antarctica. Microb Ecol 31:177–188
Yoneshigue-Valentin Y, Silva IB, Fujii MT, Yokoya NS, Pupo D, Guimarães SMPB, Martins AP,
Sanches PF, Pereira DC, Dalto AG, Souza JMC, Pereira CMP, Pellizzari F, Colepicolo P (2013)
Marine macroalgal diversity in Admiralty Bay, King George Island, South Shetlands Islands,
Antarctica. Ann Act Rep:140–148
Zaneveld JS (1966) Vertical zonation of Antarctic and Subantarctic benthic marine algae. Antarct
J US 1:211–213
Zielinski K (1990) Bottom macroalgae of Admiralty Bay (King George Island, South Shetland
Island, Antarctic). Polar Res 11:95–131
Chapter 3
Biogeographic Processes Influencing
Antarctic and sub-Antarctic Seaweeds

Ceridwen I. Fraser, Adele Morrison, and Pamela Olmedo Rojas

Abstract  Antarctica has long been seen as biologically isolated, surrounded by the
vast Southern Ocean and its circumpolar oceanographic currents and fronts and
home to many endemic species. New evidence demonstrates, however, that buoyant
seaweeds can cross perceived oceanographic barriers in the Southern Ocean to
reach Antarctic coasts. These macroalgal rafts can carry diverse passengers, includ-
ing marine invertebrates and other, non-buoyant seaweeds. The stark differences
between Antarctic and sub-Antarctic near-coastal ecosystems are therefore more
probably the result of environmental differences than physical isolation. Modelling
indicates that algal rafts from the sub-Antarctic could reach Antarctic coasts every
month, providing an ongoing influx of marine propagules that are poised to colonise
as the climate warms. In this chapter, we review the following: (i) the evidence for
the isolation of Antarctica, (ii) the oceanographic processes that can hinder or pro-
mote passive dispersal into Antarctic waters and (iii) the characteristics of organ-
isms that could be rafting to Antarctic coasts with buoyant macroalgae.

Keywords  Antarctic circumpolar current · Connectivity · Dispersal · Polar front ·


Stokes drift

C. I. Fraser (*) · P. Olmedo Rojas


Department of Marine Science, University of Otago, Dunedin, New Zealand
A. Morrison
Research School of Earth Sciences, and ARC Centre of Excellence
for Climate Extremes, Australian National University, Acton, ACT, Australia
e-mail: adele.morrison@anu.edu.au

© Springer Nature Switzerland AG 2020 43


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_3
44 C. I. Fraser et al.

3.1  Antarctica’s Place in the World: An Isolated Continent?

Antarctica has long been seen as biologically isolated from the rest of the world
(Clarke et al. 2005; Fraser et al. 2012; Convey et al. 2014; Chown et al. 2015). The
opening of the Drake Passage between South America and Antarctica, around
41  million years ago (Scher and Martin 2006), and the subsequent onset of the
Antarctic Circumpolar Current (AAC), around 30  Ma (Scher et  al. 2015), geo-
graphically isolated Antarctica. Along with declines in atmospheric CO2 in the
Cenozoic, the onset of the ACC enhanced cooling and glaciation of the Antarctic
(DeConto and Pollard 2003), resulting in marked reductions in diversity in some
marine groups (Clarke 1990). The ACC is the world’s largest ocean current, about
23,000 km long and up to 2000 km wide in some areas (Constantin and Johnson
2016); connects the Indian, Pacific and Atlantic Oceans; and extends up to 4–5-km
depth (Marynets 2019). There are several well-recognised circumpolar thermal
fronts in the Southern Ocean, including the Subantarctic Front and the Antarctic
Polar Front (see next section). These oceanic features, encircling Antarctica,
enhance the stark environmental differences between Antarctic and more northern
(sub-Antarctic or temperate) ecosystems. Biologically, the Southern Ocean with
its strong, eastward-­flowing ACC and circumpolar fronts represents a major hur-
dle for some organisms to cross (Patarnello et  al. 1996; Fraser et  al. 2012;
Gonzalez-Wevar et  al. 2012), particularly for passive dispersers. The observed
high-level endemism in the Southern Ocean marine biota (Barnes et al. 2006) and
the relatively low diversity and high endemism of terrestrial organisms in
Antarctica have largely been considered to result from long periods of evolution in
isolation (Fraser et al. 2012). However, we are increasingly recognising that the
Southern Ocean is not an impermeable biological barrier; dispersal events into the
Antarctic do occur (Clarke et al. 2005; Barnes et al. 2006) and at higher frequen-
cies than previously thought (Fraser et al. 2017; Fraser et al. 2018b). The distinc-
tive Antarctic biota is thus probably a result of adaptation to extreme environmental
conditions, rather than an inability of passively dispersing organisms to reach
the region.

3.1.1  A
 daptations of Terrestrial Organisms to Antarctic
Conditions

Only around 0.2–0.3% of Antarctic land is currently ice-free (Convey and Stevens
2007; Burton-Johnson et al. 2016). Antarctic terrestrial organisms, restricted to
this ice-free land, have apparently persisted in Antarctica for millions of years
and have adapted to the extreme conditions including low temperatures, low win-
ter photoperiod and scarce food supply (Convey and Stevens 2007). Antarctic
terrestrial biodiversity is restricted mainly to microinvertebrates, bryophytes and
lichens; only two vascular plants occur, both restricted to the maritime Antarctica
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 45

(Ochyra et al. 2008). Lichens and mosses are the dominant flora of Antarctica –
there are ca 386 and 111 species of lichens and mosses identified, respectively
(Øvstedal and Lewis Smith 2001; Ochyra et  al. 2008). Lichens have been
described as the organisms best adapted to Antarctic conditions (Ochyra et  al.
2008) and have been reported to be photosynthetically active under suboptimal
temperatures (Kappen 2000). Antarctic mosses have also shown remarkable
physiological adaptations to both water ­availability and solar radiation. For
example, mosses can survive under long periods of both desiccation and submer-
gence (Wasley et  al. 2006), including being frozen by glaciers for hundreds
(Cannone et al. 2017) and even thousands (Roads et al. 2014) of years, despite
water availability having been shown to be a key factor underpinning the abun-
dance and distribution of Antarctic terrestrial organisms (Kennedy 1993; Convey
and Stevens 2007; Robinson et al. 2018). The Antarctic terrestrial fauna is domi-
nated by microarthropods, such as springtails and mites (Convey and Stevens
2007). Glycerol in cell membranes might help arthropods inhabiting polar condi-
tions to survive freezing temperatures (Teets and Denlinger 2014). Additionally,
the small size of terrestrial organisms could assist with minimising moisture loss
in windy environments and maximising the use of limited nutrients (Kappen
et al. 1995).

3.1.2  A
 daptations of Marine Organisms to Antarctic
Conditions

The contemporary marine biota of Antarctica has been strongly shaped by glacial
cycles (Allcock and Strugnell 2012; Fraser et al. 2012) and mainly comprises fish,
macroalgae and filter feeders such as sponges, small crustaceans, molluscs and
anemones (Clayton 1994; Griffiths 2010). Some taxa such as decapods, sharks and
skates are poorly represented in Antarctica (Aronson and Blake 2015). Up to 90%
of the Antarctic fishes and marine invertebrates (Barnes et al. 2006) and 35% of
macroalgae (Clayton 1994; Gómez 2015) are endemic to the region. Whereas some
Antarctic algal species are also present in nearby northern areas (such as Adenocystis
utricularis, Gigartina skottsbergii, Monostroma hariotii or Iridaea cordata, which
are found in both Antarctica and South America: Gómez 2015), others are unique
to the region but are closely related to taxa elsewhere. Sunlight is highly limited in
the Antarctic marine environment, but Antarctic macroalgae can survive at low
temperatures, photosynthesise under very low light conditions and can store
organic compounds to use in dark periods (Wiencke et al. 2007). Other marine spe-
cies also show adaptations to their extreme environment. Antarctic notothenioid
fish have developed antifreeze glycoproteins, which lower the internal freezing
point of most of their fluids, preventing freezing in sub-zero Antarctic waters
(DeVries 1988).
46 C. I. Fraser et al.

3.1.3  Evidence for Dispersal of Organisms into the Antarctic

Transoceanic dispersal mainly occurs either by active dispersal (e.g. swimming and
flying) or through passive dispersal such as through transport with ocean or wind
currents, or ‘hitch-hiking’ with larger animals, driftwood or seaweed (Muñoz et al.
2004; Gillespie et  al. 2012; Moon et  al. 2017). Some microbes appear to have
reached Antarctica via aerial dispersal with wind (Vincent 2000), and small arthro-
pods also appear capable of dispersing with wind (Hawes et  al. 2007), although
long-distance dispersal of arthropods across the Southern Ocean via this mechanism
seems unlikely (Pugh 2003). Recently, Fraser et al. (2018b) reported that southern
bull kelp, Durvillaea antarctica, which grows in the sub-Antarctic but not in the
Antarctic, had travelled south across the Southern Ocean. These were the longest
biological rafting events ever recorded, >20,000 km, and were apparently driven by
strong winds and storms that pushed the kelp across Southern Ocean fronts (see
below). Similarly, in the last couple of decades, king crabs (Neolithodes yaldwyni)
have been found on the Antarctic continental shelf (Smith et al. 2012) and might
represent invasions from deeper water, although there has been some debate as to
whether the crabs are new immigrants (Thatje et al. 2005) or have been long-term
residents of the Antarctic and have simply gone undetected (Griffiths et al. 2013).
With global warming, many organisms are migrating poleward (Hickling et al.
2006; Chen et al. 2011; Fraser et al. 2012). Invasive (anthropogenically transported)
and naturally dispersing non-native species represent a major threat to Antarctic
biota (Frenot et al. 2005; Chown et al. 2012; Chown et al. 2015; Duffy et al. 2017).
Understanding how permeable the Southern Ocean ‘barrier’ is to dispersal of differ-
ent sorts of organisms will help us to understand how Antarctic biodiversity might
be affected by future colonisations (Fraser et al. 2017; Wauchope et al. 2019).

3.2  P
 hysical Oceanographic Processes Influencing
Movement of Seaweeds into or out of the Antarctic

There has been a long-held perception that the fronts of the ACC act as oceano-
graphic barriers preventing movement of passively dispersing marine biota south-
ward into Antarctic waters (Hunt et  al. 2016). Fronts are sharp transitions in
temperature, salinity and biogeochemical properties and align with the narrow, east-
ward currents of the ACC. Traditionally, three primary fronts have been identified in
the Southern Ocean  – the Subantarctic Front, the Polar Front and the Southern
Antarctic Circumpolar Current Front (Orsi et  al. 1995). Traversing southward
towards Antarctica, a front appears as a near step change from warmer, saltier waters
in the north to colder, fresher waters on the southern side. The separation of warm
and cold waters across the ACC fronts has cultivated the conceptual picture that
there is very little mixing or transport across the fronts. However, this is an overly
simplistic view of circulation in the Southern Ocean. Recent analysis of higher-­
resolution observations shows that the ACC has a more complicated structure with
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 47

numerous fronts that merge and split around Antarctica (Sokolov and Rintoul 2009).
Mesoscale variability is also rich in the Southern Ocean, and eddies and jets that
meander in time can move objects away from the direction of the main currents
(Lehahn et  al. 2011). An additional process that has recently been recognised as
extremely important for dispersal of floating surface material is non-linear a­ dvection
by surface waves, known as Stokes drift (Fraser et al. 2018b; Dobler et al. 2019).
Below we expand upon the most important processes influencing movement of sea-
weeds in the Southern Ocean: the mean northward drift of Ekman transport, tran-
sient north and south movement by eddies and storm-driven Stokes drift, in addition
to the predominantly zonal movement by the large-scale horizontal ocean currents.

3.2.1  Ekman Transport

The westerly winds over the Southern Ocean, also known as the ‘Roaring Forties’,
are the strongest average winds on the planet, with annual average speeds up to
10 m s−1 (Lin et al. 2018). The winds blow from west to east over a wide latitude
band covering approximately 30–65°S. Due to the Coriolis effect, which deflects
motion to the left in the Southern Hemisphere, the eastward wind stress on the
ocean drives a northward surface movement referred to as Ekman transport. Ekman
transport in the Southern Ocean extends to a depth of around 100 m and is surface
intensified. Observed northward speeds are up to ~2 cm s-1 at the surface (Lenn and
Chereskin 2009), which in isolation would result in floating objects moving more
than 600 km, or 6° of latitude northward in 1 year. Although the Southern Ocean
fronts provide a visible separation between the sub-Antarctic islands and Antarctica,
it is the constant northward drift of Ekman transport that represents the biggest
obstacle to southward movement of floating seaweeds.

3.2.2  Eddies

The Southern Ocean fronts are dynamically unstable, forming ubiquitous mesoscale
eddies and transient meanders in the ACC jets (Thompson Andrew 2008). Drifting
objects can be trapped inside coherent ring-like eddies and transported over long
distances away from the large-scale ocean currents (Lehahn et  al. 2011). In the
Southern Ocean, eddies are ~10–100 km across and commonly last for longer than
4 months, over which time they can travel north or south by 5° or more of latitude
(Chelton et al. 2011). A drifting object would be unlikely to remain trapped for the
entire lifetime of an eddy, as eddies ‘leak’ as they stretch and interact with other
ocean circulation features (d’Ovidio et al. 2013). However, seaweeds may encoun-
ter a series of eddies over time, resulting in a net northward or southward transport.
Fraser et al. (2018b) showed that the inclusion of mesoscale variability was essen-
tial for modelled virtual particles to drift sufficiently southward to reach Antarctica.
48 C. I. Fraser et al.

3.2.3  Wave-Driven Stokes Drift

High wind speeds associated with atmospheric storms over the Southern Ocean
generate an intense wave climate and make it consistently the roughest ocean on
earth (Young 1999). The non-linear nature of surface ocean waves results in a net
advection of floating objects in the direction of the waves, known as Stokes drift.
While the average direction of waves in the Southern Ocean is eastward, the same
as the average wind direction, atmospheric storms result in large variability of the
wind direction and associated wave direction. Seaweeds can be transported south-
wards across fronts by large, individual storm events, during which the Stokes drift
is frequently as large as 0.3 m s−1 (Rascle and Ardhuin 2013). Fraser et al. (2018b)
showed that the influence of wave-driven Stokes drift is necessary for floating
objects to drift sufficiently southward to reach Antarctica (Fig. 3.1).

3.2.4  Surface Currents

The large-scale horizontal circulation in the open Southern Ocean is predominantly


eastward and therefore cannot in isolation drive floating objects southward.
Nonetheless, the ocean circulation is important to consider, as it largely determines
the locations where seaweeds approach Antarctica. While eddy processes and

Fig. 3.1  Simulated drift particle trajectories from South Georgia, as described in Fraser et  al.
(2018a, b). More than 4.2 million virtual particles were released from South Georgia (red dot) and
advected for 3 years with simulated surface ocean velocities where (a) eddies, but not Stokes drift,
were included; (b) Stokes drift, but not eddies, was included; and (c) both Stokes drift and eddies
were included. Blue lines show trajectories for all particles located on the Antarctic shelf after
3 years. Brown lines show trajectories for a random sample (0.1%) of particles not satisfying this
condition. Black lines show positions of major fronts. Red arrows show the four major pathways
of approach to Antarctica, two aligned with southward deviations of the ACC and two associated
with the southward movement of the Weddell and Ross Gyres
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 49

Stokes drift are responsible for moving floating objects southward across the ACC
fronts, the large-scale meanders of the ACC and the subpolar gyres set the most
likely pathways of approach to Antarctica (Fraser et al. 2018b).
The eastward ACC dominates the ocean circulation in the Southern Ocean. The
ACC plays an important role in dispersing species zonally around the Southern
Ocean (Waters 2008) and in rapidly moving drifting objects to longitudes where
shorter routes to Antarctica are possible. The circumpolar path of the ACC is steered
by the underlying bathymetry and continental gateways. There are two close
approaches of the ACC to the Antarctic continental slope – along the West Antarctic
Peninsula and offshore from Wilkes Land in East Antarctica. In the recent Lagrangian
study of Fraser et al. (2018b), these two locations were preferred routes for south-
ward drifting particles to first make contact with the continent.
Two additional locations with high influx of floating objects to the Antarctic
coast are associated with the subpolar gyres (Fraser et al. 2018b). The Weddell and
Ross Gyres are clockwise circulation features between Antarctica and the ACC,
driven by Ekman divergence at the intersection of the westerly winds to the north
and easterly winds to the south. At the eastern edge of the Weddell and Ross Gyres
(~50°E and ~120°W, respectively), the flow is southward towards Antarctica. If
seaweeds are driven sufficiently southward across the ACC fronts through a combi-
nation of eddy and wave-driven processes (Fig. 3.2), the subpolar gyres can connect
dispersal pathways from the southern edge of the ACC to the Antarctic coast. Once
seaweeds reach the coast of Antarctica, circumpolar connectivity around the coast-
line is enabled by the westward Antarctic Coastal and Antarctic Slope Currents.

Fig. 3.2  Illustration of eddies and Stokes drift in the Southern Ocean. White lines show long-term
average fronts in the Southern Ocean. Purple colours show a snapshot of surface ocean current
speed, and green arrows show a snapshot of the additional surface velocity from wave-driven
Stokes drift. (All data are from sources listed in Fraser et al. 2018b)
50 C. I. Fraser et al.

3.3  H
 itch-Hiking to the Antarctic: Passengers on Seaweed
Rafts

Buoyant macroalgae have long been recognised as potential vectors for the trans-
port of diverse other taxa (Thiel and Gutow 2005), including invertebrates (e.g.
Helmuth et al. 1994; Miranda and Thiel 2008; Nikula et al. 2010) and non-buoyant
seaweeds (e.g. Edgar 1987; Fraser et al. 2013; López et al. 2018). Indeed, entire
coastal communities can potentially travel long distances at sea via rafting with
seaweeds (Fraser et al. 2011). With new evidence that buoyant macroalgae can tra-
verse perceived oceanic ‘barriers’ and reach the Antarctic coast (Fraser et al. 2018b),
rafting events have the potential to enable diverse non-native species to reach
Antarctica. Currently, these taxa seem unable to establish, perhaps because the fre-
quent ice scour of rocky shores in Antarctica precludes establishment of the large
kelps these species require for habitat and food (Fraser et  al. 2009; Fraser et  al.
2018b). Antarctic coastal waters are warming, however, and some parts of the
Antarctic Peninsula could be comparable to present-day sub-Antarctic environ-
ments by the end of the century (Griffiths et al. 2017), facilitating establishment of
non-native species brought by kelp rafts.

3.3.1  Characteristics of Rafting Species

Shortly after detachment, there is a decline in the number of invertebrates associated


with kelp holdfasts (Edgar 1987; Hobday 2000; Gutow et al. 2009) – many indi-
viduals ‘jump ship’, including those that are commonly found on rafts at sea and are
apparently well suited to rafting (Miranda and Thiel 2008; Gutow et  al. 2009).
There may be bet-hedging benefits for those species that have individuals that either
leave the detached kelp or remain, with departing individuals enhancing their
chances of remaining in suitable kelp habitats, and remaining individuals enhancing
their chances of colonising new territory (Gutow et al. 2009). Some species also
colonise already detached, drift macroalgae, enhancing raft diversity after the initial
decline (Edgar 1987).
Many marine invertebrates have pelagic larvae that can disperse only over rela-
tively short distances (hundreds to thousands of metres: Kinlan and Gaines 2003),
so a key benefit of dispersing with macroalgal rafts is the capacity to disperse longer
distances. Rafts can endure for hundreds of days, travelling up to tens of thousands
of kilometres (Fraser et al. 2011, 2018b), and macroalgal rafts provide both habitat
and food to many of their passengers (Thiel 2003a). Taxa that are long-lived, or that
brood offspring, are thus most likely to be able to take advantage of dispersal via
long-distance rafting journeys  – direct-developing organisms have been noted to
make up an increasing proportion of the rafting community the longer a raft is at sea
(Thiel 2003a). Examples of such taxa include peracarid crustaceans (amphipods,
isopods and tanaids) and pulmonate gastropods (Thiel and Gutow 2005). In some
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 51

Fig. 3.3  Peracarid crustaceans including isopods and amphipods can brood their young within
kelp holdfasts, and are well suited to survive long rafting journeys. (a) The boring isopod Limnoria
stephenseni, (b) the amphipod Parawaldeckia kidderi, (c) a group of Parawaldeckia amphipods in
a tunnel in a dissected holdfast of Durvillaea antarctica bull kelp, (d) beached Durvillaea antarc-
tica bull kelp raft, with holdfast. (Photos by Ceridwen Fraser)

cases, these bore into kelp holdfasts, creating tunnels in which they are largely pro-
tected from exposure to harsh conditions in the open ocean and in which they can
nurture subsequent generations (e.g. the isopod genus Limnoria: Thiel 2003b)
(Fig. 3.3). Some other taxa that have been found associated with macroalgal rafts
that have dispersed long distances in the Southern Hemisphere include chitons, lim-
pets, bivalves, echinoderms and pycnogonids (Fraser et al. 2011), highlighting the
potential for macroalgal rafts to carry diverse new taxa to the Antarctic.

3.3.2  P
 rocesses Affecting Establishment of New Taxa
in the Antarctic

Simulated dispersal of surface particles (e.g. buoyant macroalgal rafts) from sub-­
Antarctic locations suggests that the vast majority are driven eastward and eventu-
ally northward, away from Antarctica. The proportions of particles able to reach the
Antarctic coast in simulations were relatively small – from 0.0001% to 0.1915%
(Fraser et al. 2018b). Estimates based on empirical surveys of kelp at sea suggest,
52 C. I. Fraser et al.

however, that there are tens of millions of macroalgal rafts adrift in the Southern
Ocean at any time; for Durvillaea antarctica, for example, there are estimated to be
about 70 million detached individuals at sea, with around 20 million of those still
bearing intact holdfasts (Smith 2002). If such estimates are correct and assuming
most rafts complete their journeys within 3 years (Fraser et al. 2018b), there could
be between a few and several thousand sub-Antarctic-origin D. antarctica rafts
reaching the Antarctic coast each month. For example, for releases from South
Georgia, approximately 0.19% of simulated particles reached the Antarctic coast
within 3 years (Fraser et al. 2018b), and this proportion of 70 million, over a period
of 36 months, amounts to close to 4000 per month, on average. Indeed, the finding
of D. antarctica on Antarctic beaches in 2017 (Fraser et al. 2018b) was not a one-off
event; D. antarctica samples had been collected from Antarctic waters previously,
but their significance was not immediately recognised. In 1989, for example, a spec-
imen was collected from Arthur Harbour at Anvers Island off the Antarctic Peninsula
(Herbarium of the University of California Berkeley, accession UC157330). Despite
this evidence for long-term, persistent transport of marine propagules from the sub-­
Antarctic to Antarctica, the large kelps that dominate the sub-Antarctic near-shore
ecosystems (such as D. antarctica and Macrocystis pyrifera) and many of the inver-
tebrates associated with them have not yet been established in Antarctica.
There could be a range of factors that influence the chance of establishment of
these large macroalgae and their associated epifauna in Antarctica. The most obvi-
ous of these is ice scour – neither D. antarctica nor M. pyrifera currently occurs in
areas heavily affected by iceberg scour or sea ice (Fraser et al. 2009; Macaya and
Zuccarello 2010; Fraser 2012). Most Antarctic coasts are likely to be affected by ice
scour long into the future, but warming near the Antarctic Peninsula over coming
decades could make shallow marine environments in that region somewhat compa-
rable to contemporary sub-Antarctic environments by the end of this century
(Griffiths et al. 2017). If areas along the Peninsula, for example, bays and channels,
are partly protected from ice scour in this warmer environment, some large sub-­
Antarctic macroalgae – and, subsequently, their associated epibionts – might have
an opportunity to establish.
Another consideration is the viability of dispersing seaweeds and other taxa on
reaching Antarctica. The viability of kelp rafts reaching Antarctica has yet to be
tested, but the two D. antarctica specimens found at King George Island in 2017
(Fraser et al. 2018b) were both male and had mature antheridia that appeared likely
to be viable. Furthermore, prior research indicates that the duration of survival and
viability of rafting macroalgae is greater in cooler, high-latitude waters versus
warmer, lower-latitude waters (Rothäusler et al. 2009). Establishment of dispersing
macroalgae can also be influenced by the presence or absence of competing conspe-
cific (Waters et  al. 2013; Fraser et  al. 2018a) or heterospecific (Valentine and
Johnson 2004) seaweeds. For Antarctic intertidal ecosystems, however, such com-
petition seems unlikely to play a major role in preventing the establishment of sub-­
Antarctic-­origin macroalgae, as Antarctic rocky shores generally are not densely
blanketed in algal cover (Fig.  3.4). Nonetheless, should ice scour become less
­frequent in some areas, competitive processes could influence which macroalgae
can establish and survive on exposed rocky shores.
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 53

Fig. 3.4  Antarctic rocky shores are starkly different to sub-Antarctic shores. In the former, fre-
quent ice scour removes all but the hardiest organisms, and seaweeds tend to be small and/or have
biphasic life cycles. In the sub-Antarctic, large brown macroalgae create habitat for diverse other
taxa including invertebrates and other seaweeds. (a, b) show sub-Antarctic shores dominated by
Durvillaea antarctica bull kelp; (c, d) show Antarctic shores lacking large seaweeds; (e, f) show
small, hardy seaweeds that can occur in both the sub-Antarctic and Antarctic: Adenocystis utricu-
laris (e) and Bostrychia intricata (f). White scale bars represent approximately 20 cm. (Photos A,
B, E and F by Ceridwen Fraser, sub-Antarctic Marion Island and New Zealand; C by Carlos
Olavarria; D by Emma Newcombe, South Shetland Islands, Antarctica)  (Figure modified from
Fraser et al. 2012).
54 C. I. Fraser et al.

3.4  Concluding Remarks

Antarctica has long been considered biologically isolated, but we now recognise
that the Southern Ocean and its fronts are not impermeable barriers to southward
dispersal of drifting objects. While there are no direct large-scale ocean currents
crossing the fronts, there are a number of both oceanic and directly wind-driven
processes that can move floating objects such as buoyant seaweed rafts north or
south across the fronts. Fraser et al. (2018b) showed that the combination of both
transient mesoscale variability and wave-driven Stokes drift is necessary for such
frontal traverses. Once floating objects have crossed to the southern edge of the
ACC, the large-scale horizontal ocean currents determine the locations of final
approach to Antarctica. Of the four primary regions where objects approach the
Antarctic coast (Fig 3.1), the region considered most feasible for future establish-
ment of sub-Antarctic macroalgae is the Antarctic Peninsula and nearby islands,
where warming is forecast to be most rapid. Should such establishment occur, there
could be drastic changes to Antarctic near-shore marine ecosystems, as drift mac-
roalgae can transport diverse other taxa including invertebrates and other seaweeds.

Acknowledgements  We thank Richard Moe for alerting us to the existence of the D. antarctica
specimen from Antarctica in the UC Berkeley herbarium and Erasmo Macaya for discussions. CIF
was funded by a Rutherford Discovery Fellowship from the Royal Society of New Zealand (RDF-­
UOO1803),  AKM was supported by an Australian Research Council DECRA Fellowship
(DE170100184), and POR was supported by a Strategic Doctoral Scholarship from the Division of
Sciences at the University of Otago.

References

Allcock AL, Strugnell JM (2012) Southern Ocean diversity: new paradigms from molecular ecol-
ogy. Trends Ecol Evol 27:520–528
Aronson RB, Blake DB (2015) Global climate change and the origin of modern benthic communi-
ties in Antarctica. Integr Comp Biol 41:27–39
Barnes DKA, Hodgson DA, Convey P, Allen CS, Clarke A (2006) Incursion and excursion of
Antarctic biota: past, present and future. Glob Ecol Biogeogr 15:121–142
Burton-Johnson A, Black M, Fretwell P, Kaluza-Gilbert J (2016) An automated methodology for
differentiating rock from snow, clouds and sea in Antarctica from Landsat 8 imagery: a new rock
outcrop map and area estimation for the entire Antarctic continent. Cryosphere 10:1665–1677
Cannone N, Corinti T, Malfasi F, Gerola P, Vianelli A, Vanetti I et al (2017) Moss survival through
in situ cryptobiosis after six centuries of glacier burial. Sci Rep 7:4438
Chelton DB, Schlax MG, Samelson RM (2011) Global observations of nonlinear mesoscale
eddies. Prog Oceanogr 91:167–216
Chen I-C, Hill JK, Ohlemüller R, Roy DB, Thomas CD (2011) Rapid range shifts of species asso-
ciated with high levels of climate warming. Science 333:1024–1026
Chown SL, Huiskes AHL, Gremmen NJM, Lee JE, Terauds A et al (2012) Continent-wide risk
assessment for the establishment of nonindigenous species in Antarctica. Proc Natl Acad Sci
USA 109:4938–4943
Chown SL, Clarke A, Fraser CI, Cary SC, Moon KL, McGeoch MA (2015) The changing form of
Antarctic biodiversity. Nature 522:431–438
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 55

Clarke A (1990) Temperature and evolution: Southern Ocean cooling and the Antarctic marine
fauna. In: Kerry KR, Hempel G (eds) Antarctic ecosystems. Springer, Berlin, Heidelberg,
pp 9–22
Clarke A, Barnes DKA, Hodgson DA (2005) How isolated is Antarctica? Trends Ecol Evol 20:1–3
Clayton MN (1994) Evolution of the Antarctic marine benthic algal flora. J Phycol 30:897–904
Constantin A, Johnson RS (2016) An exact, steady, purely azimuthal flow as a model for the
Antarctic Circumpolar Current. J Phys Oceanogr 46:3585–3594
Convey P, Stevens MI (2007) Antarctic biodiversity. Science 317:1877–1878
Convey P, Chown SL, Clarke A, Barnes DKA, Bokhorst S, Cummings V, Ducklow HW, Frati F
et al (2014) The spatial structure of Antarctic biodiversity. Ecol Monogr 84:203–244
d’Ovidio F, De Monte S, Penna AD, Cotté C, Guinet C (2013) Ecological implications of eddy
retention in the open ocean: a Lagrangian approach. J Phys A Math Theor 46:254023
DeConto RM, Pollard D (2003) Rapid Cenozoic glaciation of Antarctica induced by declining
atmospheric CO2. Nature 421:245–249
DeVries AL (1988) The role of antifreeze glycopeptides and peptides in the freezing avoidance of
antarctic fishes. Comp Biochem Physiol B Comp Biochem 90:611–621
Dobler D, Huck T, Maes C, Grima N, Blanke B, Martinez E, Ardhuin F (2019) Large impact of
Stokes drift on the fate of surface floating debris in the South Indian Basin. Mar Pollut Bull
148:202–209
Duffy GA, Coetzee BWT, Latombe G, Akerman AH, McGeoch MA, Chown SL (2017) Barriers
to globally invasive species are weakening across the Antarctic. Divers Distrib 23:982–996
Edgar GJ (1987) Dispersal of faunal and floral propagules associated with drifting Macrocystis
pyrifera plants. Mar Biol 95:599–610
Fraser CI (2012) The impacts of past climate change on sub-Antarctic nearshore ecosystems. Pap
Proc R Soc Tasman 146:89–93
Fraser CI, Nikula R, Spencer HG, Waters JM (2009) Kelp genes reveal effects of subantarctic sea
ice during the Last Glacial Maximum. Proc Natl Acad Sci U S A 106:3249–3253
Fraser CI, Nikula R, Waters JM (2011) Oceanic rafting by a coastal community. Proc R Soc B
278:649–655
Fraser CI, Nikula R, Ruzzante DE, Waters JM (2012) Poleward bound: biological impacts of
Southern Hemisphere glaciation. Trends Ecol Evol 27:462–471
Fraser CI, Zuccarello GC, Spencer HG, Salvatore LC, Garcia GR, Waters JM (2013) Genetic
affinities between trans-oceanic populations of non-buoyant macroalgae in the high latitudes
of the Southern Hemisphere. PLoS One 8:e69138
Fraser CI, Kay GM, Plessis MD, Ryan PG (2017) Breaking down the barrier: dispersal across the
Antarctic Polar Front. Ecography 40:235–237
Fraser CI, Davies ID, Bryant D, Waters JM (2018a) How disturbance and dispersal influence intra-
specific structure. J Ecol 106:1298–1306
Fraser CI, Morrison AK, Hogg AM, Macaya EC, van Sebille E, Ryan PG, Padovan A, Jack C,
Valdivia N, Waters JM (2018b) Antarctica’s ecological isolation will be broken by storm-driven
dispersal and warming. Nat Clim Change 8:704–708
Frenot Y, Chown SL, Whinam J, Selkirk PM, Convey P, Skotnicki M, Bergstrom DM (2005)
Biological invasions in the Antarctic: extent, impacts and implications. Biol Rev 80:45–72
Gillespie RG, Baldwin BG, Waters JM, Fraser CI, Nikula R, Roderick GK (2012) Long-distance
dispersal: a framework for hypothesis testing. Trends Ecol Evol 27:47–56
Gómez I (ed) (2015) Flora marina antártica: patrimonio de biodiversidad. Ediciones Kultrún,
Valdivia, Chile, p 243
González-Wevar CA, Díaz A, Gerard K, Cañete JI, Poulin E (2012) Divergence time estimations
and contrasting patterns of genetic diversity between Antarctic and southern South America
benthic invertebrates. Rev Chil Hist Nat 85:445–456
Griffiths HJ (2010) Antarctic marine biodiversity–what do we know about the distribution of life
in the Southern Ocean? PLoS One 5:e11683
Griffiths HJ, Whittle RJ, Roberts SJ, Belchier M, Linse K (2013) Antarctic crabs: invasion or
endurance? PLoS One 8:e66981
56 C. I. Fraser et al.

Griffiths HJ, Meijers AJS, Bracegirdle TJ (2017) More losers than winners in a century of future
Southern Ocean seafloor warming. Nat Clim Change 7:749–754
Gutow L, Giménez L, Boos K, Saborowski R (2009) Rapid changes in the epifaunal community
after detachment of buoyant benthic macroalgae. J Mar Biol Assoc UK 89:323–328
Hawes TC, Worland MR, Convey P, Bale JS (2007) Aerial dispersal of springtails on the Antarctic
Peninsula: implications for local distribution and demography. Antarct Sci 19:3–10
Helmuth B, Veit RR, Holberton R (1994) Long-distance dispersal of a subantarctic brooding
bivalve (Gaimardia trapesina) by kelp rafting. Mar Biol 120:421–426
Hickling R, Roy DB, Hill JK, Fox R, Thomas CD (2006) The distributions of a wide range of
taxonomic groups are expanding polewards. Glob Chang Biol 12:450–455
Hobday AJ (2000) Persistence and transport of fauna on drifting kelp (Macrocystis pyrifera (L.)
C. Agardh) rafts in the Southern California Bight. J Exp Mar Biol Ecol 253:75–96
Hunt GL, Drinkwater KF, Arrigo K, Berge J, Daly KL, Danielson S, Daase M, Hop H, Isla E et al
(2016) Advection in polar and sub-polar environments: impacts on high latitude marine eco-
systems. Prog Oceanogr 149:40–81
Kappen L (2000) Some aspects of the great success of lichens in Antarctica. Antarct Sci 12:314–324
Kappen L, Sommerkorn M, Schroeter B (1995) Carbon acquisition and water relations of lichens
in polar regions—potentials and limitations. The Lichenologist 27:531–545
Kennedy AD (1993) Water as a limiting factor in the Antarctic terrestrial environment: a biogeo-
graphical synthesis. Arct Alp Res 25:308–315
Kinlan BP, Gaines SD (2003) Propagule dispersal in marine and terrestrial environments: a com-
munity perspective. Ecology 84:2007–2020
Lehahn Y, d’Ovidio F, Lévy M, Amitai Y, Heifetz E (2011) Long range transport of a quasi isolated
chlorophyll patch by an Agulhas ring. Geophys Res Lett 38.
Lenn Y-D, Chereskin TK (2009) Observations of Ekman currents in the Southern Ocean. J Phys
Oceanogr 39:768–779
Lin X, Zhai X, Wang Z, Munday DR (2018) Mean, variability, and trend of Southern Ocean wind
stress: role of wind fluctuations. J Climate 31:3557–3573
López BA, Macaya EC, Rivadeneira MM, Tala F, Tellier F, Thiel M (2018) Epibiont communities
on stranded kelp rafts of Durvillaea antarctica (Fucales, Phaeophyceae)—Do positive interac-
tions facilitate range extensions? J Biogeogr 45:1833–1845
Macaya EC, Zuccarello GC (2010) Genetic structure of the giant kelp Macrocystis pyrifera along
the southeastern Pacific. Mar Ecol Prog Ser 420:103–112
Marynets K (2019) On the modeling of the flow of the Antarctic Circumpolar Current. Monatsh
Math 188:561–565
Miranda L, Thiel M (2008) Active and passive migration in boring isopods Limnoria spp.
(Crustacea, Peracarida) from kelp holdfasts. J Sea Res 60:176–183
Moon KL, Chown SL, Fraser CI (2017) Reconsidering connectivity in the sub-Antarctic. Biol Rev
92:2164–2181
Muñoz J, Felicísimo ÁM, Cabezas F, Burgaz AR, Martínez I (2004) Wind as a long-distance dis-
persal vehicle in the Southern Hemisphere. Science 304:1144–1147
Nikula R, Fraser CI, Spencer HG, Waters JM (2010) Circumpolar dispersal by rafting in two sub-
antarctic kelp-dwelling crustaceans. Mar Ecol Prog Ser 405:221–230
Ochyra R, Lewis Smith RI, Bednarek-Ochyra H (2008) The illustrated moss flora of Antarctica.
Cambridge University Press, Cambridge, p 685
Orsi AH, Whitworth T, Nowlin WD (1995) On the meridional extent and fronts of the Antarctic
Circumpolar Current. Deep Sea Res I Oceanogr Res Papers 42:641–673
Øvstedal DO, Lewis Smith RI (2001) Lichens of Antarctica and South Georgia: a guide to their
identification and ecology. Cambridge University Press, Cambridge, UK, p 413
Patarnello T, Bargelloni L, Varotto V, Battaglia B (1996) Krill evolution and the Antarctic ocean
currents: evidence of vicariant speciation as inferred by molecular data. Mar Biol 126:603–608
Pugh PJA (2003) Have mites (Acarina: Arachnida) colonised Antarctica and the islands of the
Southern Ocean via air currents? Polar Rec 39:239–244
3  Biogeographic Processes Influencing Antarctic and sub-Antarctic Seaweeds 57

Rascle N, Ardhuin F (2013) A global wave parameter database for geophysical applications. Part
2: model validation with improved source term parameterization. Ocean Model 70:174–188
Roads E, Longton RE, Convey P (2014) Millennial timescale regeneration in a moss from
Antarctica. Curr Biol 24:R222–R223
Robinson SA, King DH, Bramley-Alves J, Waterman MJ, Ashcroft MB, Wasley J, Turnbull JD
(2018) Rapid change in East Antarctic terrestrial vegetation in response to regional drying. Nat
Clim Change 8:879–884
Rothäusler E, Gómez I, Hinojosa IA, Karsten U, Tala F, Thiel M (2009) Effect of temperature
and grazing on growth and reproduction of floating Macrocystis spp. (Phaeophyceae) along a
latitudinal gradient. J Phycol 45:547–559
Scher HD, Martin EE (2006) Timing and climatic consequences of the opening of Drake Passage.
Science 312:428–430
Scher HD, Whittaker JM, Williams SE, Latimer JC, Kordesch WEC, Delaney ML (2015) Onset
of Antarctic Circumpolar Current 30 million years ago as Tasmanian Gateway aligned with
westerlies. Nature 523:580–583
Smith SDA (2002) Kelp rafts in the Southern Ocean. Glob Ecol Biogeogr 11:67–69
Smith CR, Grange LJ, Honig DL, Naudts L, Huber B, Guidi L, Domack E (2012) A large popula-
tion of king crabs in Palmer Deep on the west Antarctic Peninsula shelf and potential invasive
impacts. Proc R Soc Lond [Biol] 279:1017–1026
Sokolov S, Rintoul SR (2009) Circumpolar structure and distribution of the Antarctic Circumpolar
Current fronts: 1. Mean circumpolar paths. J Geohys Res Oceans 114:C11018
Teets NM, Denlinger DL (2014) Surviving in a frozen desert: environmental stress physiology of
terrestrial Antarctic arthropods. J Exp Biol 217:84–93
Thatje S, Anger K, Calcagno JA, Lovrich GA, Pörtner H-O, Arntz WE (2005) Challenging the
cold: crabs reconquer the Antarctic. Ecology 86:619–625
Thiel M (2003a) Rafting of benthic macrofauna: important factors determining the temporal suc-
cession of the assemblage on detached macroalgae. Hydrobiologia 503:49–57
Thiel M (2003b) Reproductive biology of Limnoria chilensis: another boring peracarid species
with extended parental care. J Nat Hist 37:1713–1726
Thiel M, Gutow L (2005) The ecology of rafting in the marine environment. II. The rafting organ-
isms and community. Oceanogr Mar Biol 43:279–418
Thompson AF (2008) The atmospheric ocean: eddies and jets in the Antarctic Circumpolar
Current. Philos Trans Royal Soc A 366:4529–4541
Valentine JP, Johnson CR (2004) Establishment of the introduced kelp Undaria pinnatifida follow-
ing dieback of the native macroalga Phyllospora comosa in Tasmania, Australia. Mar Freshw
Res 55:223–230
Vincent WF (2000) Evolutionary origins of Antarctic microbiota: invasion, selection and ende-
mism. Antarct Sci 12:374–385
Wasley J, Robinson SA, Lovelock CE, Popp M (2006) Some like it wet biological characteristics
underpinning tolerance of extreme water stress events in Antarctic bryophytes. Funct Plant
Biol 33:443–455
Waters JM (2008) Driven by the West Wind Drift? a synthesis of southern temperate marine bio-
geography, with new directions for dispersalism. J Biogeogr 35:417–427
Waters JM, Fraser CI, Hewitt GM (2013) Founder takes all: density-dependent processes structure
biodiversity. Trends Ecol Evol 28:78–85
Wauchope HS, Shaw JD, Terauds A (2019) A snapshot of biodiversity protection in Antarctica.
Nat Commun 10:946
Wiencke C, Clayton MN, Gómez I, Iken K, Lüder UH, Amsler CD, Karsten U, Hanelt D, Bischof
K, Dunton K (2007) Life strategy, ecophysiology and ecology of seaweeds in polar waters. Rev
Environ Sci Bio/Tech 6:95–126
Young IR (1999) Seasonal variability of the global ocean wind and wave climate. Int J Climatol
19:931–950
Chapter 4
Detached Seaweeds as Important Dispersal
Agents Across the Southern Ocean

Erasmo C. Macaya, Fadia Tala, Iván A. Hinojosa, and Eva Rothäusler

Abstract  After detachment from their substratum, many seaweeds can float or drift
at the mercy of currents and winds, thereby facilitating their dispersal and connec-
tivity. In Antarctica only one species possess floating structures (gas-filled vesicles),
the brown seaweed Cystosphaera jacquinotti. However, other species such as
Durvillaea antarctica and Macrocystis pyrifera that form abundant forests around
the sub-Antarctic islands can also remain at the sea surface once detached, provid-
ing a potential dispersal mechanism not only for the seaweeds but also for the asso-
ciated biota. Additionally, recent reports indicate that floating D. antarctica can
reach the Antarctic continent from sub-Antarctic regions. Herein, we collect

E. C. Macaya (*)
Laboratorio de Estudios Algales (ALGALAB), Departamento de Oceanografía, Universidad
de Concepción, Concepción, Chile
Millennium Nucleus Ecology and Sustainable Management of Oceanic Island (ESMOI),
Coquimbo, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: emacaya@oceanografia.udec.cl
F. Tala
Departamento de Biología Marina, Facultad de Ciencias del Mar, Universidad Católica del
Norte, Coquimbo, Chile
Centro de Investigación y Desarrollo Tecnológico en Algas, Facultad de Ciencias del Mar,
Universidad Católica del Norte (CIDTA-UCN), Coquimbo, Chile
e-mail: ftala@ucn.cl
I. A. Hinojosa
Millennium Nucleus Ecology and Sustainable Management of Oceanic Island (ESMOI),
Coquimbo, Chile
Departamento de Ecología, Facultad de Ciencias y Centro de Investigación en Biodiversidad
y Ambientes sustentables (CIBAS), Universidad Católica de la Santísima Concepción,
Concepción, Chile
e-mail: ihinojosa@ucsc.cl
E. Rothäusler
Centro de Investigaciones Costeras - Universidad de Atacama (CIC - UDA), Avenida
Copayapu 485, Copiapó, Chile
e-mail: eva.rothausler@uda.cl

© Springer Nature Switzerland AG 2020 59


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_4
60 E. C. Macaya et al.

i­nformation about floating and drifting seaweeds in Antarctica, but also their biol-
ogy, physiology, and distribution within the sub-Antarctic sources. Up to now, only
a few species have been recorded floating in Antarctica, and mainly during the first
Antartic explorations. So far, most of the studies on detached seaweeds only high-
light their importance, when already  stranded  and serving as carbon sources for
benthic communities. However,  some seaweed  species are  able to handle pres-
ent sea surface conditions in Antarctica and thus in the future when higher tempera-
tures,  less ice and more available substrate are available, they might be able to
frequently travel and colonize this region, thereby representing an important disper-
sal mode.

Keywords  Floating seaweed · Connectivity · Drifting · Rafting · Stranding


The two most striking vegetable productions of this island are a noble seaweed, called
Sargassum jacquinotii, and a Lichen. The first of these was not found attached, but floating
in the ocean among the ice, by which it was sometimes much mutilated. Though belonging
to a highly variable order, it is a perfectly distinct as well as conspicuous species. (James
Clark Ross 1847)

4.1  Introduction

The description above by the British Royal Navy Officer Sir James Ross onboard of
the HMS Erebus during the Antarctic exploration (January 1843) it is one of the first
records of floating seaweeds in Antarctica. In his report, Ross also mention the sur-
geon of the HMS Chanticleer, William Webster, who was the first naturalist in col-
lecting and describing such species at Deception Island: “the most common seaweed
was found floating. It was of a pale chocolate colour, stem and branches flat, quar-
ter of an inch in breadth, leaves equitant, thin, delicate, four of five inches long, and
at the base of each was a spherical air-cell the size of a grape. The mode of repro-
duction appeared to be from a cluster of buds appended to the terminal branches”
Webster (1834).
The seaweed observed by Webster and Ross is known today as Cystosphaera
jacquinotti and is endemic to the Antarctic. This brown seaweed species possesses
“pneumatocysts” (the size of a grape, according to Webster’s description) which are
floating structures and allow the seaweed to remain at the surface once being detached
from their primary substratum (Fig. 4.1). Although there are abundant records of
floating seaweeds around the world (e.g. Thiel and Gutow 2005a; Rothäusler et al.
2012; Macaya et al. 2016 and references therein), polar regions appear poorly rep-
resented.  This may be due to the harsh environmental conditions onboard of a
research vessel, the limited accessibility to many regions, and the isolation of
Antarctica from the rest of the Southern Ocean. That is probably why most of the
information comes from microalgal aggregates (Cefarelli et al. 2011; Assmy et al.
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 61

Fig. 4.1  The endemic Cystosphaera jacquinotti, black arrows indicate the gas filled structures
“pneumatocysts”. Right, a floating raft observed in Fildes Bay, King George Island. (Photos by
Erasmo C. Macaya)

2013; Katlein et al. 2015; Belt et al. 2018). As a result, few records of floating sea-
weeds have been reported in Antarctica, indicating only their importance as drifting
and stranding seaweeds, because they can serve as a carbon source for benthic com-
munities (e.g. Fischer and Wiencke 1992; Norkko et  al. 2004; Braeckman et  al.
2019; see also Chap. 8 by Quartino et al.). However recent records of non-native
positively buoyant kelps (Durvillaea antarctica and Macrocystis pyrifera)  that
stranded on South Shetland Islands opened the discussion regarding the isolation of
Antarctica (Fraser et al. 2018; Avila et al. 2020). This is especially important in the
context of future warming scenarios, where long-distance dispersal might facilitate
connectivity with sub-Antarctic seaweed sources. In fact, long-distance dispersal
and connectivity within these vast areas of open oceans have been already demon-
strated for D. antarctica and Macrocystis pyrifera, with single haplotypes having a
wide sub-Antarctic distribution (Fraser et al. 2009; Macaya and Zuccarello 2010).
At the same time, these seaweeds serve as important dispersal vehicles for associ-
ated flora and fauna (Nikula et al. 2010; Wichmann et al. 2012; Nikula et al. 2013;
Cumming et al. 2014; Macaya et al. 2016), considering the large number of floating
D. antarctica rafts in the Southern Ocean (Smith 2002) (see also Chap. 3 by Fraser
et al.). In this chapter we give information about floating, drifting and stranded sea-
weeds in Antarctica in relation to their abundance and distribution. We also address
whether the extreme  environmental conditions (irradiance and temperature) can
62 E. C. Macaya et al.

affect seaweed physiology, growth and thus survival mainly in floating conditions.
In addition, we give potential sources of seaweeds that may have the chance to
travel into Antarctica.

4.2  Detached Seaweeds in Antarctica

A total of 39 seaweeds species, including 3 Chlorophyta, 14 Ochrophyta, and 22


Rhodophyta–have been recorded drifting, stranding or floating in Antarctica or
crossing the Antarctic Polar Front APF (Table 4.1). From the 95 records included
herein, 75% reported about observations or studies of drifting seaweeds, followed
by a 17% of floating seaweeds, while only an 8% reported about stranded seaweeds.
To our knowledge, so far only brown but no red or green seaweed species have been
found floating in Antarctica. The brown seaweeds  species  include Cystosphaera
jacquinotti, Himantothallus grandifolius, Desmarestia anceps and Adenocystis
utricularis. Among them, only C. jacquinotti possesses floating structures and is
also the most frequent species observed. For instance, abundant seaweed  pieces
were recorded floating during 1908–1909 in the Neumayer and Peltier Channels as
well as in the Bransfield and Gerlache Straits (Gain 1912). Fertile floating individu-
als  were observed in Harmony Cove (Neushul 1963) and pieces up to 90 cm  in
length were spotted among ice in Cockburn Island (Ross 1847). In contrast, A. utric-
ularis owes buoyancy when mature thalli become filled with air  and individuals
have been found floating abundantly within the Fildes Bay (E. Macaya pers. obs).
Floating thalli but also holdfasts of H. grandifolius were observed within Admiralty
Bay (Zemko et al. 2015). Large numbers of floating kelps such as Macrocystis pyr-
ifera and Durvillaea antarctica were found floating south of the APF (Fraser et al.
2017) and floating D. antarctica  pieces were recorded in the interior pass of
Deception Island (Gain 1912). Several records of floating seaweed were also carried
out by James Cook during his voyage towards the South Pole during 1770’s, he
described floating seaweeds at latitude 70°S, longitude 108°W: “We continued our
course to the south, and passed a piece of weed covered with barnacles, which a
brown albatross was picking off”. Those rafts might have been afloat for a while,
because of the barnacle presence (Thiel and Gutow 2005a).
Among the drifting red seaweeds, most of the species belong to the order
Ceramiales and Gigartinales, with seven and five species recorded, respectively.
Attached to the spines of the sea urchins Sterechnus neumayeri (Amsler et al. 1999;
Schwarz et  al. 2003) or  accumulated in hollows,  a bundant biomass of drifting
Phyllophora antarctica were observed at 15–30 m depth in Cape Evans, McMurdo
Sound (Norkko et al. 2004). Interestingly, S. neumayeri uses drift pieces of P. ant-
arctica and Iridaea cordata as defense against its main predator, the anemone
Isotealia antarctica. At the same time, this retention of drifting seaweed pieces indi-
rectly supports to maintain the seaweed populations, if pieces are fertile and still
within the photic zone (Amsler et al. 1999). The most common drifting brown sea-
weeds belong to the order Desmarestiales. For instance individuals of Desmarestia
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 63

Table 4.1  Detached seaweeds reported drifting, floating or stranded in Antarctica. Floating kelps
crossing the Antarctic Polar Front are also included
Algal-­
Phylum/order Species status Locality Reference
Chlorophyta
Prasiolales Prasiola crispa Drift Fildes Bay, King (Fischer and Wiencke
George Island 1992)
Ulothricales Monostroma Drift Admiralty Bay, (Oliveira et al. 2009)
hariotii King George
Island
Fildes Bay, King (Fischer and Wiencke
George Island 1992)
Stranded Fildes Bay, King (Westermeier et al.
George Island 1992)
Ulvales Ulva intestinalis Drift Whalers Bay, (Clayton et al. 1997)
Deception Island
Ochrophyta
Ascoseirales Ascoseira mirabilis Drift Fildes Bay, King (Müller et al. 1990;
George Island Sato et al. 1992; Tada
et al. 1996)
Drift Admiralty Bay, (Zielinski 1990)
King George
Island
Chrysomeridales Antarctosaccion Drift South Bay, (Gallardo et al. 1999)
applanatum Livingston Island
Desmarestiales Desmarestia anceps Drift Fildes Bay, King (Fischer and Wiencke
George Island 1992; Tada et al.
1996)
DeLaca Island (Amsler et al. 2012)
Signy Island, (Richardson 1979;
South Orkney Brouwer et al. 1995)
Islands
Admiralty Bay, (Zielinski 1990)
King George
Island
Floating, Bransfield Strait; (Gain 1912)
stranded Deception Island;
King George
Island
Desmarestia Drift Deception Island (Lastra et al. 2014)
antarctica
Fildes Bay, King (Fischer and Wiencke
George Island 1992)
Desmarestia Drift Deception Island (Lastra et al. 2014)
menziesii
Signy Island, (Brouwer 1996)
South Orkney
Islands
(continued)
64 E. C. Macaya et al.

Table 4.1 (continued)
Algal-­
Phylum/order Species status Locality Reference
Fildes Bay, King (Fischer and Wiencke
George Island 1992)
Admiralty Bay, (Zielinski 1990)
King George
Island
Desmarestia sp. Stranded Deception Island (Gain 1912)
Himantothallus Floating Admiralty Bay, (Zemko et al. 2015)
grandifolius King George
Island
Drift Admiralty Bay, (Zielinski 1981, 1990)
King George
Island
Signy Island, (Brouwer 1996)
South Orkney
Islands
Ectocarpales Adenocystis Drift, Fildes Bay, King (Westermeier et al.
utricularis George Island 1992; Tada et al.
1996)
Floating Fildes Bay, King (Erasmo Macaya pers.
George Island obs.)
Chordaria linearis Drift Admiralty Bay, (Oliveira et al. 2009)
King George
Island
Petalonia fascia Drift Whalers Bay, (Clayton et al. 1997)
Deception Island
Scytosiphon Drift Whalers Bay, (Clayton et al. 1997)
lomentaria Deception Island
Utriculidium Fildes Bay, King (Müller et al. 1992)
durvillei George Island
Fucales Cystosphaera Floating Deception Island (Hooker 1844)
jacquinotii
Cockburn Island (Ross 1847)
Antarctic (Montagne 1842;
Peninsula Skottsberg 1907)
Half Moon Island (Neushul 1963)
Fildes Bay, King (Gain 1912; Macaya
George Island E. Personal
Observation; Weykam
et al. 1996)
Gerlache Strait (De Wildeman 1935)
South Bay, Macaya E. Personal
Doumer Island Observation
Drift Admiralty Bay, (Zielinski 1981, 1990;
King George Oliveira et al. 2009)
Island
(continued)
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 65

Table 4.1 (continued)
Algal-­
Phylum/order Species status Locality Reference
Drift Ardley Bay, King (Fischer and Wiencke
George Island 1992)
Drift, Half Moon Island (Neushul 1965)
stranded
Durvillaea Floating, Deception Island, (Gain 1912)
antarctica stranded King George
Island
Stranded Livingtone Island (Avila et al. 2020)
Floating South of New (Hooker 1844)
Zealand 65°S
South of the (Fraser et al. 2017)
Antarctic Polar
Front
Drift Livingston and (Pellizzari et al. 2017)
Elephant Islands
Stranded Fildes Bay, King (Fraser et al. 2018)
George Island
Laminariales Macrocystis Floating South of the (Fraser et al. 2017)
pyrifera Antarctic Polar
Front
Stranded Deception Island (Avila et al. 2020)
Rhodophyta
Balliales Ballia callitricha Drift South Bay, (Gallardo et al. 1999)
Livingston Island
Bangiales Porphyra Drift Cuverville Island (Clayton et al. 1997)
plocamiestris
Bonnemaisoniales Delisea pulchra Drift Admiralty Bay, (Oliveira et al. 2009)
King George
Island
Ceramiales Georgiella Drift Fildes Bay, King (Fischer and Wiencke
confluens George Island 1992)
Myriogramme Drift Livingston Island (Gallardo et al. 1999)
manginii
Deception Island (Gain 1912)
Signy Island, (Brouwer 1996)
South Orkney
Islands
Neuroglossum Drift Fildes Bay, King (Fischer and Wiencke
delesseriae George Island 1992)
Pantoneura Drift Signy Island, (Brouwer 1996)
plocamioides South Orkney
Islands
South Bay, (Gallardo et al. 1999)
Livingston Island
(continued)
66 E. C. Macaya et al.

Table 4.1 (continued)
Algal-­
Phylum/order Species status Locality Reference
Paraglossum Drift Admiralty Bay, (Oliveira et al. 2009)
lancifolium King George
Island
Phycodrys Drift Fildes Bay, King (Fischer and Wiencke
antarctica George Island 1992)
Phycodrys Drift Fildes Bay, King (Weykam et al. 1996)
quercifolia George Island
Gigartinales Cystoclonium Drift Byers Peninsula, (Hommersand et al.
obtusangulum Livingston Island 2009)
Gigartina Drift Fildes Bay, King (Fischer and Wiencke
skottsbergii George Island 1992)
Admiralty Bay, (Oliveira et al. 2009)
King George
Island
Iridaea cordata Drift Cape Evans, (Schwarz et al. 2003)
McMurdo Sound
Fildes Bay, King (Fischer and Wiencke
George Island 1992)
McMurdo Sound (Amsler et al. 1999)
Admiralty Bay, (Oliveira et al. 2009)
King George
Island
Stranded Fildes Bay, King (Westermeier et al.
George Island 1992)
Notophycus Drift Admiralty Bay, (Oliveira et al. 2009)
fimbriatus King George
Island
Phyllophora Drift Cape Evans, (Miller and Pearse
antarctica McMurdo Sound, 1991; Schwarz et al.
Ross Island 2003; Norkko et al.
2004)
McMurdo Sound (Amsler et al. 1999)
Gracilariales Curdiea racovitzae Drift Fildes Bay, King (Fischer and Wiencke
George Island 1992)
Halymeniales Pachymenia Drift Admiralty Bay, (Oliveira et al. 2009)
orbicularis King George
Island
Palmariales Palmaria decipiens Drift Half Moon Island (Neushul 1965)
Foster Bay, (Lastra et al. 2014)
Deception Island
Fildes Bay, King (Fischer and Wiencke
George Island 1992)
(continued)
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 67

Table 4.1 (continued)
Algal-­
Phylum/order Species status Locality Reference
Drift, Admiralty Bay, (Zielinski 1981;
stranded King George Westermeier et al.
Island 1992; Oliveira et al.
2009)
Plocamiales Trematocarpus Drift Fildes Bay, King (Fischer and Wiencke
antarcticus George Island 1992)
Plocamium Drift Fildes Bay, King (Fischer and Wiencke
cartilagineum George Island 1992)
Admiralty Bay, (Oliveira et al. 2009)
King George
Island
Signy Island, (Brouwer 1996)
South Orkney
Islands
Plocamium hookeri Drift South Bay, (Gallardo et al. 1999)
Livingston Island
Rhodymeniales Hymenocladiopsis Drift South Bay, (Gallardo et al. 1999)
prolifera Livingston Island

anceps and D. menziesii were observed at 10 to 40  m depth in Admiralty


Bay,  where  icebergs  provoked their detachment (Zielinski 1990). Huge  drifting
patches of D. anceps with up to 60 m3 were recorded between 5 and 20 m depth in
Fildes Bay. These, detached individuals represent an important habitat for a variety
of invertebrates such as gastropods (Nacella concinna), giant isopods (Glyptonotus
antarcticus) and gammaridean amphipods (Tada et al. 1996) (Fig. 4.2).
Similarly, high quantities of drifting D. anceps detached by ice and wave action
accumulated in hollows at Signy Island, South Orkney (Brouwer 1996). When
detached  this species continued  to grow and remained healthy for over
44 weeks, without changing its palatability (Brouwer 1996; Amsler et al. 2012). To
date, only 3 species of drifting green seaweeds have been reported in Antarctica,
namely: Monostroma hariotti, Prasiola crispa and the non-native Ulva intestinalis
(Fischer and Wiencke 1992; Clayton et al. 1997; Oliveira et al. 2009). Large inflated
thalli of U. intestinalis were found drifting in Whalers Bay (Deception Island),
which is probably a recent introduction, caused by shipping activities (Clayton et al.
1997), also this species has been reported as a common fouling organism in other
latitudes (Blomster et al. 1998). However, introduction via floating dispersal cannot
be ruled out because U. intestinalis can form massive floating mats, that are known
to tolerate severe winter conditions (Bäck et al. 2000) together with a rapid acclima-
tion to changes in salinity, nutrients and light (Cohen and Fong 2004). Similarly,
free floating Ulva linza acclimated rapidly to changing light conditions by develop-
ing effective mechanisms to cope with excessive irradiation (Häder et al. 2001).
Detached seaweeds washed ashore (Fig. 4.3) have been studied at different sites
with in Antarctica. Lastra et al. (2014) estimated 1545 tons of seaweeds that arrived
68 E. C. Macaya et al.

Fig. 4.2  Drifting seaweeds accumulated in hollows, Fildes Bay, King George Island. (Photo by
Ignacio Garrido)

to the shore in Foster Bay (Deception Island) within the period from November to


July, and only 3% of the detached seaweeds accumulated as wrack along the inter-
tidal. In Admiralty Bay, 1643 tons of stranded seaweeds were estimated between
February and November 1979, and especially after storms (Rakusa-Suszczewski
1980; Zielinski 1981). Stranded seaweeds remnants decompose with contact to air,
and within 20  d, more than 50% were transformed  to organic matter (Rakusa-­
Suszczewski 1980), a process that  becomes accelerated  by winds and waves
(Zielinski 1981). The wrack deposition can be highly variable with more than 50%
of replacement at each tidal cycle (Lastra et al. 2014), probably caused by the inter-
action between wind direction and coastal orientation (Rakusa-Suszczewski 1980;
Zielinski 1981; Rakusa-Suszczewski 1995; Lastra et al. 2014). Slow decomposition
rates due to the extreme environment, however facilitates the longevity of detached
seaweeds and constitute an important carbon source for higher trophic levels via the
detrital food web (Fischer and Wiencke 1992; Schwarz et al. 2003; see also Chap 8
by Quartino et al.).
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 69

Fig. 4.3  Stranded seaweeds at Fildes Bay, King George Island. Low left, Desmarestia antarctica
as epiphyte of stranded Curdiea racovitzae. (Photos by Erasmo C. Macaya)

4.3  Abiotic Factors Influencing Floating Seaweeds

Solar radiation (280–700 nm) and sea surface temperature (SST) are the main abi-
otic factors that determine the persistence and thus dispersal ability of floating sea-
weeds (Hobday 2000; Rothäusler et al. 2009, 2012; Graiff et al. 2013; Tala et al.
2016, 2019). In Antarctica and sub-Antarctic islands, there is a strong seasonal light
regime, especially at latitudes >70°S, where several months of darkness in winter
and of complete daylight in summer with up to 1700 μmol m−2 s−1of PAR, 44 W m−2
of UV-A, and 2.3 W m−2 of UV-B can prevail (Quartino et al. 2005; Zacher et al.
2007, 2009; see also Chap 7 by Huovinen and Gómez). Also SST varies from
warmer sub-Antarctic waters (4  °C to 14  °C) to colder, icier Antarctic waters
(−1.8 °C and 2.2 °C off the Antarctic Peninsula) (Drew and Hastings 1992; Klöser
et al. 1993; Kang et al. 2002; Mélice et al. 2003). Thus, depending on the latitude or
season when detached, and/or when occasionally become pushed through the APF,
seaweeds are confronting extreme different conditions during rafting.
In situ experiments carried out with two tethered kelp species in Tierra del Fuego
at 54°S, showed for the temperate M. pyrifera that its growth capacity is favored by
high light levels (~1000 μmol m−2 s−1) and moderate SSTs (~10 °C) in summer but
the high growth rates cannot be maintained in low light (~700 μmol m−2 s−1) com-
bined with low SST (~6  °C) in winter (Tala et  al. 2016). In contrast, the
70 E. C. Macaya et al.

sub-Antarctic D. antarctica showed a reverse pattern with positive growth in winter


despite the low light and temperature conditions. Apparently, the potential for raft-
ing dispersal in D. antarctica appears to be less dependent on seasonal variations in
environmental conditions (Tala et  al. 2019). Clearly, sub-Antarctic floating sea-
weeds perform well in cooler waters and under low solar radiation than their tem-
perate counterparts (Rothäusler et al. 2009; Graiff et al. 2013; Tala et al. 2016, 2019).
Seaweeds that are distributed in sub-Antarctic islands and Tierra del Fuego such
as D. antarctica and M. pyrifera typically grow in environments with large fluctua-
tions in seawater temperature and have thus a broad performance breadth (Eggert
2012). Consequently, if floating conspecifics have the chance to become pushed
through the APF (Fraser et al. 2017; Fraser et al. 2018) they possibly can acclimate
to cold-water temperatures more easily and thereby contributing to long distance
dispersal (Fraser et  al. 2018). However, M. pyrifera may fail to establish in the
colder, icier Antarctic environment because it was shown that they lost biomass at
54°S in winter probably due to low light availability (Tala et al. 2016). Hence, abi-
otic conditions in polar waters >60 °S might temporarily, especially in winter, be
unfavorable for some cold-temperate floating seaweeds. This might be the reason
why M. pyrifera (> 1 m) has been reported in coexistence with the endemic Antarctic
H. grandifolius under the extreme environmental conditions in South Georgia
(Barnes et al. 2006). At the same latitude (54°S), but at Tierra del Fuego, floating
D. antarctica did not show biomass losses in winter and steadily grew, which was
not the case in summer (Tala et  al. 2016). As a result its floating time exceeded
200 d in winter but in summer it dropped to 90 d (Tala et al. 2019). This suggests
that D. antarctica may survive and be successful when entering into icier SSTs at
the  APF.  Indeed, two mature thalli of D. antarctica were found stranded at the
Antarctic Peninsula (62°S) in summer 2017 with goose barnacles attached (Fraser
et al. 2018). Molecular studies showed that these seaweed pieces originated from
sub-Antarctic Islands and therefore the dispersal distances represented the longest
rafting events ever recorded, with trajectories between 20,000 and 25,000  km
(Fraser et  al. 2018). This suggests that these pieces started their journey as rafts
consisting of several individuals because it was estimated that they travelled a maxi-
mum of 2 years and a minimum of 489 d in order to reach Antarctica from their
origins (Fraser et al. 2018). During this very long pelagic journey, rafts experienced
substancial fluctuations in the abiotic and biotic environment, thereby probably los-
ing biomass.
It is known that nutrient abundance stimulates seaweed growth and thus it has
been inferred to affect the persistence time of floating seaweeds (Rothäusler et al.
2012 and references therein). When seaweeds become occasionally pushed through
the APF and float in the open waters of the Southern Ocean, which are defined as
high nutrient environments, their growth and thus survival is unlikely to be restricted.
The same is true when seaweeds arrive to the coastal waters of the Antarctic
Peninsula, where high levels of nitrate and phosphate are present at the sea surface
throughout the year (reviewed in Zacher et al. 2009). There is still, however, very
little information available on the impact of nutrients on floating seaweeds.
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 71

4.4  Biotic Factors Affecting Floating Seaweeds

Large numbers of hitchhiking organisms have been reported from floating seaweeds
(Khalaman and Berger 2006; Vandendriessche et  al. 2007a; Wichmann et  al.
2012; Abé et al. 2013; Macaya et al. 2016; López et al. 2018). Many of these organ-
isms are grazers e.g. amphipods and isopods that mostly feed on vegetative parts
from the rafts, thereby removing photosynthetic tissue, and thus contribute to their
demise (Vandendriessche et  al. 2007b; Rothäusler et  al. 2009; Rothäusler et  al.
2018). But also grazing by e.g. snails and sea urchins and fouling by bryozoans,
goose barnacles, and epiphytic seaweeds can limit their persistence at the sea sur-
face (Rothäusler et al. 2011b; Graiff et al. 2016; Rothäusler et al. 2018; see also
Chap. 17 by Amsler et al. about algal-grazers interactions in Antarctic seaweeds).
Especially, the overgrowth with sessile organisms can reduce buoyancy, because
epibionts contribute to an increase in the specific gravity of the floating seaweed
(Graiff et al. 2016). These biotic interactions are particularly important for seaweed
rafts because they will directly affect the persistence (and dispersal potential) of
rafts at the sea surface.
The destructive effect of herbivorous grazers on their rafts is dependent on water
temperature (Vandendriessche et al. 2007a; Rothäusler et al. 2009, 2018). In meso-
cosms but also in field studies it was shown that low temperatures (5–15 °C) slowed
amphipod and isopod consumption on floating M. pyrifera, Fucus vesiculosus, and
Ascophyllum nodosum, and that these seaweeds can compensate for grazer induced
tissue losses (Vandendriessche et  al. 2007a; Rothäusler et  al. 2009, 2018). For
instance, at 5 °C, F. vesiculosus and A. nodosum kept in mesocosms gained 3 times
their initial weight while afloat, but the weight gain was less in the presence of iso-
pods (Vandendriessche et al. 2007b). This is probably because herbivore metabo-
lism, which is strongly controlled by temperature (O’Connor 2009), slows down at
low SSTs and seaweed growth is favored under benign environmental conditions.
Similarly, in cold sub polar waters around Iceland (64°N), Ingólfsson (1998)
reported that fronds of A. nodosum, which were accidentally left afloat for 43 d, did
not show any signs of decay and were still in perfect condition. Consequently, float-
ing seaweeds are expected to be less grazed under cold SSTs.
Rafts serve as substratum for many sessile epibionts, including e.g. encrusting
bryozoans and goose barnacles (Hinojosa et al. 2006; Rothäusler et al. 2011a; Graiff
et al. 2016). Bryozoan growth and thus additional weight generally increase with
seaweed floating time, thereby negatively affecting seaweed persistence (Rothäusler
et al. 2011b for M. pyrifera at 30°S), which can finally result in their decay and sink-
ing (Graiff et al. 2016). However, recently, it was shown that maximum bryozoan
cover on tethered M. pyrifera (30°S) was reached earlier in spring/summer than in
autumn/winter (Graiff et  al. 2016). In a study done with natural floating rafts of
M. pyrifera (30°S), the colonization by bryozoans increased with decreasing lati-
tudes, which coincides with warmer SSTs (Rothäusler et al. 2011a). This implies
that bryozoan growth is slowed down at lower SSTs. Hence, seaweeds carrying
epibionts and traveling in cold and icy sub-Antarctic or Antarctic waters may stay
72 E. C. Macaya et al.

afloat for substantially longer periods, which is also underlined by the fact that
D. antarctica thalli beached at Antarctica had goose barnacles attached (Fraser et al.
2018; E. Macaya Pers. Obs.). Preliminary observations on floating rafts of C. jac-
quinotti within the Antarctic Peninsula revealed very few epibionts (E.  Macaya
pers. obs.).
Moreover, there is evidence from molecular studies, showing that hitchhiking
organisms, such as crustaceans, mollusks, seaweed parasites, and non-buoyant sea-
weeds, can travel vast distances (Nikula et al. 2010, 2013; Fraser and Waters 2013;
Fraser et al. 2013; Boo et al. 2014; Guillemin et al. 2014; Macaya et al. 2016). Some
of them even have a circum sub-Antarctic distribution, as has been reported for two
holdfast dwelling peracarid crustaceans (Nikula et al. 2010). However, there are no
studies from high latitudes determining whether these hitchhikers affect the floating
persistence of their rafts. Probably, under favorable conditions such as prevailing in
Tierra del Fuego, sub-Antarctic islands, and Antarctica, seaweeds can continuously
grow, and thus provide a long lasting substratum and food source for their associ-
ated hitchhikers, which in turn may have a decreased metabolic rate. Particularly,
high abundances of amphipods (up to 300,000 individuals m2) have been recorded
in benthic Antarctic seaweeds stands, such as Desmarestia antarctica, D. menziesii
and D. anceps (Huang et al. 2007; see also Chap. 17 by Amsler et al.), some of these
species have been found floating or drifting (Table 4.1) and might represent a dis-
persal vehicle for associated hitchhikers.

4.5  P
 hysiology of Floating and Drifting Seaweeds:
Traspassing Thermal Barriers

Our knowledge about the physiological, reproductive and growth performance of


Antarctic seaweeds comes mainly from benthic populations and/or from seaweeds
kept  under experimental conditions. Similarly, the performance of floating sea-
weeds at high latitudes has been almost exclusively studied under mesocosm condi-
tions. Generally, in Antarctic benthic  seaweeds,  photosynthetic and bio-optical
properties as well as UV and/or  temperature stress tolerances are linked to their
vertical distribution and biogeographic affinity (Huovinen and Gómez 2013; Gómez
et al. 2019). These seaweeds are defined as shade-adapted, and hence are photosyn-
thetically and metabolically prepared to cope with extreme irradiance conditions
but also metabolically with very low temperatures (Gómez et al. 2009; Huovinen
and Gómez 2013; Rautenberger et  al. 2015; Gómez et  al. 2019). Hence,  during
spring/summer, when irradiances, day length, and light penetration are suddenly
increasing, they can become stressful for many benthic seaweeds but also for their
community associated (Zacher et al. 2009). This is especially the case for subtidal
species, which are growing in a more stable environment compared to intertidal
species.
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 73

Upon detachment, floating seaweeds suffer a sudden environmental change from


the benthic (shaded) to the pelagic enviroment  (sunny). Thus, depending on the
season and latitude when detached, floating seaweeds but also their associated
hitchhikers might respond differently (e.g. growth, physiological acclimation and
reproduction) (Rothäusler et al. 2012; Macaya et al. 2016 and references therein).
At high latitudes (54°S), floating rafts of M. pyrifera and D. antarctica displayed
different short-term physiological adjustments to floating conditions depending on
the season, which are apparently associated to their biogeographic affinity and
growth strategies (Tala et al. 2016). During winter, M. pyrifera lost biomass although
seaweeds maintained their maximum quantum yield (Fv/Fm) and increased their pig-
ment levels, presumably because of light limitation and to support some biomass
production. Opposite responses such as positive biomass changes, photoregulation,
decreasing Fv/Fm and pigment levels were found during summer (Tala et al. 2016).
Durvillaea antarctica might support better the winter conditions that prevail at high
latitudes because positive biomass changes were found during the first three months
afloat and only small variations in Fv/Fm, pigment, and phlorotannin levels. However,
in summer, tissue deterioration and antioxidant activity increased, which were
reflected in biomass loss (Tala et  al. 2016, 2019). Therefore, at high latitudes,
D. antarctica can stay longer afloat (>200 days) during winter than during summer
(90 days).

4.5.1  Out of Antarctic: Is it Physiologically Feasible?

Even though there is no specific information about the physiological and reproduc-
tive  responses of floating seaweed in and out Antarctica, important transoceanic
routes have been identified through phylogeographic studies at high latitudes (Fraser
et al. 2010; Coyer et al. 2011; Fraser et al. 2013; Boo et al. 2014; Moon et al. 2017),
confirming the capacities of floating and non-buoyant seaweeds to acclimate during
long journeys (see also Chap. 3 by Fraser et al.). The successful dispersal of floating
seaweeds over short or long distances requires an efficient physiological acclima-
tion (e.g. photosynthetic efficiency adjust, high antioxidant activity, production of
defensive metabolites) not only to maintain growth and reproductive capacity, but
also to deal with the herbivore pressure and epibiont load. When rafting is success-
ful, seaweeds can extend their distributional ranges according to their physiologi-
cal and/or reproductive tolerances and/or abiotic and biotic interactions (Thiel and
Gutow 2005b; Nikula et al. 2010; Coyer et al. 2011; Waters et al. 2013; Batista et al.
2018; López et al. 2018).
Performance and tolerance of benthic seaweed populations can give us an indica-
tion of how they may respond under floating or drifting conditions. The optimum
temperatures for photosynthesis in Antarctic seaweeds range between 5 and 15 °C
(Wiencke et al. 1993; Eggert and Wiencke 2000), thus floating seaweeds entering
Antarctic from Tierra del Fuego or sub-Antarctic sources might be able to sur-
vive. Temperature requirements for growth and survival of 15 Antarctic red seaweed
74 E. C. Macaya et al.

species were studied  by Bischoff-Bäsmann and Wiencke (1996). Consequently,


if eurythermal species (with a broad performance breadth) become pushed through
the APF they can acclimate more easily to more temperate temperatures more easily
and can support large distance dispersal. Hence  an “out of Antarctic” scenario,
could be supported at least by the high environmental tolerance showed by some
cold-adapted Antarctic seaweeds. In fact, various species display different mecha-
nisms to support elevated temperatures and high UVR /PAR conditions (Huovinen
and Gómez 2013; Rautenberger et al. 2015). In experimental conditions, UVR tol-
erance is improved in sensitive species modulated by temperature via an efficient
damage repair of the photosynthetic apparatus instead of increasing photo-­protective
metabolites or antioxidant activity (Rautenberger et  al. 2015). Therefore, chemi-
cally based tolerance mechanisms are not inducible as has been reported for some
Desmarestiales (Flores-Molina et al. 2016). In fact, high constitutive concentrations
of soluble and insoluble phlorotannins (a type of phenolics found mainly in brown
seaweeds) have been determined irrespective of the UV levels at which these sea-
weeds are  exposed in their habitat (Rautenberger et  al. 2015). In brown sea-
weeds,  phlorotannins have different biological functions such as UV shielding,
ROS scavenging, herbivore deterrence, and cell wall formation. Therefore, species
that maintain high  phlorotannin concentrations are  better prepared to face vari-
able  environmental conditions (see also Chap. 18 by Gómez and Huovinen).
However, due to an extended period of low light during the polar winter, and turbid
seawater during summer and autumn following sea-ice melting and phytoplankton
blooms, Antarctic seaweeds are strongly shade adapted, which could impair their
physiological performance during rafting and limiting the migration to lower
latitudes.
The endemic floating species Cystosphaera jacquinotii is commonly found
between 5 to 30 m depth, usually in more exposed areas and disappear south of 66°S
(Neushul 1963; Wulff et al. 2009). Adenocystis utricularis with its vesicle saccate-­
like morphology are abundant in the intertidal (Valdivia et al. 2014) and have also
been recorded floating (see above). Photosynthetic characteristics showed that
C. jacquinotti has a low saturating irradiance (Ek = 37 μmol photons m−2 s−1) and
stress tolerance to UVR and enhanced temperature, while A. utricularis showed
high saturating irradiance (Ek  =  137 μmol photons m−2  s−1) and stress tolerance
(Gómez et al. 2019). Both saturating irradiance levels match with the subtidal and
intertidal prevailing light conditions. Similar patterns were detected in seaweeds
from King George Island, with saturation irradiances of 55 μmol m−2 s−1 in subtidal
and 120  μmol  m−2  s−1 in eulittoral species (Huovinen and Gómez 2013). For
instance, photosynthesis of Desmarestia antarctica, D. anceps and Gymnogongrus
antarcticus was sensitive when exposed to higher light conditions, especially to
UVB (Bischof et al. 1998). In this context, seaweeds from subtidal environments
that form drifting patches can deteriorate quickly, thus contributing to an important
carbon source for higher trophic levels.
High concentrations of soluble phlorotannins are found in C. jacquinotii (>5% of
DW in both vegetative and reproductive tissues), acting as UVR shield and herbi-
vore deterrent (Iken et al. 2007; Huovinen and Gómez 2015; see also Chap. 17 by
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 75

Amsler et  al. and Chap. 18 by Gómez and Huovinen), thereby contributing to a
higher floating persistence.
Environmental factors that affect seaweed physiology and growth under benthic
and floating conditions, can also modify the reproductive capacity. Considering the
short-life and dispersal distances of seaweed propagules (Santelices 1990), the abil-
ity to remain reproductively competent during rafting and when arriving on a new
site is crucial for a successful dispersal and colonization. Long floating times
together with high temperature and radiation conditions decreas the availability of
propagules and can increas the disintegration of reproductive tissues such as shown
for D. antarctica (Tala et  al. 2016, 2019), M. pyrifera (Macaya et  al. 2005;
Hernández-Carmona et al. 2006; Rothäusler et al. 2011b) and Hormosira banksii
(McKenzie and Bellgrove 2008). Rafting and long distance dispersal might be facil-
itated in monoecious species or in dioecious species (e.g. D. antarctica and C. jac-
quinotti) when male and female individuals are traveling together and arriving to a
new suitable habitat (Lizée-Prynne et al. 2016).
Presumably the establishment of non-native species within the Antarctic shores is
prevented by the physiological and reproductive tolerance ability to extreme envi-
ronmental conditions rather than transport. Low temperatures, tides and wave
action, ice-scouring, salinity changes, sediment and detritus accumulation may limit
the establishment of foreign taxa on hard substrates (Campana et al. 2009). Biotic
interactions associated to local herbivores and competition can also determine the
colonization success of the colonization. However, under a warming climate sce-
nario, more sub-Antarctic and temperate species might be able to grow and survive
at higher latitudes. However, if the substrate availability is low at the new site of
arrival or the the habitat is highly diverse, only strong competitors may have the
chance to establish (Stachowicz et al. 2002). The density-dependent blocking mech-
anism  of conspecifics has been described to prevent the colonization of floating
D. antarctica, due to low space for new recruits and the low contribution of new
gametes in relation to those produced by local individuals (Waters et  al. 2013).
Future warming scenarios might facilitate colonization by floating seaweeds, reduc-
ing competition with local species and increasing substrate availability.

Acknowledgements  We thank the editors for inviting us to contribute to this review. Financial
support was received through Fondap-IDEAL grant 15150003 to ECM and FONDECYT grant
1131023 to FT and ECM, CONICYT PAI (Chile) 79160069 to FT.

References

Abé H, Komatsu T, Kokubu Y, Alabsi N, Rothäusler E, Shishido H, Yoshizawa S, Ajisaka T (2013)


Invertebrate fauna associated with floating Sargassum horneri (Fucales: Sargassaceae) in the
East China Sea. Spec Diver 18(1):75–85
Amsler CD, McClintock JB, Baker BJ (1999) An antarctic feeding triangle: defensive interactions
between macroalgae, sea urchins, and sea anemomes. Mar Ecol Prog Ser 183:105–114
76 E. C. Macaya et al.

Amsler CD, Mcclintock JB, Baker BJ (2012) Palatability of living and dead detached Antarctic
macroalgae to consumers. Antarct Sci 24(6):589–590
Assmy P, Ehn JK, Fernández-Méndez M, Hop H, Katlein C, Sundfjord A, Bluhm K, Daase M,
Engel A, Fransson A, Granskog MA, Hudson SR, Kristiansen S, Nicolaus M, Peeken I, Renner
AHH, Spreen G, Tatarek A, Wiktor J (2013) Floating ice-algal aggregates below melting Arctic
Sea ice. PLoS One 8(10):e76599
Avila C, Angulo-Preckler C, Martín-Martín RP, Figuerola B, Griffiths HJ, Waller CL (2020)
Invasive marine species discovered on non–native kelp rafts in the warmest Antarctic island.
Sci Rep 10:1639
Bäck S, Lehvo A, Blomster J (2000) Mass occurrence of unattached Enteromorpha intestinalis on
the Finnish Baltic Sea Coast. Ann Bot Fenn 37:155–161
Barnes DKA, Linse K, Waller C, Morely S, Enderlein P, Fraser KPP, Brown M (2006) Shallow
benthic fauna communities of South Georgia Island. Polar Biol 29:223–228
Batista MB, Batista AA, Franzan SP, Simionatto PP, Lima ST, Velez-Rubio G, Scarabino F,
Camacho O, Schmitz C, Martinez A (2018) Kelps’ long-distance dispersal: role of ecologi-
cal/oceanographic processes and implications to marine forest conservation. Diversity 10(11)
Belt ST, Brown TA, Smik L, Assmy P, Mundy CJ (2018) Sterol identification in floating
Arctic sea ice algal aggregates and the Antarctic sea ice diatom Berkeleya adeliensis. Org
Geochem 118:1–3
Bischof K, Hanelt D, Wiencke C (1998) UV-radiation can affect depth-zonation of Antarctic mac-
roalgae. Mar Biol 131:597–605
Bischoff-Bäsmann B, Wiencke C (1996) Temperature requirements for growth and survival of
Antarctic Rhodophyta. J Phycol 32(4):525–535
Blomster J, Maggs CA, Stanhope MJ (1998) Molecular and morphological analysis of
Enteromorpha intestinalis and E. compressa (Chlorophyta) in the British isles. J Phycol
34(2):319–340
Boo GH, Mansilla A, Nelson W, Bellgrove A, Boo SM (2014) Genetic connectivity between trans-­
oceanic populations of Capreolia implexa (Gelidiales, Rhodophyta) in cool temperate waters
of Australasia and Chile. Aquat Bot 119:73–79
Braeckman U, Pasotti F, Vázquez S, Zacher K, Hoffmann R, Elvert M, Marchant H, Buckner C,
Quartino ML, Mác Cormack W, Soetaert K, Wenzhöfer F, Vanreusel A (2019) Degradation of
macroalgal detritus in shallow coastal Antarctic sediments. Limnol Oceanogr 64(4):1423–1441
Brouwer PEM (1996) Decomposition in situ of the sublittoral Antarctic macroalga Desmarestia
anceps Montagne. Polar Biol 16(2):129–137
Brouwer PEM, Geilen EFM, Gremmen NJM, van Lent F (1995) Biomass, cover and zonation
patterns of sublittoral macroalgae at Signy Island, South Orkney Islands, Antarctica. Bot Mar
38:259–270
Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML, Wiencke C (2009) Drivers
of colonization and succession in polar benthic macro- and microalgal communities. Bot Mar
52(6):655–667
Cefarelli AO, Vernet M, Ferrario ME (2011) Phytoplankton composition and abundance in relation
to free-floating Antarctic icebergs. Deep-Sea Res II Top Stud Oceanogr 58:1436–1450
Clayton MN, Wiencke C, Klöser H (1997) New records of temperate and sub Antarctic marine
benthic macroalgae from Antarctica. Polar Biol 17(2):141–149
Cohen RA, Fong P (2004) Physiological responses of a bloom-forming green macroalga to short-­
term change in salinity, nutrients, and light help explain its ecological success. Estuaries
27(2):209–216
Coyer JA, Hoarau G, Van Schaik J, Luijckx P, Olsen JL (2011) Trans-Pacific and trans-Arctic
pathways of the intertidal macroalga Fucus distichus L. reveal multiple glacial refugia and
colonizations from the North Pacific to the North Atlantic. J Biogeogr 38(4):756–771
Cumming RA, Nikula R, Spencer HG, Waters JM (2014) Transoceanic genetic similarities
of kelp-associated sea slug populations: long-distance dispersal via rafting? J Biogeogr
41(12):2357–2370
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 77

De Wildeman É (1935) Observations sur la algues rapporlées par l’expedition animctiq’ue de la


“Belgica”. Résullanls du voyage de la “Belgica” en 897-1899. Ansvers, p. 47
Drew E, Hastings R (1992) A year-round ecophysiological study of Himantothallus grandifolius
(Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31(3–4):262–277
Eggert A (2012) Seaweed responses to temperature. In: Wiencke C, Bischof K (eds) Seaweed
biology: novel insights into ecophysiology, ecology and utilization. Springer-Verlag, Berlin,
pp 47–66
Eggert A, Wiencke C (2000) Adaptation and acclimation of growth and photosynthesis of five
Antarctic red algae to low temperatures. Polar Biol 23(9):609–618
Fischer G, Wiencke C (1992) Stable carbon isotope composition, depth distribution and fate of
macroalgae from the Antarctic Peninsula region. Polar Biol 12(3):341–348
Flores‐Molina MR, Rautenberger R, Muñoz P, Huovinen P, Gómez I (2016) Stress tolerance of the
endemic Antarctic brown alga Desmarestia anceps to UV radiation and temperature is medi-
ated by high concentrations of phlorotannins. Photochem Photobiol 92(3):455–466
Fraser CI, Waters JM (2013) Algal parasite Herpodiscus durvillaeae (Phaeophyceae: Sphacelariales)
inferred to have traversed the Pacific Ocean with its buoyant host. J Phycol 49(1):202–206
Fraser CI, Nikula R, Spencer HG, Waters JM (2009) Kelp genes reveal effects of subantarctic sea
ice during the Last Glacial Maximum. P Natl Acad Sci USA 106(9):3249–3253
Fraser CI, Thiel M, Spencer HG, Waters JM (2010) Contemporary habitat discontinuity and his-
toric glacial ice drive genetic divergence in Chilean kelp. BMC Evol Biol 10:203
Fraser CI, Zuccarello GC, Spencer HG, Salvatore LC, Garcia GR, Waters JM (2013) Genetic
affinities between trans-oceanic populations of non-buoyant macroalgae in the high latitudes
of the southern hemisphere. PLoS One 8(7):e69138
Fraser CI, Kay GM, Md P, Ryan PG (2017) Breaking down the barrier: dispersal across the
Antarctic Polar Front. Ecography 40(1):235–237
Fraser CI, Morrison AK, Hogg AM, Macaya EC, van Sebille E, Ryan PG, Padovan A, Jack C,
Valdivia N, Waters JM (2018) Antarctica’s ecological isolation will be broken by storm-driven
dispersal and warming. Nat Clim Chang 8(8):704
Gain L (1912) La flore algologique des régions antarctiques et subantarctiques.—Deuxième
Expédition Antarctique Française (1908–1910) commandée par le Dr Jean Charcot: sciences
naturelles. Sci Nat Doc Sci-Paris:156–218
Gallardo T, Pérez-Ruzafa IM, Flores-Moya A, Conde F (1999) New collections of Benthic marine
algae from Livingston and Deception Islands (South Shetland Islands) and Trinity Island
(Bransfield Strait). Bot Mar 42:61–69
Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U, Quartino ML, Dunton K, Wiencke C
(2009) Light and temperature demands of marine benthic microalgae and seaweeds in polar
regions. Bot Mar 52(6):593–608
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27
Graiff A, Karsten U, Meyer S, Pfender D, Tala F, Thiel M (2013) Seasonal variation in floating
persistence of detached Durvillaea antarctica (Chamisso) Hariot thalli. Bot Mar 56(1):3–14
Graiff A, Pantoja JF, Tala F, Thiel M (2016) Epibiont load causes sinking of viable kelp rafts: sea-
sonal variation in floating persistence of giant kelp Macrocystis pyrifera. Mar Biol 163(9):191
Guillemin M-L, Valero M, Faugeron S, Nelson W, Destombe C (2014) Tracing the trans-pacific
evolutionary history of a domesticated seaweed (Gracilaria chilensis) with archaeological and
genetic data. PLoS One 9(12):e114039
Häder D-P, Lebert M, Helbling EW (2001) Effects of solar radiation on the Patagonian mac-
roalga Enteromorpha linza (L.) J. Agardh — Chlorophyceae. J Photochem Photobiol B Biol
62(1):43–54
Hernández-Carmona G, Hughes B, Graham MH (2006) Reproductive longevity of drifting kelp
Macrocystis pyrifera (Phaeophyceae) in Monterey Bay, USA. J Phycol 42(6):1199–1207
78 E. C. Macaya et al.

Hinojosa IA, Boltaña S, Macaya E, Ugalde P, Valdivia N, Vásquez N, Newman W, Thiel M (2006)
Geographic distribution and description of four pelagic barnacles along the south east Pacific
coast of Chile – a zoogeographical approximation. Rev Chil Hist Nat. 78: 603–614.
Hobday A (2000) Age of drifting Macrocystis pyrifera (L.) C.  Agardh rafts in the Southern
California Bight. J Exp Mar Biol Ecol 253:97–114
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52(6): 509–534
Hooker JD (1844) The botany of the Antarctic voyage of H.  M. discovery ships Erebus and
Terror in the years 18391843, vol. 1, Flora Antarctica, Part 1 Botany of Lord Auklands Group
and Campbell’s Island, Part 2 Botany of Fuegia, The Falklands, Kerguelens Land. Reeve
Brothers, London
Huang Y, Amsler M, McClintock J, Amsler C, Baker B (2007) Patterns of gammarid amphipod
abundance and species composition associated with dominant subtidal macroalgae along the
western Antarctic Peninsula. Polar Biol 30:1417–1430
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36(9):1319–1332
Huovinen P, Gómez I (2015) UV Sensitivity of vegetative and reproductive tissues of two Antarctic
brown algae is related to differential allocation of phenolic substances. Photochem Photobiol
91(6):1382–1388
Iken K, Amsler CD, Hubbard JM, McClintock JB, Baker BJ (2007) Allocation patterns of phloro-
tannins in Antarctic brown algae. Phycologia 46(4):386–395
Ingólfsson A (1998) Dynamic of macrofaunal communities of floating seaweed clumps off west-
ern Iceland: a study of patches on the surface of the sea. J Exp Mar Biol Ecol 23:119–137
Kang J-S, Kang S-H, Lee JH, Lee S (2002) Seasonal variation of microalgal assemblages at a fixed
station in King George Island, Antarctica, 1996. Mar Ecol Prog Ser 229:19–32
Katlein C, Fernández-Méndez M, Wenzhöfer F, Nicolaus M (2015) Distribution of algal aggre-
gates under summer sea ice in the Central Arctic. Polar Biol 38:719–731
Khalaman V, Berger VY (2006) Floating seaweeds and associated fauna in the White Sea.
Oceanology 46:827–833
Klöser H, Ferreyra G, Schloss I, Mercuri G, Laturnus F, Curtosi A (1993) Seasonal variation of
algal growth conditions in sheltered Antarctic bays: the example of Potter Cove (King George
Island, South Shetlands). J Mar Syst 4(4):289–301
Lastra M, Rodil IF, Sánchez-Mata A, García-Gallego M, Mora J (2014) Fate and processing of mac-
roalgal wrack subsidies in beaches of Deception Island, Antarctic Peninsula. J Sea Res 88:1–10
Lizée-Prynne D, López B, Tala F, Thiel M (2016) No sex-related dispersal limitation in a dioecious,
oceanic long-distance traveller: the bull kelp Durvillaea antarctica. Bot Mar 59(1):39–50
López BA, Macaya EC, Rivadeneira MM, Tala F, Tellier F, Thiel M (2018) Epibiont communities
on stranded kelp rafts of Durvillaea antarctica (Fucales, Phaeophyceae)–do positive interac-
tions facilitate range extensions? J Biogeogr 45(8):1833–1845
Macaya EC, Zuccarello GC (2010) DNA Barcoding and genetic divergence in the giant kelp
Macrocystis (Laminariales). J Phycol 46(4):736–742
Macaya EC, Boltana S, Hinojosa IA, Macchiavello JE, Valdivia NA, Vásquez NR, Buschmann
AH, Vásquez JA, Vega JMA, Thiel M (2005) Presence of sporophylls in floating kelp rafts
of Macrocystis spp. (Phaeophyceae) along the Chilean Pacific coast. J Phycol 41(5):913–922
Macaya EC, López B, Tala F, Tellier F, Thiel M (2016) Float and raft: role of buoyant seaweeds in
the phylogeography and genetic structure of non-buoyant associated flora. In: Hu ZM, Fraser
C (eds) Seaweed phylogeography. Springer, pp 97–130
McKenzie PF, Bellgrove A (2008) Dispersal of Hormosira banksii (Phaeophyceae) via detached
fragments: reproductive viability and longevity. J Phycol 44(5):1108–1115
Mélice J, Lutjeharms J, Rouault M, Ansorge I (2003) Sea-surface temperatures at the sub-­Antarctic
islands Marion and Gough during the past 50 years. S Afr J Sci 99(7–8):363–366
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 79

Montagne JFC (1842) Prodromus generum specierumque phycearum novarum, in itinere ad polum
antarcticum...ab illustri Dumont d’Urville peracto collectarum, notis diagnosticis tantum huc
evulgatarum, descriptionibus verò fusioribus nec non iconibus analyticis iam iamque illustran-
darum. Paris apud Gide, editorem, Parisiis, pp. 1–16
Moon KL, Chown SL, Fraser CI (2017) Reconsidering connectivity in the sub-Antarctic. Biol Rev
92(4):2164–2181
Müller DG, Westermeier R, Peters A, Boland W (1990) Sexual reproduction of the Antarctic
brown alga Ascoseira mirabilis (Ascoseirales, Phaeophyceae). Bot Mar 33:251–255
Müller DG, Ramírez ME, Westermeier R (1992) Utriculidium durvillei (Bory?) Skottsberg en isla
Rey Jorge, Antártica. Ser Cien INACH 42:47–52
Neushul M (1963) Reproductive morphology of Antarctic kelps. Bot Mar 5:19–24
Neushul M (1965) Diving observations of subtidal Antarctic marine vegetation. Bol Mar
8(2–4):234–243
Nikula R, Fraser CI, Spencer HG, Waters JM (2010) Circumpolar dispersal by rafting in two sub-
antarctic kelp-dwelling crustaceans. Mar Ecol Prog Ser 405:221–230
Nikula R, Spencer HG, Waters JM (2013) Passive rafting is a powerful driver of transoceanic gene
flow. Biol Lett 9(1):20120821
Norkko A, Thrush SF, Cummings VJ, Funnell GA, Schwarz A-M, Andrew NL, Hawes I (2004)
Ecological role of Phyllophora antarctica drift accumulations in coastal soft-sediment com-
munities of McMurdo Sound, Antarctica. Polar Biol 27(8):482–494
O’Connor MI (2009) Warming strengthens an herbivore-plant interaction. Ecology 90:388–398.
https://doi.org/10.1890/08-0034.1
Oliveira EC, Absher TM, Pellizzari FM, Oliveira MC (2009) The seaweed flora of Admiralty Bay,
King George Island, Antarctic. Polar Biol 32(11):1639–1647
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa LH,
Colepicolo P (2017) Diversity and spatial distribution of seaweeds in the South Shetland
Islands, Antarctica: an updated database for environmental monitoring under climate change
scenarios. Polar Biol 40:1671–1685
Quartino ML, Zaixso HE, Boraso de Zaixso AL (2005) Biological and environmental characteriza-
tion of marine macroalgal assemblages in Potter Cove, South Shetland Islands, Antarctica. Bot
Mar 48(3):187–197
Rakusa-Suszczewski S (1980) Environmental conditions and the functioning of Admiralty Bay
(South Shetland Islands) as part of the near shore Antarctic ecosystem. Pol Polar Res 1:11–27
Rakusa-Suszczewski S (1995) Flow of matter in the Admiralty Bay area, King George Island,
Maritime Antarctic. Proc NIPR Symp Polar Biol 8:101–113
Rautenberger R, Huovinen P, Gómez I (2015) Effects of increased seawater temperature on UV
tolerance of Antarctic marine macroalgae. Mar Biol 162(5):1087–1097
Richardson MG (1979) The distribution of Antarctic marine macro-algae related to depth and
substrate. Br Antarct Surv Bull 49:1–13
Ross JC (1847) A voyage of discovery and research in the Southern and Antarctic Regions, during
the Years 1839–43, vol 2. John Murray
Rothäusler E, Gómez I, Hinojosa IA, Karsten U, Tala F, Thiel M (2009) Effect of temperature
and grazing on growth and reproduction of floating Macrocystis spp.(Phaeophyceae) along a
latitudinal gradient. J Phycol 45(3):547–559
Rothäusler E, Gómez I, Hinojosa IA, Karsten U, Miranda L, Tala F, Thiel M (2011a) Kelp rafts in
the Humboldt Current: interplay of abiotic and biotic factors limit their floating persistence and
dispersal potential. Limnol Oceanogr 56(5):1751–1763
Rothäusler E, Gomez I, Hinojosa IA, Karsten U, Tala F, Thiel M (2011b) Physiological perfor-
mance of floating giant kelp Macrocystis pyrifera (Phaeophyceae): latitudinal variability in the
effects of temperature and grazing. J Phycol 47(2):269–281
Rothäusler E, Gomez I, Karsten U, Tala F, Thiel M (2011c) Physiological acclimation of floating
Macrocystis pyrifera to temperature and irradiance ensures long-term persistence at the sea
surface at mid-latitudes. J Exp Mar Biol Ecol 405(1–2):33–41
80 E. C. Macaya et al.

Rothäusler E, Gutow L, Thiel M (2012) Floating seaweeds and their communities. In: Wiencke C,
Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and utilization.
Springer-Verlag, Berlin, pp 359–380
Rothäusler E, Reinwald H, López BA, Tala F, Thiel M (2018) High acclimation potential in float-
ing Macrocystis pyrifera to abiotic conditions even under grazing pressure–a field study. J
Phycol 54(3):368–379
Santelices B (1990) Patterns of reproduction dispersal and recruitment in seaweeds. Oceanogr Mar
Biol Ann Rev 28:177–276
Sato T, Sakurai H, Takasaki A, Watanabe K, Hirano Y (1992) Underwater observation of Antarctic
fishes and invertebrates with a note on collecting and transportation techniques for research and
exhibition in the aquarium. Ser Cient INACH 42:95–103
Schwarz A-M, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photo-
synthesis near the southern global limit for growth; Cape Evans, Ross Sea, Antarctica. Polar
Biol 26(12):789–799
Skottsberg K (1907) Zur Kenntnis der subantarktischen und antarktischen Meeresalgen.
I.  Phaeophyceen. In: Nordenskjöld O (ed) Wissenschaftliche Ergebnisse der Schwedischen
Südpolar-Expedition 1901–1903, vol 4: 1. Lithographisches Institut des Generalstabs,
Stockholm, pp 1–172
Smith SDA (2002) Kelp rafts in the Southern Ocean. Glob Ecol Biogeogr 11(1):67–69
Stachowicz JJ, Fried H, Osman RW, Whitlatch RB (2002) Biodiversity, invasion resistance, and
marine ecosystem function: reconciling pattern and process. Ecology 83(9):2575–2590
Tada S, Sato T, Sakurai H, Arai H, Kimpara I, Kodama M (1996) Benthos and fish community asso-
ciated with clumps of submerged drifting algae in Fildes Bay, King George Island, Antarctica.
Proc NIPR Symp Polar Biol 9:243–251
Tala F, Gómez I, Luna-Jorquera G, Thiel M (2013) Morphological, physiological and reproductive
conditions of rafting bull kelp (Durvillaea antarctica) in northern-central Chile (30º S). Mar
Biol 160(6):1339–1351
Tala F, Velásquez M, Mansilla A, Macaya EC, Thiel M (2016) Latitudinal and seasonal effects on
short-term acclimation of floating kelp species from the South-East Pacific. J Exp Mar Biol
Ecol 483:31–41
Tala F, López BA, Velásquez M, Jeldres R, Macaya EC, Mansilla A, Ojeda J, Thiel M (2019) Long-­
term persistence of the floating bull kelp Durvillaea antarctica from the South-East Pacific:
potential contribution to local and transoceanic connectivity. Mar Environ Res 149:67–79
Thiel M, Gutow L (2005a) The ecology of rafting in the marine environment. I. The floating sub-
strata. Oceanogr Mar Biol Ann Rev 42:181–263
Thiel M, Gutow L (2005b) The ecology of rafting in the marine environment. II. The rafting organ-
isms and community. Oceanogr Mar Biol Ann Rev 43:279–418
Valdivia N, Díaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9(6)
Vandendriessche S, Messiaen M, O’Flynn S, Vincx M, Degraer S (2007a) Hiding and feeding in
floating seaweed: floating seaweed clumps as possible refuges or feeding grounds for fishes.
Estuar Coast Shelf Sci 71(3–4):691–703
Vandendriessche S, Vincx M, Degraer S (2007b) Floating seaweed and the influences of tem-
perature, grazing and clump size on raft longevity—a microcosm study. J Exp Mar Biol Ecol
343(1):64–73
Waters JM, Fraser CI, Hewitt GM (2013) Founder takes all: density-dependent processes structure
biodversity. Trends Ecol Evol 28(2):78–85
Webster W (1834) Narrative of a voyage to the southern Atlantic Ocean in the years 1828, 29, 30,
performed in HM sloop Chanticleer under the command of the late Captain Henry Foster, FRS
& c. vol II. London, Richard Bentley, 1–398
Westermeier R, Gómez I, Rivera PJ, Müller DG (1992) Macroalgas marinas antárticas: distribu-
ción, abundancia y necromasa en isla Rey Jorge, ShetIand del Sur, Antártica Ser. Cient INACH
42:21–34
4  Detached Seaweeds as Important Dispersal Agents Across the Southern Ocean 81

Weykam G, Gómez I, Wiencke C, Iken K, Klöser H (1996) Photosynthetic characteristics and C:N
ratios of macroalgae from King George Island (Antarctica). J Exp Mar Biol Ecol 204:1–22
Wichmann CS, Hinojosa IA, Thiel M (2012) Floating kelps in Patagonian Fjords: an important
vehicle for rafting invertebrates and its relevance for biogeography. Mar Biol 159(9):2035–2049
Wiencke C, Rahmel J, Karsten U, Weykam G, Kirst GO (1993) Photosynthesis of marine macroal-
gae from Antarctica: light and temperature requirements. Bot Acta 106(1):78–87
Wulff A, Iken K, Quartino M, Al-Handal A, Wiencke C, Clayton M (2009) Biodiversity, biogeog-
raphy and zonation of marine benthic micro-and macroalgae in the Arctic and Antarctic. Bot
Mar 56:491–507
Zacher K, Roleda MY, Hanelt D, Wiencke C (2007) UV effects on photosynthesis and DNA in
propagules of three Antarctic seaweeds (Adenocystis utricularis, Monostroma hariotii and
Porphyra endiviifolium). Planta 225(6):1505–1516
Zacher K, Rautenberger R, Hanelt D, Wulff A, Wiencke C (2009) The abiotic environment of polar
marine benthic algae. Bot Mar 52(6):483–490
Zemko K, Pabis K, Siciński J, Błażewicz-Paszkowycz M (2015) Diversity and abundance of isopod
fauna associated with holdfasts of the brown alga Himantothallus grandifolius in Admiralty
Bay, Antarctic. J Pol Polar Res 36(4):405–415
Zielinski K (1981) Benthic macroalgae of Admiralty Bay (King George Island, South Shetland
Islands) and circulation of algal matter between the water and the shore. J Pol Polar Res
2(3–4):71–94
Zielinski K (1990) Bottom macroalgae of the Admiralty Bay (King George Island, South Shetlands,
Antarctica). J Pol Polar Res 11:95–131
Chapter 5
Biogeography of Antarctic Seaweeds
Facing Climate Changes

Franciane Pellizzari, Luiz Henrique Rosa, and Nair S. Yokoya

Abstract The seaweed biogeography and diversity in remote areas, such as


Antarctica, should be reassessed considering the population shifts induced by global
changes. This chapter addresses the hypothesis that ecological isolation can be dis-
rupted and that biogeographical distribution of some species could be altered by
thermohaline changes, which in turn would alter the dispersal patterns of macroal-
gae driven by severe meteorological and oceanographic events. Algal growth and
distribution are limited by physical and biological processes, acting as sensitive
bioindicators of changes or abrupt oscillations in the environmental regimes. In
addition, Antarctica represents a natural laboratory highly susceptible to the climate
changes, and the monitoring of their ecosystems may help to predict their potential
effects beyond the Southern Ocean. Another fundamental issue is to understand the
increase in species richness due to the cryptic and alien species, considering shifts
in their biogeographic distribution. The large-scale patterns of some of these species
reported for Antarctica may provide clues to reevaluate aspects of endemism, bio-
logical corridors, ecotone, and expansion of geographical distribution of algal
assemblages facing climate changes, reinforcing the hypothesis that these isolated
ecosystems will become gradually more connected.

Keywords  Alien species · Biogeographic processes · Deception Island ·


Endemism · Seaweed diversity

F. Pellizzari (*)
Universidade Estadual do Paraná, Campus Paranaguá, Laboratório de Ficologia e Qualidade
de Água Marinha, Paranaguá, Brazil
e-mail: franciane.pellizzari@unespar.edu.br
L. H. Rosa
Universidade Federal de Minas Gerais, Departamento de Microbiologia,
Belo Horizonte, Brazil
e-mail: ihrosa@icb.ufmg.br
N. S. Yokoya
Instituto de Botânica, Núcleo de Pesquisa em Ficologia, São Paulo, Brazil
e-mail: nyokoya@hotmail.com

© Springer Nature Switzerland AG 2020 83


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_5
84 F. Pellizzari et al.

5.1  The Abiotic Setting of the Southern Ocean

The Southern Ocean (SO) biota distribution is the result of major geological, ocean-
ographic, and climate changes during the last 50 Ma, and there is a paradox between
the broad distributions of some species and their inherently poor dispersal capacity,
although marine biogeographic studies supported by molecular studies are increas-
ingly revealing examples of this paradox and indicating that long-distance dispersal
of macroalgae is possible (Fraser et al. 2013).
The seaweed processes in the Antarctic and sub-Antarctic regions are strongly
limited and correlated with environmental parameters (Wiencke and Amsler 2012).
In contrast with the Arctic Ocean, the Southern Ocean has no land bridge to temper-
ate regions since the late Mesozoic and has been further separated from the southern
continents by the Antarctic Circumpolar Current (ACC) since ca. 26 Ma (Kirst and
Wiencke 1995). The eastward movement of the ACC or the West Wind Drift (WWD)
had strongly defined the oceanography in the Southern Hemisphere and conse-
quently the diversity, biogeography, and ecology of seaweeds in all the adjacent
coastal regions (Orsi et al. 1995; Huovinen and Gómez 2012). Thus, biogeographic
barriers and ecological corridors are fundamental concepts to understand macroeco-
logical and evolutionary processes, since that ocean circulation, considering present
and past patterns of continental drift, may isolate or connect groups of marine
organisms, including seaweeds.
The Southern Ocean encircling Antarctica, whose natural boundary is the
Antarctic Convergence, is known as the most extreme environment on the Earth.
Considering ocean circulation in the SO, the ACC is permeated by the Southern
Antarctic Circumpolar Front, located between the Polar Front (50°–60°S) and sur-
roundings of the Antarctic coastline, representing the disjunction of Antarctic waters
and resulting in distinct thermal stratifications for each zone (Orsi et  al. 1995;
Sanches et al. 2016). In the southern hemisphere, two circumpolar fronts define the
boundaries of the cold-temperate region (Fig. 5.1a): (1) the Antarctic Polar Front
(APF or Antarctic Convergence), the southern limit, characterized by cold surface
waters (ranging 3–5°C), and (2) the Subtropical Front (STF or Subtropical
Convergence), the northern limit that separates subtropical water in the north from
sub-Antarctic water, showing ca. of the 10°C and 15°C surface temperature, during
winter and summer, respectively (Huovinen and Gómez 2012). Thus the ACC,
being the largest and deepest current in the world, was considered in several past
studies as a physical barrier that could limit passive dispersal of new species to
Antarctica (Wells et al. 2011; Pritchard et al. 2012; Sanches et al. 2016).
In a macroscale, Longhurst (2007) identified four “biogeochemical” provinces in
the Southern Ocean considering physical and hydrological properties of these cur-
rents: South Subtropical Convergence, sub-Antarctic water ring, Antarctic, and
Polar Southern provinces. On a regional scale, the dynamics of ice formation and/or
its presence as drifting sheets or icebergs, substrate type, and current regime are the
main factors that set biogeographical boundaries for biota.
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 85

Fig. 5.1 (a) Antarctica contextualized in the Southern Ocean, showing continental proximities
and major oceanographic fronts: subtropical front (STF), Antarctic Polar Front (APF), and
Antarctic Circumpolar Current (ACC). (b) Detailed map indicating the geographical position of
the SSI South Shetland Islands, including Deception, WAP Western Antarctic Peninsula, and EAP
Eastern Antarctic Peninsula. (Photo by Franciane Pellizzari)

5.2  Biogeographic Patterns

Due to the isolation, the SO presents higher endemism of seaweeds, although


lower richness than some sub-Antarctic islands and continental areas from the
Southern Atlantic and Pacific. The Eastern Antarctic Peninsula (EAP) and south-
ern regions from the Western Antarctic Peninsula (WAP) present low species rich-
ness when compared to the South Shetland Islands (SSI – Fig. 5.1b) (Pellizzari
et al. 2017) (see also Chap. 2 by Oliveira et al.). However, Wulff et al. (2009),
comparing the biodiversity, biogeography, and zonation of seaweeds from the
Arctic and Antarctic regions, reported that the diversity and number of endemic
species are higher in Antarctica than in the Arctic. Wiencke and Clayton (2002)
mentioned that about a third of the 120 species recorded to the Antarctic region
are endemic (Phaeophyceae 44%, Rhodophyta 32%, and Chlorophyta 18%).
These differences are explained by the biogeographical histories of both regions
(Wiencke and Clayton 2002).
86 F. Pellizzari et al.

Temperature is the most important factor controlling biogeographic distribution


of macroalgae, and then shifts in their distribution are inevitable during periods of
global changes. However, other factors must be considered. For example, rafting of
floating organisms and objects appears to have a critical role in facilitating long-­
distance dispersal and structuring intertidal ecosystems in Antarctica (Fraser et al.
2013). Many species that lack obvious transoceanic dispersal ability (e.g., brood-
ing invertebrates, small non-buoyant macroalgae, and terrestrial vertebrates) are
nonetheless evidently able to disperse long distances attached in floating wood or
detached kelp (see also Chap. 3 by Fraser et al.). Also, aiming to avoid generalized
biogeography of the SO benthos, Griffiths et al. (2009) studied distributional data
of mollusks, cheilostome, cyclostome, and bryozoans (species richness, rates of
endemism, patterns of radiation) in 29 provinces in the South American, South
African, Tasmanian, New Zealand, sub-Antarctic, and Antarctic regions. The
authors reported high species numbers in New Zealand, Tasmania, and South
Africa and low species richness in South America. In contrast, no difference was
found in richness between the east and west parts of the Southern Ocean (called
“single functional unit”). Besides evidence of strong faunal links between the
Antarctica and South America, biogeographical regions in the SO differ depending
upon the class of organism, which is being considered. Linse et al. (2006) indicated
that patterns of diversity and endemism were very different between the bivalves
and gastropods. Moreover, Pierrat et  al. (2013), in a comparative study among
echinoids, bivalves, and gastropods from the Antarctic, sub-Antarctic, and cold-
temperate areas, reported a significant advantage of the BSN (bootstrapped span-
ning network) procedure in the identification of faunal similarities among
biogeographical regions and transitional areas, considering the following faunal
provinces: (1) New Zealand, (2) southern South America, (3) east sub-Antarctic
islands, (4) West Antarctica, and (5) East Antarctica. Strong faunal relationships
perfectly match the flows of the ACC and Antarctic Coastal currents, suggesting
strong connections and groupings between bioregions. However, Antarctic and
sub-Antarctic regions are regarded as distinct biogeographical regions, with pat-
terns driven by a small number of widely distributed species (Griffiths and Waller
2016). Gutt et  al. (2016) studied macroepibenthic communities in Weddell Sea,
Bransfield Strait, and Drake Passage and suggested that large-scale patterns in
pelagic organisms segregate three eco-regions possibly correlated with ice pres-
ence and depth. Keith et  al. (2011), using a geographically weighted regression
(GWR) model, determined that some environmental conditions could predict
genus richness in macroalgae and confirmed the hypothesis that environmental
tolerances influence species distributions more strongly at higher latitudes, whereas
biotic interactions play a more prominent role in the tropics. This chapter discusses
in that extent shifts in seaweed diversity and distribution in the SO may be also
associated to the rapid climatic changes, resulting in higher connections or disrup-
tion of these previously established provinces.
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 87

5.3  S
 eaweed Assemblages: Are Antarctic Seaweed Diversity
and Richness Changing?

Shifts in seaweed distribution were important during periods of changing tempera-


tures as demonstrated in the geological past during the last glaciation (Crame 1994).
Similar changes may be expected during the current period of global warming with
a significant impact on marine communities. In addition, Antarctic macroalgal com-
munities will presumably expand north and also southward due to less strong ice
abrasion and improved light conditions, and the new ice-free areas appearing due to
glacier retreat will be probably colonized by seaweeds (Quartino et  al. 2013;
Wiencke et  al. 2014 and studies summarized in Pellizzari et  al. 2017) (see also
Chap. 8 by Quartino et al.).
Climate changes in Antarctic Peninsula are resulting in shifts of marine popula-
tions (summarized in Ducklow et al. 2013). These authors suggested that macroal-
gal assemblages show lower biomass and diversity in the southern region of WAP
when compared to the northern assemblages. Additionally, they mentioned that the
assemblages of the northern WAP seem to be expanding to the south as the annual
sea ice declines.
Physico-chemical changes, mainly due to increased surface seawater tempera-
tures (SST) in the SO, result in a potential driver to shifts in seaweed assemblages
(Müller et al. 2009; Sangil et al. 2012; Wernberg et al. 2016; Pellizzari et al. 2017).
Based on a multivariate analysis of Antarctic and sub-Antarctic seaweed distribu-
tion focusing in an evaluation of the role of the ACC, Sanches et al. (2016) found
differences in the distributional patterns of species and genera in relation to the
zones and sections connected with South America (1: north), influenced by the Ross
Sea Gyre (2: Western Antarctica Peninsula) and by the Weddell Sea Gyre (3: Eastern
Antarctica Peninsula). The predicted latitudinal gradient of species richness was
confirmed, and two main biogeographical clusters were defined: (1) subtropical and
sub-Antarctica and (2) Antarctic Peninsula and surroundings of continental
Antarctic. These differences in diversity patterns between both regions suggest the
existence of a species distributional flux. This may result either from natural disper-
sion or due to global changes, suggesting that Antarctica may not be as isolated and
as suggested in Chap 3 of this book and by Fraser et al. (2019).
Recently, high seaweed richness among the South Shetland Islands (SSI), high-
lighted by six new records for the area, four cryptogenic species and two putative
new species, was reported (Pellizzari et al. 2017). It is noteworthy that the 104 spe-
cies of benthic marine algae listed by these authors represent ~82% of all seaweeds
described to Antarctica. The majority of these new records are cryptic species, rep-
resented by small specimens and inconspicuous forms, and their occurrence may
easily pass unnoticed until a large-scale sampling and a seasonal program have been
implemented (Fig. 5.2) (see also Chap. 2 by Oliveira et al.). Dubrasquet et al. (2018)
studied the diversity of red algae and their distribution along the WAP and SSI
through a molecular-assisted revision, identifying significant differences among
assemblages of SSI and northern part of the WAP (at ≈63°S), central part of the
88 F. Pellizzari et al.

Fig. 5.2  Activities carried out within the frame of the Antarctic Program supported by the
Brazilian Navy. (Photos by Franciane Pellizzari)

WAP (at ≈64°S), and central-southern part of the WAP (at ≈67°S). Apparently,
these subregions do not correspond to the classical bioregions reported in previous
studies, suggesting that the Bransfield Strait cannot be regarded as a strong barrier
for red macroalgae.
The South Shetland Islands and interface islands between Antarctic and sub-
Antarctic regions were previously reported to have transitional biota features, which
are most likely related with the composition and distribution of benthic marine com-
munities (Pellizzari et al. 2017). The seaweed richness found in the SSI (ca. 60°S)
was higher than in Adelaide Island and Terra Nova Bay (Ross Sea, above latitude
70°S), according to Cormaci et al. (1992) and Mystikou et al. (2014). In addition,
Pellizzari et al. (2017) considered Livingston and King George islands as hotspots
among the SSI.
Recent changes in the composition and structure of algal assemblages from
Antarctica are featured by new records, including green algae and cyanobacteria
identified unprecedentedly for the area, followed by detection of cryptogenic spe-
cies and reduction of endemism (Pellizzari et al. 2017; see also Chap. 2 by Oliveira
et al.). Cyanobacteria are opportunistic organisms that can inhabit diversified and
extreme environments, and they have been regarded as indicators of changes in
seawater pH and temperature (Sangil et al. 2012). Antarctic cyanobacteria are con-
spicuous and well studied in lake assemblages (Taton et al. 2006). However, incipi-
ent information in marine ecosystems is available. Pellizzari et al. (2017) identified
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 89

six species of filamentous Cyanobacteria as epiphytes on large seaweeds, occurring


along the intertidal mats of Deception Island. They also reported new record of
opportunistic Chlorophyta, including a cryptogenic species. It has been suggested
that general diversity may be decreasing due to the effects of climate changes (sum-
mary in Pibernat et al. 2007; Wiencke and Amsler 2012). However, this pattern is
not reported in some studies (synopsis in Pellizzari et al. 2017), considering that
seaweed communities seem to be in “transit” for a higher connectivity and distribu-
tion expansion. Further taxonomical studies supported by molecular analyses and
biogeographical models are needed to infer more consistent information regarding
effects of global changes in seaweed diversity.

5.4  The Physiological Bases of Macroalgal Shifts

Antarctic seaweed species are adapted to extreme environmental conditions; how-


ever, they can be highly sensitive to abrupt changes in meteorological and oceano-
graphical settings, suggesting that the distribution, diversity, and richness of these
organisms can be used as suitable sentinels of the effects of the climate change.
In general, Antarctic macroalgae are very well adapted to low seawater tempera-
tures. Kirst and Wiencke (1995) suggested that changes in lipid composition of
membranes are related to genetic adaptation to extreme conditions. Moreover,
endemic Antarctic species show high photosynthetic efficiency and high P:R ratios
at low temperatures (0–5°C), which allow them to exhibit high growth rates and
positive carbon balance at depths close to 30–40-m locations (Gómez et al. 1997;
Wiencke et al. 2007) (see also Chap. 9 by Deregibus et al.). Additionally, endemic
species exhibit upper survival temperatures (USTs) between 9°C and 13°C. The red
alga Georgiella confluens exhibits even lower temperature demands, growing only
at 0°C, but not at 5°C, and exhibits an UST of 11°C (Wiencke et al. 2014). In con-
trast, few endemic Antarctic species have higher temperature demands. For exam-
ple, the brown alga Ascoseira mirabilis grows up to 10°C but exhibits a low UST
similar to the above-mentioned species. The red algae Gymnogongrus antarcticus
and Phyllophora ahnfeltioides grow up to temperatures of 10°C or even 15°C, but
exhibit considerably higher USTs of 19°C or 22°C, respectively. So, the latter three
species could theoretically occur even at temperate sites (Bischoff-Bäsmann and
Wiencke 1996). The highest temperature demands of Antarctic species were dem-
onstrated in Antarctic cold-temperate green algae. These species grow up to 10°C,
15°C, or 20°C and exhibit USTs between 19 and 26 (28)°C (Wiencke et al. 2014).
These temperature requirements reflect the preferential occurrence of these species
along supra- and eulittoral zones, where temperatures may vary considerably.
Responses to climate changes are particularly rapid and strong in marine habitats
especially in high latitudes and in the intertidal and shallow subtidal zones, where
species are often growing at their upper temperature tolerance limits (Hoegh-­
Guldberg and Bruno 2010; Sorte et al. 2010). In this context, temperature require-
ments for growth and survival of endemic Antarctic species and species distributed
90 F. Pellizzari et al.

in the Antarctic cold-temperate region have been extensively tested in laboratory.


However, results from laboratory experiments should be interpreted with caution
since growth and survival of seaweeds in their natural habitat depend on a complex
interaction of physical, chemical, and biological factors. In fact, warming at coastal
Antarctic waters can alter light regime due to increased turbidity and shortening in
the ice cover period, with consequences not well understood for algae adapted to
short periods of favorable light conditions and extended periods of darkness
(Wiencke et al. 2007) (see also Chap. 7 by Huovinen and Gómez). On the other
hand, the inflow of melt water in summer, due to enhanced glacier melting or
increased terrestrial runoff, has also considerable effects on the salinity and tem-
perature regime in inshore waters down to 20-m depth (Klöser et al. 1993; Zacher
et al. 2009). In other instance, inorganic nutrient concentrations may be high and in
general not limiting for seaweeds at any time of the year in the Antarctica (Sanches
et al. 2016).
It is well known that species that fails to acclimatize physiologically or evolve
genetically to increasing temperature will probably move into another habitats or
become extinct (Parmesan 2006; Thomas 2010; Jueterbock et al. 2013). According
to Guo et  al. (2005), the most effective and informative method of predicting
declines or disappearances and/or expansion of nonnative species is by monitoring
boundary conditions of very isolated populations, as is the case in Antarctic sea-
weed assemblages. In all, large-scale climate-driven changes have recently been
observed in maritime Antarctic (Macaya et al. 2016; Pellizzari et al. 2017; López
et al. 2018; Fraser et al. 2019; Diaz Tapia et al. 2018), which can favor the prolifera-
tion and extend the growing season of warm affinity species, inducing local domi-
nance of introduced species and causing adverse effect on cold-adapted species
(López et al. 2018).

5.5  D
 eception Island: A Case Study of Opportunistic, Alien,
Cryptic and Cryptogenic Species

Deception Island (62°57’S; 60°38’W), South Shetland Islands (SSI  – Fig.  5.1b),
affords unique ecosystems essential to understanding the impacts of global warm-
ing due to conditions of a highly temperature-sensitive environment. This island is
also an excellent model to work in a semi-enclosed environment that can be moni-
tored with long-term instrumentation while being free from the blocking effects of
the prevalence of ice and icebergs even during the winter (Smith et al. 2003). The
ATCM XXXV (2012) defines Deception Island as an ecosystem of high relevance
associated to a unique terrestrial flora, including at least 18 species that have not
been recorded elsewhere in Antarctica (e.g., the communities associated with geo-
thermal areas of the Antarctic flowering plant, pearlwort, Colobanthus quitensis).
All these areas are under conservation strategies (classified in ASPAs and ASMAs).
Due to its scientific importance, the area offers a rare opportunity to study the effects
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 91

Fig. 5.3 (a) Deception Island (SSI) aerial view. (Photo: DAE, Brazilian Navy); (b) hydrothermal
vents; (c) volcanic fumaroles; (d) Spongomorpha arcta (Chlorophyta); (e) filamentous
Cyanobacteria epiphytized by diatoms, opportunistic taxa dominating the benthic algae assem-
blages in the area; (f) lake along the island cone. (Photos by Franciane Pellizzari)

of environmental changes, in order to contrast them with the dynamics and recovery
from natural disturbance. On the other side, the long history of human activity
(since c.1820), including exploration sailing, whaling, war aviation, and scientific
research (e.g., the Norwegian whaling station; the British secret base of the World
War II, “Base B”), makes this island a site of high historical value. The island was
previously a whaling station, and the presence of research stations (Argentina and
Spain) is also since decades ago. About 11 localities were designated as Antarctic
Specially Protected Area (ASPA).
Deception Island is a circular-shaped volcano, with a linear glaciated coastline,
forming semi-enclosed environments showing intense geothermal activity. Black
sand beaches and outcrops of volcanic rock form the hard substrate (Fig. 5.3a–c, f).
Deception Island exhibits widely varying microclimates with a diameter of ca.
12 km; the center of the island was formed by a huge eruption that flooded the sea
to form a large bay (Port Foster). The bay has a narrow entrance (Neptune’s
Bellows), and shortly after the pass, there is a cove (Whalers Bay).
The fumarolic emissions and thermal springs generally occur along a principal
fracture encompassing Fumarole Bay, Telefon Bay, and Pendulum Cove.
Temperatures of fumarolic discharges as high as 110°C have been recorded depend-
ing on tidal cycle. These fumarolic emissions and thermal springs along the main
fracture result in temperature and pH anomalies (Fig. 5.3a–c), and the bottom water
temperatures in shallow areas could rise to 8°C (Pellizzari et al. 2017).
92 F. Pellizzari et al.

Thus, mainly due to the unique physical and chemical features besides long
human presence history, including current tourism, Deception Island (Antarctic
Specially Managed Area No. 4) is exceptionally vulnerable to the impacts of intro-
duced nonnative species. The human visitation, the mild climate (compared to other
polar areas), protected and safest natural harbor, and the presence of geothermal
heated sites may catalyze more introduced marine and terrestrial species than any
other Antarctic locations.
The diversity of macroalgae from Deception was early described by Clayton
et al. (1997), Gallardo et al. (1999), and Pellizzari et al. (2017). Two putative new
species, Prasiola sp. and Callophyllis sp., occur in Deception Island and deserve
elucidation due to their cryptic features (Pellizzari et al. 2017). Besides, the widely
distributed species Ulvella viridis, Spongomorpha arcta, and Rhizoclonium ambig-
uum are also conspicuous in the island. Pellizzari et  al. (2017) also reported the
predominance of filamentous, opportunistic, and widely distributed or cosmopoli-
tan cyanobacteria and green algae, high diatom epiphytism, and it was one of the
locations where the bipolar and cryptogenic species were reported (Fig. 5.3d, e).
Monostroma grevillei, whose introduction possibilities lie in the anthropic activity
(whaling and/or tourism), is also a bipolar species, although with some degree of
latitudinal disruption. Clayton (1994) reported the first occurrence of Petalonia fas-
cia for Deception Island and suggested that ships could introduce the species, men-
tioning also that the intertidal and subtidal seaweeds from Port Foster belong
probably to widely distributed cold-adapted species. At least 20 seaweed species in
Antarctica are broadly distributed, e.g., the red alga Plocamium cartilagineum, the
brown alga Petalonia fascia, and the green alga Ulothrix flacca (Wiencke and
Clayton 2002). It is possible that such species may be recent invaders from temper-
ate regions (Clayton 1994). Two decades later, Pellizzari et al. (2017) observed the
intertidal diversity permeated by filamentous algae belonging mainly to
Cyanobacteria and typical eurythermal Chlorophyta such as Cladophora albida and
Rhizoclonium riparium. Following McCarthy et  al. (2019), many Northern
Hemisphere species have become invasive in southern temperate sites, such as Port
Phillip Bay (Melbourne, Australia). Pathways to the Southern Hemisphere are
clearly viable for some species of algae, polychaetes, and sponges that can inhabit
polar waters. Besides, several bipolar species have been reported, including micro-
organisms, macrofauna, and macroalgae, for example, M. grevillei (Pellizzari et al.
2017). Other studies have demonstrated the presence of cryptic species in Antarctica
(Billard et al. 2015 and synopsis in Dubrasquet et al. 2018), and Deception Island is
one of the target places where cryptic and hidden species must be investigated.
Crustose calcareous algae belonging to Corallinaceae (Pseudolithophyllum sp.
and Lithophyllum subantarcticum) and Hapalidiaceae (Clathromorphum obtectu-
lum, Lithothamnion granuliferum, Synarthrophyton patena, and two unidentified
species of the Phymatolithon/Mesophyllum complex) have been reported along the
Shetland Islands (listed in Pellizzari et al. 2017) (Fig. 5.4a, f). However, in Deception
Island, these calcareous algae are absent, probably due to lower pH values, confirm-
ing the island as a natural laboratory to study responses in species facing extreme
abiotic changes. Moreover, the macroalgae from Deception Island host a rich
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 93

Fig. 5.4  Some species sampled along the SSI. (a) Clathromorphum (calcareous Rhodophyta), (b)
fertile Iridaea cordata, (c) transverse section of I. cordata, (d) Delesseriaceae, (e) cystocarp of
Myriogramme mangini, (f) Nacella concinna (the Antarctic limpet) grazing over benthic algae.
(Photos by Franciane Pellizzari)

associated fungal diversity composed of endemic, cold-adapted, or cosmopolitan


taxa. Fungal species of the genera Aspergillus, Cryptococcus, Debaryomyces,
Meyerozyma, Penicillium, and Pseudogymnoascus have been isolated from algae
such as Adenocystis utricularis, Monostroma hariotii, and Adenocystis sp. and
Pyropia endiviifolia (Godinho et  al. 2013; Furbino et  al. 2014). The studies also
emphasize that the fate of endemic or cold-adapted macroalgal species under chang-
ing Antarctic environment will affect also their associated microbiota.

5.6  R
 eevaluating Eco-Regions, Isolation, and Endemism
in the Southern Ocean

According to Wiencke and Tom Dieck (1990), many endemic Antarctic species
show a lower range of thermal tolerance compared to Arctic species. Arctic species
have been isolated for a much less time than Antarctic species. However, some of
these species were also listed in Falkland Islands (Clayton 2003), South Georgia
(Wells et al. 2011), and Macquarie Island (Ricker 1987), e.g., Desmarestia menzie-
sii, G. confluens, and Myriogramme manginii. Ramírez and Santelices (1991)
reported D. anceps for Chile. Monostroma hariotii Gain was registered in
Argentinean Patagonia (Boraso de Zaixso 2003, 2013), South Georgia (Wells et al.
94 F. Pellizzari et al.

2011), and the Falkland Islands (Clayton 2003). Iridaea cordata (Fig. 5.4b, c) and
Geminocarpus geminatus also occurs in the sub-Antarctic islands and Tierra del
Fuego. Ballia callitricha and Adenocystis utricularis also occur in New Zealand and
Australia, suggesting higher connectivity among the assemblages and as already
discussed in this chapter.
Studies using molecular markers (Hommersand et  al. 2009; Medeiros 2013;
Billard et  al. 2015; Ocaranza-Barrera et  al. 2018; Dubrasquet et  al. 2018) have
revealed that cryptic species previously listed in both maritime Antarctica and South
America, e.g., Plocamium, Ulothrix, Gigartina, and Iridaea, are in fact distinct
taxonomic entities. John et  al. (1994) reported 127 species of benthic algal flora
from South Georgia (54°S). Comparing the results of this sub-Antarctic island with
the list from the SSI (Pellizzari et al. 2017, Fig. 5.5a–d), 40% of the species co-­
occurred in both locations, suggesting high similarity between marine flora of dis-
tinct eco-regions or provinces (Sanches et al. 2016). Wells et al. (2011) described
the intertidal and subtidal benthic seaweed diversity of South Georgia revealing that
63% of these species list co-occur in the SSI. Palmaria decipiens and Iridaea cor-
data (Fig. 5.4), conspicuous Antarctic red algae, occur from the Ross Sea to sub-
Antarctic islands (Wiencke and Clayton 2002). However, I. cordata was also

Fig. 5.5 (a) King George Island (SSI) threatened by global changes. (b) Weddell seal in a drifted
bed of conspicuous Palmaria decipiens, (c) Iridaea cordata, Adenocystis utricularis, Ascoseira
mirabilis, (d) Monostroma hariotii. (Photos by Franciane Pellizzari)
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 95

reported in Chile and recently in Argentinean Patagonia (Ramírez and Santelices


1991; Boraso de Zaixso 2013), and P. decipiens was also recorded in New Zealand
(Nelson 2012), Macquarie Island (Ricker 1987), and South Georgia (Wells
et al. 2011).
Advances in molecular techniques have allowed estimating more precisely phy-
logenetic relationships, levels of differentiation, and divergence time between popu-
lations from these continents (Ocaranza-Barrera et al. 2018). According to Papenfuss
(1964), the seaweed flora of the sub-Antarctic islands and the Antarctic region
totaled 550 species. Approximately 75% of these species were distributed in the
sub-Antarctic islands (Clayton 1994). Wulff et al. (2009) suggested that, after the
endemics, the second largest seaweed group in Antarctic includes those occurring
both in the Antarctic region and on sub-Antarctic islands and Tierra del Fuego (see
also Chap. 2 by Oliveira et al.). Some of these species may be examples of distribu-
tional extension northward; besides few studies are supported by molecular data
(Hommersand et al. 2009; Fraser et al. 2013; Medeiros 2013; Billard et al. 2015;
Pellizzari et al. 2017; Ocaranza-Barrera et al. 2018). Fraser et al. (2013) studied the
genetic affinities between transoceanic populations of Adenocystis utricularis and
Bostrychia intricata in southern Chile, New Zealand, and several sub-Antarctic
islands (disjoint regions connected oceanographically by the ACC). The authors
observed divergent clades for both species, but close phylogenetic relationships –
even shared haplotypes – among populations separated by large oceanic distances.
Despite not being particularly buoyant, A. utricularis and B. intricata showed
genetic signatures of recent dispersal across vast oceanic distances, presumably by
attachment to floating substrata (e.g., wood, buoyant large macroalgae, or still, hull
fouling, or ballast water), although the genetic data indicate the possibility of cryp-
tic species within both taxa (see also Chap. 4 by Macaya et al.).
For algae present on both sides of the ACC, e.g., Gigartina skottsbergii, two
hypotheses have been discussed: (1) they correspond to recent immigrants from
adjacent continents, or (2) they have evolved in situ surviving the effects of the last
glaciations. The haplotype networks of G. skottsbergii suggested that there is some
evolutionary divergence between populations and a disjoint distribution, a pattern
that exceeds its dispersal capacity (Billard et al. 2015; see also Chap 6 by Guillemin
et  al.). In accordance to Hommersand et  al. (2009), the monophyletic clades of
G. skottsbergii may correspond to two cryptic species: (1) G. skottsbergii distrib-
uted in the sub-Antarctic islands and in the southern coast of Chile and (2) a new
species, still to be formally described and named, that occurs in the Antarctic
Peninsula, South Shetland, and South Orkney Islands. The divergence time between
these two cryptic species suggested that algae with limited dispersal capabilities
were able to cross the Scotia Sea after separation of the continents, potentially via a
stepping stone process through the volcanic arc of islands. This strongly supports
the idea that the last Quaternary glaciations induced marked bottlenecks that were
followed by rapid colonization events (Billard et  al. 2015; see also Chap. 6 by
Guillemin et al.).
Regarding other connections in the SO, the seaweed assemblage of Southern
Australia (Victoria-Tasmania Region) has been featured by high species richness
96 F. Pellizzari et al.

(Kerswell 2006) – endemic (40–77%) and temperate (17–45%) species – whereas


species with tropical (4.5– 9.7%) and polar (1.5–5.1%) affinities are much less rep-
resented. Wernberg et al. (2011a) used a seaweed database from herbarium records,
sampled in Australia since the 1940s, and documented changes in communities’
distribution in Indian and Pacific Oceans, related with rapid warming over the past
five decades. For the coasts of New Zealand, regarded as areas of high seaweed
diversity (Kerswell 2006), a recent list suggested that among the 770 currently
known seaweed recorded, 265 are endemic (34%) and 22 alien (2.9%). Rhodophyta
accounts around 60% of the total taxa. Overall, a reduction in endemism (see also
Chap. 2 by Oliveira et al.) and higher species “transit” among assemblages from
Antarctica, sub-Antarctic, and Southern America have been suggested, probably
requiring increasing seawater temperatures over a longer period of time (Sjøtun
et al. 2015).
Although temperature increases may not directly affect the latitudinal distribu-
tion of Antarctic key structural seaweeds in the short term (Müller et al. 2009), they
may modify the presence of these species indirectly via changes in sea ice dynam-
ics, which enhances the hard substrate availability, besides changes on thermoha-
line and circulation patterns. Stammerjohn et al. (2008) reported that the sea ice
cover along the southern portion of the Western Antarctic Peninsula (WAP)
advances nearly 2 months later in the winter and retreats approximately 1 month
earlier in the spring. In this context, the duration of the sea ice season around the
Antarctic Peninsula decreased by 4 days per year from 1979/1980 to 2011/2012
(Hughes and Ashton 2017). However, the sea ice surrounding Antarctica has
increased in extent and concentration from the late 1970s, when satellite-based
measurements began, until 2015. While this increased ice cover trend is modest, it
is surprising given the general warming of the global climate (Fig.  5.6a, b) and
despite some climate models – that incorporate a better understanding of the pro-
cesses affecting the region – usually simulating a tendency of sea ice decrease in
the mid and long term. In contrast, in the Artic, the sea ice cover has exhibited
pronounced declines over the same period, consistent with global climate model
simulations. However, and due to the oscillating patterns of the sea ice cover
recently in the Antarctic shelves, this random behavior in distinct Antarctic areas
has presented an enigma for global climate change science. In addition, significant
differences between the Southern and Arctic Oceans resulted from SST opposite
anomalies (Fig. 5.6c).
Recent data suggest that the SO seems to be barely sensitive to greenhouse
gases (GHG), resulting in a current lower warming levels of SST than observed
in other areas (NASEM 2017). However, changes in the deep ocean can have
important effects due to significant upwelling (NASEM 2017). Other factors that
may play a role in this conundrum include tropical Pacific and Atlantic telecon-
nections, variability in the wind, and ocean circulation that circumnavigate
Antarctica. However, an understanding of the mechanisms and processes driving
sea ice variability, as well SST trends, is limited by the lack of long and homog-
enous records.
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 97

Fig. 5.6 (a) Antarctic sea ice extent: line is the median of ice edge between 1981 and 2010 and the
white contour, in 2019. A decrease has been observed since January 2016. (Image Source: National
Snow & Ice Data Center http://nsidc.org/data/seaice_index). (b) Southern Ocean sea ice extends
anomalies. (Between 2010 and 2016 the trend line is slightly positive http://nsidc.org/data/seaice_
index). (c) SST anomalies of the Arctic and Southern Oceans. (Source: NOAA NOMADS http://
nomad1.ncep.noaa.gov/cgi-bin/pdisp_sst.sh)

5.7  C
 oncluding Remarks: Prospects for the Future Marine
Flora of the Southern Ocean

Shifts in species distribution are being reported worldwide as a result of changes in


connectivity patterns, mediated by anthropic activity, climate changes, and increased
species “transit” even towards remote areas. Dispersal refugia for Antarctic and sub-
Antarctic seaweeds (see Billard et al. 2015) are also concepts that contribute to the
better understanding of the dispersal mechanisms and recolonization processes dur-
ing biogeographic expansion in the Southern Ocean and connections (see also Chap.
6 by Guillemin et al.). Thus, seaweed assemblages in Antarctica, and in the Southern
Hemisphere, will be modified not only by large-scale geological and paleoclimatic
processes, but also by long-distance dispersal events (see also Chap. 3 by Fraser
et al.).
98 F. Pellizzari et al.

Changes in meteorological and oceanographic patterns are rearranging the bio-


geography of ecologically important species (Wernberg et al. 2010, 2011a, b). The
impacts on Antarctic seaweed assemblages probably will include (i) shifts in disper-
sal patterns and invasive potential with impacts on local richness and diversity, (ii)
changes in primary productivity and biogeochemical fluxes, (iii) changes in the
genetic pool with consequences for phenotypic plasticity and adaptive responses
due thermohaline shift and habitat losses, and (iv) changes in the intensity and
direction of biological interactions. Considering these projections, the threat to
polar unique lineages, regarded as climatic relics, will increase.
One major issue to accurately interpret the effects of abiotic and biotic changes
on the Southern Ocean marine flora is the scarce basic knowledge of seaweed dis-
tribution over large areas. Although some advances have been made, the studies are
spatially fragmented due mostly to difficulties to sample in polar/extreme sites,
impairing suitable estimations on their potential genetic loss due to environmental
shifts. Thus, in a macroscale, changes in the distributional patterns of the Antarctic
seaweed assemblages highlight the essential importance of long-term monitoring
programs along the West Antarctic Peninsula, probably the area experiencing the
most rapid regional warming. These initiatives will improve the knowledge of these
natural laboratories and models for further comparative biogeographic studies in the
Southern Ocean.
Finally, Antarctic macroalgae can be regarded as suitable bioindicators of global
changes in virtue not only of their specialized physiological mechanisms to cope
with low temperatures and limited light conditions, but also by their high adaptabil-
ity to extreme conditions in their habitat. These remote assemblages can provide
essential clues to understand the resilience potential of organisms living in the edge
of their adaptive windows and are fundamental to reinforce the need for maintaining
global database aiming to integrate and normalize abiotic/biotic metadata (see also
Chap. 2 by Oliveira et al.).

Acknowledgements The authors thank PROANTAR (Brazilian Antarctic Program


557030/2009-9, 407588/2013-2 and 442258/2018-2), INCT Criosfera 2, Brazilian Navy (Polar
Ship Almirante Maximiano–H41), Brazilian Air Force, MMA (Ministry of Environment), MCTIC
(Ministry of Science, Technology and Innovation), CNPq (National Council of Research and
Development), Fundação Araucária–Government of Paraná, CAPES, FAPEMIG, and FNDC.

References

ATCM XXXV (2012) Final report. Deception Island management package: management plan for
Antarctic specially managed area No 4 Deception Island, South Shetland Islands, Antarctica
Billard E, Reyes J, Mansilla A, Faugeron S, Guillemin M-L (2015) Deep genetic divergence between
austral populations of the red alga Gigartina skottsbergii reveals a cryptic species endemic to
the Antarctic continent. Polar Biol 38:2021. https://doi.org/10.1007/s00300-015-1762-4
Bischoff-Bäsmann B, Wiencke C (1996) Temperature requirements for growth and survival of
Antarctic Rhodophyta. J Phycol 32:525–535
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 99

Boraso de Zaixso AL (2003) Algas marinas de la Patagonia: una guía ilustrada. Fundación de
Historia Natural Félix de Azara, Argentina
Boraso de Zaixso AL (ed) (2013) Elementos para el estudio de las macroalgas de Argentina, Con
colaboración de J.M. Zaixso. 1a ed. – Comodoro Rivadavia, Universitaria de la Patagonia.
ISBN 978-987-1937-14-1
Clayton MN (1994) Evolution of the Antarctic marine benthic algal flora. Rev J Phycol 30:897–904
Clayton MN (2003) Falkland Islands seaweed survey, Shackleton Scholaship Fundation, Monash
University, Victoria 3800, Australia
Clayton MN, Wiencke C, Klöser H (1997) New records of temperate and sub-Antarctic marine
benthic macroalgae from Antarctica. Polar Biol 17:141–149. https://doi.org/10.1007/
s003000050116
Cormaci M, Furnari G, Scammacca B (1992) The benthic algal flora of Terra Nova Bay (Ross Sea,
Antarctica). Bot Mar 35:541–552. https://doi.org/10.1515/botm.1992.35.6.541
Crame JA (1994) Evolutionary history of Antarctica. In: Hempel G (ed) Antarctic science: global
concerns. Springer, Berlin, pp 188–214
Díaz Tapia P, Maggs C, Macaya EC, Verbruggen H (2018) Widely distributed red algae often
represent hidden introductions, complexes of cryptic species or species with strong phylogeo-
graphic structure. J Phycol 54:829–839. https://doi.org/10.1111/jpy.12778
Dubrasquet H, Reyes J, Sanchez RP, Valdivia N, Guillemin ML (2018) Molecular-assisted revi-
sion of red macroalgal diversity and distribution along the Western Antarctic Peninsula and
South Shetland Islands. Cryptogamie Algol 39(4):409–429. https://doi.org/10.7872/crya/v39.
iss4.2018.409
Ducklow HW, Fraser WR, Meredith MP, Stammerjohn SE, Doney SC, Martinson DG, Sailley SF,
Schofield OM, Steinberg DK, Venables HJ, Amsler CD (2013) West Antarctic Peninsula: an
ice-dependent coastal marine ecosystem in transition. Oceanography 26:190–203. https://doi.
org/10.5670/oceanog.2013.62
Fraser CI, Zuccarello GC, Spencer HG, Salvatore LC, Garcia GR et  al (2013) Genetic affini-
ties between trans-oceanic populations of non-buoyant macroalgae in the high latitudes of the
southern Hemisphere. PLoS One 8(7):e69138. https://doi.org/10.1371/journal.pone.0069138
Fraser CI, Morrison AK, McC Hogg A, Macaya EC, van Sebille E, Ryan PG, Padovan A, Jack
C, Valdivia N, Waters JM (2019) Antarctica’s ecological isolation will be broken by storm-­
driven dispersal and warming. Nat Clim Change Lett 42:475–483. https://doi.org/10.1038/
s41558-018-0209-7
Furbino LE, Godinho VM, Santiago IF, Pellizzari FM, TMA A, Zani CL, PAS J, Romanha AJ,
AGO C, LHVG G, Rosa CA, Minnis AM, Rosa LH (2014) Diversity patterns, ecology and
biological activities of fungal communities associated with the endemic macroalgae across the
Antarctic Peninsula. Microb Ecol 67:775–787
Gallardo T, Pérez-Ruzafa IM, Flores-Moya A, Conde F (1999) New collections of benthic marine
algae from Livingston and Deception Islands (South Shetland Islands) and Trinity Island
(Bransfield Strait), Antarctica. Bot Mar 42:61–69. https://doi.org/10.1515/bot.1999.009
Godinho VM, Furbino LE, Santiago IF, Pellizzari FM, Yokoya N, Rosa LH (2013) Diversity and
bioprospecting of fungal communities associated with endemic and cold-adapted macroalgae
in Antarctica. ISME 7:1434–1451
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, metabolic
carbon balance and zonation of sublittoral macroalgae from King George Island (Antarctica).
Mar Ecol Prog Ser 149:281–293
Griffiths HJ, Waller CL (2016) The first comprehensive description of the biodiversity and bioge-
ography of Antarctic and Sub-Antarctic intertidal communities. J Biogeogr 43(6):1143–1155.
https://doi.org/10.1111/jbi.12708
Griffiths HJ, Barnes DKA, Linse K (2009) Towards a generalized biogeography of the Southern
Ocean benthos. J Biogeogr 36:162–177
Guo KM, Taper M, Schoenberger B, Brandle J (2005) Spatial-temporal population dynamics
across species range: from centre to margin. Oikos 108:47–57
100 F. Pellizzari et al.

Gutt J, Alvaro MC, Barco A, Böhmer A, Bracher A, David B et al (2016) Macroepibenthic com-
munities at the tip of the Antarctic Peninsula, an ecological survey at different spatial scales.
Polar Biol 39:829–849. https://doi.org/10.1007/s00300-015-1797-6
Hoegh-Guldberg O, Bruno JF (2010) The impact of climate change on the world’s marine ecosys-
tems. Science 328:1523–1528
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52:509–534. https://doi.org/10.1515/bot.2009.081
Hughes KA, Ashton GV (2017) Breaking the ice: the introduction of biofouling organisms to
Antarctica on vessel hulls. Aquatic Conserv Mar Freshw Ecosyst 27:158–164. https://doi.
org/10.1002/aqc.2625
Huovinen P, Gómez I (2012) Cold-temperate seaweed communities of the Southern Hemisphere.
In: Wiencke C, Bischof K (eds) Seaweed biology, novel insights into ecophysiology, ecology
and utilization. Ecological studies 219. Springer, Berlin Heidelberg, pp 293–314. https://doi.
org/10.1007/978-3-642-28451-9_14
John DM, Pugh PJA, Tittley I (1994) Observations on the benthic marine algal flora of South
Georgia: a floristic and ecological analysis. Bull Nat Hist Museum 24:101–114
Jueterbock A, Tyberghein L, Verbruggen H, Coyer JA, Olsen JL, Hoarau G (2013) Climate change
impact on seaweed meadow distribution in the North Atlantic rocky intertidal. Ecol Evol
3(5):1356–1373. https://doi.org/10.1002/ece3.541
Keith SA, Kerswell AP, Connolly SR (2011) Global diversity of marine macroalgae: environmen-
tal conditions explain less variation in the tropics. Glob Ecol Biogeogr 23:517–529
Kerswell AP (2006) Global biodiversity patterns of benthic marine algae. Ecology 87:2479–2488
Kirst GO, Wiencke C (1995) Ecophysiology of polar algae. J Phycol 31:181–199
Klöser H, Ferreyra G, Schloss I, Mercuri G, Laturnus F, Curtosi A (1993) Seasonal variation of
algal growth conditions in sheltered Antarctic bays: the example of Potter Cove (King George
Island, South Shetlands). J Mar Syst 4:289–301
Linse K, Griffiths HJ, Barnes DKA, Clarke A (2006) Biodiversity and biogeography of Antarctic
and sub-Antarctic mollusca. Deep-Sea Res II 53:985–1008
Longhurst A (ed) (2007) Ecological geography of the sea, 2nd edn. Academic Press, London, p 390
López BA, Macaya EC, Rivadeneira MM, Tala F, Tellier F, Thiel M (2018) Epibiont communities
on stranded kelp rafts of Durvillaea antarctica (Fucales, Phaeophyceae). Do positive interac-
tions facilitate range extensions? J Biogeogr 45:1833–1845. https://doi.org/10.1111/jbi.13375
Macaya EC, López B, Tala F, Tellier F, Thiel M (2016) Float and raft: role of buoyant sea-
weeds in the phylogeography and genetic structure of non-buoyant associated flora. In: Hu
ZM, Fraser C (eds) Seaweed phylogeography. Springer, Dordrecht, pp  97–130. https://doi.
org/10.1007/978-94-017-7534-2_4
McCarthy AH, Peck LS, Hughes KA, Aldridge DC (2019) Antarctica: the final frontier for marine
biological invasions. Glob Chang Biol 25(7): 2221–2241. https://doi.org/10.1111/gcb.14600
Medeiros AS (2013) Macroalgae diversity of Admiralty Bay, King George Island, Antarctic
Peninsula based on DNA barcoding and other molecular markers. Thesis. São Paulo University.
http://www.teses.usp.br/teses/disponiveis/41/41132/tde-24032014-090801/
Müller R, Laepple T, Bartsch I, Wiencke C (2009) Impact of oceanic warming on the distribution
of seaweeds in polar and cold-temperate waters. Bot Mar 52:617–638. https://doi.org/10.1515/
bot.2009.080
Mystikou A, Peters AF, Asensi AO, Fletcher KI, Brickle P, van West P, Convey P, Küpper FC
(2014) Seaweed biodiversity in the south-western Antarctic Peninsula: surveying macroalgal
community com- position in the Adelaide Island/Marguerite Bay region over a 35-year time
span. Polar Biol 37:1607–1619. https://doi.org/10.1007/s00300-014-1547-1
NASEM (2017) National academies of sciences, engineering, and medicine. In: Antarctic sea ice
variability in the Southern Ocean-climate system: proceedings of a workshop. The National
Academies Press, Washington, DC. https://doi.org/10.17226/24696
Nelson WA (2012) Phylum rhodophyta: red algae. In: Gordon DP (ed) New Zealand inventory of
biodiversity. Canterbury University Press, Christchurch
5  Biogeography of Antarctic Seaweeds Facing Climate Changes 101

Ocaranza-Barrera P, González-Wevar C, Guillemin ML, Rosenfeld S, Mansilla A (2018) Molecular


divergence between Iridaea cordata (Turner) Bory de Saint-Vincent from the Antarctic
Peninsula and the Magellan Region. J App Phycol. 31:939–949. https://doi.org/10.1007/
s10811-018-1656-2
Orsi AH, Whitworth T, Nowlin WD Jr (1995) On the meridional extent and fronts
of the Antarctic Circumpolar Current. Deep-Sea Res 42:641–673. https://doi.
org/10.1016/0967-0637(95)00021W
Papenfuss GF (1964) Catalogue and bibliography of Antarctic and Subantarctic benthic marine
algae. Antar Res Ser 1:1–76
Parmesan C (2006) Ecological and evolutionary responses to recent climate change. Annu Rev
Ecol Evol Syst 37:637–669
Pellizzari F, Santos-Silva MC, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa L (2017)
Diversity and spatial distribution of seaweeds in the South Shetland Islands, Antarctica: an
updated database for environmental monitoring under climate change scenarios. Polar Biol.
https://doi.org/10.1007/s00300-017-2092-5
Pibernat RA, Ellis-Evans C, Hinghofer-Szalkay HG (eds) (2007) Life in extreme environments.
Springer, The Netherlands. https://doi.org/10.1007/978-1-4020-6285-8
Pierrat B, Sausede T, Brayard A, David B (2013) Comparative biogeography of echinoids, bivalves
and gastropods from the Southern Ocean. J Biogeogr 40:1374–1385
Pritchard HD, Ligtenberg SR, Fricker HA, Vaughan DG, van den Broeke MR, Padman L (2012)
Antarctic ice-sheet loss driven by basal melting of ice shelves. Nature 484:502–505. https://
doi.org/10.1038/nature10968
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223. https://doi.org/10.1371/journal.pone.0058223
Ramírez ME, Santelices B (1991) Catálogo de las algas marinas bentónicas de la costa temperada
del Pacífico de Sudamérica. Facultad de Ciencias Biológicas, Pontificia Universidad Católica
de Chile
Ricker RW (1987) Taxonomy and biogeography of Macquarie Island seaweeds. British Museum
(Natural History), London
Sanches PF, Pellizzari FM, Horta PH (2016) Multivariate analyses of Antarctic and sub- Antarctic
seaweed distribution patterns: an evaluation of the role of the Antarctic Circumpolar Current. J
Sea Res 110:29–38. https://doi.org/10.1016/j.seares.2016.02.002
Sangil C, Sansón M, Afonso-Carillo J, Herrera R, Rodríguez A, Martín-García L, Díaz-Villa T
(2012) Changes in sub- tidal assemblages in a scenario of warming: proliferations of ephemeral
benthic algae in the Canary Islands (eastern Atlantic Ocean). Mar Environ Res 77:120–128
Sjøtun K, Husa V, Asplin L, Sandvik AD (2015) Climatic and environmental factors influencing
occurrence and distribution of macroalgae  — a fjord gradient revisited. Mar Ecol Prog Ser
532:73–88. https://doi.org/10.3354/meps11341
Smith KL, Baldwin RJ, Kaufmann RS, Sturz A (2003) Ecosystem studies at Deception
Island, Antarctica: an overview. Deep-Sea Res II 50:1595–1609. https://doi.org/10.1016/
S0967-0645(03)00081-X
Sorte CJB, Williams SL, Carlton JT (2010) Marine range shifts and species introductions: com-
parative spread rates and community impacts. Glob Ecol Biogeogr 19:303–316
Stammerjohn SE, Martinson DG, Smith RC, Yuan X, Rind D (2008) Trends in Antarctic annual
sea ice retreat and advance and their relation to El Niño–Southern Oscillation and Southern
Annular Mode variability. J Geophys Res 113. https://doi.org/10.1029/2007jc004269
Taton A, Grubisic S, Balthasart P, Hodgson DA, Laybourn-Parry J, Wilmotte A (2006)
Biogeographical distribution and ecological ranges of benthic cyanobacteria in East Antarctic
lakes. FEMS Microbiol Ecol 57(2):272–289. https://doi.org/10.1111/j.1574-6941.2006.00110.x
Thomas CD (2010) Climate change and range boundaries. Divers Distrib 16:488–495
Wells E, Brewin P, Brickle P (2011) Intertidal and subtidal benthic seaweed diversity of South
Georgia. Norfolk, UK
102 F. Pellizzari et al.

Wernberg T, Thomsen MS, Tuya F, Kendrick GA, Staehr PA, Toohey BD (2010) Decreasing resil-
ience of kelp beds along a latitudinal temperature gradient: potential implications for a warmer
future. Ecol Lett 13:685–694
Wernberg T, Russell BD, Thomsen MS, Gurgel CFD, Bradshaw CJA, Poloczanska ES, Connell SD
(2011a) Seaweed communities in retreat from ocean warming. Curr Biol 21(21):1828–1832.
https://doi.org/10.1016/j.cub.2011.09.028
Wernberg T, Russell BD, Moore PJ, Ling SD, Smale DA, Campbell A, Coleman MA, Steinberg
PD, Kendrick GA, Connell SD (2011b) Impacts of climate change in a global hotspot for tem-
perate marine biodiversity and ocean warming. J Exp Mar Biol Ecol 400:7–16
Wernberg T, Bennett S, Babcock RC, Bettignies T, De Cure K, Depczynski M, Wilson S (2016)
Climate-driven regime shift of a temperate marine ecosystem. Science 353:169–172
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Wiencke C,
Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and utilization.
Springer-Verlag, Berlin, pp. 265–291
Wiencke C, Clayton MN (2002) Antarctic seaweeds. In: Wagele JW (ed) Synopses of the Antarctic
benthos. Lichtensein, Germany
Wiencke C, Tom Dieck IT (1990) Temperature requirements for growth and survival of macroal-
gae from Antarctica and southern Chile. Mar Ecol Prog Ser 24:157–170
Wiencke C, Clayton MN, Gómez I, Iken K, Luder UH, Amsler CD, Karsten U, Hanelt D, Bischof
K, Dunton K (2007) Life strategy, ecophysiology and ecology of seaweeds in polar waters. Rev
Environ Sci Biotechnol 6: 95–126. https://doi.org/10.1007/s11157-006-9106-z
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: De Broyer C, Koubbi P,
Griffiths HJ, Raymond B, Cd UA (eds) Biogeographic Atlas of the Southern Ocean. Scientific
Committee on Antarctic Research, Cambridge, pp 66–73
Wulff A, Iken K, Quartino ML, Al-Handal A, Wiencke C, Clayton MN (2009) Biodiversity, bio-
geography and zonation of marine benthic micro–and macroalgae in the Arctic and Antarctic.
Bot Mar 52:491–507. https://doi.org/10.1515/bot.2009.072
Zacher K, Rautenberger R, Hanelt D, Wulff A, Wiencke C (2009) The abiotic environment of polar
marine benthic algae. Bot Mar 52:483–490. https://doi.org/10.1515/bot.2009
Chapter 6
Comparative Phylogeography of Antarctic
Seaweeds: Genetic Consequences
of Historical Climatic Variations

Marie-Laure Guillemin, Claudio González-Wevar, Leyla Cárdenas,


Hélène Dubrasquet, Ignacio Garrido, Alejandro Montecinos,
Paula Ocaranza-Barrera, and Kamilla Flores Robles

Abstract  In the Southern Ocean, rapid climatic fluctuations during the Quaternary
are thought to have induced range contractions and bottlenecks, which drastically
impacted marine communities. For photosynthetic macroalgae that are restricted to
very shallow waters, survival in deepwater refugia is not possible. Comparing pat-
tern of distribution of genetic diversity using sequences of mitochondrial and chlo-
roplast markers in distinct species of green, brown and red macroalgae, we sought
to detect common responses to the effect of these glacial cycles. All the Antarctic
macroalgae were characterized by very low genetic diversity, absence of genetic
structure and significant signatures of recent population expansion. The eight stud-
ied species seem to have barely survived glacial events in situ, in a unique refugium
from which they recolonized their current distribution area. We propose that polyn-
yas or areas showing long-term geothermal activity along Antarctic continental

M.-L. Guillemin (*)


Instituto de Ciencias Ambientales y Evolutivas, Universidad Austral de Chile, Campus Isla
Teja, Valdivia, Chile
CNRS, UMI 3614 Evolutionary Biology and Ecology of Algae, Sorbonne Universités,
Roscoff, France
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: marie-laure.guillemin@uach.cl
C. González-Wevar
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Campus Isla Teja, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: claudio.gonzalez@uach.cl
L. Cárdenas
Instituto de Ciencias Ambientales y Evolutivas, Universidad Austral de Chile, Campus Isla
Teja, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile

© Springer Nature Switzerland AG 2020 103


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_6
104 M.-L. Guillemin et al.

margins or peri-Antarctic islands could be good candidate as glacial refugium, but


more variable genetic markers will be needed to precisely pinpoint its location.
Common haplotypes, scattered over hundreds or even thousands of kilometres of
coastline, point out to long-distance dispersal of fronds drifting on the strong oce-
anic currents in the region as the main mechanism of postglacial expansion.

Keywords  Glacial refugia · Genetic diversity · Last glacial maximum · Population


bottleneck · Quaternary · Southern ocean

6.1  Historical Isolation of Antarctic Marine Macroalgae

Antarctic marine macroalgae diversity, described as less diverse than other areas of
the Southern Ocean, is still characterized by high levels of endemism reaching up to
35% (Wiencke and Clayton 2002; Wiencke et al. 2014; also see chapter by Oliveira
et al. 2020 in this volume). As recorded for much of the Antarctic flora and fauna
(Clarke et al. 2005; Allcock and Strugnell 2012), the presence of various endemic
Antarctic macroalgal species has been linked to the major tectonic, oceanographic
and climatic changes that have affected the region during the last 50 million years
(Ma) (Dell 1972; Crame 1999, 2018; Mackensen 2004; Aronson et al. 2007; Moon
et al. 2017; Halanych and Mahon 2018). Indeed, the fragmentation of the continen-
tal landmasses and the initiation of the Antarctic Circumpolar Current (ACC)
(Fig.  6.1) gradually isolated the Antarctic waters from the rest of the Southern
Ocean (Crame 1999; Scher and Martin 2006; Aronson et  al. 2007; Dalziel et  al.
2013; Sijp et al. 2014; Scher et al. 2015). Currently, the Antarctic Polar Front (APF)
(Fig. 6.1) is considered as an effective barrier for many near-shore marine benthic

H. Dubrasquet · K. Flores Robles
Instituto de Ciencias Ambientales y Evolutivas, Universidad Austral de Chile, Campus Isla
Teja, Valdivia, Chile
I. Garrido
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Campus Isla Teja, Valdivia, Chile
Département de Biologie et Quebec Océan– Ocean Institute, Université Laval,
Quebec City, QC, Canada
A. Montecinos
Instituto de Ciencias Ambientales y Evolutivas, Universidad Austral de Chile, Campus Isla
Teja, Valdivia, Chile
CNRS, UMI 3614 Evolutionary Biology and Ecology of Algae, Sorbonne Universités,
Roscoff, France
P. Ocaranza-Barrera
Laboratorio de Ecosistemas Marinos Antárticos y Subantárticos, Universidad de Magallanes
(LEMAS), Punta Arenas, Chile
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 105

Fig. 6.1  Reconstruction of the Antarctic and sub-Antarctic ice coverage during the Last Glacial
Maximum (LGM; ~20,000 ka) and in the present day. Putative LGM ice extension is based on vari-
ous recent glaciological studies published for the tip of South America (McCulloch et al. 2000), the
Antarctic Peninsula and the South Shetland Islands (Simms et  al. 2011; Larter et  al. 2014;
O’Cofaigh et al. 2014) and South Georgia (White et al. 2018). Current ice extension is drawn based
on ®GoogleEarth satelital images. Arrow: Antarctic Circumpolar Current (ACC); thickness repre-
sent the strength of the ACC (Roberts et al. 2017). Dotted blue lines: Antarctic Polar Front (APF;
mean position of the Polar Front is represented) and sub-Antarctic Front (SAF) (Roberts et  al.
2017). The five sampling areas where Antarctic seaweeds were sampled are noted with back points
(more details are given in the text and in Figs. 6.3 and 6.4)

taxa (but see chapter by Oliveira et  al. 2020 in this volume), especially between
Antarctic and sub-Antarctic provinces (Clarke et al. 2005; González-Wevar et al.
2012, 2017; Poulin et  al. 2014; Billard et  al. 2015; Crame 2018; Halanych and
Mahon 2018). Non-endemic Antarctic seaweeds have classically included (1) cos-
mopolitan species reported on most Antarctic and sub-Antarctic/temperate coasts of
the Southern Ocean (e.g. Plocamium cartilagineum), (2) cold-water species with
disjoint amphiequatorial distribution (e.g. Acrosiphonia arcta and Desmarestia viri-
dis/confervoides) and (3) more broadly distributed species also occurring in peri-
and sub-Antarctic areas of the Southern Ocean as South Georgia, South Sandwich
Islands or even the southern tip of South America (e.g. Iridaea cordata, Gigartina
skottsbergii and Adenocystis utricularis) (Wiencke and Clayton 2002; Wiencke
et al. 2014).
Nevertheless, reconstruction of these Southern Ocean macroalgae species geo-
graphic distribution maps relies solely on classical taxonomy. Macroalgal anatomy
and high degree of phenotypic plasticity could lead to incorrect taxonomic classifi-
cation of specimens and molecular tools are now recognized as essential for species
determination in these organisms (Saunders 2005, 2008). Inaccurate morphological
106 M.-L. Guillemin et al.

identification and existence of cryptic species have been commonly reported in


green, red and brown macroalgae, including in Antarctica (Hommersand et al. 2009;
Moniz et al. 2012; Billard et al. 2015; Pellizzari et al. 2017; Dubrasquet et al. 2018;
Ocaranza-Barrera et al. 2019).
If few molecular studies have indeed confirmed the existence of macroalgal spe-
cies characterized by vast disjoint distributions that include Antarctica (i.e.
Acrosiphonia and Desmarestia; Olsen et al. 1993), most genetic data available sup-
port the existence of endemic Antarctic cryptic species. For example, two cryptic
species, one Antarctic and one sub-Antarctic, were detected in G. skottsbergii and
I. cordata (Hommersand et  al. 2009; Billard et  al. 2015; Ocaranza-Barrera et  al.
2019). In the same way, even if exhaustive phylogenetic studies have yet to be under-
taken for P. cartilagineum, Hommersand et al. (2009) noted that ‘unpublished rbcL
sequence analyses by S. Fredericq show that P. “cartilagineum” from Antarctica is
distinct from all other species of Plocamium investigated so far’. In G. skottsbergii,
the two cryptic species are clearly separated by the APF (Fig.  6.1) with the sub-
Antarctic species distributed in South America along the coasts of Chile and
Argentina and the Falkland Islands, while the Antarctic species is found in the
Antarctic Peninsula, the South Shetland Islands and the South Orkney Islands
(Billard et  al. 2015). Divergence time estimations between Antarctic and sub-­
Antarctic cryptic macroalgal species indicate a split between two lineages occurring
at the end of the Miocene, some 10–5 Ma ago (Hommersand et al. 2009; Billard et al.
2015; Ocaranza-Barrera et  al. 2019), long after the physical fragmentation of the
Southern Ocean continental landmasses or even to the formation of the ACC (Crame
1999; Mackensen 2004). However, the time of divergence between Antarctic and
sub-Antarctic lineages of macroalgae predates the Quaternary glacial cycles (~2 Ma).
The existence of macroalgal species (both nominal species and cryptic species
detected only when using molecular tools), entirely restricted to the Antarctic shelf,
or at most including some adjacent islands of the maritime Antarctic (i.e., South
Shetland Islands, South Orkney Islands and Balleny Islands) and offshore peri-­
Antarctic islands located south of the APF (i.e., South Georgia and Sandwich
Islands), evidenced the long evolutionary history of these organisms within the
Antarctic waters. In contrast to the hypothesis that most of the Antarctic biota could
be recent colonists, arrived after the deglaciation of the Last Glacial Maximum
(LGM) ice sheets (Clarke and Crame 1992), various Antarctic marine species,
including macroalgae, have withstood Quaternary glacial cycles in situ (Convey
et al. 2008, 2009; Allcock and Strugnell 2012).

6.2  A
 ntarctic Marine Macroalgae: Surviving Quaternary
Glacial Cycles in Situ

The onset of glaciations in Antarctica began during the Eocene-Oligocene bound-


ary, and after a slightly warmer stage (27–15 Ma) that seemed to have reduced the
extent of Antarctic ice, gradual cooling during the Miocene-Pliocene transition led
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 107

to the re-establishment of major ice sheets covering Antarctica (Kennett 1977;


Zachos et al. 2001). Later on, the Quaternary (2 Ma to 20 Ka) was characterized by
the alternation between glacial and interglacial periods that greatly affected the sea-
sonality and intensity of sea ice formation in the region (Clarke and Crame 1992;
Barker and Thomas 2004; Gersonde et  al. 2005; Kemp et  al. 2010). During the
LGM, ice sheet expansions were accompanied by eustatic drop in sea level (~120 m
lower than at present), as well as important latitudinal changes in the position of the
ACC (Fig.  6.1; McCulloch et  al. 2000; Simms et  al. 2011; Larter et  al. 2014;
O’Cofaigh et al. 2014; Roberts et al. 2017; White et al. 2018).
The development of large ice sheets generated major landscape and seascape
changes in Antarctica (Zachos et al. 2001; Thatje et al. 2005; Davies et al. 2012).
For instance, glacial reconstructions of the Antarctic Peninsula during the LGM
(20 Ka) showed stable grounding ice sheets advancing up to the continental shelf
edge (Larter et al. 2014; O’Cofaigh et al. 2014) (Fig. 6.1). Major glacial events in
Antarctica are thought to have led to mass extinctions, with ice scouring pushing
near-shore marine communities down the steep shelf slopes surrounding the
Antarctic continent (Clarke and Crame 1992; Thatje et al. 2005). Antarctic benthic
communities, especially sessile organisms with narrow depth distribution range, as
macroalgae, would have been remarkably vulnerable to continental sea ice advances
and retreats (Thatje et al. 2005; Dambach et al. 2012). Nevertheless, it seems that
grounded ice did not completely cover the Antarctic continental shelf, not even dur-
ing the glacial maxima (Klages et  al. 2017), and some isolated areas could have
acted as in situ glacial refugia. A wealth of evidences support this view, including
genetic data obtained for various terrestrial and marine organisms (Convey et  al.
2008, 2009; Allcock and Strugnell 2012), fossil evidences (Hiller et al. 1988) and
glaciological and oceanographic reconstructions (Paillard and Parrenin 2004;
Bentley et al. 2014; Larter et al. 2014; O’Cofaigh et al. 2014; Klages et al. 2017).
Several mechanisms could have sustained Antarctic shelf areas uncovered by ice
during the Quaternary glacial events. First, areas of open water (polynyas) formed
along Antarctic continental margins could have acted as in situ marine refugia
(Thatje et al. 2008). The possible existence of much more saline waters surrounding
Antarctica during glacial periods could have produced more stable and widespread
polynyas than previously postulated (Paillard and Parrenin 2004; Thatje et al. 2008).
Alternatively, in situ refugia could also be associated to volcanoes or areas of geo-
thermal activities (Fraser et al. 2014). Such geothermal refugia could be very impor-
tant in areas where clusters of long-lived volcanoes are recorded such as the tip of
the Antarctic Peninsula, the South Shetland Islands or the Ross Sea (Fraser et al.
2012, 2014) (see Chaps. 3 and 4 in this volume). Moreover, geological evidence
supports the diachrony of ice sheet extensions around Antarctica (Anderson et al.
2002) and near-shore marine organisms could ultimately have survived, hopping
from one open ice-free continental shelf area to another during glacial periods
(Thatje et al. 2008; Allcock and Strugnell 2012; Hughes et al. 2013).
At last, Antarctic near-shore macroalgae could also have endured the Quaternary
glacial events south of the APF but beyond the Antarctic continental shelf margins.
The coasts of peri-Antarctic islands have been postulated as potential glacial refugia
108 M.-L. Guillemin et al.

for some species since their distribution generally includes islands of the Scotia Arc
such as South Georgia and the South Sandwich Islands and Balleny Islands
(Wiencke and Clayton 2002; Wiencke et  al. 2014). South Georgia represents the
northern limit of distribution for many species of the Antarctic benthos (Barnes
et  al. 2006), and this area was reported as glacial refugium for some Antarctic
marine invertebrates (e.g. the Antarctic limpet Nacella concinna: González-Wevar
et al. 2013).

6.3  P
 ersistence in Multiple Isolated Glacial Refugia Versus
a Single Antarctic Refugium

Quaternary climatic oscillations dramatically affected species geographic range and


demography, especially at high latitude (Hewitt 2000, 2004; Maggs et  al. 2008;
Provan and Bennett 2008; Allcock and Strugnell 2012; Fraser et  al. 2012). The
expansion-contraction model (Provan and Bennett 2008) proposed an alternation
between the contraction of species distribution ranges during glacial advances and
subsequent rapid expansions during interglacial periods. These expansion/contrac-
tion cycles had a strong impact on the distribution of genetic variation of high lati-
tude organisms and can be detected, nowadays, using molecular tools and
phylogeographic analyses (Hewitt 2000, 2004; Provan and Bennett 2008; Allcock
and Strugnell 2012). Depending on the number and localization of glacial refugia,
different evolutionary scenarios can arise (Fig. 6.2).
For species that have survived glacial events in a unique refugium from which
they recolonized their current distribution area, a very low level of genetic diversity
and a high spatial genetic homogeneity are expected (see scenario 1 in Fig. 6.2). In
this case, the effect of genetic drift during glacial bottlenecks could be amplified by
gene surfing at expanding frontiers leading to extreme erosion of genetic diversity
and even complete fixation of pioneer alleles over huge areas (Excoffier and Ray
2008; Excoffier et al. 2009; Hallatschek and Nelson 2010). Supporting this idea,
complete genetic fixation has been recorded over distance of hundreds or even thou-
sands of kilometres for some temperate Southern Hemisphere macroalgae, with the
same mitochondrial haplotype covering all regions of the species distribution previ-
ously scoured by ice during glacial maxima (e.g. along the southern coast of Chile:
Fraser et al. 2009; Montecinos et al. 2012).
In contrast, population fragmentation in multiple disjoint glacial refugia (see sce-
nario 2 in Fig. 6.2) could lead to repeated events of divergence and diversification
(Clarke and Crame 1992; Allcock and Strugnell 2012). Indeed, small populations
could rapidly diverge due to strong genetic drift and, in addition, divergence could be
enhanced due to selection if environmental differences exist between refugia.
Divergence during isolation could be sufficient to generate reproductive barrier and
lead to speciation (Wilson et al. 2009; Allcock and Strugnell 2012; Lecointre et al.
2013). This phenomenon, classically known as the climate ‘diversity pump’ (Haffer
1969), has been proposed to be at the origin of recent evolutionary radiation events
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 109

Fig. 6.2  Putative effects of genetic bottleneck due to intense ice scouring during glacial periods
(i.e. the LGM) and postglacial range expansion following ice retreat yielding to the species current
distribution. Two distinct scenarios have been considered: (1) the existence of only one in situ
glacial refugium and (2) the existence of multiple (here three) in situ glacial refugia. Species dis-
tribution area is drawn in grey; each circle represents one individual; distinct colours represent
genetic variants (e.g. distinct haplotypes). Effects of bottleneck and postglacial expansion on the
current genetic diversity and phylogeographic structure potentially observed nowadays in species
affected by the glacial cycles are given on the right

in Antarctica (Clarke and Crame 1989; Near et al. 2012; Crame 2018). Such process
is expected to be more effective for species having limited dispersal capabilities such
as those with non-pelagic development or very short free-living larval stage (Pearse
et al. 2009). Indeed, highly differentiated genetic lineages/sister species have been
detected in various Antarctic marine organisms characterized by low dispersal capac-
ity and related to a scenario of isolation in multiple refugia during glacial events
(Wilson et al. 2009; Allcock and Strugnell 2012; Hemery et al. 2012). In macroalgae,
population fragmentation in multiple refugia has been commonly reported in the
North East Atlantic where genetic distinctiveness characterizes patchy populations
located at low latitude and inhabiting former refugial areas (Neiva et al. 2016).

6.4  A
 ntarctic Macroalgae Genetic Diversity: COI and TufA
Sequences Data Sets

Little is known about Antarctic macroalgae genetic diversity and most of the studies
have used molecular markers primarily for species identification (Hommersand
et al. 2009; Moniz et al. 2012; Mystikou et al. 2014; Pellizzari et al. 2017; Dubrasquet
110 M.-L. Guillemin et al.

et al. 2018; Ocaranza-Barrera et al. 2019) or phylogenetic inferences among taxa


including Antarctic species (Olsen et al. 1993; Peters et al. 2000). However, infor-
mation about intraspecific genetic diversity, mostly based on the acquisition of
mitochondrial sequences (coding sequences of the cytochrome c oxidase I gene,
noted COI, Guillemin et  al. 2018; Ocaranza-Barrera et  al. 2019; non-coding
sequences of the intergenic region Cox2–3, Billard et al. 2015), is building up for
some red macroalgae in the South Shetland Islands and the Antarctic Peninsula:
Curdiea racovitzae, Georgiella confluens, Gigartina skottsbergii, Iridaea cordata,
Palmaria decipiens and Plocamium cartilagineum.
Here we used the published COI data set obtained for six species of Rhodophyta
(Guillemin et al. 2018) to which we added new COI sequences for one species of
Ochrophyta (Himantothallus grandifolius) and sequences of the plastid gene tufA,
encoding for protein synthesis elongation factor Tu (EF-Tu), for one species of
Clorophyta (Monostroma hariotii) in order to compare genetic signature among
macroalgae phylum. Monostroma hariotii and H. grandifolius DNA extractions
were performed with an E.Z.N.A tissue DNA kit (Omega Bio-tek, Inc. Georgia,
USA) following the manufacturer’s instructions. Primers TufAgf4 and TufAR and
PCR reactions and programme described in Famà et al. (2002) were used to amplify
tufA fragments in M. hariotii, while a fragment of the COI gene was amplified in
H. grandifolius using the primers GazF2 and GazR2 and methodology described in
Lane et  al. (2007). PCR products were purified using the commercial kit
UltraCleanTM (MO BIO Laboratories, Carlsbad, USA) and sequenced in both
directions at the AUSTRAL-omics Core-Facility (Universidad Austral de
Chile, Chile).
Sequences of all macroalgae except H. grandifolius were obtained from five
sampling localities: two located in the South Shetland Islands (in Greenwich and
King George Island), two along the Northern part of the West Antarctic Peninsula
(WAP) near the Chilean O’Higgins Antarctic base and in Paradise Bay and one
along the central part of the WAP (Marguerite Bay) (Figs.  6.1, 6.3, and 6.4).
Himantothallus grandifolius COI data set includes 17 new sequences obtained from
Greenwich Island, King George Island, O’Higgins and Paradise Bay (GENBANK
N° MK503231–MK503247) and two sequences available in GENBANK
(HE866784, King George Island, Yang et  al. 2014; GQ368262, Terre Adelie,
Silberfeld et  al. 2010). Monostroma hariotii tufA sequences correspond to
GENBANK N° MK507414–MK507450. For each species-specific data set, we cal-
culated five diversity indices, Tajima’s D (Tajima 1989) and Fu’s Fs (Fu 1997) sta-
tistics and estimated the observed distributions of pairwise differences between
sequences using Arlequin v 3.5 (Excoffier and Lisher 2010). For Tajima’s D and
Fu’s Fs, significant departure from mutation drift equilibrium was tested using 1000
bootstrap replicates in Arlequin (Excoffier and Lisher 2010). Observed distributions
of pairwise differences between sequences were compared to estimated values
under a model of sudden demographic expansion through a generalized least squares
approach and goodness of fit was tested using 1000 permutations in Arlequin
(Excoffier and Lisher 2010). Haplotype networks were reconstructed using
NETWORK v 4.510 (Bandelt et al. 1999).
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 111

Curdiea Palmaria
racovitzae decipiens
(2) (5)
GEO GEO
(6) PRA (10) PRA
OHI OHI
Network: 42 COI (10) Network: 35 COI (4)
sequences sequences
100 km
PAR Land PAR
(14) (13)
Ice shelf

(10) (3)
MAR MAR

Fig. 6.3  Haplotype networks and pie charts showing the geographical distribution of haplotypes
for the mitochondrial genetic marker COI (632 bp) in Curdiea racovitzae and Palmaria decipiens.
In the networks, each circle represents a haplotype and its size is proportional to the frequency in
which the haplotype was encountered, and black lines correspond to one mutational step. The five
localities correspond, from north to south, to King George Island (GEO), Greenwich Island (PRA),
O’Higgins (OHI), Paradise Bay (PAR) and Marguerite Bay (MAR). The number of sequenced
individuals is given between brackets. From the six species of red algae studied (Guillemin et al.
2018), only results obtained for C. racovitzae and P. decipiens are illustrated since these two spe-
cies represent the less and the most diverse Rhodophyta sampled along the coasts of the Antarctic
Peninsula and the South Shetland Islands, respectively. Underwater photographs of specimen’s
characteristic of both species were taken by I.  Garrido; please note the impact of ice scour on
nearby sea bed P. decipiens populations. (King George Island)

6.5  B
 rown, Red and Green Macroalgae: Sharing a Common
Pattern of Glacial Impact and Postglacial
Populations Recovery?

All the analysed macroalgae are non-buoyant and commonly found in the Antarctic
waters. Nevertheless, they present some noticeable ecological differences. While
I. cordata, G. skottsbergii, C. racovitzae and P. decipiens are mostly found in the
intertidal down to the shallow subtidal, G. confluens and P. cartilagineum are gener-
ally found deeper in the subtidal, as understory of large brown macroalgae.
Monostroma hariotii presents very thin, delicate thalli, is very common in both the
intertidal and subtidal zones and is considered a pioneer species able to colonize
112 M.-L. Guillemin et al.

Fig. 6.4  Haplotype network and pie charts showing the geographical distribution of haplotypes
for the plastid genetic marker tufA (772 bp) in Monostroma hariotii and distribution of the unique
haplotype detected in the 19 COI sequences (mitochondrial marker, 619 bp, include HE866784
from King George Island and GQ368262 from Terre Adelie already published in GENBANK) in
Himantothallus grandifolius. In the network, each circle represents a haplotype and its size is
proportional to the frequency in which the haplotype was encountered; black line corresponds to
one mutational step. The five localities correspond, from north to south, to King George Island
(GEO), Greenwich Island (PRA), O’Higgins (OHI), Paradise Bay (PAR) and Marguerite Bay
(MAR). The number of sequenced individuals is given between brackets. An underwater photo-
graph of a mixed high subtidal bed of M. hariotii (arrow) and H. grandifolius (arrow heads) is
given on the top. (King George Island) (Photo by Ignacio Garrido)
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 113

areas with freshwater influence from glacier run-off, presenting heavy ice scour or
recently available for colonization after glacier retreat (Quartino et  al. 2013;
Wiencke et al. 2014). With thalli more than ten meters long, H. grandifolius is the
largest seaweed reported in Antarctica. The species, in combination with other
brown macroalgae of the genus Desmarestia, form dense beds of canopy-forming
algae, dominating the subtidal rocky shores.

6.5.1  S
 ignature of a Drastic Impact of the Last
Glacial Maximum

Regardless of their ecological differences or classification into distinct taxonomic


divisions, molecular data show that all eight macroalgae have been strongly
impacted by glacial events (Table  6.1, Figs.  6.3 and 6.4). Extremely low genetic
diversity was observed at the intraspecific level with a number of haplotypes vary-
ing between one in H. grandifolius up to only seven in the case of I. cordata and
P. decipiens (Table 6.1). These results could be related to historical changes in mac-
roalgal population size in the study area. The complete lack of diversity observed
for H. grandifolius does not allow to further test this hypothesis; however, the seven
other species presented additional results that support this idea: a starlike topology
was observed in haplotype networks (Figs.  6.3 and 6.4), mismatch distributions
were unimodal (hypothesis of sudden expansion could not be rejected when calcu-
lated for the sum of squared deviation, SSD, or Harpending’s raggedness index,
Rag, except for M. hariotii Rag for which a p < 0.05 was obtained; Table 6.1; see
also Guillemin et al. 2008 for more details on red algae species) and all but the green
alga M. hariotii presented negative values of Tajima’s D and Fu’s Fs statistics
(Table 6.1).
None to very few mutations have been accumulated in the populations since
expansion from their glacial refugium, strengthening the idea that the demographic
expansion occurred recently. We proposed (as in Billard et al. 2015; Guillemin et al.
2018; estimations based on the calculation of the parameter Tau from models of
sudden expansion and previously published mutation rate per sequence per genera-
tion for mitochondrial genes in macroalgae) that the demographic bottleneck cor-
responds to population contraction of macroalgae in refugium during the LGM and
that recolonization of the coast begun some 18,000 years ago, at most, a date con-
gruent with time of deglaciation in the area (Simms et al. 2011; O’Cofaigh et al.
2014). Similarly, other studies have related the recent expansion detected in the
southern most populations of sub-Antarctic macroalgae to population recolonizing
areas previously covered by ice after the LGM (Fraser et  al. 2009, 2012, 2013;
Montecinos et al. 2012; Billard et al. 2015).
The level of genetic diversity estimated for Antarctic macroalgae is within the
lowest reported for Antarctic marine species with broad distribution, with values of
haplotype diversity (H) ranging between 0.000 and 0.398 and nucleotide diversity
114 M.-L. Guillemin et al.

Table 6.1  Genetic diversity indices and neutrality test in six red algae, one brown alga and one
green alga sampled along the Antarctic Peninsula and South Shetland Islands. Sequences used for
analyses of all six species of Rhodophyta and Himantothallus grandifolius correspond to the
mitochondrial gene COI, while the plastid gene tufA was used for analyses of Monostroma hariotii
π Mismatch
Species N k S H Π (×102) Tajima’s D Fu’s FS distribution
Rhodophyta
 Curdiea 42 2 1 0.048 0.048 0.008 −1.120∗ −1.491 ns Unimodala
racovitzae
 Georgiella 20 4 4 0.363 0.489 0.077 −1.638∗ −1.613 ns Unimodala
confluens
 Gigartina 28 2 1 0.071 0.071 0.011 −1.151 ns −1.155 ns Unimodala
skottsbergii
 Iridaea cordata 90 7 5 0.398 0.623 0.102 −0.797 ns −2.882 ns Unimodala
 Palmaria 35 7 6 0.318 0.343 0.054 −2.103∗∗∗ −7.041∗∗∗ Unimodala
decipiens
 Plocamium 64 4 4 0.122 0.155 0.025 −1.759∗∗∗ −3.466∗∗ Unimodala
cartilagineum
Ochrophyta
 Himantothallus 19b 1 0 0.000 0.000 0.000
grandifolius
Clorophyta
 Monostroma 37c 2 1 0.378 0.378 0.049 0.846ns 1.223ns Unimodald
hariotii
N, number of sampled specimens; k, number of haplotypes; S, polymorphic sites; H, haplotype
diversity; Π, average number of nucleotide difference; π, nucleotide diversity; ns, non-significant.
All results obtained for Rhodophyta were taken from Guillemin et al. (2018). COI sequences are
of 632 base pair (bp)
∗p < 0.05
∗∗p < 0.01
∗∗∗p < 0.001
a
Hypothesis of sudden expansion could not be rejected when calculated for the sum of squared
deviation or Harpending’s raggedness index (i.e. 0.000 < SSD < 0.007; 0.179 < Rag < 0.821; all
p > 0.05). See Guillemin et al. (2018) for more detail
b
Include two sequences available in GENBANK: HE866784 from King George Island (Yang et al.
2014) and GQ368262 from Terre Adelie (Silberfeld et al. 2010). All COI sequences are of 619 bp
c
Sequences for the plastid gene tufA were all of 772 bp, except for the individual GGVMLG0575,
which was 52 bp shorter. Since the 3′ part of the tufA gene was not variable in our whole data set,
GGVMLG0575 was completed to 772 bp as in the other individuals before analyses
d
Goodness-of-fit tests for a model of sudden expansion calculated in ARLEQUIN v.3.5 for the sum
of squared deviation (SSD  =  0.005, p  =  0.0001) and for Harpending’s raggedness index
(Rag = 0.202, p = 0.459)

(π × 102) ranging between 0.000 and 0.102 (lowest values in H. grandifolius and
highest values in I. cordata; Table 6.1). Rather low level of diversity and complete
lack of genetic structure have also been reported in some widespread Antarctic
marine invertebrates as Sterechinus neumayeri (H = 0.257; π × 102 = 0.036; Díaz
et  al. 2018), Chorismus antarcticus (H  =  0.639; π  ×  102  =  0.209; Raupach et  al.
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 115

2010), Parborlasia corrugatus (H = 0.762; π × 102 = 0.118; Thornhill et al. 2008),


Euphausia superba (H  =  0.856; π  ×  102  =  1.394; Bortolotto et  al. 2011),
Nematocarcinus lanceopes (H = 0.902; π × 102 = 0.568; Raupach et al. 2010) and
Nymphon australe (H = 0.918; π × 102 = 0.657; Soler-Membrives et al. 2017). The
lowest levels of genetic diversity have been observed in benthic marine inverte-
brates with distribution restricted to shallow waters of the Antarctic shelf and have
been related to the extreme effect of ice scouring in these organisms, not able to
migrate to deeper waters during glacial maxima (Raupach et al. 2010; Díaz et al.
2018). It is then not surprising to observe genetic diversity congruent with extreme
events of mortality due to LGM ice impact in photosynthetic macroalgae, these
organisms being characterized by one of the most restricted bathymetry in the
marine realm.
Indeed, Antarctic macroalgae are anchored to the sea bottom and cannot survive
under ice, without direct sunlight, for long periods of time. Glacial refugia for these
organisms should then be limited in depth (generally dense beds of Antarctic mac-
roalgae are reported growing down to 40–50 m) and to areas where the sea ice melt
during summer. However, periods of ice-free summer as short as 2 months allow
macroalgal bed growth (Norkko et al. 2004) and living specimens of various mac-
roalgae species were retrieved in Terre Adelie under ice sheets in an area that did not
melt during summer for 2 years in a row (S. Hourdez pers. com., specimens kept in
the Museum of Natural History, Paris, France). This time period could, however, be
considered as a maximum before severe population decline. Indeed, none of the
fleshy macroalgae were considered as healthy (but half the encrusting coralline
algae were still pigmented) after 3  years of permanent sea ice coverage in the
Commonwealth Bay, a polynya affected by the grounding of a huge drifting iceberg
in 2010 (Clark et al. 2015). Moreover, Antarctic macroalgae are extremely shade
adapted and have been observed to grow at depth of 90 m (Wiencke and Clayton
2002; Wiencke et al. 2007). In other regions of the world, the existence of kelp beds
reaching depth of 200 m was even predicted using oceanographic and ecophysio-
logical models (Graham et  al. 2007; see also Spalding et  al. 2019 for reports of
macroalgae growing down to 140 m). Additionally, long-term survival of Antarctic
macroalgae in areas highly impacted by ice could be facilitated by the synthesis and
accumulation of protective compounds (Wiencke et al. 2007). Some species, such
as the green algae Acrosiphonia arcta and Prasiola spp., have developed remark-
able adaptations to cope with ice disturbance characteristic of Antarctic habitats.
Indeed, both synthesize cryoprotectants that prevent damage by ice to membranes
or enzymes and Prasiola has been reported to photosynthesize at temperatures as
low as −15 °C (Jacob et al. 1991; Raymond and Fritsen 2001; Wiencke et al. 2007).
Antarctic macroalgae could then be tougher than expected and may have survived
in small areas highly impacted by ice during the LGM.
116 M.-L. Guillemin et al.

6.5.2  One Refugium to Rule Them all

Classically, locations of glacial refugia have been inferred by comparing genetic


diversity between populations located within areas potentially free of ice during the
LGM and the ones located in recently colonized areas, previously covered by ice
(Hewitt 2000, 2004; Maggs et al. 2008; Provan and Bennett 2008). Since population
effective size is more stable in glacial refugia than in recently colonized areas, a
much higher genetic diversity is expected to characterize the former than the later
(Hewitt 2000, 2004; Maggs et al. 2008; Provan and Bennett 2008). If the particu-
larly low genetic variability observed in our species lead to the idea of one unique
refugium in Antarctica, it frustratingly does not allow pinpointing its location.
Indeed, whatever the species under study, none of our sampled populations shows a
specifically high genetic diversity (Figs. 6.3 and 6.4; see also Guillemin et al. 2018
for more details on distribution of genetic diversity of I. cordata, G. confluens,
G. skottsbergii and P. cartilagineum).
Various areas could be proposed as in situ refugia along the coast of Antarctica
for our seaweeds and these include localities close to active volcanoes (as Deception
Island, Penguin Island and Bridgeman Island in the South Shetland Islands or the
coast of the Western Ross Sea; Fraser et al. 2014) or localities where the shelf sea-
floor was free of grounded ice during glacial maxima (i.e. Alexander Island, Eastern
Amundsen Sea Embayment, Western Ross Sea, George V Land and Prydz Bay;
Klages et al. 2017). Moreover, all eight macroalgal distributions include offshore
peri-Antarctic Islands located south of the APF, such as the South Orkney Islands,
South Sandwich Islands and South Georgia (Wiencke and Clayton 2002; Wiencke
et al. 2014; Griffiths and Waller 2016). We then cannot rule out a recolonization of
the Antarctic coasts from these more northern areas. Supporting this idea, a study
based on molecular markers allowed the recent discovery of one cryptic species of
the brown alga Adenocystis utricularis endemic to South Georgia (Fraser et  al.
2013). The authors related this result to a persistence of Adenocystis in local refu-
gium along the island coasts during the LGM (Fraser et al. 2013). Antarctic samples
were not included in the study of Fraser et al. (2013) and the level of genetic diver-
gence between peri-Antarctic and Antarctic Adenocystis remains to be estimated in
order to test for a possible recolonization of the Antarctic shelf from a South Georgia
glacial refugium. However, ice coverage during the LGM strongly impacted South
Georgia (White et al. 2018) and coastal glacial refugia could have been as scarce in
the offshore peri-Antarctic Islands as at the margins of the Antarctic shelf itself
(Fig. 6.1). Indeed, the complete eradication of populations of another brown alga,
the intertidal/shallow subtidal Durvillaea antarctica, from the coasts of most peri-
Antarctic islands (in particular in Macquarie Island, Falkland Islands, South
Georgia, Marion Island and Kerguelen Islands) during the LGM has been estab-
lished (Fraser et al. 2009, 2012).
Unfortunately, in order to locate glacial refugia in the region, we will need more
variable genetic markers and a much better sampling of the coasts of the Antarctic
shelf (i.e. Alexander Island, Western Ross Sea, George V Land and Prydz Bay; see
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 117

Griffiths and Waller 2016), adjacent islands of the maritime Antarctic and offshore
peri-Antarctic Islands south of the APF where populations of macroalgae have been
reported (in particular, South Orkney Islands, South Sandwich Islands and South
Georgia, Bouvet Island, Heard Island and Balleny Islands). Sampling localities that
could have been less affected by ice scour during the Last Glacial Maximum should
be a priority.
The proposed scenario of one unique in situ refugium in Antarctica contrasts
with history of glacial perturbations in Arctic macroalgae. Indeed, based on patterns
of genetic diversity and distribution of private haplotypes, the existence of various
marine glacial refugia has been proposed in the Northern Hemisphere for these
organisms (Maggs et al. 2008; Provan and Bennett 2008; Neiva et al. 2016). Less
severe cooling in the Arctic than the Antarctic could explain, in part, these differ-
ences between hemispheres (Pointing et  al. 2015). Moreover, Arctic macroalgae
distribution during glacial cycles could have shifted along continuous coastlines
spanning a huge latitudinal gradient (Fraser et al. 2012; Neiva et al. 2016), while no
such opportunity was available for Antarctic species, bounded by the deep water of
the Southern Ocean, leading to a more thorough extinction along the Antarctic
coastline during the LGM.

6.5.3  P
 ostglacial Recolonization: Widespread Haplotypes
Drifting Around Antarctica?

Apart from their very low genetic diversity, all eight algae under study present a
common striking characteristic: one or a few haplotypes were encountered in all
localities even when situated hundreds or even thousands of kilometres apart
(Figs. 6.3 and 6.4). For all six red algae and for the green alga M. hariotii, the most
common haplotype was observed from King George Island in the South Shetland
Islands down to Marguerite Bay, a sampling point located along the central part of
the WAP (Figs. 6.3 and 6.4). These common haplotypes are spread over more than
450  km of coast and within two distinct biogeographic subregions (the South
Shetland Islands and the WAP; Linse et al. 2006; Spalding et al. 2007; Terauds et al.
2012). Shared haplotypes have been observed even between the WAP and South
Orkney Island, reaching distances greater than 1600 km, in the red alga G. skotts-
bergii (same Cox2–3 haplotype from Marguerite Bay to the South Orkney Islands;
Billard et al. 2015). The same pattern is observed in I. cordata for which the same
COI haplotype has also been sequenced along the WAP and South Orkney Island
(Guillemin M-L. unpublished data, no genetic differences detected between the
most common haplotype encountered along the WAP and three individuals from
South Orkney Island sequenced for 632 pb). For H. grandifolius, the distribution of
the unique COI haplotype encountered in our study zone could be extended to Terre
Adelie, in a sampling site located almost on the other side of the Antarctic continent
(sequence from Silberfeld et al. 2010). Himantothallus grandifolius is then the first
118 M.-L. Guillemin et al.

register of a non-buoyant species displaying a true Antarctic circumpolar distribu-


tion supported by genetic data. Complete circum-Antarctic distribution has been
described for other seaweeds, including common species of brown (Desmarestia
menziesii), green (M. hariotii and Urospora penicilliformis) and red (I. cordata,
P. cartilagineum, P. decipiens, Phyllophora antarctica and Phycodrys antarctica)
macroalgae (Wiencke et al. 2014). However, for these species, all sequences avail-
able to date have been obtained for specimens located in the South Shetland Islands
and/or the WAP (Olsen et  al. 1993; Peters et  al. 2000; Hommersand et  al. 2009;
Dubrasquet et al. 2018; Ocaranza-Barrera et al. 2019). One could thus wonder if
other common Antarctic macroalgae species will present the same pattern as
H. grandifolius and samples should be gathered from the Western Ross Sea, Terre
Adelie and Prydz Bay and sequenced in order to test this hypothesis.
In contrast to what was reported for pelagic species or benthic species exhibiting
extensive dispersal by planktonic stage (Lange et  al. 2002; Raupach et  al. 2010;
Sromek et al. 2015; Soler-Membrives et al. 2017; Caccavo et al. 2018), we believe
that the presence of widespread haplotypes in our data sets clearly reflects Antarctic
macroalgae past glacial demographic histories but not necessarily contemporary
high gene flow. Benthic macroalgae free-living stages (i.e. for all species except
buoyant ones, Macaya et al. 2016, also see chapter by Macaya et al. 2020 in this
volume) are generally extremely restricted to short-lived gametes and spores that
quickly sink a few meters, at most, from thalli of origin (Valero et  al. 2011).
However, in these organisms, large-scale spread of haplotypes was commonly
related to strong genetic drift and allelic surfing at species’ leading edges during
postglacial range expansions or after transoceanic colonization (Maggs et al. 2008;
Provan and Bennett 2008; Montecinos et  al. 2012; Fraser et  al. 2013; Guillemin
et  al. 2014; Neiva et  al. 2016). Non-buoyant macroalgal species, for which both
population genetic and phylogeographic data sets are available, show well this dis-
crepancy with, on the one hand, the presence of haplotypes distributed over thou-
sands of kilometres in area of recent colonization and, on the other hand, current
reproduction and recruitment occurring at a very local scale (in Agarophyton [for-
mer Gracilaria]: Guillemin et al. 2008, 2014, in Fucus: Neiva et al. 2016; Jueterbock
et al. 2018). Exceptional events of thalli rafting on strong oceanic currents and colo-
nizing new areas when depleted of abundant algal coverage (Waters et  al. 2013)
have been advanced to explain these patterns (Montecinos et al. 2012; Fraser et al.
2013; Guillemin et al. 2014; Neiva et al. 2016). In order to test for actual gene flow
between studied populations of macroalgae, more variable genetic markers, such as
microsatellites or SNPs (i.e. single nucleotide polymorphism), will be needed.
Indeed, highly variable microsatellite loci revealed a clear regional substructure of
populations in Phaeocystis antarctica, a brown micro algal planktonic species for
which no clear phylogeographic patterns were described before around Antarctica
using nuclear and plastid sequences (Lange et al. 2002; Gäbler-Schwarz et al. 2015).
Even if none of the study models possess floating structures, large amounts of all
eight species can easily be found cast ashore after storm events, including huge
adult specimens of H. grandifolius (authors pers. obs.). Passive transport of detached
fronds could therefore play a major role in Antarctic non-buoyant macroalgae
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 119

long-­distance dispersal, a mechanism most probably facilitated by the strong oce-


anic currents in the region (Loeb 2007). Classically, the circumpolar current (ACC,
see Fig. 6.1) was reported as a structuring force in the Southern Ocean and has been
proposed to be the main connecting current between populations of marine organ-
isms located around Antarctica (e.g. Fraser et  al. 2009; Soler-Membrives et  al.
2017). For the red algae G. skottsbergii and I. cordata, fronds rafting on the ACC
could easily explain the spread of the most common haplotypes between localities
of the South Shetland Islands, the Antarctic Peninsula and even South Orkney Island
(Billard et al. 2015; Guillemin et al. 2018; Ocaranza-Barrera et al. 2019). Indeed,
the ACC flows northeastward from the Bellingshausen Sea through the South
Shetland Islands and the South Orkney Islands, before entering the Scotia Sea (Loeb
2007). However, since it deflects from the Antarctic coasts during most of its jour-
ney around the continent (from the Weddell Sea to Prydz Bay and the Western Ross
Sea, see Loeb 2007), rafting on the ACC does not easily explain the presence of the
same COI haplotype in Terre Adelie and the Antarctic Peninsula for H. grandifolius.
It is highly probable that fronds transport between these two distant regions was
promoted by the Antarctic Costal Current (ACoC), a current that flows in a counter-
clockwise motion and located near the coast of the Antarctic shelf (Loeb 2007). The
importance of transport by the ACoC was also proposed to be at the origin of the
widespread haplotypes observed in marine species with distribution restricted to the
shallow waters of the shelf, as in the sea urchin S. neumayeri (Díaz et al. 2018).
Remarkably, a study following the track of icebergs around Antarctica shows that a
large majority do travel counterclockwise around the continental shelf and that the
ACoC can very easily connect Terre Adelie with the tip of the Antarctic Peninsula
(Stuart and Long 2011).

6.6  Concluding Remarks

Macroalgae are very important primary producers and ecosystem engineers in


Antarctic coastal ecosystems, serving as food for herbivores and detritivores and
providing habitat for many associated organisms (see Chaps. 15 and 16 of this
book). Even if these species have developed remarkable adaptations to cope with
the high level of disturbances characteristic of Antarctic swallow depth (Wiencke
et al. 2007, 2014), they could be highly vulnerable to actual threats intensifying in
the region as global warming, pollution and introduction of invasive species. As
seawater temperature rise around Antarctica, a trend particularly noticeable along
the Antarctic Peninsula, extent and duration of sea ice coverage decline and glaciers
are in retreat (Cook et al. 2016). At first sight, this scenario seems highly favourable
for Antarctic macroalgae that can rapidly colonize these newly deglaciated areas
(Quartino et  al. 2013; also see chapter by Quartino et  al. 2020 in this volume).
Indeed, in the Arctic, the long-term study of Kortsch et  al. (2012) shows that
decrease in sea ice cover can produce rapid shift in marine communities from ben-
thic habitats dominated by calcareous algae and sea anemones to dense beds of
120 M.-L. Guillemin et al.

fleshy, habitat-forming macroalgae. However, climatic changes are also associated


with higher probability of ice sheet collapse and iceberg calving. Scouring by these
drifting icebergs, which can even get stuck and ground in polynyas, may have a
highly detrimental effect on macroalgae beds (Clark et al. 2015). Climate changes
have also been related to possible alteration in large-scale ocean circulation and
gyre and eddy kinetic in the Southern Ocean, leading to breaches in the dispersal
barriers (APF and ACC) established since the Eocene/Oligocene in the region
(Barker and Thomas 2004; Fraser et al. 2018). When coupled with the intensifica-
tion of maritime transport (linked to both touristic and scientific activities) between
Antarctic and sub-Antarctic provinces, these oceanic changes could lead to an
increase in propagule pressure and introduction of sub-Antarctic macroalgae spe-
cies in the Antarctic waters. To the date, only one green alga (i.e. Ulva intestinalis)
has been reported as an established introduced alien species in the South Shetland
Islands, probably arriving as fouling on the hull of a visiting vessel (Clayton et al.
1997). However, macroalgae highly successful in recolonising sub-Antarctic coasts
after glacial retreat have recently been reported to recurrently reach the Antarctic
coasts as cast ashore living thalli (i.e. Durvillaea antarctica, Fraser et al. 2018; also
see chapter by Macaya et al. 2020 in this volume). The capacity of Antarctic mac-
roalgae to withstand competition from sub-Antarctic invaders has not been studied
yet. Nonetheless, Antarctic macroalgae are organisms highly adapted to the local
stenothermal environment. Studies report temperature optima between −2 °C and
10 °C for most Antarctic species with limitation or even failure in growth or game-
togenesis at temperature higher than 5 °C in G. confluens, G. skottsbergii, P. carti-
lagineum and Desmarestia antarctica (Wiencke et al. 2007; Wiencke et al. 2014). In
the actual context of rapid temperature increase, populations of sub-Antarctic or
even more temperate colonizers may then more easily settle along the Antarctic coasts.
Antarctic macroalgae were clearly able to cope with changes related to the cool-
ing and then freezing of the Southern Ocean waters and have survived Quaternary
glacial perturbations in situ. However, our study shows that maximal glacial events
have led to mass extinction in all eight model species, whatever their taxonomic
divisions. Will these species be able to keep pace with the current rapid environmen-
tal fluctuations? Genetic adaptation could be key in enhancing resistance and resil-
ience to climatic changes and a high level of genetic diversity has classically been
related to high population fitness, resilience and future adaptability (Reed and
Frankham 2003; Jump and Peñuelas 2005). Due to strong bottleneck during the
LGM, standing genetic diversity in Antarctic macroalgae populations seems to be
extremely depleted and an adaptive response of Antarctic algae populations to
future changes could be limited. Nevertheless, recent studies on the brown seaweed
Sargassum muticum show a complete absence of genetic diversity in all populations
of the invaded range when tested using mitochondrial sequences or even nuclear
microsatellites and only a very low level of diversity when using more than 8000
SNPs (Cheang et al. 2010; Le Cam et al. 2019). Rapid and successful invasion of
Europe and the USA was clearly not impaired by the extremely low genetic diver-
sity in S. muticum. In order to predict the possible outcome of future climatic
changes on macroalgae populations, it is now essential to use more variable genetic
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 121

markers in order to detect a subtle genetic structure along the Antarctic coasts and
point out restricted regional or local genotypes that could be of interest (see results
using SNPs in Le Cam et  al. 2019) and to experimentally test for existence and
extent of phenotypic plasticity/acclimatization potential (a mechanism that can
potentially buffer negative effects of climate change long enough to allow for popu-
lation adaptation; Jump and Peñuelas 2005) and potential for an in situ adaptive
response to climate change of these organisms.

Acknowledgments  Funding by Instituto Antártico Chileno (INACH) T_16-11 and RG_15-16


and the Centro FONDAP IDEAL no. 15150003 is gratefully acknowledged. At King George
Island, sampling of Monostroma hariotii and Himantothallus grandifolius was done with the help
of the Project Anillo ART1101 from Conicyt-Chile. The authors thank P. Brunning, J. L. Kappes,
T. Heran, Y. Henriquez and L. Vallejos for their help in the field. The authors would also like to
thank the Chilean Navy (especially the captain and crew of the ships Almirante Oscar Viel and
Lautaro), the staff from the Chilean Army in the O’Higgins base and the Air Force of Chile (FACh)
for the logistic support of our fieldwork in Subantarctica and Antarctica.

References

Allcock AL, Strugnell JM (2012) Southern Ocean diversity: new paradigms from molecular ecol-
ogy. Trends Ecol Evol 27:520–528
Anderson JB, Shipp SS, Lowe AL, Wellner JS, Mosola AB (2002) The Antarctic ice sheet during
the last glacial maximum and its subsequent retreat history: a review. Quat Sci Rev 22:49–70
Aronson RB, Thatje S, Clarke A, Peck LS, Blake DB, Wilga CD, Seibel BA (2007) Climate change
and invasibility of the Antarctic benthos. Annu Rev Ecol Evol Syst 38:129–154
Bandelt HJ, Forster P, Röhl A (1999) Median-Joining networks for inferring intraspecific phylog-
enies. Mol Biol Evol 16:37–48
Barker P, Thomas E (2004) Origin, signature and palaeoclimatic influence of the Antarctic
Circumpolar Current. Earth Sci Rev 66:143–162
Barnes DK, Hodgson DA, Convey P, Allen CS, Clarke A (2006) Incursion and excursion of
Antarctic biota: past, present and future. Glob Ecol Biogeogr 15:121–142
Bentley MJ, Ó Cofaigh CO, Anderson JB, Conway H, Davies B, Graham AG et al (2014) A
community-based geological reconstruction of Antarctic ice sheet deglaciation since the Last
Glacial Maximum. Quat Sci Rev 100:1–9
Billard E, Reyes J, Mansilla A, Faugeron S, Guillemin ML (2015) Deep genetic divergence
between austral populations of the red alga Gigartina skottsbergii reveals a cryptic species
endemic to the Antarctic continent. Polar Biol 38:2021–2034
Bortolotto E, Bucklin A, Mezzavilla M, Zane L, Patarnello T (2011) Gone with the currents:
lack of genetic differentiation at the circum-continental scale in the Antarctic krill Euphausia
superba. BMC Genet 12:32
Caccavo JA, Papetti C, Wetjen M, Knust R, Ashford JR, Zane L (2018) Along-shelf connectivity
and circumpolar gene flow in Antarctic silverfish (Pleuragramma antarctica). Sci Rep 8:17856
Cheang CC, Chu KH, Fujita D, Yoshida G, Hiraoka M, Critchley A, Choi HG, Duan D, Serisawa
Y, Ang PO Jr (2010) Low genetic variability of Sargassum muticum (Phaeophyceae) revealed
by a global analysis of native and introduced populations. J Phycol 46:1063–1074
Clark GF, Marzinelli EM, Fogwill CJ, Turney CS, Johnston EL (2015) Effects of sea-ice cover on
marine benthic communities: a natural experiment in Commonwealth Bay, East Antarctica.
Polar Biol 38:1213–1222
122 M.-L. Guillemin et al.

Clarke A, Crame JA (1989) The origin of the Southern Ocean marine fauna. In: Crame JA
(ed) Origins and evolution of the Antarctic biota, vol 47. The Geol Soc Spec Publ, London,
pp 253–268
Clarke A, Crame JA (1992) The Southern Ocean benthic fauna and climate change: a historical
perspective. Phil Trans R Soc Lond B 338:299–309
Clarke A, Barnes DKA, Hodgson DA (2005) How isolated is Antarctica? Trends Ecol Evol 20:1–3
Clayton MN, Wiencke C, Klöser H (1997) New records and sub-Antarctic marine benthic mac-
roalgae from Antarctica. Polar Biol 17:141–149
Convey P, Gibson JA, Hillenbrand CD, Hodgson DA, Pugh PJ, Smellie JL, Stevens MI (2008)
Antarctic terrestrial life–challenging the history of the frozen continent? Biol Rev 83:103–117
Convey P, Stevens M, Hodgson D, Smellie J, Hillenbrand C, Barnes D, Clarke A, Pugh P, Linse K,
Cary S (2009) Exploring biological constraints on the glacial history of Antarctica. Quat Sci
Rev 28:3035–3048
Cook AJ, Holland PR, Meredith MP, Murray T, Luckman A, Vaughan DG (2016) Ocean forcing of
glacier retreat in the western Antarctic Peninsula. Science 353:283–286
Crame JA (1999) An evolutionary perspective on marine faunal connections between southern-
most South America and Antarctica. Sci Mar 63:1–14
Crame JA (2018) Key stages in the evolution of the Antarctic marine fauna. J Biogeogr 45:986–994
Dalziel IWD, Lawver LA, Pearce JA, Barker PF, Hastie AR, Barfod DN, Schenke HW, Davis MB
(2013) A potential barrier to deep Antarctic circumpolar flow until the late Miocene? Geology
41:947–950
Dambach J, Thatje S, Rödder D, Basher Z, Raupach MJ (2012) Effects of late-Cenozoic glaciation
on habitat availability in Antarctic benthic shrimps (Crustacea: Decapoda: Caridea). PLoS One
7:e46283
Davies B, Hambrey M, Smellie J, Carrivick J, Glasser N (2012) Antarctic Peninsula ice sheet
evolution during the Cenozoic Era. Quat Sci Rev 31:30–66
Dell RK (1972) Antarctic benthos. Adv Mar Sci 10:1–216
Díaz A, Gerard K, González-Wevar C, Maturana C, Féral JP, David B, Saucède T, Poulin E (2018)
Genetic structure and demographic inference of the regular sea urchin Sterechinus neumayeri
(Meissner, 1900) in the Southern Ocean: the role of the last glaciation. PLoS One 13:e0197611
Dubrasquet H, Reyes J, Sanchez RP, Valdivia N, Guillemin ML (2018) Molecular-assisted revision
of red macroalgal diversity and distribution along the Western Antarctic Peninsula and South
Shetland Islands. Cryptogamie Algol 39:409–430
Excoffier L, Lisher H (2010) Arlequin suite ver 3.5: a new series of programs to perform popula-
tion genetics analyses under Linux and Windows. Mol Ecol Resour 10:56
Excoffier L, Ray N (2008) Surfing during population expansions promotes genetic revolutions and
structuration. Trends Ecol Evol 23:347–351
Excoffier L, Foll M, Petit RJ (2009) Genetic consequences of range expansions. Annu Rev Ecol
Evol Syst 40:481–501
Famà P, Wysor B, Kooistra WH, Zuccarello GC (2002) Molecular phylogeny of the genus Caulerpa
(Caulerpales, Chlorophyta) inferred from chloroplast tufA gene. J Phycol 38:1040–1050
Fraser CI, Nikula R, Spencer HG, Waters JM (2009) Kelp genes reveal effects of Subantarctic sea
ice during the Last Glacial Maximum. Proc Natl Acad Sci 106:3249–3253
Fraser CI, Nikula R, Ruzzante DE, Waters JM (2012) Poleward bound: biological impacts of
Southern Hemisphere glaciation. Trends Ecol Evol 27:462–471
Fraser CI, Zuccarello GC, Spencer HG, Salvatore LC, Garcia GR, Waters JM (2013) Genetic
affinities between trans-oceanic populations of non-buoyant macroalgae in the high latitudes
of the Southern Hemisphere. PLoS One 8:e69138
Fraser CI, Terauds A, Smellie J, Convey P, Chown SL (2014) Geothermal activity helps life survive
glacial cycles. Proc Natl Acad Sci 111:5634–5639
Fraser CI, Morrison AK, Hogg AM, Macaya EC, van Sebille E, Ryan PG, Padovan A, Jack C,
Valdivia N, Waters JM (2018) Antarctica’s ecological isolation will be broken by storm-driven
dispersal and warming. Nat Clim Chang 8:704
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 123

Fu YX (1997) Statistical tests of neutrality of mutations against population growth, hitchhiking


and background selection. Genetics 147:915–925
Gäbler-Schwarz S, Medlin LK, Leese F (2015) A puzzle with many pieces: the genetic structure
and diversity of Phaeocystis antarctica Karsten (Prymnesiophyta). Eu. J Phycol 50:112–124
Gersonde R, Crosta X, Abelmann A, Armand L (2005) Sea-surface temperature and sea ice dis-
tribution of the Southern Ocean at the EPILOG Last Glacial Maximum—a circum-Antarctic
view based on siliceous microfossil records. Quat Sci Rev 24:869–896
González-Wevar CA, Díaz A, Gerard K, Poulin E, Cañete JI (2012) Divergence time estimations
and contrasting patterns of genetic diversity between Antarctic and southern South America
benthic invertebrates. Rev Chil Hist Nat 85:445–456
González-Wevar CA, Saucède T, Morley SA, Chown SL, Poulin E (2013) Extinction and recolo-
nization of maritime Antarctica in the limpet Nacella concinna (Strebel, 1908) during the last
glacial cycle: toward a model of Quaternary biogeography in shallow Antarctic invertebrates.
Mol Ecol 22:5221–5236
González-Wevar CA, Hüne M, Segovia NI, Nakano T, Spencer HG, Chown SL, Saucède T,
Johnstone G, Mansilla A, Poulin E (2017) Following the Antarctic Circumpolar Current: pat-
terns and processes in the biogeography of the limpet Nacella (Mollusca: Patellogastropoda)
across the Southern Ocean. J Biogeogr 44:861–874
Graham MH, Kinlan BP, Druehl LD, Garske LE, Banks S (2007) Deep-water kelp refugia as poten-
tial hotspots of tropical marine diversity and productivity. Proc Natl Acad Sci 104:16576–16580
Griffiths HJ, Waller CL (2016) The first comprehensive description of the biodiversity and bio-
geography of Antarctic and Sub-Antarctic intertidal communities. J Biogeogr 43:1143–1155
Guillemin ML, Faugeron S, Destombe C, Viard F, Correa JA, Valero M (2008) Genetic variation in
wild and cultivated populations of the haploid–diploid red alga Gracilaria chilensis: how farm-
ing practices favor asexual reproduction and heterozygosity. Evolution 62:1500–1519
Guillemin ML, Valero M, Faugeron S, Nelson W, Destombe C (2014) Tracing the trans-Pacific
evolutionary history of a domesticated seaweed (Gracilaria chilensis) with archaeological and
genetic data. PLoS One 9:e114039
Guillemin ML, Dubrasquet H, Reyes J, Valero M (2018) Comparative phylogeography of six red
algae along the Antarctic Peninsula: extreme genetic depletion linked to historical bottlenecks
and recent expansion. Polar Biol 41:827–837
Haffer J (1969) Speciation in Amazonian forest birds. Science 165:131–137
Halanych KM, Mahon AR (2018) Challenging dogma concerning biogeographic patterns of
Antarctica and the Southern Ocean. Ann Rev Ecol Evol S 49:355–378
Hallatschek O, Nelson DR (2010) Life at the front of an expanding population. Evolution
64:193–206
Hemery LG, Eléaume M, Roussel V, Améziane GC, Steinke D, Cruaud C, Couloux A, Wilson NG
(2012) Comprehensive sampling reveals circumpolarity and sympatry in seven mitochondrial
lineages of the Southern Ocean crinoid species Promachocrinus kerguelensis (Echinodermata).
Mol Ecol 21:2502–2518
Hewitt GM (2000) The genetic legacy of the Quaternary ice ages. Nature 405:907
Hewitt GM (2004) Genetic consequences of climatic oscillations in the Quaternary. Phil Trans R
Soc Lond B 359:183–195
Hiller A, Wand U, Kämpf H, Stackebrandt W (1988) Occupation of the Antarctic Continent by
petrels during the past 35000 years: inferences from a 14C study of stomach oil deposits. Polar
Biol 9:69–77
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Marina 52:509–534
Hughes PD, Gibbard PL, Ehlers J (2013) Timing of glaciation during the last glacial cycle: evaluat-
ing the concept of a global ‘Last Glacial Maximum’(LGM). Earth Sci Rev 125:171–198
Jacob A, Kirst GO, Wiencke C, Lehmann H (1991) Physiological responses of the Antarctic green
alga Prasiola crispa ssp. antarctica to salinity stress. J Plant Physiol 139:57–62
124 M.-L. Guillemin et al.

Jueterbock A, Coyer JA, Olsen JL, Hoarau G (2018) Decadal stability in genetic variation and
structure in the intertidal seaweed Fucus serratus (Heterokontophyta: Fucaceae). BMC Evol
Biol 18:94
Jump AS, Peñuelas J (2005) Running to stand still: adaptation and the response of plants to rapid
climate change. Ecol Lett 8:1010–1020
Kemp A, Grigorov I, Pearce R, Naveira Garabato A (2010) Migration of the Antarctic Polar Front
through the mid-Pleistocene transition: evidence and climatic implications. Quat Sci Rev
29:1993–2009
Kennett JP (1977) Cenozoic evolution of Antarctic glaciation, the Circum-Antarctic ocean, and
their impact on global paleoceanography. J Geophys Res 82:3843–3860
Klages JP, Kuhn G, Hillenbrand CD, Smith JA, Graham AG, Nitsche FO, Frederichs T, Jernas PE,
Gohl K, Wacker L (2017) Limited grounding-line advance onto the West Antarctic continental
shelf in the easternmost Amundsen Sea Embayment during the last glacial period. PLoS One
12:e0181593
Kortsch S, Primicerio R, Beuchel F, Renaud PE, Rodrigues J, Lønne OJ, Gulliksen B (2012)
Climate-driven regime shifts in Arctic marine benthos. Proc Natl Acad Sci 109:14052–14057
Lane CE, Lindstrom SC, Saunders GW (2007) A molecular assessment of northeast Pacific Alaria
species (Laminariales, Phaeophyceae) with reference to the utility of DNA barcoding. Mol
Phylogenet Evol 44:634–648
Lange M, Chen YQ, Medlin LK (2002) Molecular genetic delineation of Phaeocystis species
(Prymnesiophyceae) using coding and non-coding regions of nuclear and plastid genomes. Eur
J Phycol 37:77–92
Larter RD, Anderson JB, Graham AGC, Gohl K, Hillenbrand CD, Jakobsson M et  al (2014)
Reconstruction of changes in the Amundsen Sea and Bellingshausen sea sector of the West
Antarctic ice sheet since the last glacial maximum. Quat Sci Rev 100:55–86
Le Cam S, Daguin-Thiébaut C, Bouchemousse S, Engelen AH, Mieszkowska N, Viard F (2019) A
genome-wide investigation of the worldwide invader Sargassum muticum shows high success
albeit (almost) no genetic diversity. Evol Appl 00:1–15 DOI: 10.1111/eva.12837
Lecointre G, Améziane N, Boisselier MC, Bonillo C, Busson F, Causse R et al (2013) Is the spe-
cies flock concept operational? The Antarctic shelf case. PLoS One 8:e68787
Linse K, Griffiths HJ, Barnes DK, Clarke A (2006) Biodiversity and biogeography of Antarctic and
sub-Antarctic mollusca. Deep Sea Res Pt II 53:985–1008
Loeb V (2007) Environmental variability and the Antarctic marine ecosystem. In: Vasseur DA,
McCann KS, Vasseur DA (eds) The impact of environmental variability on ecological systems.
Springer, Dordrecht, pp 197–225
Macaya EC, López B, Tala F, Tellier F, Thiel M (2016) Float and raft: role of buoyant seaweeds in
the phylogeography and genetic structure of non-buoyant associated flora. In: Hu ZM, Fraser
C (eds) Seaweed phylogeography. Springer, Dordrecht, pp 97–130
Macaya EC, Tala F, Hinojosa I, Rothäusler E (2020) Detached seaweeds as important dispersal
agents across the Southern Ocean. In: Gómez I, Huovinen P (eds) Antarctic seaweeds: diver-
sity, adaptation and ecosystem services. Springer, Cham, Switzerland, pp 59–75
Mackensen A (2004) Changing Southern Ocean paleocirculation and effects on global climate.
Antarct Sci 16:369–386
Maggs CA, Castilho R, Foltz D, Henzler C, Jolly MT, Kelly J et al (2008) Evaluating signatures of
glacial refugia for North Atlantic benthic marine taxa. Ecology 89:S108–S122
McCulloch RD, Bentley MJ, Purves RS, Hulton NRJ, Sugden DE, Clapperton CM (2000) Climatic
inferences from glacial and palaeoecological evidence at the last glacial termination, southern
South America. J Quat Sci 15:409–417
Moniz MB, Rindi F, Novis PM, Broady PA, Guiry MD (2012) Molecular phylogeny of Antarctic
Prasiola (Prasiolales, Trebouxiophyceae) reveals extensive cryptic diversity. J Phycol
48:940–955
Montecinos A, Broitman BR, Faugeron S, Haye PA, Tellier F, Guillemin ML (2012) Species
replacement along a linear coastal habitat: phylogeography and speciation in the red alga
Mazzaella laminarioides along the south east Pacific. BMC Evol Biol 12:01–97
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 125

Moon KL, Chown SL, Fraser CI (2017) Reconsidering connectivity in the sub-Antarctic. Biol Rev
92:2164–2181
Mystikou A, Peters AF, Asensi AO, Fletcher KI, Brickle P, van West P, Convey P, Küpper FC
(2014) Seaweed biodiversity in the south-western Antarctic Peninsula: surveying macroalgal
community composition in the Adelaide Island/Marguerite Bay region over a 35-year time
span. Polar Biol 37:1607–1619
Near TJ, Dornburg A, Kuhn KL, Eastman JT, Pennington JN, Patarnello T, Zane L, Fernández DA,
Jones CD (2012) Ancient climate change, antifreeze, and the evolutionary diversification of
Antarctic fishes. Proc Natl Acad Sci 109:3434–3439
Neiva J, Serrão EA, Assis J, Pearson GA, Coyer JA, Olsen JL, Hoarau G, Valero M (2016) Climate
oscillations, range shifts and phylogeographic patterns of North Atlantic Fucaceae. In: Hu ZM,
Fraser C (eds) Seaweed phylogeography. Springer, Dordrecht, pp 279–308
Norkko A, Thrush SF, Cummings VJ, Funnell GA, Schwarz AM, Andrew NL, Hawes I (2004)
Ecological role of Phyllophora antarctica drift accumulations in coastal soft-sediment com-
munities of McMurdo Sound, Antarctica. Polar Biol 27:482–494
O’Cofaigh C, Davies BJ, Livingstone SJ, Smith JA, Johnson JS, Hocking EP et  al (2014)
Reconstruction of ice-sheet changes in the Antarctic Peninsula since the Last Glacial Maximum.
Quat Sci Rev 100:87–110
Ocaranza-Barrera P, González-Wevar CA, Guillemin ML, Rosenfeld S, Mansilla A (2019)
Molecular divergence between Iridaea cordata (Turner) Bory de Saint-Vincent from the
Antarctic Peninsula and the Magellan Region. J Appl Phycol 31:939–949
Oliveira MC, Pellizzari F, Medeiros AS, Yokoya NS (2020) Diversity of Antarctic seaweeds. In:
Gómez I, Huovinen P (eds) Antarctic seaweeds: diversity, adaptation and ecosystem services.
Springer, Cham, Switzerland, pp 23-39
Paillard D, Parrenin F (2004) The Antarctic ice sheet and the triggering of deglaciations. Earth
Planet Sci Lett 227:263–271
Pearse JS, Mooi R, Lockhart SJ, Brandt A (2009) Brooding and species diversity in the Southern
Ocean: selection for brooders or speciation within brooding clades? In: Krupnik I, Lang MA,
Miller SE (eds) Smithsonian at the poles: contributions to international polar year science.
Smithsonian Institution Scholarly Press, Washington, DC, pp 181–196
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa LH,
Colepicolo P (2017) Diversity and spatial distribution of seaweeds in the South Shetland
Islands, Antarctica: an updated database for environmental monitoring under climate change
scenarios. Polar Biol 40:1671–1685
Peters AF, Ramírez ME, Rülke A (2000) The phylogenetic position of the subantarctic marine mac-
roalga Desmarestia chordalis (Phaeophyceae) inferred from nuclear ribosomal ITS sequences.
Polar Biol 23:95–99
Pointing SB, Buedel B, Convey P, Gillman L, Koerner C, Leuzinger S, Vincent WF (2015)
Biogeography of photoautotrophs in the high polar biome. Front Plant Sci 6:692
Poulin E, González-Wevar C, Díaz A, Gérard K, Hüne M (2014) Divergence between Antarctic
and South American marine invertebrates: what molecular biology tells us about Scotia Arc
geodynamics and the intensification of the Antarctic Circumpolar Current. Glob Planet Chang
123:392–399
Provan J, Bennett KD (2008) Phylogeographic insights into cryptic glacial refugia. Trends Ecol
Evol 23:564–571
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8:e58223
Quartino ML, Saravia LA, Campana GL, Deregibus D, Matula CV, Boraso AL, Momo FR (2020)
Production and biomass of seaweeds in newly ice-free areas: implications for coastal processes
in a changing Antarctic environment. In: Gómez I, Huovinen P (eds) Antarctic seaweeds: diver-
sity, adaptation and ecosystem services. Springer, Cham, Switzerland, pp 155–168
126 M.-L. Guillemin et al.

Raupach MJ, Thatje S, Dambach J, Rehm P, Misof B, Leese F (2010) Genetic homogeneity and
circum-Antarctic distribution of two benthic shrimp species of the Southern Ocean, Chorismus
antarcticus and Nematocarcinus lanceopes. Mar Biol 157:1783–1797
Raymond JA, Fritsen CH (2001) Semi purification and ice recrystallization inhibition activity of ice
active substances associated with Antarctic photosynthetic organisms. Cryobiology 43:63–70
Reed DH, Frankham R (2003) Correlation between fitness and genetic diversity. Conserv Biol
17:230–237
Roberts J, McCave IN, McClymont EL, Kender S, Hillenbrand CD, Matano R, Hodell DA, Peck
VL (2017) Deglacial changes in flow and frontal structure through the Drake Passage. Earth
Planet Sci Lett 474:397–408
Saunders GW (2005) Applying DNA barcoding to red macroalgae: a preliminary appraisal holds
promise for future applications. Phil Trans R Soc Lond B 360:1879–1888
Saunders GW (2008) A DNA barcode examination of the red algal family Dumontiaceae in
Canadian waters reveals substantial cryptic species diversity. 1. The foliose Dilsea–Neodilsea
complex and Weeksia. Botany 86:773–789
Scher HD, Martin EE (2006) Timing and climatic consequences of the opening of Drake Passage.
Science 312:428–430
Scher HD, Whittaker JM, Williams SE, Latimer JC, Kordesch WE, Delaney ML (2015) Onset
of Antarctic Circumpolar Current 30 million years ago as Tasmanian Gateway aligned with
westerlies. Nature 523:580
Sijp WP, von der Heydt AS, Dijkstra HA, Flögel S, Douglas P, Bijl PK (2014) The role of ocean
gateways on cooling climate on long time scales. Glob Planet Chang 119:1–22
Silberfeld T, Leigh JW, Verbruggen H, Cruaud C, De Reviers B, Rousseau F (2010) A multi-locus
time-calibrated phylogeny of the brown algae (Heterokonta, Ochrophyta, Phaeophyceae):
investigating the evolutionary nature of the “brown algal crown radiation”. Mol Phylogenet
Evol 56:659–674
Simms AR, Milliken KT, Anderson JB, Wellner JS (2011) The marine record of deglaciation
of the South Shetland Islands, Antarctica since the Last Glacial Maximum. Quat Sci Rev
30:1583–1601
Soler-Membrives A, Linse K, Miller KJ, Arango CP (2017) Genetic signature of Last Glacial
Maximum regional refugia in a circum-Antarctic sea spider. R Soc Open Sci 4:170615
Spalding MD, Fox HE, Allen GR, Davidson N, Ferdaña ZA, Finlayson MAX et al (2007) Marine
ecoregions of the world: a bioregionalization of coastal and shelf areas. Bioscience 57:573–583
Spalding HL, Amado-Filho GM, Bahia RG, Ballantine DL, Fredericq S, Leichter JJ, Nelson
WA, Slattery M, Tsuda RT (2019) Macroalgae. In: Loya Y, Puglise KA, Bridge TCL (eds)
Mesophotic coral ecosystems. Springer, Cham, pp 507–536
Sromek L, Lasota R, Wolowicz M (2015) Impact of glaciations on genetic diversity of pelagic
mollusks: Antarctic Limacina antarctica and Arctic Limacina helicina. Mar Ecol Prog Ser
525:143–152
Stuart KM, Long DG (2011) Tracking large tabular icebergs using the SeaWinds Ku-band micro-
wave scatterometer. Deep Sea Res Pt II 58:1285–1300
Tajima F (1989) Statistical method for testing the neutral mutation hypothesis by DNA polymor-
phism. Genetics 123:585–595
Terauds A, Chown SL, Morgan F, Peat H, Watts DJ, Keys H, Convey P, Bergstrom DM (2012)
Conservation biogeography of the Antarctic. Divers Distrib 18:726–741
Thatje S, Hillenbrand CD, Larter R (2005) On the origin of Antarctic marine benthic community
structure. Trends Ecol Evol 20:534–540
Thatje S, Hillenbrand CD, Mackensen A, Larter R (2008) Life hung by a thread: endurance of
Antarctic fauna in glacial periods. Ecology 89:682–692
Thornhill DJ, Mahon AR, Norenburg JL, Halanych KM (2008) Open-ocean barriers to disper-
sal: a test case with the Antarctic Polar Front and the ribbon worm Parborlasia corrugatus
(Nemertea: Lineidae). Mol Ecol 17:5104–5117
6  Comparative Phylogeography of Antarctic Seaweeds: Genetic Consequences… 127

Valero M, Destombe C, Mauger S, Ribout C, Engel CR, Daguin-Thiebaut C, Tellier F (2011)


Using genetic tools for sustainable management of kelps: a literature review and the example
of Laminaria digitata. Cah Biol Mar 52:467
Van Oppen MJH, Olsen JL, Stam WT, van Denhoek C, Wiencke C (1993) Arctic-Antarctic dis-
junctions in the benthic seaweeds Acrosiphonia arcta (Chlorophyta) and Desmarestia viridis/
willii (Phaeophyta) are of recent origin. Mar Biol 115:381–386
Waters JM, Fraser CI, Hewitt GM (2013) Founder takes all: density-dependent processes structure
biodiversity. Trends Ecol Evol 28:78–85
White DA, Bennike O, Melles M, Berg S, Binnie SA (2018) Was South Georgia covered by an ice
cap during the Last Glacial Maximum? In: Siegert MJ, SSR J, White DA (eds) Exploration of
subsurface Antarctica: uncovering past changes and modern processes. The Geological Society
of London, pp 49–59
Wiencke C, Clayton MN (ed) (2002) Antarctic seaweeds. ARG Gantner Verlag, KG Ruggell, p 239
Wiencke C, Clayton MN, Gómez I, Iken K, Lüder UH, Amsler CD, Karsten U, Hanelt D, Bischof
K, Dunton K (2007) Life strategy, ecophysiology and ecology of seaweeds in polar waters. Rev
Environ Sci Biotechnol 6:95–126
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: De Broyer C, Koubbi
P, Griffiths HJ, Raymond B, Dudekem DAC et al (eds) Biogeographic atlas of the Southern
Ocean. Scientific Committee on Antarctic Research, Cambridge, pp 66–73
Wilson NG, Schrödl M, Halanych KM (2009) Ocean barriers and glaciation: evidence for explosive
radiation of mitochondrial lineages in the Antarctic sea slug Doris kerguelenensis (Mollusca,
Nudibranchia). Mol Ecol 18:965–984
Yang EC, Peters AF, Kawai H, Stern R, Hanyuda T, Bárbara I, Müller DG, Strittmatter M, van
Reine WFP, Küpper FC (2014) Ligulate Desmarestia (Desmarestiales, Phaeophyceae) revis-
ited: D. japonica sp. nov. and D. dudresnayi differ from D. ligulata. J Phycol 50:149–166
Zachos J, Pagani M, Sloan L, Thomas E, Billups K (2001) Trends, rhythms, and aberrations in
global climate 65 Ma to present. Science 292:686–693
Part III
Physiology, Productivity and
Environmental Reponses
Chapter 7
Underwater Light Environment
of Antarctic Seaweeds

Pirjo Huovinen and Iván Gómez

Abstract  Antarctic seaweeds are highly shade-adapted organisms, which can pho-
tosynthesize under very dim light. This remarkable characteristic allows them colo-
nizing over 30  m depths and surviving extended dark periods during the polar
winter. On the other hand, they are well equipped to cope with high light stress,
which points to a trade-off between shade adaptation and efficient UV stress toler-
ance. Optical properties of water determine both the underwater light climate for
photosynthesis and the risk of seaweeds for UV exposure in their habitats. Thus,
understanding the natural (spatial, temporal) and anthropogenic-driven changes in
spectral transparency of water and factors governing it is fundamental in evaluating
the state of seaweeds under current and future environmental scenarios. In the pres-
ent chapter the aspects related to the optical properties determining the underwater
habitat of Antarctic seaweeds are summarized, along with the potential changes in
water optics as a result of climate change, ozone depletion and other environmental
and emerging threats, and their interactions.

Keywords  Antarctic costal waters · Light climate · Ozone depletion ·


Photosynthesis · Pollution · UV radiation · Water column optics

7.1  Introduction

Aquatic organisms in the polar regions have adapted to survive wide seasonal
changes in the light field of their habitats (McMinn and Martin 2013). The polar
winter accompanied by ice cover leads to extended periods of dim light conditions
or even darkness (Berge et al. 2015), while in spring-summer, water column receives
increasing levels of solar radiation, which is furthermore enhanced after icebreak

P. Huovinen (*) · I. Gómez


Instituto de Ciencias Marinas y Limnológicas, Facutad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: pirjo.huovinen@uach.cl; igomezo@uach.cl

© Springer Nature Switzerland AG 2020 131


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_7
132 P. Huovinen and I. Gómez

and also during episodes of ozone depletion leading to higher risk of UV exposure
during austral spring (Frederick et al. 1989; Neale et al. 1998). In order to withstand
such widely changing conditions, Antarctic seaweeds are characterized by remark-
able photobiological adaptations to grow and photosynthesize at irradiances as low
as 10 μmol m−2 s−1, allowing them surviving extended dark periods during winter
and colonizing over 30 m depths. On the other hand, they present efficient mecha-
nisms to tolerate high UV stress (Gómez et al. 2009; Karsten et al. 2009; Gómez and
Huovinen 2015).
Underwater light climate is governed by the optical properties that define the
depth limits for photosynthesis and hence the vertical distribution of seaweeds as
well as the risk for exposure to detrimental UV levels in Antarctic coastal waters
(see Chap. 11 by Gómez and Huovinen). Thus, understanding the natural (temporal,
spatial) and anthropogenic-driven changes in water optics and factors controlling it
is essential, not only in evaluating aquatic primary production and UV risk in
aquatic habitats (Holm-Hansen et  al. 1993; Cullen and Neale 1994) but also for
predictions of the status of these ecosystems under current and future scenarios of
global climate change (Vincent and Belzile 2003). The knowledge on light attenua-
tion and the potential impact, e.g., of glacier-derived freshwater input, are also
needed in order to explain the spatial variation in primary production and carbon
fluxes with far-reaching implications for the benthic communities with a potential
to modify the biogeochemical gradients in the whole Antarctic coastal system (see
review by Sulzberger et al., 2019, and Chap. 8 by Quartino et al.). In the present
chapter, we summarize the aspects related to the optical characteristics determining
the underwater habitat of Antarctic seaweeds with implications for their photobiol-
ogy, along with the potential changes in water optics as a result of interactions with
climate change factors and other environmental and emerging threats.

7.2  Optics of Antarctic Coastal Waters

7.2.1  Light in Aquatic Environment

In natural waters, sunlight may be either absorbed or scattered. Principles of under-


water optics have been described and reviewed by Hargreaves (2003), Kirk (2011),
and Mobley (2015). In summary, the main components responsible for light absorp-
tion have been described: (1) water itself, (2) colored dissolved organic matter
(CDOM), (3) living photosynthetic organisms (phytoplankton), and (4) nonliving
organic and inorganic particulate material. Pure water absorbs and thus attenuates
light increasingly in red (>550 nm) and infrared (IR) bands, while absorption by
CDOM (mainly yellow dissolved humic substances with diverse intrinsic proper-
ties, e.g., molecular weight, origin, age; see Hessen and Tranvik, 1998) increases
strongly towards blue and UV bands. Phytoplankton attenuates photosynthetically
active radiation (PAR) through their light-absorbing photosynthetic pigments and
7  Underwater Light Environment of Antarctic Seaweeds 133

also by causing scattering of light as a result of their particulate form. Nonliving


particulate matter attenuates light strongly through scattering, whereas their light
absorption (increasing towards blue and UV wavelengths) is weaker (for details, see
Hargreaves 2003, Kirk 2011 and Mobley 2015).
Underwater light climate varies strongly in different types of water bodies
according to their inherent optical properties (IOPs), i.e., those describing aquatic
medium (such as absorption coefficient and volume scattering function) indepen-
dent of the light field (thus measurable in a water sample). Together with IOPs,
environmental conditions (e.g., radiance, solar angle, waves) partially influence
apparent optical properties (AOPs, such as reflectance and diffuse attenuation coef-
ficient), which can be used to describe water optics through in situ measurements
(see Kirk 2011; Mobley 2015). In an optically homogeneous water column, the
downward irradiance (Ed) diminishes in an exponential manner with depth (z); thus
diffuse attenuation coefficient for downward irradiance (Kd) can be calculated
according to Kirk (2011) as:

K d = −dln Ed ( z,λ ) / dz

Consequently, a depth where 1% or 10% of the subsurface irradiance is available
can be estimated from Kd as:

=zz1 4=
.6 / K d and zz10 2.3 / K d

The upper water column down to a depth where 1% of PAR remains is generally
defined as the euphotic zone, which can range from the upmost water layer to hun-
dred meters. For example, in saline-alkaline lakes and turbid estuaries, z1% can reach
less than half and one meter, respectively (Oduor and Schagerl 2007; Coljin et al.
1987), while in some oceanic areas (e.g., Sargasso Sea in the Atlantic and the coast
of Hawaii in the Pacific), light can penetrate more than hundred of meters (Tyler
1975; Bienfang et al. 1984). Penetration of UV wavelengths in the water column
can also range widely, from few centimeters (e.g., in small humic lakes; Scully and
Lean 1994; Huovinen et al. 2003) to dozens of meters in highly transparent oligo-
trophic waters (e.g., z1% around 40 m in Sargasso Sea; Smith and Baker 1979).

7.2.2  Light Climate in Antarctic Waters

An extensive study around the Antarctic Peninsula in mid-1980s by Mitchell and


Holm-Hansen (1991) revealed euphotic zone of approximately 75 m in the offshore
waters in Drake Passage, while in coastal areas PAR attenuated to 1% within the
upper 15 m. The clear open water areas were characterized by negligible sediment
load and terrigenous material, which can be explained by the scarcity of terrestrial
vegetation (the ice-free landscape is largely dominated by tundra-type vegetation
134 P. Huovinen and I. Gómez

like moss and lichen; Convey 2010) and hence a low contribution of allochtonous
organic matter from the catchment. In contrast, shallow near-shore waters around
the South Shetland Islands with glaciers showed high turbidity due to meltwater
runoff and related increase of organic matter decreasing transparency (Mitchell and
Holm-Hansen 1991). In fact, long-term measurements from Potter Cove (King
George Island, South Shetland Islands) indicate that increased light attenuation due
to strong sedimentation in areas impacted by glacier retreat is modifying the growth
conditions for seaweeds and their upper distribution limits (Deregibus et al. 2016;
see also Chap. 9 by Deregibus et al.). Also in Fildes Bay, another well-studied area
in King George Island, biomass of subtidal seaweeds displays a glacier-related gra-
dient (Valdivia et al. 2015). In Bransfield and Gerlache Straits, seasonal variation in
light penetration was observed: in January, light was strongly attenuated (z1% around
4.6 m based on Kd (488 nm)) in coastal areas due to massive phytoplankton blooms,
which disappeared by February–March leading to deeper (over 30 m) light penetra-
tion, reducing also the spatial (onshore-offshore) gradient in water transparency
(Mitchell and Holm-Hansen 1991). Strong spatial heterogeneity was also observed
by Figueroa (2002) with euphotic zone ranging 9–58 m in open waters of Gerlache
and Bransfield Straits. On the other hand, in Drake Passage where no blooms
occurred, the waters remained clear (z1% 46–76  m based on Kd (488  nm)) from
December to March (Mitchell and Holm-Hansen 1991). In coastal waters of Fildes
Bay, z1% PAR ranging 19–34 m has been reported (Huovinen et al. 2016) (Figs. 7.1
and 7.2). In Potter Cove, higher water transparency in November–December (z1%
PAR over 25 m) has been shown to decrease in January–February, light attenuating
already in the upper meters in areas influenced by meltwater (Klöser et al. 1993).
Here, in winter and early spring, light attenuation is mainly regulated by phyto-
plankton, while towards summer, particulate matter from increasing meltwater con-
tributes strongly (up to 47%) (Schloss and Ferreyra 2001).
During the open water season, relatively high UV transparency has been mea-
sured in Fildes Bay (Huovinen et al. 2016), although spatial variation between sites
and with depth is observed (Fig. 7.1). UV (z10%) penetration has been shown to range
here from approximately 3 to 6 m for UV-B305nm to 8 to 19 m for UV-A395nm (Huovinen
et al. 2016). Based on theoretical estimations (tropospheric ultraviolet and visible
(TUV) model) for summer solstice and mean Kd of the bay area, sunlight levels at
z10% could reach 0.13 W m−2 for UV-B (at 5 m), 4.4 W m−2 for UV-A (at 10 m), and
170 μmol m−2 s−1 for PAR (at 15 m) (Fig. 7.2). In Potter Cove, penetration depth
(z10%) of up to 8 m has been reported for UV-B radiation. Subsurface (10 cm) levels
of up to 1.5 and 26 W m−2 of UV-B (280–320 nm) and UV-A (320–400 nm), respec-
tively, were measured. Corresponding levels at 2 m depth were decreased to 0.26
and 8.6 W m−2. PAR levels at corresponding depths were 1178 and 515 μmol m−2 s−1
(Quartino et al. 2005; Zacher et al. 2007a). It should be noted that these authors
considered 320 nm as the limit between UV-B and UV-A bands, resulting in mark-
edly higher UV-B levels than when limiting the band to 315 nm (see Huovinen et al.
2006). Maximal penetration depths (z10%) for UV-B (305 nm) and UV-A (340 nm)
of 8  m and 11  m, respectively, have been reported for Bransfield and Gerlache
Straits and Palmer Station (Helbling et al. 1995; Figueroa 2002). In Bellinghausen
7  Underwater Light Environment of Antarctic Seaweeds 135

Fig. 7.1  Fildes Bay (King George Island, Maritime Antarctica (a). Variation of spectral irradiance
with depth in Fildes Bay based on measurement with the hyperspectral radiometer RAMSES-­
ACC2-­UV–vis  (Trios Optical Sensors, Oldenburg, Germany) (b).  Spatial (with depth and dis-
tance) variation of Kd at 395 nm (m−1) determined at 1 m intervals from measurements with the
radiometer PUV-2500 (Biospherical Instruments Inc., USA) in six areas of Fildes Bay and visual-
ized with Ocean Data View software (Schlitzer R., Ocean Data View, odv. Awi.de, 2015) (Adpated
from Huovinen et al. 2016) (c)

Sea, even higher penetration depths for UV-B (z10% 310 nm around 12 m) and UV-A
(340 nm around 17 m) have been measured. Here, UV-B radiation could be detected
down to 60–70 m (Smith et al. 1992).
Potential for DNA damage at depths ranging 3.5–16  m (based on z10% DNA
weighted irradiance) have been reported for Antarctic waters (Huot et  al. 2000;
Buma et  al. 2001). Under ozone hole the effective UV-B penetration depth was
found to increase by 7 m (Smith et al. 1992). For the estimations on net UV impact
on aquatic organisms, knowledge on spectral attenuation is needed as it varies with
depth (see Figs.  7.1b and 7.2), and shifts in underwater irradiance spectra have
implications for repair and other processes that depend on spectral light composi-
tion (Neale 2000; Williamson et al. 2001). For example, a higher ratio (mean around
1.0) between z10% 305 and 340  nm has been reported in Antarctic waters under
136 P. Huovinen and I. Gómez

Fig. 7.2  Estimation of irradiance levels at different water depths, based on irradiance levels during
summer solstice in Fildes Peninsula (62°S, 50°W) derived from Tropospheric Ultraviolet and
Visible (TUV 5.3; Madronich and Flocke (1999)) model (https://www2.acom.ucar.edu/modeling/
tropospheric-ultraviolet-and-visible-tuv-radiation-model) and on Kd values (mean of 305 and
313  nm for UV-B; mean of 320, 340, 380, and 395  nm for UVA) of the bay area (data from
Huovinen et al. 2016). Penetration depths z1% (solid arrows) and z10% (open arrows) are indicated.
Representative seaweeds for different depths are shown. (Photos by Ignacio Garrido except
Urospora by Iván Gómez)

ozone hole conditions than in other oceanic areas (without ozone depletion) (mean
0.54) (Tedetti and Sempéré 2006).
In polar regions, seasonal ice cover plays an important role in governing the
underwater light climate (Lesser et  al. 2004; Fritsen et  al. 2011; Taskjelle et  al.
2016). In McMurdo Sound (Ross Sea), 0.16% of incident irradiance (mainly blue-­
green band) was transmitted through a 2-m-thick ice (Schwarz et  al. 2003) and
0.05% (0.2–0.6 μmol m−2 s−1) reached bottom (23 m) (Robinson et al. 1995). Under
ice cover, water was highly transparent with Kd for PAR 0.09 m−1 (Schwarz et al.
2003). In scenarios of global climate change, the importance of sea-ice duration has
been highlighted and is regarded, together with light penetration and sedimentation,
a major driver structuring the shallow Antarctic benthos (Clark et al. 2017). Thinner
ice cover and its shorter duration results in enhanced exposure to solar radiation for
extended periods, which can bring consequences for photosynthesis (Runcie and
7  Underwater Light Environment of Antarctic Seaweeds 137

Riddle 2007)‚ zonation patterns of seaweeds (Campana et  al. 2009; Clark et  al.
2017) and increase UV damage (Fountoulakis et al. 2014). Timing of ice breakup
may be crucial for the establishment of canopy-forming seaweed communities in
spring (Johnston et al. 2007). When large brown algae are established, their cano-
pies can markedly reduce light levels reaching their understory species, and it can
be also altered by tidal fluctuations (Huovinen and Gómez 2011). In Antarctic
coastal waters, the presence of large brown algae growing at depths as shallow as
5 m can affect considerably the incident irradiance at deeper locations where sea-
weeds coexist with abundant populations of understory species, especially red algae
(Klöser et al. 1996; Gómez et al. 2019; see also Chap. 11 by Gómez and Huovinen).

7.3  A
 daptations of Antarctic Seaweeds to Extreme
Light Conditions

7.3.1  Photosynthetic Shade Adaptation of Antarctic Seaweeds

The strong seasonality and turbidity of coastal waters (Zacher et al. 2009) imply
severe constraints for photosynthesis of Antarctic seaweeds (see Wiencke et  al.
2009). They can overall be characterized by very low requirements of light and
constitutively high efficiency of photosynthesis, allowing them colonizing deep
habitats with low light availability and coping with periods of darkness (reviewed
by Gómez et al. 2009). For example, some Antarctic crustose red algae (corallines)
in the Ross Sea live permanently under ice cover and remain in darkness for several
months in winter (Schwarz et al. 2005). In fact, shade adaptation is considered a
metabolic prerequisite that allows survival under wide gradient of light (Weykam
et al. 1996; Huovinen and Gómez 2013; Gómez and Huovinen 2015).
Photosynthetic adaptations to cope with low light availability allow Antarctic
seaweeds supplying their light requirements over a broad depth distribution (Gómez
et al. 1997). The growth and photosynthesis especially of the endemic brown sea-
weeds has been shown not to be limited even at deep locations close to 30–40 m
(Drew 1977; Weykam et  al. 1996; Schwarz et  al. 2003). Although some species
from King George Islands collected at shallow waters displayed higher saturated net
photosynthesis (Pmax) than species from deeper sites, their photosynthetic efficiency
(α) and light demands for photosynthesis (Ek) did not vary with depth (Weykam
et al. 1996). These photosynthetic characteristics have been confirmed more recently
with chlorophyll fluorescence techniques (Huovinen and Gómez 2013). Also photo-
synthetic efficiency (αETR) has been found to be high along the depth gradient
(0–30 m) (Gómez and Huovinen 2015) in spite of relatively high water transparency
(PAR at 20  m depth around 50  μmol  m−2  s−1) (Huovinen et  al. 2016; Fig.  7.2).
Despite their wide vertical distribution (from 5 m downwards), lower distribution
limit (close to 30 m) of Antarctic seaweeds such as Desmarestia mensiezii, D. anceps,
Palmaria decipiens, and Gigartina skottsbergii coincides with the depth of
138 P. Huovinen and I. Gómez

compensation irradiance level (Klöser et al. 1996; Gómez et al. 1997; Deregibus
et al. 2016). Also, although it can be found at depths close to 5–10 m, Himantothallus
grandifolious starts to dominate only at depth below 30 m when substrate and com-
petitive balance with other large Desmarestiales, e.g., D.  anceps, are favorable
(Zielinski 1990; Klöser et  al. 1996). Thus, the vertical distribution patterns of
Antarctic seaweeds cannot be explained by light limitation alone, but rather in com-
bination with other biotic (e.g., herbivory, competition) and abiotic (e.g., substrate
characteristics, ice-induced perturbations, water movement) factors (Klöser et  al.
1996; Iken et al. 1998; Amsler et al. 2011; see Chap. 13 by Valdivia and Chap. 17
by Amsler et al.).

7.3.2  Tolerance of Antarctic Seaweeds to High PAR and UV

Antarctic coastal waters are often characterized by high water transparency with
light (PAR) penetration (z1%) ranging 19–34 m (Huovinen et al. 2016; Figs. 7.1 and
7.2). In coastal waters of King George Island, PAR has been found to penetrate
down to 40 m depth (Klöser et al. 1993), and levels around 50 μmol m−2 s−1 have
been measured at 30 m (Gómez et al. 1997). This contrasts with the light conditions
in some Arctic near-shore waters, e.g., in the Beaufort Sea with muddy bottoms and
estuarine characteristics where PAR penetration is markedly lower (z1% 3–11 m; in
offshore waters >4.6 m) (Dunton et al. 2009). After Antarctic icebreak in spring-­
summer, seaweeds experience sudden increase of underwater light; however, water
optics may present strong variation due to turbidity, e.g., from glacier melting or
freshwater runoff (Klöser et  al. 1993; Deregibus et  al. 2016; see Chap. 9 by
Deregibus et al.). Under these conditions, seaweeds exhibit a suite of physiological
mechanisms to cope with high levels of solar radiation. One of them is dynamic
photoinhibition, downregulation of photosynthesis under high solar radiation, dis-
sipating excess absorbed energy as heat in photosystem II (PSII) (Adams III et al.
2006). This protective mechanism has been reported in Antarctic seaweeds exposed
to natural solar irradiation, showing decreased photosynthetic activity around solar
noon followed by recovery towards evening (Hanelt et al. 1994). Thus, the capacity
to withstand high solar radiation allows these shade-adapted organisms thriving also
at environments where light levels exceed their requirements for saturation of pho-
tosynthesis, such as in intertidal and shallow waters (Huovinen and Gómez 2013;
Gómez et al. 2019).
Together with high PAR levels, seaweeds can also be exposed to harmful levels
of UV radiation in their habitats (Huovinen and Gómez 2013, Huovinen et al. 2016;
Figs.  7.1 and 7.2). Deleterious effects of UV-B radiation on aquatic organisms,
including seaweeds, are widely recognized (reviewed by Holm-Hansen et al. 1993;
Vincent and Neale 2000; Karsten et al. 2009; Burritt and Lamare 2016). UV-B radi-
ation can directly affect cellular components (e.g., nucleus, chloroplast), their ultra-
structure and processes, as well as target important biomolecules, such as DNA
(leading, e.g., to formation of cyclobutane dimers (CPDs) and 6–4 photoproducts
7  Underwater Light Environment of Antarctic Seaweeds 139

that interfere with replication) (Mitchell and Karentz 1993), proteins (e.g., D1 pro-
tein in PSII), and photosynthetic pigments (Gerber and Häder 1992). It can also
interfere with uptake of nutrients (Döhler et al. 1991) and metabolism of fatty and
amino acids (Goes et  al. 1994, 1995). It may cause oxidative stress by inducing
production of reactive oxygen species (ROS, such as singlet-oxygen, hydroxyl, and
superoxide radicals) that are damaging to biomolecules and can cause, e.g., peroxi-
dation of lipids (Bischof and Rautenberger 2012). ROS may also be formed in
aquatic environment when UV interacts with DOM (Kieber et al. 2003) (see also
Sect. 7.4.1). Disturbance or damage to important cellular components and biomol-
ecules can impair biochemical and physiological processes, such as photosynthesis
(through effect on pigments, enzymes, photosynthetic apparatus, etc.), growth, and
reproduction (reviewed in Bischof et al. 2006).
The magnitude of final harmful impact depends on the balance between pro-
duced damage and the efficiency of protective, e.g., UV-shielding compounds such
as mycosporine-like amino acids (MAAs) in red algae and phlorotannins in brown
algae (see Chap. 18 by Gómez and Huovinen) and repair (e.g., light-induced repair
of DNA damage involving photolyase enzyme) mechanisms to mitigate it (Karentz
et al. 1991; Mitchell and Karentz 1993; Vincent and Roy 1993; Nuñez-Pons et al.
2018). Antioxidants (e.g., carotenoids, phlorotannins, enzymes such as superoxide
dismutase and glutathione peroxidase) furthermore serve as defense mechanism
through ROS scavenging (Bischof and Rautenberger 2012). Also, downregulation
of PSII (dynamic photoinhibition) has been proposed as a protective mechanism,
not only against high PAR, but also against UV radiation, contributing to the physi-
ological tolerance of seaweeds (Bischof et al. 2006).
Early life stages of seaweeds with a thinner wall are more vulnerable to UV dam-
age than multicellular adult stages. Marked DNA damage and physiological stress
has been reported in propagules of Antarctic seaweeds upon UV exposure (Roleda
et  al. 2006; Zacher et  al. 2007b), although they also show capacity for recovery
(Navarro et al. 2019) and UV-screening compounds (Roleda et al. 2006; see also
Chap. 10 by Navarro et al.). Furthermore, adult thalli of intertidal species display
high tolerance to light stress as they have to cope with high levels of solar radiation,
especially during low tidal levels (Cruces et al. 2013). Here, dynamic photoinhibi-
tion seems to play a key role as a protective mechanism (Hanelt et  al. 1994).
Interestingly, high photosynthetic tolerance has also been reported for subtidal spe-
cies that are not exposed to high UV levels in their habitat (deeper than 20  m)
(Huovinen and Gómez 2013). Especially large endemic brown algae have shown
marked UV stress tolerance over their broad vertical distribution (5–30 m), which
has been related to efficient morph-functional mechanisms and constitutively high
levels of phlorotannins (Gómez and Huovinen 2015; Flores-Molina et  al. 2016;
Gómez et al. 2019; see also Chaps. 11 and 18 by Gómez and Huovinen). Overall,
UV sensitivity of seaweeds at their different life stages is considered one of the
ecologically important factors defining their vertical zonation and distribution limits
(Bischof et al. 1998; Wiencke et al. 2000, 2006; see Chap. 10 by Navarro et al.). In
contrast to short-term laboratory experiments, organisms in their natural habitats are
influenced by complex interactions of multiples environmental factors, which may
140 P. Huovinen and I. Gómez

mitigate or potentiate the adverse impact of UV radiation (see Sect. 7.4). For exam-
ple, a global field study examining UV impact on marine benthic community level
found that any effect disappeared after few months during succession (Wahl et al.
2004). However, recent reports based on meta-analyses point to overall negative UV
impact at all trophic levels (summarized by Williamson et al. 2019).

7.4  C
 onsequences for Light Field Under Current
and Future Threats

7.4.1  Ozone Depletion

The deleterious effects of solar radiation are mainly associated with the UV wave-
bands, i.e., UV-C (100–280 nm), UV-B (280–315 nm), and UV-A (315–400 nm)
radiation, with greater effectiveness for biological damage occurring towards
shorter wavelengths (Setlow 1974; Cullen et al. 1992). The most damaging UV-C
waveband is absorbed by oxygen and ozone in the atmosphere thus not reaching the
earth’s surface. On the other hand, UV-A waveband is not absorbed by atmospheric
ozone, while UV-B waveband is partially absorbed and its levels are affected by
changes in the thickness of the ozone layer (Frederick et al. 1989; Madronich et al.
1998). Based on radiation amplification factors (RAF), it has been estimated that
1% ozone depletion leads to approximately 2–3% increase in DNA-damaging UV-B
dose, i.e., weighted with DNA action spectra (Madronich 1994). More than four
decades ago, Molina and Rowland (1974) discovered the destructing effect of chlo-
rofluorocarbons (CFCs) for ozone layer. A decade later, Farman et  al. (1985)
reported for the first time the “ozone hole” over the Antarctic. Although the biologi-
cal risk from UV radiation also in aquatic environments had already been recog-
nized (Smith and Baker 1979; Calkins and Thordardottir 1980), stratospheric ozone
depletion and consequently enhanced levels solar UV-B radiation started to gain
importance as one of the major anthropogenic-driven threats for Antarctic biota
thereafter (Holm-Hansen et  al. 1993; Weiler and Penhale 1994). After several
decades of alarming trends, recent reports indicate that the ozone layer over the
Antarctic is showing recovery (Solomon et  al. 2016, 2017; summarized by Bais
et al. 2018, 2019). Since 2000, the observed recovery seems to be in agreement with
a decreasing amount of ozone-depleting substances. Although global trend over all
latitudes is masked due to strong variability (Chipperfield et al. 2017), recovering
trend has been reported for some areas, including Antarctica (Kuttippurath and Nair
2017). However, the complex interactions of stratospheric ozone with climate
change effects in the atmosphere and ocean lead to uncertainties in estimating the
timescale for recovery, and changes (either increase or decrease) in UV levels are
estimated to vary in different regions (summarized by Bais et  al. 2018, 2019).
Together with greenhouse gases, ozone depletion has led to latitudinal shifts of cli-
mate (by moving the Southern Annular Mode (SAM) towards a more positive
7  Underwater Light Environment of Antarctic Seaweeds 141

Fig. 7.3  Overview of the major components related to underwater optics and processes occurring
under changing environmental context in Antarctic coastal waters, with implications for photobiol-
ogy of Antarctic seaweeds. See text for details and references

phase), which results in the strengthing of the westerly winds over Antarctic, with
impact on aquatic ecosystems, e.g., their mixed layer depth (Bais et  al. 2019;
Williamson et al. 2019). Because climate change also modifies surface reflectivity
(through changes in snow and ice cover), clouds, and aerosols, it together with
ozone layer plays key role in defining the UV exposure of Antarctic ecosystems in
future (Bais et al. 2019).
Under episodes of ozone depletion during austral spring, band ratios can be mod-
ified resulting in enhanced UV-B exposure in proportion to UV-A/PAR, i.e., affect-
ing the balance between UV damage and photorepair (Mitchell and Karentz 1993),
thus leading to a higher risk for Antarctic ecosystems. In addition to direct adverse
effects on organisms (see Sect. 7.3.2 in the present chap), enhanced UV levels imply
elevated potential for photochemical reactions between UV radiation and CDOM,
leading to photodegradation of CDOM and hence increasing water transparency
(Morris and Hargreaves 1997), bioavailability of complexed contaminants (e.g.,
heavy metals), and nutrients (reviewed by Zepp 2003), formation of biologically
damaging products (e.g., singlet oxygen) (reviewed by Kieber et  al. 2003), and
stimulating bacterial growth (Kieber et al. 1989). Overall, in recent years increasing
attention has been paid to the role of UV radiation in a variety of biogeochemical
processes (e.g., carbon cycling, enhanced photodegradation) and how global change
142 P. Huovinen and I. Gómez

is altering them (summarized by Sulzberger et  al. 2019; Williamson et  al. 2019)
(Fig. 7.3).

7.4.2  Regional Warming

Among the major consequences of warming are the retreat of glaciers and reduction
of ice cover (Vaughan and Doake 1996; Cook et  al. 2005; see also Chap. 1 by
Gómez and Huovinen). Ice sheets play an important role in global carbon cycle
(Wadham et  al. 2019), and in the Antarctic where marine biota was evolved in
response to a massive glaciation since 30  Ma (Crame 1992; Clayton 1994), the
enhanced melting will have far-reaching consequences for the biogeochemical pro-
cesses in vast coastal areas (Constable et al. 2014; Sulzberger et al. 2019). Under the
strongest forcing scenario of IPCC, climate change-driven expansion of ice-free
areas is estimated to be around 25% by the end of the century in the Antarctic,
mainly in the WAP (Amesbury et al. 2017; Lee et al. 2017), where summer snow
melting has currently reached its highest intensity over last 1000 years (Abram et al.
2013). New ice-free areas provide habitats that can be colonized by seaweeds
(Quartino et al. 2013; see also Chap. 8 by Quartino et al.). On the other hand, these
areas are often characterized by high turbidity (low light penetration) as a result of
sediment runoff (see Fig. 7.3), which can modify the zonation patterns of some spe-
cies towards shallower waters (Deregibus et al. 2016) and cause reduction of sea-
weed productivity (Jerosch et  al. 2019; see Chap. 9 by Deregibus et  al.). In the
Arctic, the growth of kelp Laminaria solidungula has been found to depend directly
on the water transparency, which is strongly governed by resuspension of sediments
especially during increased frequency of storm events (Dunton et  al. 2009).
Increased turbidity can furthermore interfere with disinfection of pathogens and
other microbiomes by UV radiation in surface waters with consequences hitherto
not well understood (Williamson et  al. 2017, 2019; see Chap. 14 by Gaitan and
Schmid).
In polar regions, high permafrost temperatures have been registered, with poten-
tial impact on global climate through emissions of CO2 and release of methane. In
the Antarctic, temperature increased by 0.37 °C in zones with continuos permafrost
during 2007–2016 (IPCC 2019). Increased input of organic matter from catchment
as a result of glacier melting can also lead to enhanced nutrient (e.g., nitrates, iron)
transfer to coastal areas (Hodson et  al. 2017; Ducklow et  al. 2018), potentially
increasing phytoplankton biomass and thus decreasing light penetration. Increased
nutrient availability can also modify the responses of seaweeds to environmental
stress, e.g., mitigating adverse effects of UV radiation and metals as observed in
sub-Antarctic kelps (Huovinen et al. 2010). Elevated temperature is also known to
improve the efficiency of repair processes, leading to higher UV tolerance of
Antarctic seaweeds (Rautenberger et  al. 2015). Increase of Fe from runoff may
cause oxidative stress in Antarctic seaweeds (González et al. 2017).
7  Underwater Light Environment of Antarctic Seaweeds 143

7.4.3  Feedback with Other Emergent Threats

Growing evidence on the presence of anthropogenic pollution in the Antarctic,


especially in the WAP region, is changing our vision of a pristine environment
(reviewed by Bargagli 2008). In spite of natural barriers (Antarctic Circumpolar
Current and atmospheric circulation) and remoteness, these ecosystems receive pol-
lution, e.g., persistent organic pollutants (POPs), from other regions through long-­
range atmospheric transport (LRAT), in addition to local sources (Bengtson Nash
2011; Vecchiato et al. 2015; Khairy et al. 2016). In fact, growing human activities
(e.g., waste incineration, sewage effluents, fuel combustion) within this region are
leaving their marks in the environment (reviewed by Bargagli 2008). Emerging con-
taminants, such as microplastics, have also already been reported in the Antarctic
environment (Waller et al. 2017; Lacerda et al. 2019). In this context, more acces-
sible areas such as the Antarctic Peninsula are increasingly threatened by human
impact (e.g., from scientific stations and visitors), which is evidenced by chemical
contamination including heavy metals (de Moreno et al. 1997; Farías et al. 2002;
Amaro et al. 2015; Padeiro et al. 2016; Chu et al. 2019) and polycyclic aromatic
hydrocarbons (PAHs) (Na et al. 2011; Préndez et al. 2011) in the sediments, water,
and snow in these ecosystems. The fate of environmental contaminants can be dras-
tically modified under changing climate and ozone depletion scenarios implying
multiple and complex consequences (Wrona et  al. 2006; Schiedek et  al. 2007;
Noyes et al. 2009; Grannas et al. 2013; Galic et al. 2017; Sulzberger et al. 2019)
(Fig. 7.3). For instance, increased melting of cryospheric environments can release
chemical contaminants that have been stored in ice and snow for prolonged periods
(Wania and Westgate 2008; Grannas et al. 2013; Hauptmann et al. 2017). The risk
of enhanced levels and availability of contaminants like POPs and mercury due to
climate change is receiving increased attention (IPCC 2019; Sulzberger et al. 2019).
Certain light-absorbing impurities (LAIs) deposited on snow and ice surfaces can
induce complex feedback processes that further accelerate melting (Lutz et al. 2014;
Tedesco et al. 2016; Huovinen et al. 2018), potentially leading to unpredicted pollu-
tion levels in the surrounding environments. The fate of LAIs reaching aquatic envi-
ronment and their potential for underwater light attenuation is still largely unknown.
On the other hand, interaction of certain environmental contaminants (such as
PAHs) with UV radiation can lead directly or via ROS to their photomodification
(photooxidation, photolysis) resulting in photodegradation or formation of photo-
products with different characteristics and effects (including lower toxicity) as the
original compound. They can also achieve higher toxicity (phototoxicity) through
photosensitization reactions when they interact with UV radiation (reviewed by
Björn and Huovinen 2015; Sulzberger et al. 2019) (Fig. 7.3). Therefore, even when
PAH levels in the environment are low, such as those reported for King George
Island (Na et  al. 2011; Préndez et  al. 2011), their potential for UV-induced or
UV-enhanced toxicity under ozone depletion implies an unpredicted risk. UV radia-
tion can also contribute to generation of microplastics (summarized by Andrady and
Pandey 2019) and interact with other emerging contaminants (e.g., pharmaceutical
144 P. Huovinen and I. Gómez

and personal care products such as sunscreens) (see Björn and Huovinen 2015; Bais
et al. 2019; Williamson et al. 2019).
Since 1980s, the oceans as carbon sink have removed 20–30% of CO2 from
anthropogenic emissions, leading to increased ocean acidification (summarized in
Gattuso and Hansson 2011). Such conditions are corrosive to marine organisms,
causing a decline (3.9% during 1998–2014 in the Southern Ocean) in calcification
rates in skeleton- and shell-forming species (IPCC 2019). Ocean systems at high
latitudes are especially vulnerable to experience changes of pH due to their lower
buffering capacity (McNeil and Matear 2008; Jewett and Romanou 2017). Decreased
pH can increase UV transparency of water through photodegradation of CDOM
(Schindler et al. 1996; Yan et al. 1996) (Fig. 7.3). In systems affected by freshening
due to local glacier melt, acidification can be exacerbated increasing calcium car-
bonate corrosivity (Evans et al. 2014). To date, studies carried out in calcareous and
non-calcareous seaweeds indicate that responses to acidification depend on the
taxonomic status, biogeographic location, prevailing metabolism of species (e.g.,
algae with carbon concentrating mechanisms versus those that rely on CO2 trans-
port), the type of experimental approach and time of responses, etc. (Hurd et  al.
2009; Roleda and Hurd 2012; Hofmann and Bischof 2014). Overall, interaction of
acidification, climate change, and UV radiation has been found to impact negatively
some calcifying seaweed species (e.g., by reducing structural UV protection)
(Russell et  al. 2011) or affecting photosynthesis (Gao and Häder 2017). In the
Antarctic seaweeds Desmarestia anceps and D. menziesii, changes in protein and
lipid contents were observed when the algae were exposed to combined treatments
of pH and temperatures (Schram et  al. 2017); however, these conditions did not
affect markedly other physiological parameters such as photosynthesis and phloro-
tannin content (Schoenrock et al. 2015).

7.5  Concluding Remarks

Global change together with increasing anthropogenic impact lead to complex sce-
narios for Antarctic coastal waters: enhanced turbidity from glacier and ice melt
impedes the light penetration, but earlier ice breakup in late winter-spring increases
underwater light levels during extended open-water period, whereas UV radiation
and acidification can lead to degradation of organic matter and hence increase light
penetration. Under favorable light conditions in the water column, increased nutri-
ent load from terrestrial and glacier runoff can stimulate phytoplankton growth
causing blooms with negative impact on water clarity and light availability for ben-
thic productivity. Therefore, complex implications for seaweeds are to be expected,
ranging from beneficial to harmful, depending on local processes as well as the
structure and function of the seaweed communities. How other feedbacks related,
e.g., with transient and persistent contaminants, ocean acidification and local fresh-
ening will impact the underwater light climate in the Antarctic and their conse-
quences for the biota are difficult to predict. Overall, the scarcity of long-term data
7  Underwater Light Environment of Antarctic Seaweeds 145

that connects the changes in climate and habitat with ecosystem responses has been
identified among the major gaps when predicting the climate change impact on
Antarctic cryosphere and its associated marine realm (Fountain et  al. 2012).
Similarly, susceptibility of seaweeds to the predicted changes in the Antarctic habi-
tat and their photobiological adaptations need to be addressed from molecular to
systemic scales. In fact, the novel molecular approaches are likely to improve our
understanding of the mechanisms that cold-adapted organisms display to cope with
the projected environmental variability (Lyon and Mock 2014).

Acknowledgments  The funding provided by CONICYT (through Projects Anillo ART1101,


FONDECYT 1161129 and FONDAP 15150003) and by Instituto Antártico Chileno (INACH;
Grant T-20-09) to carry out our research in the Antarctic is greatly acknowledged. We are also
grateful to the members of our research group in Universidad Austral de Chile and the staff of the
Instituto Antártico Chileno for their invaluable cooperation.

References

Abram NJ, Mulvaney R, Wolff EW, Triest J, Kipfstuhl S et al (2013) Acceleration of snow melt
in an Antarctic Peninsula ice core during the twentieth century. Nat Geosci 6:404–411. https://
doi.org/10.1038/NGEO1787
Adams WW III, Zarter CR, Mueh KE, Amiard V, Demmig-Adams B (2006) Energy dissipation
and photoinhibition: a continuum of photoprotection. In: Demmig-Adams B, Adams WW III,
Mattoo AK (eds) Photoprotection, photoinhibition, gene regulation, and environment. Springer,
The Netherlands, pp 49–64
Amaro E, Padeiro A, Mão de Ferro A, Mota AM, Leppe M et al (2015) Assessing trace element
contamination in Fildes Peninsula (King George Island) and Ardley Island. Ant Mar Poll Bull
15:523–527
Amesbury MJ, Roland TP, Royles J, Hodgson DA, Convey P et al (2017) Widespread biological
response to rapid warming on the Antarctic Peninsula. Curr Biol 27:1616–1622. https://doi.
org/10.1016/j.cub.2017.04.034
Amsler CD, Iken K, McClintock JB, Beaker BJ (2011) Defenses of polar macroalgae against her-
bivores and biofoulers. In: Wiencke C (ed) Biology of polar benthic algae. Walter de Gruyter
GmBH & Co, KG, Berlin, New York, pp 101–120
Andrady AL, Pandey KK, Heikkilä (2019) Interactive effects of solar UV radiation and climate
change on material damage. Photochem Photobiol Sci 18:804–825. https://doi.org/10.1039/
c8pp90065e
Bais AF, Lucas RM, Bornman JF, Williamson CE, Sulzberger B et al (2018) Environmental effects
of ozone depletion, UV radiation and interactions with climate change: UNEP environmen-
tal effects assessment panel, update 2017. Photochem Photobiol Sci 17:127–179. https://doi.
org/10.1039/c7pp90043k
Bais AF, Bernhard G, McKenzie RL, Aucamp PJ, Young AJ et al (2019) Ozone-climate interac-
tions and effects on solar ultraviolet radiation. Photochem Photobiol Sci 18:602–640. https://
doi.org/10.1039/c8pp90059k
Bargagli R (2008) Environmental contamination in Antarctic ecosystems. Sci Total Environ
400:212–226
Bengtson Nash S (2011) Persistent organic pollutants in Antarctica: current and future research
priorities. J Environ Monit 13:497–504. https://doi.org/10.1039/c0em00230e
146 P. Huovinen and I. Gómez

Berge J, Renaud P, Darnis G, Cottier F, Last K et al (2015) In the dark: a review of ecosystem
processes during the Arctic polar night. Prog Oceanogr 139:258–271. https://doi.org/10.1016/j.
pocean.2015.08.005
Bienfang PK, Szyper JOP, Okamoto MY, Noda EK (1984) Temporal and spatial variability of
phytoplankton in a subtropical ecosystem. Limnol Oceanogr 29:527–539
Bischof K, Rautenberger R (2012) Seaweed responses to environmental stress: reactive oxygen
and antioxidative strategies. In: Wiencke C, Bischof K (eds) Seaweed biology: novel insights
into ecophysiology, ecology and utilization. Springer, Berlin, Heidelberg, pp 109–132
Bischof K, Hanelt D, Wiencke C (1998) UV-radiation can affect depth-zonation of Antarctic mac-
roalgae. Mar Biol 131:597–605
Bischof K, Gómez I, Molis M, Hanelt U, Karsten D et al (2006) Ultraviolet radiation shapes sea-
weed communities. Rev Environ Sci Biotechnol 5:141–166
Björn LO, Huovinen P (2015) Phototoxicity. In: Björn LO (ed) Photobiology: the science of light
and life, 3rd edn. Springer, New York, pp 335–345
Buma AGJ, de Boer MK, Boelen P (2001) Depth distributions of DNA damage in Antarctic marine
phyto-and bacterioplankton exposed to summertime ultraviolet radiation. J Phycol 37:200–208
Burritt DJ, Lamare MD (2016) The cellular responses of marine algae and invertebrates to ultra-
violet radiation, alone and in combination with other common abiotic stressors. In: Solan M,
Whiteley NM (eds) Stressors in the marine environment. Oxford University Press, Oxford,
pp 117–134
Calkins J, Thordardottir T (1980) The ecological significance of solar UV radiation on aquatic
organisms. Nature 283:563–566
Campana GL, Zacher K, Fricke A, Molis M, Wulff A et al (2009) Drivers of colonization and suc-
cession in polar benthic macro-and microalgal communities. Bot Mar 52:655–667
Chipperfield MP, Bekki S, Dhomse S, Harris NRP, Hassler B et al (2017) Detecting recovery of the
stratospheric ozone layer. Nature 549:211–218. https://doi.org/10.1038/nature23681
Chu WL, Dang NL, Kok YY, Yap KSI, Phang SM, Convey P (2019) Heavy metal pollution in
Antarctica and its potential impact on algae. Pol Sci 20:75–83. https://doi.org/10.1016/j.
polar.2018.10.004
Clark GF, Stark JS, Palmer AS, Riddle MJ, Johnston EL (2017) The roles of sea-ice, light and sedi-
mentation in structuring shallow Antarctic benthic communities. PLoS One 12(1):e0168391.
https://doi.org/10.1371/journal.pone.0168391
Clayton MN (1994) Evolution of the Antarctic benthic algal flora. J Phycol 30:897–904
Coljin F, Admiraal W, Baretta JW, Ruardij P (1987) Primary production in a turbid estuary, the
Ems-Dollard: field and model studies. Cont Shelf Res 7:1405–1409
Constable AJ, Melbourne-Thomas J, Corney SP, Arrigo KR, Barbraud C et  al (2014) Climate
change and Southern Ocean ecosystems. I: how changes in physical habitats directly affect
marine biota. Glob Chang Biol 20:3004–3025. https://doi.org/10.1111/gcb.12623
Convey P (2010) Terrestrial biodiversity in Antarctica–recent advances and future challenges.
Polar Sci 4(2):135–147
Cook AJ, Fox AJ, Vaughan DG, Ferrigno JG (2005) Retreating glacier fronts on the Antarctic
Peninsula over the past half-century. Science 308:541–544
Crame JA (1992) Evolutionary history of the polar regions. Hist Biol 6:37–60
Cruces E, Huovinen P, Gómez I (2013) Interactive effects of UV radiation and enhanced tempera-
ture on photosynthesis, phlorotannin induction and antioxidant activities of two sub-Antarctic
brown algae. Mar Biol 160:1–13
Cullen JJ, Neale PJ (1994) Ultraviolet radiation, ozone depletion, and marine photosynthesis.
Photosynth Res 39:303–320
Cullen JJ, Neale PJ, Lesser MP (1992) Biological weighting function for the inhibition of phyto-
plankton photosynthesis by ultraviolet radiation. Science 258:646–650
de Moreno JEA, Gerpe MS, Moreno VJ, Vodopivez C (1997) Heavy metals in Antarctic organisms.
Polar Biol 17:131–140
7  Underwater Light Environment of Antarctic Seaweeds 147

Deregibus D, Quartino ML, Campana GL, Momo FR, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in Potter
Cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166
Döhler G, Hegmeier E, Grigoleit E, Krause KD (1991) Impact of solar UV radiation on uptake of
15
N-ammonia and 15N-nitrate by marine diatoms and natural phytoplankton. Biochem Physiol
Pflanzen 187:293–303
Drew EA (1977) Physiology of photosynthesis and respiration in some Antarctic marine algae.
Brit Ant Surv Bull 46:59–76
Ducklow HW, Stukel MR, Eveleth R, Doney SC, Jickells T et al (2018) Spring–summer net com-
munity production, new production, particle export and related water column biogeochemical
processes in the marginal sea ice zone of the Western Antarctic Peninsula 2012–2014. Phil
Trans R Soc A 376:20170163
Dunton KH, Schonberg SV, Funk DW (2009) Interannual and spatial variability in light attenu-
ation: evidence from three decades of growth in the arctic kelp, Laminaria solidungula. In:
Proceedings of Smithsonian at the Poles Symposium, Smithsonian Institution, Washington,
DC, 3–4 May 2007. Smithsonian Institute Scholarly Press, Washington, DC, pp 271–284
Evans W, Mathis JT, Cross JN (2014) Calcium carbonate corrosivity in an Alaskan inland sea.
Biogeosciences 11:365–379
Farías S, Pérez Arisnabarreta S, Vodopivez C, Smichowski P (2002) Levels of essential and poten-
tially toxic metals in Antarctic macro algae. Spectrochim Acta B 57:2133–2140
Farman JC, Gardiner BG, Shanklin JD (1985) Large losses of total ozone in Antarctica reveal
seasonal ClOx/NOx interaction. Nature 315:207–210
Figueroa FL (2002) Bio-optical characteristics of Gerlache and Bransfield Strait waters during an
Antarctic summer cruise. Deep-Sea Res II Top Stud Oceanogr 49:675–691
Flores-Molina MR, Rautenberger R, Muñoz P, Huovinen P, Gómez I (2016) Stress tolerance of the
endemic Antarctic brown alga Desmarestia anceps to UV radiation and temperature is medi-
ated by high concentrations of phlorotannins. Photochem Photobiol 92:455–466
Fountain AG, Campbell JL, Schuur EAG, Stammerjohn SE, Williams MW, Ducklow HW (2012)
The disappearing cryosphere: impacts and ecosystem responses to rapid cryosphere loss.
Bioscience 62:405–415
Fountoulakis I, Bais AF, Tourpali K, Fragkos K, Misios S (2014) Projected changes in solar UV
radiation in the Arctic and sub-Arctic Oceans: effects from changes in reflectivity, ice transmit-
tance, clouds, and ozone. J Geophys Res Atmosp 119:8073–8090
Frederick JE, Snell HE, Haywood EK (1989) Solar ultraviolet radiation at the earth’s surface.
Photochem Photobiol 50:443–450
Fritsen CH, Wirthlin ED, Momberg DK, Lewis MJ, Ackley SF (2011) Bio-optical properties of
Antarctic pack ice in the early austral spring. Deep-Sea Res II 58:1052–1061. https://doi.
org/10.1016/j.dsr2.2010.10.028
Galic N, Grimm V, Forbes VE (2017) Impaired ecosystem process despite little effects on popula-
tions: modeling combined effects of warming and toxicants. Glob Chang Biol 23:2973–2989.
https://doi.org/10.1111/gcb.13581
Gao K, Häder DP (2017) Effects of ocean acidification and UV radiation on marine photosynthetic
carbon fixation. In: Kumar M, Ralph P (eds) Systems biology of marine ecosystems. Springer,
Cham, pp 235–250
Gattuso JP, Hansson L (2011) Ocean acidification: background and history. In: Gattuso JP,
Hansson L (eds) Ocean acidification. Oxford University Press, Oxford, pp 1–27
Gerber S, Häder DP (1992) UV effects on photosynthesis, proteins and pigmentation in the flag-
ellate Euglena gracilaris: biochemical and spectroscopic observations. Biochem Syst Ecol
20:485–492
Goes JI, Handa N, Taguchi S, Hama T (1994) Effect of UV-B radiation on the fatty acid composi-
tion of the marine phytoplankter Tetraselmis sp.: relationship to cellular pigments. Mar Ecol
Prog Ser 114:259–274
148 P. Huovinen and I. Gómez

Goes JI, Handa N, Taguchi S, Hama T, Saito H (1995) Impact of UV radiation on the production
pattern and composition of dissolved free and combined amino acids in marine phytoplankton.
J Plakton Res 17:1337–1362
Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic
Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation? PLoS
One 10(8):e0134440. https://doi.org/10.1371/journal.pone.0134440
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, daily carbon
balance and zonation of sublittoral macroalgae from King George Island (Antarctica). Mar
Ecol Prog Ser 148:281–293
Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U et al (2009) Light and temperature demands
of marine benthic micro-algae and seaweeds in the polar regions. Bot Mar 52:593–608
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27. https://doi.org/10.1016/j.pocean.2018.03.013
González PM, Deregibus D, Malanga G, Campana GL, Zacher K et al (2017) Oxidative balance
in macroalgae from Antarctic waters. Possible role of Fe. J Exp Mar Biol Ecol 486:379–386.
https://doi.org/10.1016/j.jembe.2016.10.018
Grannas AM, Bogdal C, Hageman KJ, Halsall C, Harner T et  al (2013) The role of the global
cryosphere in the fate of organic contaminants. Atmos Chem Phys 13:3271–3305. https://doi.
org/10.5194/acp-13-3271-2013
Hanelt D, Jaramillo J, Nultsch W, Senger S, Westermeier R (1994) Photoinhibition as a regulative
mechanisms of photosynthesis in marine algae of Antarctica. Ser Cient INACh 44:67–77
Hargreaves BR (2003) Water column optics and penetration of UVR. In: Helbling EW, Zagarese
HE (eds) UV effects in aquatic organisms and ecosystems. Royal Society of Chemistry,
Cambridge, pp 59–105
Hauptmann AL, Sicheritz-Ponten T, Cameron KA, Baelum J, Plichta DR et  al (2017)
Contamination of the Arctic reflected in microbial metagenomes from Greenland ice sheet.
Environ Res Lett 12:074019
Helbling EW, Marguet ER, Villafañe VE, Holm-Hansen O (1995) Bacterioplankton viability in
Antarctic waters as affected by solar ultraviolet radiation. Mar Ecol Prog Ser 126:293–298
Hessen DO, Tranvik LJ (eds) (1998) Aquatic humic substances: ecology and biogeochemistry.
Springer, Berlin Heidelberg
Hodson A, Nowak A, Sabacka M, Jungblut A, Navarro F et al (2017) Climatically sensitive trans-
fer of iron to maritime Antarctic ecosystems by surface runoff. Nat Commun 8:14499
Hofmann LC, Bischof K (2014) Ocean acidification effects on calcifying macroalgae. Mar Ecol
Prog Ser 22:261–279
Holm-Hansen O, Lubin D, Helbling WE (1993) Ultraviolet radiation and its effects on organisms
in aquatic environments. In: Young AR, Björn LO, Moan J, Nultsch W (eds) Environmental UV
photobiology. Plenum Press, New York, pp 379–425
Huot Y, Jeffrey WH, Davis RF, Cullen JJ (2000) Damage to DNA in bacterioplankton: a model
of damage by ultraviolet radiation and its repair as influenced by vertical mixing. Photochem
Photobiol 72:62–74
Huovinen P, Gómez I (2011) Spectral attenuation of solar radiation in Patagonian fjord and coastal
waters and implications for algal photobiology. Cont Shelf Res 31:254–259
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332. https://doi.org/10.1007/
s00300-013-1351-3
Huovinen PS, Penttilä H, Soimasuo MR (2003) Spectral attenuation of solar ultraviolet radiation
in humic lakes in Central Finland. Chemosphere 51:205–214
Huovinen P, Gómez I, Lovengreen C (2006) A five-year study of solar ultraviolet radiation in
Southern Chile (39° S): potential impact on physiology of coastal marine algae? Photochem
Photobiol 82:515–522
Huovinen P, Leal P, Gómez I (2010) Interacting effects of copper, nitrogen and ultraviolet radiation
on the physiology of three south Pacific kelps. Mar Freshw Res 61:330–341
7  Underwater Light Environment of Antarctic Seaweeds 149

Huovinen P, Ramírez J, Gómez I (2016) Underwater optics in sub-Antarctic and Antarctic coastal
ecosystems. PLoS One 11(5):e0154887. https://doi.org/10.1371/journal.pone.0154887
Huovinen P, Ramírez J, Gómez I (2018) Remote sensing of albedo-reducing snow algae and impu-
rities in the Maritime Antarctica. ISPRS J Photogram Rem Sens 146:507–517
Hurd CL, Hepburn CD, Currie KI, Raven JA, Hunter KA (2009) Testing the effects of ocean acidi-
fication on algal metabolism: considerations for experimental designs. J Phycol 45:1236–1251
Iken K, Quartino ML, Barrera-Oro E, Palermo J, Wiencke C, Brey T (1998) Trophic relations
between macroalgae and herbivores. Ber Polarforsch 299:258–262
IPCC (2019) Special report on the ocean and cryosphere in a changing climate. https://www.ipcc.
ch/srocc/download-report/
Jerosch K, Scharf FK, Deregibus D, Campana GL, Zacher K et al (2019) Ensemble modeling of
Antarctic macroalgal habitats exposed to glacial melt in a Polar Fjord. Front Ecol Evol 7:207.
https://doi.org/10.3389/fevo.2019.00207
Jewett L, Romanou A (2017) Ocean acidification and other ocean changes. In: Wuebbles DJ,
Fahey DW, Hibbard KA, Dokken DJ, Stewart BC et al (eds) Climate science special report:
fourth national climate assessment, vol I. U.S. Global Change Research Program, Washington,
DC, pp 364–392
Johnston EL, Connell SD, Irving AD, Pile AJ, Gillanders BM (2007) Antarctic patterns of shallow
subtidal habitat and inhabitants in Wilke’s Land. Polar Biol 30:781–788
Karentz D, McEuen FS, Land MC, Dunlap WC (1991) Survey of mycosporine-like amino acid
compounds in Antarctic marine organisms: potential protection from ultraviolet exposure. Mar
Biol 108:157–166
Karsten U, Wulff A, Roleda MY, Müller R, Steinhoff FS et al (2009) Physiological responses of
polar benthic algae to ultraviolet radiation. Rev Bot Mar 52:639–654. https://doi.org/10.1515/
BOT.2009.077
Khairy MA, Luek JL, Dickhut R, Lohmann R (2016) Levels, sources and chemical fate of per-
sistent organic pollutants in the atmosphere and snow along the western Antarctic Peninsula.
Environ Pollut 216:304–313. https://doi.org/10.1016/j.envpol.2016.05.092
Kieber DJ, McDaniel J, Mopper J (1989) Photochemical source of biological substrates in seawa-
ter: implications for carbon cycling. Nature 341:637–639
Kieber DJ, Peake BM, Scully NM (2003) Reactive oxygen species in aquatic ecosystems. In:
Helbling EW, Zagarese HE (eds) UV effects in aquatic organisms and ecosystems. Royal
Society of Chemistry, Cambridge, pp 251–290
Kirk JTO (ed) (2011) Light and photosynthesis in aquatic ecosystems, 3rd edn. Cambridge
University Press, Cambridge
Klöser H, Ferreyra G, Schloss I, Mercuri G, Laturnus F, Curtosi A (1993) Seasonal variation of
algal growth conditions in sheltered Antarctic bays: the example of Potter Cove (King George
Island, south Shetlands). J Mar Syst 4:289–301
Klöser H, Quartino ML, Wiencke C (1996) The distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiologia 333:1–17
Kuttippurath J, Nair PJ (2017) The signs of Antarctic ozone hole recovery. Sci Rep 7:585. https://
doi.org/10.1038/s41598-017-00722-7
Lacerda ALF, Rodrigues LS, van Sebille E, Rodrigues FL, Ribeiro L et  al (2019) Plastics in
sea surface waters around the Antarctic Peninsula. Sci Rep 9:3977. https://doi.org/10.1038/
s41598-019-40311-4
Lee JR, Raymond B, Bracegirdle TJ, Chadès I, Fuller RA et  al (2017) Climate change drives
expansion of Antarctic ice-free habitat. Nature 547:49. https://doi.org/10.1038/nature22996
Lesser MP, Lamare MD, Barker MF (2004) Transmission of ultraviolet radiation through the
Antarctic annual sea ice and its biological effects on sea urchin embryos. Limnol Oceanogr
49:1957–1963
Lutz S, Anesio AM, Villar J, Benning LG (2014) Variations of algal communities cause
darkening of a Greenland glacier. FEMS Microbiol Ecol 89:402–414. https://doi.
org/10.1111/1574-6941.12351
150 P. Huovinen and I. Gómez

Lyon BR, Mock T (2014) Polar microalgae: new approaches towards understanding adaptations to
an extreme and changing environment. Biology 3:56–80
Madronich (1994) Increases in biologically damaging UV-B radiation due to stratospheric ozone
reductions: a brief review. Arch Hydrobiol Beih Ergebn Limnol 43:17–30
Madronich S, McKenzie RL, Björn LO, Caldwell MM (1998) Changes in biologically active ultra-
violet radiation reaching the earth’s surface. J Photochem Photobiol B Biol 46:5–19
Madronich S, Flocke S (1999) The role of solar radiation in atmospheric chemistry. In: Boule P
(ed) Environmental photochemistry. Springer-Verlag, Berlin, pp 1–26
McMinn A, Martin A (2013) Dark survival in a warming world. Proc R Soc B 280:20122909.
https://doi.org/10.1098/rspb.2012.2909
McNeil BI, Matear RJ (2008) Southern Ocean acidification: a tipping point at 450-ppm atmo-
spheric CO2. PNAS 105(48):18860–18864
Mitchell BG, Holm-Hansen O (1991) Bio-optical properties of Antarctic Peninsula waters: dif-
ferentiation from temperate ocean models. Deep-Sea Res 38:1009–1028
Mitchell DL, Karentz D (1993) The induction and repair of DNA photodamage in the environ-
ment. In: Young AR, Björn LO, Moan J, Nultsch W (eds) Environmental UV photobiology.
Plenum Press, New York, pp 345–377
Mobley CD (2015) Underwater light. In: Björn LO (ed) Photobiology: the science of light and life,
3rd edn. Springer, New York Springer, pp 77–84
Molina MJ, Rowland RS (1974) Stratospheric sink for chlorofluoromethanes: chlorine atom cata-
lyzed destruction of ozone. Nature 249:810–812
Morris DP, Hargreaves BP (1997) The role of photochemical degradation of dissolved organic car-
bon in regulating the UV transparency of three lakes on the Pocono Plateau. Limnol Oceanogr
42:239–249
Na G, Liu C, Wang Z, Ge L, Ma X, Yao Z (2011) Distribution and characteristic of PAHs in snow
of Fildes Peninsula. J Environ Sci 23:1445–1451
Navarro NP, Huovinen P, Gómez I (2019) Photosynthetic characteristics of geographically disjunct
seaweeds: a case study on the early life stages of Antarctic and Subantarctic species. Prog
Oceanogr 174:28–36. https://doi.org/10.1016/j.pocean.2018.11.001
Neale PJ (2000) Spectral weighting functions for quantifying effects of UV radiation in marine
ecosystems. In: de Mora S, Demers S, Vernet M (eds) Effects of UV radiation in the marine
environment, vol 10. Cambridge University Press, Environmental Chemical Series, pp 72–100
Neale PJ, Davis RF, Cullen JJ (1998) Interactive effects of ozone depletion and vertical mixing on
photosynthesis of Antarctic phytoplankton. Nature 392:585–589
Noyes PD, McElwee MK, Miller HD, Clark BW, Van Tiem LA et al (2009) The toxicology of
climate change: environmental contaminants in a warming world. Rev Env Int 35:971–986
Nuñez-Pons L, Avila C, Romano G, Verde C, Giordano D (2018) UV-protective compounds in
marine organisms from the Southern Ocean. Rev Mar Drugs 16:336. https://doi.org/10.3390/
md16090336
Oduor S, Schagerl M (2007) Phytoplankton primary productivity characteristics in response to
photosynthetically active radiation in three Kenyan Rift Valley saline-alkaline lakes. J Plankton
Res 29:1041–1050. https://doi.org/10.1093/plankt/fbm078
Padeiro A, Amaro E, dos Santos MMC, Araujo MF, Gomes SS et al (2016) Trace element contami-
nation and availability in the Fildes Peninsula, King George Island, Antarctica. Environ Sci:
Processes Impacts 18:648–657
Préndez M, Barra C, Toledo C, Richter B (2011) Alkanes and polycyclic aromatic hydrocarbons
in marine surficial sediment near Antarctic stations at Fildes Peninsula. King George Island
Antarctic Sci 23:578–588
Quartino M, Zaixso H, Boraso de Zaixso A (2005) Biological and environmental characterization
of marine macroalgal assemblages in Potter Cove, South Shetland Islands, Antarctica. Bot Mar
48:187–197
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223. https://doi.org/10.1371/journal.pone.0058223
7  Underwater Light Environment of Antarctic Seaweeds 151

Rautenberger R, Huovinen P, Gómez I (2015) Effects of increased seawater temperature on UV


tolerance of Antarctic marine macroalgae. Mar Biol 162:1087–1097
Robinson D, Arrigo K, Iturriaga R, Sullivan C (1995) Microalgal light-harvesting in extreme low-­
light environments in McMurdo Sound, Antarctica. J Phycol 31:508–520
Roleda MY, Hurd CL (2012) Seaweed responses to ocean acidification. In: Wiencke C, Bischof K
(eds) Seaweed biology: insights into ecophysiology, ecology and utilization. Springer, Berlin,
pp 407–431
Roleda MY, Wiencke C, Lüder UH (2006) Impact of ultraviolet radiation on cell structure,
UV-absorbing compounds, photosynthesis, DNA damage and germination in zoospores of
Arctic Saccorhiza dermatodea. J Exp Bot 57:3847–4856
Runcie JW, Riddle MJ (2007) Photosynthesis of marine macroalgae in ice-covered and ice free
environments in East Antarctica. Eur J Phycol 41:223–233
Russell BD, Passarelli CA, Connell SD (2011) Forecasted CO2 modifies the influence of light in
shaping subtidal habitat. J Phycol 47:744–752
Schiedek D, Sundelin B, Readman JW, Macdonald RW (2007) Interactions between climate
change and contaminants. Rev Mar Poll Bull 54:1845–1856
Schindler DW, Curtis PJ, Parker BP, Stainton MP (1996) Consequences of climate warming and
lake acidification for UV-B penetration in North American boreal lakes. Nature 379:705–708
Schloss IR, Ferreyra GA (2001) Primary production, light and vertical mixing in Potter Cove, a
shallow bay in the maritime Antarctic. Polar Biol 25:41–48
Schram JB, Schoenrock KM, McClintock JB, Amsler CD, Angus RA (2017) Ocean warming and
acidification alter Antarctic macroalgal biochemical composition but not amphipod grazer
feeding preferences. Mar Ecol Prog Ser 581:45–56
Schwarz AM, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photo-
synthesis near the southern global limit for growth; Cape Evans, Ross Sea, Antarctica. Polar
Biol 26:789–799
Schwarz A-M, Hawes I, Andrew N, Mercer S, Cummings V, Thrush S (2005) Primary production
potential of non-geniculate coralline algae at Cape Evans, Ross Sea, Antarctica. Mar Ecol Prog
Ser 294:131–140
Schoenrock KM, Schram JB, Amsler CA, McClintock JB, Angus RA (2015) Climate change
impacts on overstory Desmarestia spp. from the western Antarctic Peninsula. Mar Biol
162:377–389
Scully NM, Lean DRS (1994) The attenuation of ultraviolet radiation in temperate lakes. Arch
Hydobiol Ergebn Limnol 43:135–144
Setlow RB (1974) The wavelengths in sunlight effective in producing skin cancer: a theoretical
analysis. Proc Natl Acad Sci U S A 71:3363–3366
Smith RC, Baker KS (1979) Penetration of UV-B and biologically effective dose-rates in natural
waters. Photochem Photobiol 29:311–323
Smith RC, Prézelin BB, Baker KS, Bidigare RR, Boucher NP et al (1992) Ozone depletion: ultra-
violet radiation and phytoplankton biology in Antarctic waters. Science 255:952–959
Solomon S, Ivy DJ, Kinnison D, Mills MJ, Neely RR, Schmidt A (2016) Emergence of heal-
ing in the Antarctic ozone layer. Science 353(6296):269–274. https://doi.org/10.1126/sci-
ence.aae0061
Solomon S, Ivy D, Gupta M, Bandoro J, Santer B et al (2017) Mirrored changes in Antarctic ozone
and stratospheric temperature in the late 20th versus early 21st centuries. J Geophys Res Atmos
122:8940–8950. https://doi.org/10.1002/2017JD026719
Sulzberger B, Austin AT, Cory RM, Zepp RG, Paul ND (2019) Solar UV radiation in a changing
world: roles of cryosphere-land-water-atmosphere interfaces in global biogeochemical cycles.
Photochem Photobiol Sci 18:747–774. https://doi.org/10.1039/c8pp90063a
Taskjelle T, Hudson SR, Granskog MA, Nicolaus M, Lei R et  al (2016) Spectral albedo and
transmittance of thin young Arctic sea ice. J Geophys Res Oceans 121:540–553. https://doi.
org/10.1002/2015JC011254
Tedesco M, Doherty S, Fettweis X, Alexander P, Jeyaratnam J, Stroeve J (2016) The darkening of
the Greenland ice sheets: trends, drivers, and projections (1981–2100). Cryosphere 10:477–496
152 P. Huovinen and I. Gómez

Tedetti M, Sempéré R (2006) Penetration of ultraviolet radiation in the marine environment. A


review. Photochem Photobiol 82:389–397
Tyler JE (1975) The in situ quantum efficiency of natural phytoplankton populations. Limnol
Oceanogr 20:976–980
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental gradients in a marine macroalgal-dominated subtidal community on the Western
Antarctic Peninsula. PLoS One 10(9):e0138582. https://doi.org/10.1371/journal.pone.0138582
Vaughan DG, Doake CSM (1996) Recent atmospheric warming and retreat of ice shelves on the
Antarctic Peninsula. Nature 379:328–331
Vecchiato M, Argiriadis E, Zambon S, Barbante C, Toscano G et  al (2015) Persistent Organic
Pollutants (POPs) in Antarctica: occurrence in continental and coastal surface snow. Microchem
J 119:75–82. https://doi.org/10.1016/j.microc.2014.10.010
Vincent WF, Belzile C (2003) Biological UV exposure in the polar oceans: Arctic-Antarctic com-
parisons. In: Huiskes AHL, Gieskes WWC, Rozema J, Schorno RML, van der Vies SM, Wolff
WJ (eds) Antarctic biology in a global context. Backhuys Publishers, Leiden, The Netherlands,
pp 176–181
Vincent WF, Neale PJ (2000) Mechanisms of UV damage to aquatic organisms. In: de Mora S,
Demers S, Vernet M (eds) The effects of UV radiation in the marine environment, vol. 10.
Cambridge University Press, Cambridge, pp 149–176
Vincent WF, Roy S (1993) Solar ultraviolet-B radiation and aquatic primary production: damage,
protection, and recovery. Environ Rev 1:1–12
Wadham JL, Hawkings JR, Tarasov L, Gregoire LJ, Spencer RGM et al (2019) Ice sheets matter for
the global carbon cycle. Nat Commun 10:3567. https://doi.org/10.1038/s41467-019-11394-4
Wahl M, Molis M, Davis A, Dobretsov S, Dürr ST et al (2004) UV effects that come and go: a global
comparison of marine benthic community level impacts. Glob Chang Biol 10:1962–1972.
https://doi.org/10.1111/j.1365-2486.2004.00872.x
Waller CL, Griffiths HJ, Waluda CM, Thorpe SE, Loaiza I et al (2017) Microplastics in the Antarctic
marine system: an emerging area of research. Review. Sci Total Environ 598:220–227. https://
doi.org/10.1016/j.scitotenv.2017.03.283
Wania F, Westgate JN (2008) On the mechanisms of mountain cold-trapping of organic chemicals.
Environ. Sci Technol 42:9092–9098
Weiler CS, Penhale PA (eds) (1994) Ultraviolet radiation in Antarctica: measurements and biologi-
cal effects, vol 62. American Geophysical Union, Antarctic Research Series, Washington DC
Weykam G, Gómez I, Wiencke C, Iken K, Klöser H (1996) Photosynthetic characteristics and C:
N ratios of macroalgae from King George Island (Antarctica). J Exp Mar Biol Ecol 204:1–22
Wiencke C, Gómez I, Pakker H, Flores-Moya A, Altamirano M et al (2000) Impact of UV-radiation
on viability, photosynthetic characteristics and DNA of brown algal zoospores: implications for
depth zonation. Mar Ecol Prog Ser 197:217–229
Wiencke C, Roleda MY, Gruber A, Clayton MN, Bischof K (2006) Susceptibility of zoospores to
UV radiation determines upper depth distribution limit of Arctic kelps: evidence through field
experiments. J Ecol 94:455–463
Wiencke C, Gómez I, Dunton K (2009) Phenology and seasonal physiological performance of
polar seaweeds. Bot Mar 52:585–592
Williamson CE, Neale PJ, Grad G, de Lange HJ, Hargreaves BR (2001) Beneficial and detrimental
effects of UV on aquatic organisms: implications of spectral variation. Ecol Appl 11:1843–1857
Williamson CE, Madronich S, Lal A, Zepp RG, Lucas RM et al (2017) Climate change-induced
increases in precipitation are reducing the potential for solar ultraviolet radiation to inactivate
pathogens in surface waters. Sci Rep 7:13033
Williamson CE, Neale PJ, Hylander S, Rose KC, Figueroa FL et al (2019) The interactive effects
of stratospheric ozone depletion, UV radiation, and climate change on aquatic ecosystems.
Photochem Photobiol Sci 18:717–746. https://doi.org/10.1039/c8pp90062k
Wrona FJ, Prowse TD, Reist JD, Hobbie JE, Levesque LMJ et al (2006) Effects of ultraviolet radi-
ation and contaminant-related stressors on Arctic freshwater ecosystems. Ambio 35:388–401
7  Underwater Light Environment of Antarctic Seaweeds 153

Yan ND, Keller W, Scully NM, Lean DRS, Dillon PJ (1996) Increased UV-B penetration in a lake
owing to drought-induced acidification. Nature 381:141–143
Zacher KD, Hanelt D, Wiencke C, Wulff A (2007a) Grazing and UV radiation effects on an
Antarctic intertidal microalgal assemblage: a long-term field study. Polar Biol 30:1203–1212
Zacher K, Roleda MY, Hanelt D, Wiencke C (2007b) UV effects on photosynthesis and DNA
in propagules of three Antarctic seaweeds (Adenocystis utricularis, Monostroma hariotii and
Porphyra endiviifolium). Planta 225:1505–1516
Zacher K, Rautenberger R, Hanelt D, Wulff A, Wiencke C (2009) The abiotic environment of polar
marine benthic algae. Bot Mar 52:483–490
Zepp RG (2003) Solar UVR and aquatic carbon, nitrogen, sulfur and metals cycles. In: Helbling
EW, Zagarese HE (eds) UV effects in aquatic organisms and ecosystems. Royal Society of
Chemistry, Cambridge, pp 137–184
Zielinski K (1990) Bottom macroalgae of the Admiralty Bay (King George Island, Antarctica). Pol
Polar Res 11:95–131
Chapter 8
Production and Biomass of Seaweeds
in Newly Ice-Free Areas: Implications
for Coastal Processes in a Changing
Antarctic Environment

María L. Quartino, Leonardo A. Saravia, Gabriela L. Campana,


Dolores Deregibus, Carolina V. Matula, Alicia L. Boraso,
and Fernando R. Momo

Abstract  The Antarctic rocky coasts are mainly colonized by extensive seaweed
communities, which play key roles as food resource, habitat, and refuge for many
benthic and pelagic organisms. Due to climate warming, Antarctic marine ecosys-
tems are being affected by glacier retreat opening new habitats, e.g., newly ice-free
areas that can be colonized by macroalgae. As a consequence, primary production
and fate of macroalgae are changing in these new polar environments. In these eco-
systems, the carbon production, especially from large brown algae, is an important
food source to the benthic invertebrate communities mainly when other resources
are scarce. Thus, in new areas colonized by seaweeds, the trophic structure and
biogeochemical fluxes can vary considerably. Moreover, when seaweeds die or are
removed by water movement, ice scouring, or storms, they are detached, frag-
mented, and degraded, incorporating and releasing particulate and dissolved organic
matter to the coastal food webs, i.e., they support a large fraction of the secondary
production of the benthos. The present chapter is a review of the knowledge on
seaweed biomass and production in the coastal Antarctic ecosystem opening a dis-
cussion on the role of these organisms as main energy sources in, e.g., small fjords
and glacier-influenced sites, impacted by recent climatic change.

Keywords  Carbon flux · Glacier retreat · Ice-free areas · Potter Cove · Seaweed
production

M. L. Quartino (*)
Departamento de Biología Costera, Instituto Antártico Argentino (IAA),
Buenos Aires, Argentina
Museo Argentino de Ciencias Naturales B Rivadavia (MACN), Buenos Aires, Argentina
e-mail: lquartino@dna.gov.ar

© Springer Nature Switzerland AG 2020 155


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_8
156 M. L. Quartino et al.

8.1  Introduction: Seaweeds in Coastal Marine Ecosystems

Worldwide, macroalgae are essential components of coastal waters providing habi-


tat‚ nursery and refuge for fish and numerous mobile and sessile invertebrates (Hurd
et al. 2014) and can potentially also play a role in providing refuge from climate
change driven stressors, e.g., ocean acidification (Hurd 2015; Krause-Jensen et al.
2016; Wahl et al. 2018; see also Chap. 11 by Gómez and Huovinen). Furthermore,
large brown algae, e.g., species of order Laminariales (known as “kelps”), are
regarded as ecosystem engineers that could significantly modify the habitat charac-
teristics such as water velocity, light penetration, and physic-chemical properties of
seawater (Jones et al. 1994; Dawson et al. 2010), while creating an understory con-
dition favorable for species adapted to, e.g., low light intensity (Steneck et al. 2002).
The macroalgal biomass is an important component of the functioning of the
marine ecosystem; a large portion of seaweeds is not consumed by herbivores but
returns to the environment as decaying organic matter (Cebrián 2004). These large

L. A. Saravia
Instituto de Ciencias, Universidad de General Sarmiento (UNGS),
Los Polvorines, Buenos Aires, Argentina
e-mail: lsaravia@campus.ungs.edu.ar
G. L. Campana
Departamento de Biología Costera, Instituto Antártico Argentino (IAA),
Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: gcampana@dna.gov.ar
D. Deregibus
Departamento de Biología Costera, Instituto Antártico Argentino (IAA),
Buenos Aires, Argentina
Consejo Nacional de Investigaciones Científicas y Técnicas, (CONICET),
Buenos Aires, Argentina
e-mail: dderegibus@dna.gov.ar
C. V. Matula
Departamento de Biología Costera, Instituto Antártico Argentino (IAA),
Buenos Aires, Argentina
e-mail: ucv@mrecic.gov.ar
A. L. Boraso
Instituto de Desarrollo Costero, Universidad Nacional de la Patagonia San Juan Bosco
(UNPSJB), Comodoro Rivadavia, Argentina
F. R. Momo
Instituto de Ciencias, Universidad de General Sarmiento (UNGS),
Los Polvorines, Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: fmomo@campus.ungs.edu.ar
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 157

quantities of seaweed biomass can follow different paths: they can be detached from
the hard bottom and drift by the water movements (Baring et al. 2018). Also the
whole thalli can drift and slowly be fragmented and be thrown by the waves to the
coastline. On the beach they can be accumulated and, once degraded, enter again to
the system by the tides sinking to the bottom of the seabed where they decompose
(Krause-Jensen and Duarte 2016). Thus, this algal debris can supply important
amounts of organic matter to benthic ecosystems and provide food, directly for
detritivores or indirectly, by stimulating bacterial metabolism (Rossi et al. 2013).
Primary production is the formation of organic matter (carbon) in a photosyn-
thetic organism per time unit (Israel 1995). In general, ecological studies of primary
production refer to “net primary production” (NPP), which is that portion of gross
primary production from photosynthesis that remains after consumption via respira-
tion. This may represent different processes depending on the methodology used to
measure it. In most studies, net primary production represents the dry mass of plant
matter produced per unit area per unit time (Reed et al. 2008).
Seaweed biomass can be measured using destructive and nondestructive sam-
pling. The first one corresponds to the collection of algal biomass in terms of grams
by unit area (g m−2). The nondestructive sampling implies the use of underwater
images or video, which mainly permit to gain information on algal cover. Coverage
data must be converted into biomass, usually through linear regression to estimate
finally biomass (Allison 2004). Several improvements have been proposed to this
method, taking into account the species-specific and nonlinear nature of the rela-
tionship between cover and biomass due to the different algal architectures. Thus, a
better prediction of biomass would be obtained by fitting a multivariate generalized
additive model (GAM) considering all species present in each quadrat (Pedersen
et al. 2019). Various studies have focused on quantifying the macrophyte production
to assess its contribution to the marine ecosystem worldwide. For example, mac-
roalgae and seagrasses store about 0.4% and 16% of their net primary production in
the sediments, respectively (Duarte and Cebrián 1996), and some of the excess
organic matter can be exported to adjacent water column (Barrón and Duarte 2009;
Wada and Hama 2013). Duarte and Cebrián (1996) reported that the export of
organic matter accounts, on average, for 25% of the net primary production of mac-
roalgae. Thus, macroalgae can be regarded as the most extensive and productive
primary producers in coastal zones around the world. It was estimated that they
cover about 3.5 million km2 and can account for a global net primary production of
about 1.5 Tg Cyr−1 (Krause-Jensen and Duarte 2016).

8.2  M
 acroalgae and Carbon Fluxes in Antarctic
Coastal Areas

In contrast to the global algal production, there is not enough information on mac-
roalgal production in Antarctica. In Antarctic marine ecosystems, macroalgae are
one of the major primary producers that grow and develop associated to rocky
158 M. L. Quartino et al.

substrates of different slope and size (Klöser et  al. 1996; Wiencke and Amsler
2012). In general, Antarctic benthic macroalgae are distributed from the intertidal
down to the lower subtidal zone at 40 m depths and present a clear zonation: an
intertidal fringe is usually abraded by winter ice floes, it is mainly colonized by
opportunistic species, and it is devoid of large perennial macroalgae. The upper
subtidal is dominated by the brown seaweeds Desmarestia menziesii J.Agardh and
Ascoseira mirabilis Skottsberg under strong turbulence. The central subtidal (above
15 m) is mainly dominated by Desmarestia anceps Montagne under moderate tur-
bulence, and it is gradually replaced by Himantothallus grandifolius (Gepp and
Gepp) Zinova. This latter species becomes dominant at deeper waters, where turbu-
lence is usually negligible (Klöser et  al. 1996; Quartino et  al. 2005; Wulff et  al.
2009; see Chap 11 by Gómez and Huovinen).
Antarctic macroalgae are thought to deliver huge amounts of organic matter to
the Antarctic coastal food webs (Fischer and Wiencke 1992; Gómez et  al. 2009,
Marina et al. 2018). Furthermore, they are a food source for numerous mobile con-
sumers, including invertebrates and some nototheniid fish (Iken et al. 1998; Amsler
et al. 2005; Barrera Oro et al. 2019). Sublittoral rocky shores dominated by mac-
roalgae have been regarded as important areas hosting high values of abundance and
biomass of organisms (Amsler et  al. 1995; Brouwer et  al. 1995; Quartino and
Boraso de Zaixso 2008; Wulff et al. 2009).
One of the main macroalgal carbon contributions to the ecosystem is probably
the great amounts of biomass scattered on the coasts, an accumulation of allochtho-
nous organic debris known as “wrack” that is commonly observed in the intertidal
zone. In the maritime Antarctic zone, this type of organic matter has been recog-
nized as an important source of carbon for diverse consumers (Zenteno et al. 2019).
Grange and Smith (2013) reported drifting macroalgae as very abundant in some
fjord basins compared to the open shelf. These authors observed substantial abun-
dances of drifting algae ranging from ~8 to 130 cm m−2.
Once on the coast, the stranded algae can also be used by seals and sea elephants
as resting places. Westermeier et al. (1992) recorded a mean value of 22 kg m−2 cor-
responding to the highest abundance of drifting macroalgae along the coastline of
Maxwell Bay. In Admiralty Bay, the amount of macroalgae deposited on the beach
was estimated to be 104 kg m−2 dry weight (Rakusa-Suszczewski 1995). The organic
matter of the decomposed algae mixed with excrement of seabirds (guano) reenters
the marine ecosystem by the waves, winds, and tide action. Leaching of nutrients to
the environment and subsequently to the coastal waters, are the final stages of mac-
roalgal decomposition, after which the dissolved nutrients fertilize the shore through
the tidal beach runoff with nitrate, nitrite, ammonia, and phosphorous (Nedzarek
and Rakusa-Suszczewski 2004). Thus, the volumes of algal debris can reflect the
high productivity normally assigned to Antarctic seaweeds populating different
regions across the Antarctic Peninsula.
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 159

8.3  Macroalgal Biomass Studies in Antarctica

In the decade 1980–1990s and early the 2000s, several studies described the bio-
mass (standing crop) of macroalgae at the Antarctic benthos, particularly around the
West Antarctic Peninsula (Table 8.1). These early works implied frequently the use
of “destructive sampling,” which allowed a more accurate assessment of the identity
and biomass of each taxa (DeLaca and Lipps 1976; Miller and Pearse 1991; Amsler
et al. 1995; Brouwer et al. 1995; Quartino et al. 2001, 2005). These studies showed
that in the northern portion of the Western Antarctic Peninsula and adjacent islands,
seaweeds dominate shallow benthic communities on hard substrates, covering over
80% of the seafloor, with standing biomass levels in the range of 5–10 kg m−2 wet
weight (Wiencke and Amsler 2012, Gómez et al. 2009), a number comparable to
temperate kelp forests (Mann 1972; Duggins 1988).
During the following decades, many benthic community studies were conducted
using underwater photo and/or video (nondestructive sampling) (Quartino et  al.
2013; Sahade et  al. 2015; Valdivia et  al. 2015; Lagger et  al. 2018; Jerosch et  al.
2019). These methodologies facilitate fast sampling and are a great advantage in

Table 8.1  Maximum biomass (g m−2, wet or dry weight) determined for four species of macroalgae
reported by different surveys (depths between brackets when provided by the authors)
Desmarestia Desmarestia Himantothallus Iridaea
Locality anceps menziesii grandifolius cordata Reference
Wet weight
(1) – 2050 – – Richardson (1979)
(2) 5660 (11 m) 1850 (11 m) 1250 (11 m) – Brouwer et al. (1995)
(3) – 3440 Cormaci et al. (1996)
(4–5 m)
(4) a 6044 (20 m) 6737 (5 m) 10,336 (20 m) 2554.60 Quartino and Boraso
(0 m) de Zaixso (2008)
(4) b 3900 (5 m) 3470 (3 m) 923.8 C. Matula, personal
(3 m) communication
4198 (10 m) 3975 (5 m) 575 (5 m)
Dry weight
(5) – 800 (4 m) 600 (12 m) – DeLaca and Lipps
(1976)
(2) 1000 (11 m) 460 (11 m) 240 (11 m) – Brouwer et al. (1995)
3300 (5 m)
(4) a 671 (20 m) 749 (5 m) 1152 (20 m) 255.36 Quartino and Boraso
(0 m) de Zaixso (2008)
(4) b 821 (5 m) 730.5 (3 m) 194.5 C. Matula, personal
(3 m) communication
883.8 (10 m) 836.85 (5 m) 121 (5 m)
Localities: (1) Borge Bay; South Orkney Islands (60° 43′S, 45° 36′W); (2) Signy Island, South
Orkney Islands (60° 42′S 45° 36′W); (3) Ross Sea (74° 30′S, 165° 30′E); (4) Potter Cove, South
Shetland Islands (62° 14′S, 58° 38′W), a, sampling done in 1994–1996, and b, sampling done in
2015–2016; (5) Ansvers Island (62° 46′S, 64° 04′W)
160 M. L. Quartino et al.

this extreme and rough environment. Values recorded by these methods are mainly
percent coverage of macroalgae by unit area. Although cover data are valuable, it is
always convenient to transform it into biomass in order to calculate and quantify the
algal production as mentioned previously.

8.4  T
 he Ecosystem of Potter Cove: An Outstanding
Case Study

Potter Cove (62° 14′S, 58° 40′W), a small fjord at the Isla 25 de Mayo/King Gorge
Island, South Shetland Islands, has become an ideal place to carry out biological,
geological, and oceanographic studies. This cove is divided into two characteristic
areas, the mouth (or outer part) and an inner part separated by a shallow sill (Klöser
et al. 1996). Both areas are biologically distinctive: the rocky shores of the outer
part of Potter Cove host a high biomass of macroalgae (Klöser et al. 1996; Quartino
et al. 2005), whereas the inner cove has one of the largest concentrations of benthic
filter feeders found in Antarctica (Tatián et al. 2004).
In fact, the first surveys on macroalgae had identified an important benthic algal
community associated with the presence of hard bottom substrate and light avail-
ability (Quartino et al. 2001, 2005). In most of the studied sites, abundance of large
Desmarestiales was observed from 5 down to 30 m depth, in a continuous vertical
zonation. This high algal biomass was detected on rocky coastal areas close to the
mouth of the cove, where the more transparent, oceanic water prevailed (Quartino
et al. 2001, 2005).
Macroalgal production of the most abundant species decreased during summer
months (Quartino and Boraso de Zaixso 2008) following the seasonal strategy
exhibited normally by “season anticipators,” which start their growth under short-­
day conditions in late winter/spring, often under the sea-ice, reaching their maxi-
mum biomass in summer months (Wiencke et  al. 2009; Wiencke et  al. 2014).
Quartino and Boraso de Zaixso (2008) calculated an average standing stock of 792,
84 ton and an algal production of 1401 ton during summer season for the whole
Potter Cove. This study provided the first estimation of seaweed production in
Antarctica, combining field biomass data with experimental growth rates calculated
by Wiencke (1990a, b) and Gómez and Wiencke (1997). Additionally, an estimation
of their decomposition was included to describe the macroalgal mass balance in this
Antarctic fjord (Fig. 8.1). Due to the usually low pelagic primary productivity in
Potter Cove (Schloss and Ferreyra 2002), the study suggested that macroalgae were
the main carbon source, supporting a large fraction of the secondary production of
the benthic system (Tatián et al. 2004; Sahade et al. 2004; Quartino et al. 2008).
The continuous monitoring of this marine ecosystem has permitted to survey the
changes in the physiognomy of the cove along the last three decades. One of the
most singular observations was the progressive melting of the Fourcade Glacier
(Rückamp et al. 2011), which surrounds Potter Cove. The retreat of this glacier has
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 161

Fig. 8.1  Pathways of macroalgae production in Potter Cove. The average macroalgal standing
stock (792 ton) was calculated for the summer season considering the entire cove area (Quartino
and Boraso de Zaixso 2008). The production (1401 ton) and the flux of biomass to the ecosystem
(1370 ton) were calculated using only the dominant macroalgal species (Quartino et  al. 2008).
Detached macroalgae are expected to enter the inner cove via the prevailing cyclonic water circula-
tion (Schloss and Ferreyra 2002). The decomposed fraction was calculated using a rate of
0.0016 day−1, estimated by Brouwer (1996) for Desmarestia anceps

been uncovering new hard bottom ice-free areas available for benthic colonization,
especially macroalgae (Quartino et  al. 2013). These newly ice-free areas are
extremely disturbed, presenting a considerable alteration of the water column (due
to the increase of sediment input and salinity changes particularly during the sum-
mer melting season) and on the ice disturbance patterns (Eraso and Domínguez
2007; Quartino et al. 2013). In fact, the photographs and video records revealed a
well-developed macroalgal community even in close proximity to the retreating
glacier where the sediment load was high. In these sites, the increase in the sediment
runoff reduces the light penetration and can constitute a constraint for photosynthe-
sis (Schloss et al. 2012; Wiencke and Amsler 2012; Quartino et al. 2013). Particularly,
this can result in a change in the vertical distribution of those species adapted to
dark conditions, which probably will move to shallower depths adjusting their light
requirements in this new environment with higher sediment load (Deregibus et al.
2016; see also Chap. 9 by Deregibus et al.).
Thus, under these optical conditions, development and vertical distribution of the
macroalgae can be strongly affected by the high turbidity (Deregibus et al. 2016; see
162 M. L. Quartino et al.

also Chap. 7 by Huovinen and Gómez). Additionally macroalgal colonization was


negatively impacted by the ice disturbance caused by ice blocks originated from the
glacier (Quartino et  al. 2013). Nevertheless some species such as the red alga
Palmaria decipiens (Reinsch) Ricker and the green alga Monostroma hariotii Gain
thrive in the most disturbed sites. It was also shown that Gigartina skottsbergii
Setchell and Gardner and Himantothallus grandifolius were adapted to harsh abi-
otic conditions, but the red alga G. skottsbergii seemed to be more sensitive to the
external stress (González et al. 2017). The vertical distribution of the subtidal spe-
cies in the inner part of the cove did not fit with the typical macroalgal zonation
reported for other sites around King George Island (Huovinen and Gómez 2013 and
Chap 11 by Gómez and Huovinen); in Potter Cove algal species usually found at
greater depth were observed growing at shallow waters. The main results showed
that (1) the complexity of the macroalgal community was positively correlated to
the elapsed time from the ice retreat: newly ice-free areas closer to the outer side of
the cove presented mature macroalgal communities dominated by perennial and
large brown algae of the genus Desmarestia, (2) algal development depended on the
optical conditions and the sediment input with some species being limited by light
availability, (3) macroalgal colonization was negatively affected by ice disturbance,
and (4) the colonization was determined by the size and type of substrate and by the
slope of the seafloor (Fig. 8.2).
In this new warming scenario, ice melting is probably the primary cause of
changes in the macroalgal communities in the inner Cove, and it could be mediated
by different associated phenomena (Fig. 8.3). Glacier retreat originates newly ice-­
free areas providing hard substrate available for benthic colonization, which can be
positively related to increased diversity, richness, and macroalgal cover at different
spatial scales (Valdivia et al. 2014). In addition, glacier melting increases ice scour-
ing, sediment runoff in the water column and enhances turbidity having a negative
effect on these ecological parameters because algae do not have enough light avail-
able for photosynthesis. Consequently, macroalgae may shift their vertical distribu-
tion increasing overlapping and competition, resulting in a negative effect (Fig. 8.3).
Additionally a complementary PAR model was performed with the variables affect-
ing the annual PAR availability in this climate change context (see Chap 9 by
Deregibus et al.).

8.5  A
 Dynamic Growth Model for Antarctic Macroalgae
Under a Fast-Changing Environment

Dynamic growth models have been developed mainly for farmed macroalgae
(Hadley et al. 2015; Zhang et al. 2016; Broch and Slagstad 2012; Duarte and Ferreira
1997) and also to determine the growth of species in eutrophicated sites and under
algal blooms at mid latitudes (Ren et  al. 2014; Perrot et  al. 2014). Despite their
recognized great importance as primary producers (Hurd et al. 2014), there are no
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 163

Fig. 8.2  Three different situations observed in the newly ice-free areas of Potter Cove. S1 corre-
sponded to sites unaffected by sedimentation and dominated by large brown algae. S2 corre-
sponded to a site dominated by the red algae Palmaria decipiens in close proximity to a retreating
glacier with high sediment inflow and high ice disturbance. S3 represented a newly ice-free area in
the inner side of the cove close to the glacier and with influence of sediment runoff from small
creeks. In this site the coexistence of macroalgae and macrofauna was observed. Macroalgae were
classified according to their life history as annuals, pseudo perennials and perennials. Symbols +
and – correspond to the grade of intensity. (Adapted from Quartino et al. (2013))

dynamic models for Antarctic macroalgae. In this line, we recently described a new
model adapted to the particular conditions of the Antarctic environment (Guillaumot
et al. 2018). The model describes the biomass dynamics of algal assemblages with-
out considering a particular species (Fig. 8.4) and can also be applied to dominant
species that do not experience interspecific competition. The outputs of this approach
can be integrated in carbon flux models of Antarctic coastal ecosystems. The
assumptions of the model are that the growth is not limited by nutrient availability
(Drew and Hastings 1992; Ducklow et al. 2007; Wiencke and Amsler 2012) and that
substrate availability, light, and ice scouring should be regarded as the main envi-
ronmental variables (Quartino et al. 2013). Considering that the levels of irradiance
that reach the bottom is mediated by depth, suspended sediments, and incident radi-
ation, the model takes into account the high variability in incident light and turbid-
ity, two factors with high seasonal fluctuations in Antarctic. The estimations use a
daily time sequence, assuming that the numbers of days available for growth are
164 M. L. Quartino et al.

Fig. 8.3  Conceptual model on the effects of glacier retreat on structural attributes of the macroal-
gal community. (Adapted from Quartino et al. (2013))

Fig. 8.4  Overview of the dynamic growth model of Antarctic macroalgae and the main variables
entered in the algorithm. (Adapted from Gómez and Huovinen (2015))
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 165

determined by the period when the zone considered is free of ice. Then the model
equations would be:

B ( t + 1) = B ( t ) + µmax Λ ( t ) Γ ( t ) B ( t ) − D ( t ) B ( t ) − M ( t )

where B(t) is the biomass; μmax is the maximum intrinsic growth rate; Λ(t) and Γ
represent the effects of light and temperature on growth, respectively; and D(t) and
M(t) are the biomass dependent mortality rates associated with herbivory and ice
scouring respectively.
Each of these terms is also dependent on other environmental characteristics
defined as:

 I 
I 1− I 
Λ = e 0 
I0

where I0 is the optimum irradiance and I is the light available at the seabed that will
have more dependencies:

I = I s e − Kd Z

where Is is the irradiance at the sea surface, Kd is the extinction coefficient (Kirk
1994), and Z is the depth. The incident light Is must be recorded in situ and averaged
over several years to represent the mean conditions, while Kd is also affected by the
turbidity produced by glacier melting (Deregibus et al. 2016; see also Chap. 9 by
Deregibus et al.).
The dependence of growth on temperature is modeled as a function of an optimal
temperature range (Duarte et al. 2003):

2.0 (1 + β ) XT
Γ=
X + 2.0 β XT + 1.0
2
T

where Tmax is the upper temperature above which growth ceases and β is an adjust-
ment parameter. XT is defined as:

T − Topt
XT =
Topt − Tmax

where Topt is the optimal temperature for growth and T is the average daily water
temperature which must be supplied from local measurements or estimated. Even
although this kind of mechanistic model requires experimental and field data to be
calibrated and validated, it can be a useful tool to provide predictions of macroalgal
biomass in different scenarios of climate change.
166 M. L. Quartino et al.

8.6  S
 eaweed Production in Present and Future
Warming Scenarios

In recent decades, the ice shelves of the WAP have changed rapidly and conse-
quently a significant retreat was observed on both sides of the Peninsula (Cook et al.
2005). In response to relatively small environmental change, some ecosystems can
undergo abrupt transformation (Clarke et al. 2013). For example, decreases in the
spatial and temporal extent of sea ice will affect the underwater light environment
during winter and early spring altering photosynthesis of organisms adapted to spe-
cific light regimes (See Chap. 9 by Deregibus et al.). Moreover, a shorter perma-
nence of sea ice and enhanced sediment inflow from the glaciers will modify the
environment reducing light availability for algal photosynthesis and can also affect
survival and germination rates directly (Arakawa and Matsuike 1992; Chapman and
Fletcher 2002; Eriksson and Johansson 2005). A significant change in the annual
light budget is likely to have major consequences for benthic ecosystems in shallow
waters: the timing of annual sea ice breakup affects the composition of benthic com-
munities, primarily due to a change in the available light on an annual basis (Clarke
et al. 2013; see also Chap. 7 by Huovinen and Gómez). Thus, global warming is
expected to cause an earlier sea-ice melting, which will likely induce tipping points
for many areas at shallow depths, probably causing ecosystems to shift from pre-
dominantly heterotrophic to predominantly autotrophic, the later state dominated by
macroalgal production (Clarke et al. 2013; see also Chap. 15 by Momo et al. and
Chap. 16 by Ortiz et al.).
Linking macroalgal decomposition and the trophic network in the future coastal
scenarios, the colonization of macroalgae in the new areas probably will change the
structure and composition of the whole ecosystem. As it was mentioned before, this
situation is being detected in Potter Cove, where the increasingly higher coloniza-
tion of macroalgae in newly ice-free areas is providing carbon sources (food) for the
second fraction of the trophic web. In this system sediment retains the nitrogen and
carbon released from the macroalgae degradation process, whereas a relatively
small amount of macroalgal-derived nutrients is released back into the overlying
water, with a negative feedback on phytoplankton production in the water column
(Braeckman et al. 2019). These authors suggested that this “sink effect” or “food
bank” would also explain the prolonged food availability in the euphotic Antarctic
sedimentary benthos in the form of microphytobenthic biomass, at least as long as
the overlying waters are not too turbid or ice covered (Mincks et al. 2005).
A recent study in Potter Cove using an ensemble model predicted a reduction of
macroalgal summer production under increasing sedimentation (Jerosch et  al.
2019). Thus, the projection of macroalgal distribution in a near future scenario
shows a dramatic reduction of summerly macroalgal productivity inside the cove,
while the glacier continues melting and increasing the discharge of sediments.
However, this process could be mitigated in part by an increased colonization and
productivity of macroalgae in shallow newly ice-free hard bottom areas inside the
cove (Quartino et al. 2013; Deregibus et al. 2016; Campana et al. 2018).
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 167

8.7  Future Prospects

Even though macroalgae are the most productive marine macrophytes worldwide,
they have been excluded from “blue carbon” assessments as they typically grow in
habitats that are not considered to accumulate large stocks of organic carbon
(Krause-Jensen and Duarte 2016). Recently, more studies have highlighted the
potential of marine vegetation as a sink for anthropogenic C emissions (known as
“Blue Carbon”) suggesting that marine macroalgae can sequestrate anthropogenic
CO2 (Chung et al. 2011; Queirós et al. 2019). So far, there are also some few reports
of the presence of macroalgal carbon in marine sediments, suggesting that the pres-
ence of macroalgal carbon may be widespread, extending from shallow to deep-sea
sediments and from polar to tropical regions. It has been also reported that this type
of carbon can be found across a broad range of depths into the sediment, from sur-
face and subsurface layers down to deeper than a hundred metres into the sediment
(Krause-Jensen et al. 2018).
Despite the evidence of macroalgae as carbon sinks, the rates and magnitude of
this process has not been estimated precisely. Some studies have delivered a first-­
order estimate of the contribution of macroalgae to carbon sequestration from burial
in coastal sediments and its export to the deep sea (Duarte and Cebrián 1996;
Krause-Jensen and Duarte 2016). This last step of macroalgal C sequestration is
particularly important because the exchange of carbon with the atmosphere is pre-
cluded over extended time periods, even after being remineralized (Krause-Jensen
and Duarte 2016).
In all, meltwater influences coastal areas where macroalgae are the dominant
species covering most of the rocky bottoms of many coastal areas along the
WAP. However, there is still little knowledge on the macroalgal carbon production
in Antarctic shallow waters and its fate under changing environmental conditions.
Therefore, it would be necessary to implement monitoring programs to measure the
expansion of these organisms in other locations of the WAP to estimate at large
scale the fraction of macroalgae production sink in the sediment, the budget of mac-
roalgal production, their decomposition, and further export to other marine
ecosystems.

Acknowledgments  Special thanks to Silvia Rodríguez for her valuable collaboration and techni-
cal assistance at the Coastal Biology Department of Instituto Antártico Argentino. We also thank
the support of the diving and logistic teams at Carlini Station. This study was performed with the
financial support of Instituto Antártico Argentino/Dirección Nacional del Antártico, ANPCyT-­
DNA (PICTO 2010-116, PICT 2017-2691). This chapter also presents an outcome of the interna-
tional Research Network IMCONet funded by the Marie Curie Action IRSES IMCONet (FP7
IRSES, Action No. 318718) and of CoastCarb (Funding ID 872690, H2020-MSCA-RISE-2019 -
Research and Innovation Staff Exchange).
168 M. L. Quartino et al.

References

Allison G (2004) The influence of species diversity and stress intensity on community resistance
and resilience. Ecol Monogr 74(1):117–134
Amsler CD, Rowley RJ, Laur DR et al (1995) Vertical distribution of Antarctic peninsular mac-
roalgae: cover, biomass and species composition. Phycologia 34(5):424–430
Amsler CD, Iken K, McClintock JB et al (2005) Comprehensive evaluation of the palatability and
chemical defenses of subtidal macroalgae from the Antarctic Peninsula. Mar Ecol Prog Ser
294:141–159
Arakawa H, Matsuike K (1992) Influence on insertion of zoospores, germination, survival,
and maturation of gametophytes of brown algae exerted by sediments. Nippon Suis Gakkai
58:619–625
Baring RJ, Fairweather PG, Lester RE (2018) Nearshore drift dynamics of natural versus artificial
seagrass wrack. Estuar Coast Shelf Sci 202:164–171
Barrera-Oro E, Moreira E, Seefeldt MA et al (2019) The importance of macroalgae and associ-
ated amphipods in the selective benthic feeding of sister rockcod species Notothenia rossii and
N. coriiceps (Nototheniidae) in West Antarctica. Polar Biol 42(2):317–334
Barrón C, Duarte CM (2009) Dissolved organic matter release in a Posidonia oceanica meadow.
Mar Ecol Prog Ser 374:75–84
Braeckman U, Pasotti F, Vazquez S, Zacher K, Hoffmann R, Elvert M, Marchant H, Buchner C,
Quartino ML et al (2019) Degradation of macroalgal detritus in shallow coastal Antarctic sedi-
ments. Limnol Oceanogr 64:1423–1441
Broch OJ, Slagstad D (2012) Modelling seasonal growth and composition of the kelp. J Appl
Phycol 24:759–776
Brouwer PE (1996) Decomposition in situ of the sublittoral Antarctic macroalga Desmarestia
anceps Montagne. Polar Biol 16(2):129–137
Brouwer PE, Geilen EFM, Gremmen NJ et  al (1995) Biomass, cover and zonation pattern of
sublittoral macroalgae at Signy Island, South Orkney Islands, Antarctica. Bot Mar 38:259–270
Campana GL, Zacher K, Deregibus D et al (2018) Succession of Antarctic benthic algae (Potter
Cove, South Shetland Islands): structural patterns and glacial impact over a four-year period.
Polar Biol 41(2):377–396
Cebrián J (2004) Role of first-order consumers in ecosystem carbon flow. Ecol Lett 7(3):232–240
Chapman AS, Fletcher RL (2002) Differential effects of sediments on survival and growth of
Fucus serratus embryos (Fucales, Phaeophyceae). J Phycol 38:894–903
Chung IK, Beardal J, Mehta S et al (2011) Using marine macroalgae for carbon sequestration: a
critical appraisal. J Appl Phycol 23(5):877–886
Clarke GF, Stark JS, Johnston EL, Runcie JW, Goldsworthy PM, Raymond B, Riddle MJ (2013)
Light-driven tipping points in polar ecosystems. Glob Chang Biol 19(12):3749–3761
Cook AJ, Fox AJ, Vaughan DG et al (2005) Retreating glacier fronts on the Antarctic Peninsula
over the past half-century. Science 308(5721):541–544
Cormaci M, Furnari G, Scammacca B, Alongi G (1996) Summer biomass of a population of Iridaea
cordata Iridaea cordata (Gigartinaceae, Rhodophyta) from Antarctica. In: Lindstrom SC,
Chapman DJ (eds) Proceedings of the 15th International Seaweed Symposium. Hydrobiologia
326/327:267–272
Dawson TP, Rounsevell MD, Kluvánková-Oravská T et  al (2010) Dynamic properties of com-
plex adaptive ecosystems: implications for the sustainability of service provision. Biodivers
Conserv 19(10):2843–2853
DeLaca TE, Lipps JH (1976) Shallow-water marine associations, Antarctic Peninsula. Antarct J
11:12–20
Deregibus D, Quartino ML, Campana GL et al (2016) Photosynthetic light requirements and verti-
cal distribution of macroalgae in newly ice-free areas in Potter Cove, South Shetland Islands,
Antarctica. Polar Biol 39(1):153–166
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 169

Drew EA, Hastings RM (1992) A year-round ecophysiological study of Himantothallus grandifo-


lius (Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31:262–277
Duarte CM, Cebrián J (1996) The fate of marine autotrophic production. Limnol Oceanogr
41(8):1758–1766
Duarte P, Ferreira JG (1997) A model for the simulation of macroalgal population dynamics and
productivity. Ecol Model 98:199–214
Duarte P, Meneses R, Hawkins AJS, Zhu M, Fang J, Grant J (2003) Mathematical modelling
to assess the carrying capacity for multi-species culture within coastal waters. Ecol Model
168(1–2):109–143
Ducklow HW, Hansell DA, Morgan JA (2007) Dissolved organic carbon and nitrogen in the
Western Black Sea. Mar Chem 105(1–2):140–150
Duggins DO (1988) The effects of kelp forests on nearshore environments: biomass, detritus,
and altered flow. In: Van Blaricom GR, Estes JA (eds) The community ecology of sea otters.
Ecological studies (analysis and synthesis), vol 65. Springer, Berlin, pp 192–201
Eraso A, Domınguez MA (2007) Physicochemical characteristics of the subglacier discharge in
Potter Cove, King George Island, Antarctica. In: Tyk A, Stefaniak K (eds) Karst and cryokarst,
vol 45. Studies of the Faculty of Earth Sciences, University of Silesia, Sosnowiec, pp 111–122
Eriksson BK, Johansson G (2005) Effects of sedimentation on macroalgae species-specific
responses are related to reproductive traits. Oecologia 143:438–448
Fischer G, Wiencke C (1992) Stable carbon isotope composition, depth distribution and fate of
macroalgae from the Antarctic Peninsula region. Polar Biol 12:341–348
Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic
Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation?. PLoS
One 10(8):e0134440
Gómez I, Wiencke C (1997) Seasonal growth and photosynthetic performance of Antarctic mac-
roalga Desmarestia menziesii (Phaeophyceae) cultivated under fluctuating Antarctic day-
lengths. Bot Acta 110:25–31
Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U, Quartino ML, Dunton K, Wiencke C
(2009) Light and temperature demands of marine benthic micro-algae and seaweeds in the
polar regions. Bot Mar 52:593–608
González PM, Deregibus D, Malanga G et  al (2017) Oxidative balance in macroalgae from
Antarctic waters. Possible role of Fe. J Exp Mar Biol Ecol 486:379–386
Grange LJ, Smith CR (2013) Megafaunal communities in rapidly warming fjords along the West
Antarctic Peninsula: hotspots of abundance and beta diversity. PLoS One 8(12):e77917
Guillaumot C, Aguera A, Danis B, Deregibus D, Quartino ML, Saravia LA (2018) Growth model
of Antarctic macroalgae in a fast-changing environment. Abstract Proceedings of Scar Open
Science Conference, Davos, Switzerland, p 863
Hadley S, Wild-Allen K, Johnson C et al (2015) Modeling macroalgae growth and nutrient dynam-
ics for integrated multi-trophic aquaculture. J Appl Phycol 27(2):901–916
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
Hurd CL (2015) Slow‐flow habitats as refugia for coastal calcifiers from ocean acidification. J
Phycol 51(4):599–605
Hurd CL, Harrison PJ, Bischof K, Lobban CS (eds) (2014) Seaweed ecology and physiology.
Cambridge University Press, Cambridge
Iken K, Quartino ML, Barrera-Oro E, Palermo J, Wiencke C, Brey T (1998) Trophic relations
between marcoalgae and herbivores. Rep Polar Res 299:201–206
Israel A (1995) Determinacion de la produccion primaria en macroalgas marinas. In: Alveal
K, Ferrario ME, Oliveira EC, Sar E (eds) Manual de métodos ficológicos. Universidad de
Concepción, Concepción, pp 795–823
Jerosch K, Scharf FK, Deregibus D et al (2019) Ensemble modelling of Antarctic macroalgal habi-
tats exposed to glacial melt in a polar fjord. Front Ecol Evol 7(207):1–16
Jones CG, Lawton JH, Shachak M (1994) Organisms as ecosystem engineers. Oikos 69:373–386
170 M. L. Quartino et al.

Kirk JT (ed) (1994) Light and photosynthesis in aquatic ecosystems. Cambridge University Press,
Cambridge
Klöser H, Quartino ML, Wiencke C (1996) Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiology 333:1–17
Krause-Jensen D, Duarte CM (2016) Substantial role of macroalgae in marine carbon sequestra-
tion. Nat Geosci 9(10):737–743
Krause-Jensen D, Marbà N, Sanz-Martin M et al (2016) Long photoperiods sustain high pH in
Arctic kelp forests. Sci Adv 2(12):e1501938
Krause-Jensen D, Lavery P, Serrano O, Marbà N, Masque P, Duarte CM (2018) Sequestration of
macroalgal carbon: the elephant in the blue carbon room. Biol Lett 14:20180236
Lagger C, Nime M, Torre L et al (2018) Climate change, glacier retreat and a new ice-free island
offer new insights on Antarctic benthic responses. Ecography 40:1–12
Mann KH (1972) Ecological energetics of the seaweeds zone in a marine bay on the Atlantic coast
of Canada. II productivity of seaweeds. Mar Biol 14:199–209
Marina TI, Salinas V, Cordone G et al (2018) The food web of Potter Cove (Antarctica): complex-
ity, structure and function. Estuar Coast Shelf Sci 200:141–151
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48
Mincks SL, Smith CR, DeMaster DJ (2005) Persistence of labile organic matter and microbial
biomass in Antarctic shelf sediments: evidence of a sediment food bank. Mar Ecol Prog Ser
300:3–19
Nedzarek A, Rakusa-Suszczewski S (2004) Decomposition of macroalgae and the release of nutri-
ent Admiralty Bay, King George, Antarctica. Polar Biosci 17:26–35
Pedersen EJ, Miller DL, Simpson GL, Ross N (2019) Hierarchical generalized additive models in
ecology: an introduction with mgcv. Peer J 7:e6876
Perrot T, Rossi N, Ménesguen A et al (2014) Modelling green macroalgal blooms on the coasts of
Brittany, France to enhance water quality management. J Mar Syst 132:38–53
Quartino ML, Boraso de Zaixso AL (2008) Summer macroalgal biomass in Potter Cove, South
Shetland Islands, Antarctica: its production and flux to the ecosystem. Polar Biol 31:281–294
Quartino ML, Klöser H, Schloss IR et al (2001) Biomass and associations of benthic marine mac-
roalgae from the inner Potter Cove (King George Island, Antarctica) related to depth and sub-
strate. Polar Biol 24(5):349–355
Quartino ML, Zaixso HE, Boraso de Zaixso AL (2005) Biological and environmental characteriza-
tion of marine macroalgal assemblages in Potter Cove, South Shetland Islands, Antarctica. Bot
Mar 48:187–197
Quartino ML, Boraso de Zaixso AL, Momo FR (2008) Macroalgal production and the energy
cycle of Potter Cove. In: Wiencke C, Ferreyra GA, Abele D, Marenssi S (eds) The Antarctic
ecosystem of Potter Cove, King-George Island (Isla 25 de Mayo). Synopsis of research per-
formed 1999–2006 at the Dallmann Laboratory and Jubany Station, p 68–74
Quartino ML, Deregibus D, Campana GL et  al (2013) Evidence of macroalgal colonization
on newly ice-free areas following glacial retreat in Potter Cove (South Shetland Islands),
Antarctica. PLoS One 8:e58223
Queirós AM, Stephens N, Widdicombe S et al (2019) Connected macroalgal‐sediment systems:
blue carbon and food webs in the deep coastal ocean. Ecol Monogr 89(3):e01366
Rakusa-Suszczewski S (1995) The hydrography of Admiralty Bay and its inlets, coves and lagoons
(King George Island, Antarctica). Pol Polar Res 16:61–70
Reed DC, Rassweiler A, Arkema KK (2008) Biomass rather than growth rate determines variation
in net primary production by giant kelp. Ecology 89(9):2493–2505
Ren JS, Barr NG, Scheuer K et al (2014) A dynamic growth model of macroalgae: application in
an estuary recovering from treated wastewater and earthquake-driven eutrophication. Estuar
Coast Shelf Sci 148:59–69
Richardson MG (1979) The distribution of Antarctic marine macro-algae related to depth and
substrate. Br Antarct Surv Bull 49:1–13
8  Production and Biomass of Seaweeds in Newly Ice-Free Areas: Implications… 171

Rossi F, Gribsholt B, Gazeau F et al (2013) Complex effects of ecosystem engineer loss on benthic
ecosystem response to detrital macroalgae. PLoS One 8(6):e66650
Rückamp M, Braun M, Suckro S et al (2011) Observed glacial changes on the King George Island
ice cap, Antarctica, in the last decade. Glob Planet Change 79:99–109
Sahade R, Tatián M, Esnal GB (2004) Reproductive ecology of the ascidian Cnemidocarpa ver-
rucosa Cnemidocarpa verrucosa at Potter Cove, South Shetland Islands, Antarctica. Mar Ecol
Prog Ser 272:131–140
Sahade R, Lagger C, Torre L et  al (2015) Climate change and glacier retreat drive shifts in an
Antarctic benthic ecosystem. Sci Adv 1:e1500050
Schloss IR, Ferreyra GA (2002) Primary production, light and vertical mixing in Potter Cove, a
shallow bay in the maritime Antarctic. Polar Biol 25(1):41–48
Schloss IR, Abele D, Moreau S et  al (2012) Response of phytoplankton dynamics to 19-year
(1991–2009) climate trends in Potter Cove (Antarctica). J Mar Syst 92:53–66
Steneck RS, Graham MH, Bourque BJ et al (2002) Kelp forest ecosystems: biodiversity, stability,
resilience and future. Environ Conserv 29(4):436–459
Tatián M, Sahade M, Esnal GB (2004) Diet components in the food of Antarctic ascidians living at
low levels of primary production. Antarct Sci 16:123–128
Valdivia N, Díaz M, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all around:
scale-dependent spatial variation in rocky coastal communities of Fildes Peninsula, King
George Island, Antarctica. PLoS One 9(6):e100714
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental stress gradients in a marine macroalgal-dominated subtidal community on the
Western Antarctic Peninsula. PLoS One 10(9):e0138582. 
Wada S, Hama T (2013) The contribution of macroalgae to the coastal dissolved organic matter
pool. Estuar Coast Shelf Sci 129:77–85
Wahl M, Schneider Covachã S, Saderne V et al (2018) Macroalgae may mitigate ocean acidifica-
tion effects on mussel calcification by increasing pH and its fluctuations. Limnol Oceanogr
63(1):3–21
Westermeier R, Gómez I, Rivera PJ, Müller DG (1992) Macroalgas marinas antárticas: distribu-
ción, abundancia y necromasa en Isla Rey Jorge, Shetland del Sur, Antártica. Ser Cient INACH
42:21–34
Wiencke C (1990a) Seasonality of brown macroalgae from Antarctica a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600
Wiencke C (1990b) Seasonality of red and green macroalgae from Antarctica—long-term culture
study under fluctuating Antarctic daylengths. Polar Biol 10:601–607
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Seaweed biol-
ogy, novel insights into ecophysiology, ecology and utilization. Springer, Berlin, Heidelberg,
pp 265–291
Wiencke C, Gómez I, Dunton K (2009) Phenology and seasonal physiological performance of
polar seaweeds. Bot Mar 52:585–592
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: De Broyer C, Koubbi P,
Griffiths HJ, Ramond B, Udekem d’Acoz C, Van de Putte A, Danis B, David B, Grant S, Gutt
J et  al (eds) Biogeographic atlas of the Southern Ocean. Scientific Committee on Antarctic
Research, Cambridge, pp 66–73
Wulff A, Iken K, Quartino ML et al (2009) Biodiversity, biogeography and zonation of marine
benthic micro- and macroalgae in the Arctic and Antarctic. Bot Mar 52(6):491–507
Zenteno L, Cárdenas L, Valdivia N et al (2019) Unraveling the multiple bottom-up supplies of an
Antarctic nearshore benthic community. Prog Oceanogr 174:55–63
Zhang J, Wu W, Ren JS, Lin F (2016) A model for the growth of mariculture kelp Saccharina
japonica in Sanggou Bay, China. Aquac Environ Interact 8:273–283
Chapter 9
Carbon Balance Under a Changing Light
Environment

Dolores Deregibus, Katharina Zacher, Inka Bartsch, Gabriela L. Campana,


Fernando R. Momo, Christian Wiencke, Iván Gómez, and María L. Quartino

Abstract The natural environment of Antarctic seaweeds is characterized by


changing seasonal light conditions. The ability to adapt to this light regime is one of
the most important prerequisites for their ecological success. Thus, the persistence
of seaweeds depends on their capacity to maintain a positive carbon balance (CB)

D. Deregibus (*)
Departamento de Biología Costera, Instituto Antártico Argentino, Buenos Aires, Argentina
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), Buenos Aires,
Argentina
e-mail: dderegibus@dna.gov.ar
K. Zacher · I. Bartsch · C. Wiencke
Alfred-Wegener-Institute, Helmholtz Centre for Polar and Marine Research,
Bremerhaven, Germany
e-mail: katharina.zacher@awi.de; inka.bartsch@awi.de; Christian.wiencke@awi.de
G. L. Campana
Departamento de Biología Costera, Instituto Antártico Argentino, Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: gcampana@dna.gov.ar
F. R. Momo
Instituto de Ciencias, Universidad de General Sarmiento, Los Polvorines, Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: fmomo@campus.ungs.edu.ar
I. Gómez
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: igomezo@uach.cl
M. L. Quartino
Departamento de Biología Costera, Instituto Antártico Argentino, Buenos Aires, Argentina
Museo Argentino de Ciencias Naturales “B. Rivadavia”, Buenos Aires, Argentina
e-mail: lquartino@dna.gov.ar

© Springer Nature Switzerland AG 2020 173


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_9
174 D. Deregibus et al.

for buildup of biomass over the course of the year. A positive CB in Antarctica
occurs only during the ice-free period in spring and summer, when p­ hotosynthetically
active radiation (PAR, 400–700 nm) penetrates deeply into the water column. The
accumulated carbon compounds during this period are stored and remobilized to
support metabolism for the rest of the year.
Over the last decades climate warming has induced a severe glacial retreat in
Antarctica and has opened newly ice-free areas. Increased sediment runoff, and
reduced light penetration due to melting during the warmer months, may lead to a
negative CB with changes in the vertical distribution of seaweeds. Furthermore,
warmer winters and springs result in earlier sea-ice melt, causing an abrupt increase
in light, compensating the reduction in PAR in summer or increasing the annual
light budget. Studies performed in Potter Cove, Isla 25 de Mayo/King George
Island, reveal that algae growing in newly ice-free areas did not acclimate to the
changing light conditions. Lower or even negative CB values in areas close to the
glacier runoff seem to be primarily dependent on the incoming PAR that finally
determines the lower distribution limit of seaweeds. The present chapter discusses
how carbon balance respond to the changing Antarctic light environment and its
potential implications for the fate of benthic algal communities.

Keywords  Metabolic carbon balance · Glacier runoff · Light requirements ·


Photosynthesis · Photosynthetic acclimation · Turbidity

9.1  Introduction

9.1.1  Climate Change in the Antarctic Coastal Zone

Rapid regional warming will continue to be one of the major forcing elements in
Antarctica during this century (Hendry et al. 2018; IPCC 2014, 2019). The Western
Antarctic Peninsula (WAP) is a hotspot of climate warming evidenced by a rapid
increase in air temperature and, as a consequence, strong sea-ice decline and fast
glacial retreat (Cook et al. 2005; Meredith and King 2005; Stammerjohn et al. 2008;
Turner et al. 2009). This region is also highly impacted by ice disturbance, and the
rates of iceberg scour have been shown to increase with consequences still not well
understood for the benthic life (Barnes and Souster 2011; Barnes et al. 2014). Due
to a high density of research stations, accessibility, and mild climate, the WAP
coastal areas are natural laboratories for studying how ecosystems respond to rapid
climate change (Constable et al. 2014; Deregibus et al. 2017; Lagger et al. 2018). In
fact, a high diversity of benthic assemblages concentrate in the upper 70 m zone,
which is subject to diverse physical perturbations.
9  Carbon Balance Under a Changing Light Environment 175

A notorious phenomenon in coastal zones of the WAP is the formation of “newly


ice-free areas” as a consequence of glacier retreat (Cook et al. 2005; Rückamp et al.
2011; Quartino et al. 2013). In these areas new substrates become available for ben-
thic colonization (Constable et al. 2014; Lagger et al. 2017, 2018). In parallel, the
enhanced glacier melting leads to increased turbidity and decreases salinity in the
water column (Barnes and Conlan 2007; Campana et al. 2009; Schloss et al. 2012;
Bers et al. 2013; Grange and Smith 2013). Moreover, ice-scouring events can be
intensified (Barnes and Souster 2011; Barnes et  al. 2018). These climate-driven
rapid shifts and concomitant new stressors (e.g., sedimentation, ice disturbance) can
have complex consequences for the functioning of the coastal ecosystems (Gutt
et al. 2015). Thus, various hypotheses have arisen inspiring numerous studies on
their effects on coastal ecosystems (Smale and Barnes 2008; Gutt et  al. 2011;
Schofield et al. 2010; Torre et al. 2012; Ducklow et al. 2013; Barnes et al. 2014;
Moon et al. 2015; Sahade et al. 2015; Valdivia et al. 2015). The question whether
rising air temperatures could cause changes in the light environment in Antarctic
coastal zones and their effect on primary producers is central to understand the
future benthic primary production in these coastal ecosystems (Zacher et al. 2009;
Clark et al. 2013, 2017). In this context, given that photosynthesis in coastal areas is
the most important biological process affected by light and due to its vital impor-
tance in neritic areas, seaweeds can be considered as sentinel taxa to evaluate these
changes.

9.1.2  A
 ntarctic Seaweeds and the Changing Light
Environment

The availability of light for photosynthesis and growth is the major factor governing
depth distribution of seaweeds (Lüning 1990). At the southern distribution limit of
seaweeds in the Antarctic, the polar night lasts for about 4 months. Sea-ice cover
extends the period of hibernal darkness even more (reviewed in Zacher et al. 2009).
Besides phytoplankton blooms, circulation and wind have major importance in lim-
iting light penetration into the water column (Schloss et al. 2012). Thus, Antarctic
seaweeds are well adapted to cope with extended periods of darkness showing low
light requirements for photosynthesis (Ek) and high photosynthetic efficiencies (α).
In addition, a number of species have also the ability to cope with high light condi-
tions in summer (Wiencke 1990; Gómez et al. 1997, 2009; Wiencke and Amsler
2012; see also Chap. 7 by Gómez and Huovinen).
Seasonality strongly determines the fluctuations in abiotic factors, which affect
the physiological and ecological performance of seaweeds (Wiencke and Amsler
2012; Marcías et al. 2017). In Antarctica, adaptation to the seasonality of the light
regime is a fundamental prerequisite for the ecological success of seaweeds
(Wiencke et al. 2011). There are two different growth strategies, i.e., season antici-
pators and season responders sensu Kain (1989). The season anticipators start grow-
176 D. Deregibus et al.

ing in late winter/spring. In contrast, the season responders start growth and
reproduction later in spring and summer when the light conditions are favorable
(Wiencke and Amsler 2012; see also Chap. 10 by Navarro et al.). However, both
groups are highly dependent on the light availability in spring and summer when
primary production mostly occurs as the solar angle is high and fast ice is absent
(Miller and Pearse 1991; Wiencke et al. 2006; Runcie and Riddle 2012; see also
Chap. 7 by Huovinen and Gómez).
Changes in light availability do not only have natural causes; besides the high
seasonal variability in Antarctica, changes in external variables that influence sea-
weeds are due to anthropogenic reasons (e.g., higher regional air temperatures due
to the CO2 emissions) (Vaughan et al. 2003; Turner et al. 2013). Especially in coastal
areas, climate change may significantly shrink the annual light budget available for
benthic primary producers due to an enhancement of sedimentation and a decrease
in the duration of the fast ice season (Clark et al. 2013; Quartino et al. 2013). Algae
growing in newly ice-free areas are subjected to a reduction of light penetration,
which constitutes a constraint for photosynthesis (Schloss et al. 2012; Wiencke and
Amsler 2012; Deregibus et  al. 2016; González et  al. 2017; see also Chap. 8 by
Quartino et al.). Furthermore, glacier melting also causes accumulation of sediment
on the seafloor, which could affect the attachment of benthic algae (Johnston et al.
2007). On the other hand, beneficial effects of sediment input could be considered
during spring and summer by attenuating high PAR, which may inhibit the recruit-
ment of macroalgal species (Graham 1996; Hanelt 1996; Hanelt et al. 1997) or may
protect against harmful UV radiation (Roleda et al. 2009). However, previous stud-
ies have shown a positive relationship between the degree of light penetration and
the complexity of the macroalgal community (in terms of diversity and the presence
of large perennial species) in the newly ice-free areas (Quartino et al. 2013). Thus,
the time of occurrence of the thawing is extremely relevant for seaweeds as favor-
able light conditions for algal growth are constrained to only a few months of the
year during the bright light season (Wiencke and Amsler 2012; see also Chap. 7 by
Huovinen and Gómez). Here, the maintenance of positive carbon balance is proba-
bly one of the most significant physiological adjustments of seaweeds to cope with
these changing light conditions (Gómez et al. 1997; Deregibus et al. 2016).

9.1.3  C
 arbon Balance: Concepts and Methodological
Challenges

The production of organic matter via photosynthesis using carbon dioxide, water,
and sunlight is known as primary production, while primary productivity is the rate
at which energy is converted to organic substances by photosynthetic organisms
(Hurd et al. 2014). The rates of primary productivity refer to the efficiency at which
solar energy is used to fix inorganic carbon and create biomass, and hence it is an
essential parameter reflecting the ecosystem function (Falkowski and Raven 1997).
9  Carbon Balance Under a Changing Light Environment 177

The carbon balance (CB) equalizes the assimilated carbon (C) during photosyn-
thesis in relation to the C that is lost due to respiration (Wiencke and Amsler 2012).
Photosynthesis versus irradiance (P-E) curves describes how the photosynthetic rate
(e.g., net O2 evolution) varies with increasing irradiance. Integrating data on daily
changes of in situ irradiance with the P-E-derived parameters, Ek, the ­photosynthetic
capacity (Pmax), and dark respiration (Deregibus et al. 2016), it is possible to calcu-
late the daily metabolic CB as an indicator of the physiological ability to live at a
certain depth (Hanelt and Figueroa 2012). In coastal areas not affected by glacial
melt, PAR can penetrate into the water column as deep as 30 m (1% depth) during
late winter-spring, still allowing a positive CB (Gómez et al. 1997; see also Chap. 7
by Huovinen and Gómez).
The lower distribution limit of algae is determined by their capacity to main-
tain a positive CB to build up biomass (Hanelt and Figueroa 2012). During this
period a positive CB replenishes the energy budget to be used to survive the long
periods of darkness during the rest of the year (Wiencke et al. 2011; Deregibus
et al. 2016). Several studies indicate that the daily exposure time to light is more
important than the intensity of light for macroalgal productivity in coastal areas
(Dennison and Alberte 1985; Matta and Chapman 1991; Gómez et  al. 1997).
Direct associations between increases in turbidity and decreases in macroalgal
productivity have been reported from a variety of systems worldwide (Airoldi
2003; Anthony et  al. 2004; Spurkland and Iken 2011; Pritchard et  al. 2013).
Similarly, changes in the productivity of seaweeds in relation to sea ice variations
have been reported in Antarctic (Clark et al. 2013) and Arctic (Krause-Jensen and
Duarte 2014) assemblages. Apparently, direct and indirect effects of climate
change on light availability will have dramatic effects on the annual macroalgal
CB, which could result in changes in benthic primary productivity (Runcie and
Riddle 2012; Bartsch et al. 2016; Deregibus et al. 2016; Gómez et al. 2019; see
also Chap. 8 by Quartino et al.).
Various mathematical models have related irradiance to photosynthesis in order
to estimate primary productivity (Jassby and Platt 1976; Nelson and Siegrist 1987;
Henley 1993; Jones et al. 2014). However, the photosynthetic parameters α, Ek, and
Pmax derived from P-E curves may differ depending on the selected fitting model
(Smith 1936; Steele 1962; Webb et al. 1974; Jassby and Platt 1976; Cullen 1990;
Frenette et al. 1993; Henley 1993). Thus, productivity estimates using distinct mod-
els may not be comparable and even lead to erroneous conclusions (Frenette et al.
1993; Deregibus et  al. 2016). Therefore, the use of best mathematical fit will
improve considerably the quality of our productivity estimations (Jassby and Platt
1976; Nelson and Siegrist 1987).
The primary productivity of seaweeds can be estimated in several ways, e.g.,
through changes in net weight, through growth, in situ productivity measurements,
or measurements of dissolved oxygen production (Gómez et al. 2009; Runcie et al.
2009; Runcie and Riddle 2011; Hanelt and Figueroa 2012). These different method-
ologies show constraints in their ability to accurately describe the primary produc-
178 D. Deregibus et al.

tivity in marine environments (Runcie and Riddle 2012). In the WAP, Gómez et al.
(1997) did an initial effort to calculate the seaweed primary productivity through the
calculation of the daily metabolic CB using the amount of in situ daily hours when
algae are light saturated. Their findings could explain well the vertical zonation of
seaweeds according to the light regime at the WAP (Gómez et al. 1997).

9.2  Carbon Balance: A Case Study in Potter Cove

Over the last years, numerous studies have been performed in Potter Cove, Isla 25
de Mayo/King George Island, which is a distinctive example of an area with a spe-
cial focus on studies related to the impacts of climate-forced glacier retreat across
an entire shallow water ecosystem.
Adjacent to Carlini Station, Potter Cove is surrounded by the Fourcade Glacier.
Over the last decades, this glacier has been retracting and melt water inflow in the
cove increased (Eraso and Dominguez 2007; Rückamp et al. 2011) (Fig. 9.1). Newly
ice-free areas in costal shallow areas were formed and are still appearing (Quartino
et al. 2013; Lagger et al. 2017, per obs). In this cove, increase in sedimentation has
negatively affected species and changed benthic communities (Torre et  al. 2012;
Quartino et al. 2013; Pasotti et al. 2014; Sahade et al. 2015; Campana et al. 2018;
Meredith et al. 2018; Jerosch et al. 2019; see Chap. 8 by Quartino et al.). As a con-
sequence of glacier melting, a spatial gradient developed along the cove, providing
the opportunity to analyze the primary productivity under a natural abiotic gradient
(Deregibus et  al. 2016). For the first time, the primary productivity of seaweeds
growing in these new areas in a gradient of glacial influence was estimated. High
“glacier influence” was defined as a decrease in light penetration in the water col-

Fig. 9.1.  Small creek originated by glacier melting during summer months causing a discharge of
turbid fresh water filled with terrigenous sediments into the inner zone of Potter Cove. (Adapted
from Deregibus (2017) with permission (photos by Dolores Deregibus))
9  Carbon Balance Under a Changing Light Environment 179

umn (increased turbidity) due to sediment input. Improvements on the calculations


were proposed through a novel and more accurate approximation of seaweed CB
calculations with the inclusion of in situ continuous-light measurements (Deregibus
et al. 2016).

9.2.1  Light Availability

The major new achievement and challenge was to install and perform continuous
underwater PAR measurements along a depth gradient (5, 10, 20, and 30 m) in sev-
eral areas along the turbidity gradient in summer 2010, for the first time in an
Antarctic environment. Since then, daily continuous PAR measurements are
obtained at 10  m depth and are utilized in combination with the photosynthetic
parameters to estimate the CB of seaweeds over the seasons and along turbidity
gradients (Deregibus 2017).
The light loggers (for details see Deregibus et  al. 2016) were installed in an
upright position, on a specially constructed concrete base and are periodically
secured by SCUBA divers at the respective sites (Fig. 9.2). The PAR data could then
be used to plot irradiance vs time with a time interval of 15 min (Fig. 9.3).
As an example, in Fig. 9.3, we present the mean PAR records of one selected
week in January 2010. At 5 m depth, seaweeds were exposed to PAR for 12 h (mean
daily irradiance of 22.2 μmol photons m−2 s−1). At 10 m depth, PAR values showed
an abrupt decrease indicating a very low daily light budget below 5 m in summer in
this area. Also, the daily exposure time to sunlight PAR decreased with increasing
depths shortening the photoperiod (Fig. 9.3).
Despite high solar angle, the question rises how the investigated seaweeds may
survive and grow under these conditions in Potter Cove. It became evident that it is
insufficient to consider not only summer periods to assess productivity but also
annual irradiance (Fig. 9.4), and carbon budgets need to be calculated. Accordingly,

Fig. 9.2  Underwater PAR measurements: (a) Odyssey Logger Sensor for continuous PAR mea-
surements. Setup (concrete base and adjustable arm) that holds the sensor. (b) Diver exchanging
the sensor. (Adapted from Deregibus (2017) (photos by Dolores Deregibus))
180 D. Deregibus et al.

Fig. 9.3  Daily course of

Irradiance (µmol photons m-2 s-1)


100
irradiance in newly ice-free 5m
area in Potter Cove at 5 10 m
and 10 m depth between
Jan 15 and 22, 2010 (Data
were averaged over seven
50
continuous days). (Adapted
from Deregibus et al.
(2016))

0
0:00 4:00 8:00 12:00 16:00 20:00 0:00
Time of the day

Fig. 9.4.  Annual course of incident irradiance measured at Potter Cove, Isla 25 de Mayo/King
George Island. Daily mean of PAR was integrated over single months in a newly ice-free area dur-
ing 2015

whole-year continuous PAR measurements are being performed to estimate the


light climate over an entire year – and between years in Potter Cove (e.g. Fig. 9.4).
During 2015, which serves as an example here, there was an evident presence of
fast ice in winter leading to very low mean values of daily PAR, followed by an
abrupt increase of the PAR intensity once the pack-ice broke up in spring (Fig. 9.4).
Whole-year underwater in situ irradiances show that the most productive season
may start as early as October with highest mean daily PAR values. This is in con-
trast to our expectations as during the warmer summer months (December to
March), the light intensity was lower than in spring, mainly due to the input of sedi-
ment of terrestrial origin into the water column. On an interannual scale, these val-
ues may differ considerably (Deregibus unpublished). Similar results were found
for the Arctic in Kongsfjorden (Spitsbergen), a fjord system highly affected by melt-
ing glaciers (Bartsch et al. 2016; Pavlov et al. 2019). Overall, our long-term, year-­
round PAR measurements can be further used to estimate the annual primary
productivity of Antarctic seaweeds in coastal areas subjected to extreme changes in
light availability.
9  Carbon Balance Under a Changing Light Environment 181

9.2.2  Photosynthetic Acclimation

The endemic brown alga Himantothallus grandifolius (A.  Gepp and E.  S. Gepp)
Zinova and the red alga Palmaria decipiens (Reinsch) Ricker, presented here as
examples, were sampled at 5 and 10  m in the vicinity of the light logger.
Photosynthetic oxygen evolution and dark respiration were measured under labora-
tory conditions, and photosynthesis versus irradiance curves (P-E curves) were
obtained and used to calculate photosynthetic parameters with the hyperbolic tan-
gent function of Jassby and Platt (1976). This function, which was proved to be the
best fitting model to the experimental data, is expressed as:

P = Pmax ∗ tanh (α E / Pmax ) + R



where P is the photosynthetic rate, Pmax is the maximum photosynthetic rate, tanh is
the hyperbolic tangent, α is the initial slope of the curve at low irradiance, E is the
incident irradiance, and R is the dark respiration rate (Fig. 9.5).
The hyperbolic tangent function (Jassby and Platt 1976) was found to be the best
fit for our data compared to other commonly used models (e.g., Webb et al. 1974).
Jassby and Platt (1976) found that their hyperbolic tangent equation proved to be
the best overall fit to 200 types of datasets in a comparison of various equations
(Jassby and Platt 1976; Jones et al. 2014).
In Potter Cove, photosynthetic parameters such as Pmax, α, and Ek of seaweeds
growing in newly ice-free areas with variable glacial influence and at different
depths were generally quite similar. This indicates a low acclimation potential of
photosynthesis to different irradiance regimes (Deregibus et al. 2016). The absence
of photoacclimation in Antarctic seaweeds living under different? light availability
has been reported previously (Gómez et al. 2009). Apparently, this is a characteristic
that forms part of suite of photobiological features conferring the extreme shade
adaptation of these organisms, developed primarily to cope with dim light in
autumn-winter (Gómez and Huovinen 2015). In fact, similar values of Pmax and α
were reported at different depths for H. grandifolius (Drew and Hastings 1992;

Fig. 9.5.  Photosynthesis-irradiance (P-E) curves of Palmaria decipiens (a) and Himantothallus
grandifolius (b) in a newly ice-free area in Potter Cove, which represents a typical P-E curve for a
shade-adapted species. Generally, the oxygen production under light saturation is lower in H. gran-
difolius than in P. decipiens. Both species did not show any sign of photoinhibition even under
highest irradiances of 800 μmol photons m−2 s−1. (Adapted from Deregibus et al. (2016))
182 D. Deregibus et al.

Gómez et al. 1997), and in P. decipiens, Gigartina skottsbergii Setchell and Gardner
Trematocarpus antarcticus (Hariot) Frederic and Moe, and Desmarestia anceps
Montagne (Gómez et al. 1997). Interestingly, other studies, carried out in the Ross
Sea, Antarctica (Schwarz et al. 2003), in Greenland (Kühl et al. 2001), and in the
Arctic (Krüger 2016) using chlorophyll fluorescence, suggest a greater acclimation
potential of seaweeds to low light at deeper depths as reflected in their lower Ec
(photosynthetic compensation point), Ek, and rETRmax values. Furthermore, a recent
study reports different photosynthetic response to low light between algae growing
in Antarctica and in the Subantarctic (Navarro et al. 2019).

9.2.3  Daily Carbon Balance of Seaweeds

Calculations of daily net CB (mg C g−1 FW day−1) integrated the photosynthetic


parameters derived from the P-E curves and the incident underwater irradiances
according to the following formula:

  E1 
  
E
P = Pmax ∗ tanh  α  average 2  + R
  Pmax 
  
  

Fig. 9.6  Example of the daily metabolic carbon balance of two selected Antarctic seaweeds grow-
ing at different depths in a highly turbid newly ice-free area in Potter Cove during summer 2010.
Values correspond to an overall net gain or loss of C during 24 h. (Adapted after Deregibus et al.
(2016))
9  Carbon Balance Under a Changing Light Environment 183

where “average (E1:E2)” is the average of two incident irradiances between time 1
and time 2. This formula improved the calculation of the net oxygen production
along an entire day (Fig. 9.6) and shows that in an area with high turbidity due to
sedimentation, the CB was only positive for P. decipiens at 5 m depth with a mean
value of 0.2 (±0.4) mg C g−1 FW day−1, but negative at 10 m. In the case H. grandi-
folius, it was negative for both depths (Fig. 9.6). Negative values were mainly due
to the low light availability and a reduction of the time to which algae were exposed
to light (Deregibus et al. 2016). The results agree with the in situ distribution of both
species: they grow only to a maximum depth of 10 m in this highly disturbed area
(Deregibus et al. 2016; Deregibus 2017).

9.3  N
 ew Scenarios and Their Implications for Algal
Photosynthesis

Carbon balance is the most conclusive parameter to understand and explain the
zonation patterns in Antarctic seaweeds (Gómez et  al. 2009). This parameter is
directly related to light availability as it is lower or negative in more turbid areas
and/or at deeper depths compared to shallower depths and areas with low turbidity.
Isla 25 de Mayo/King George Island is characterized by a marked increase in the
atmospheric temperature in recent decades (Ferron et al. 2004; Schloss et al. 2012;
Bers et  al. 2013). Falk and Sala (2016) indicated an anticipation of the thawing
period towards spring and an extension towards autumn in this area. The latter, and
the fact that turbidity has increased not only in summer but also in spring in Potter
Cove over the last years (Schloss et al. 2012; Deregibus et al. 2016), and that in
some cases negative CB have also been measured in spring (Deregibus 2017) raises
additional questions. For example: What would happen if the thawing season starts
earlier due to increased air temperatures, leading also to negative CB in spring?
Would the storage compounds produced in summer be enough to support the energy
requirements the rest of the year?
Conversely, it should also be considered that warmer winters and springs lead to
earlier fast ice melting (Schloss et al. 2012; Deregibus et al. 2017), which causes an
abrupt increase of light, probably compensating the reduction of PAR in summer or
even significantly increasing the annual light budget for seaweeds (Johnston et al.
2007; Clark et al. 2013). This last assumption leads to additional questions: Will the
total CB values become really lower under a future warming? Or will there be com-
pensation due to a higher light availability in late winter and early spring (less or no
fast ice cover)?
If the annual photon doses do not sustain growth and reproduction at a certain
area, seaweeds will not survive (Runcie and Riddle 2012). Furthermore, light not
only serves as a source of energy for seaweeds, but it is also an environmental signal
inducing changes in processes which are dependent on daylength as trigger signal
184 D. Deregibus et al.

Climate Change and Underwater Light Variations

Fast ice
Temperature
duration and
increase
extension decrease

Glacier Turbidity Underwater Seaweed


retreat increase Irradiance
Community

Changes in wind
directions and
intensities

Fig. 9.7  Conceptual model on the impact of direct and indirect climate change factors on the
underwater irradiance in newly ice-free areas at the Western Antarctic Peninsula. (Adapted from
Quartino et al. (2013) and Deregibus (2017))

(Hanelt and Figueroa 2012). The seasonal development of seaweeds must be tuned
to the strong seasonality of the light conditions (Wiencke et al. 2011). This is very
important as variations in PAR would not only affect the primary energy source but
could also alter the information to modulate the morphogenetic development and
the other signals for growth and reproduction. It is key and maybe relevant to eluci-
date the growth seasons that are being more/most? affected by global warming
­during the year, as higher temperatures in winter impact differently on the light
availability than in spring and summer (Wiencke and Amsler 2012).
The conceptual model in Fig 9.7 describes the ways through which these envi-
ronmental changes could be affecting the light availability for the benthic autotro-
phic organisms. Environmental conditions directly affect, and are reflected, in the
underwater annual PAR budget. Light changes are mediated by a series of factors:
firstly, glacier melting increases the amount of sediments in the water column and
enhances turbidity. This factor has most impact on benthic primary producers as it
directly and negatively reduces light availability, which is a constraint for photosyn-
thesis. Secondly, the decrease in fast ice duration may increase the open water
period and thereby the availability of underwater light in winter and spring. Thirdly,
wind, depending on its intensity and direction, has a significant impact on sediment
resuspension and distribution processes of the sediments limiting light penetration.
This model is complementary to the general conceptual model shown in Chap. 8 by
Quartino et  al., on the environmental factors affecting the coastal macroalgal
community.
9  Carbon Balance Under a Changing Light Environment 185

9.4  Concluding Remarks and Future Prospects

So far, glacial retreat has opened new space (hard substratum) in the inner Potter
Cove and seaweeds have colonized and persisted in newly ice-free areas (Quartino
et al. 2013; Campana et al. 2018). Although newly ice-free areas are highly affected
by the glacial influence, several seaweed species grow under high sedimentation
showing their exceptional ability to survive and successfully reproduce under such
conditions (Becker et al. 2011; Quartino et al. 2013; Deregibus et al. 2016).
The realization that seaweeds are the most productive marine macrophytes and
identified as very relevant contributors to global blue carbon sequestration (Hill
et al. 2011; Krause-Jensen and Duarte 2016) has consigned them in a key role at a
global scale. Considering that the spatial distribution of the seaweed community has
expanded to the inner side of Potter Cove, it is likely that the increase in seaweed
biomass leads to an enhanced production in this area (Quartino et al. 2013), with
cascading effects to the rest of the food web (Marina et al. 2018). This raises the
question of how changes in seaweed productivity could affect the rest of the coastal
ecosystem in Potter Cove. In this context, given that seaweeds in Potter Cove are in
a constant tradeoff between extending their distribution into newly ice-free areas
and being affected by climate change, an interest to reveal whether this expansion
will give way to a persistence and maturation of these communities in these new
areas exists.
Due to climate change in polar ecosystems, shifts from predominantly heterotro-
phic to autotrophic states of shallow polar seabeds have been predicted (Bartsch
et al. 2016; Clark et al. 2013, 2017). Kortsch et al. (2012) and Scherrer et al. (2018)
reported marked community shifts with abrupt and persistent increase in macroalgal
cover in the Arctic following the extension in the ice-free period. Krause-Jensen
et  al. (2012) also reported substantial increases in the productivity and maximal
depth distribution of kelps in Greenland. Furthermore, in the Arctic it was also
stated that longer ice-free seasons have extended the growth season of seaweeds
favoring an increased diversity of algae and macrozoobenthos (Paar et al. 2016). In
this context, it is also expected that further glacier retreat will continue to favor
seaweed colonization in new coastal areas potentially resulting in higher productiv-
ity and carbon sequestration (Clark et al. 2013; Krause-Jensen and Duarte 2016).
The future projection in case study systems such as Potter Cove would be to
continue with the long-term monitoring of multiple abiotic factors and primary pro-
ductivity calculations. Knowledge on the minimum light requirements for seaweeds
and on their ecophysiological characteristics are needed to better understand and
predict macroalgal survival and possible changes in primary productivity, distribu-
tion, and depth zonation in areas affected by glacial melting due to climate change.
Continuous PAR measurements are currently being performed in the newly ice-free
areas of Potter Cove, which is a unique data set in terms of time span (more than five
years) and continuity in a gradient of glacial influence in Antarctica.
186 D. Deregibus et al.

Finally, as a way to understanding the regional variability and its impacts on the
biota, these monitoring programs require international collaboration including inte-
grated actions among different Antarctic stations (Deregibus et al. 2017).

Acknowledgments  The fieldwork described in this chapter has been performed at Carlini Station-­
Dallmann Laboratory within the framework of the scientific collaboration existing between
Instituto Antártico Argentino/Dirección Nacional del Antártico and the Alfred-Wegener-Institute,
Helmholtz Centre for Polar and Marine Research (AWI). The research was supported by Grants
from DNA-IAA (PICTA 7/2008–2011) and ANPCyT-DNA (PICTO 0116/2012–2015, PICT
2017-­
­ 2691). These studies were also supported by MINCYT-BMBF Program (AL/17/06-
01DN18024) and Conicyt-Chile (Center FONDAP IDEAL 15150003). We are especially grateful
to the scientific, logistic, and diving groups of Carlini Station. We thank Carolina Matula for her
contribution to improve the annual light figure. We gratefully acknowledge financial support by the
AWI. We thank Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, for allow-
ing the reuse of material published in Deregibus 2017 Thesis. The present research also presents
an outcome of the EU project IMCONet (FP7 IRSES, Action No. 319718) and EU project
CoastCarb, Funding ID 872690, H2020-MSCA-RISE-2019  – Research and Innovation Staff
Exchange.

References

Airoldi L (2003) The effects of sedimentation on rocky coast assemblages. In: Atkinson RJA,
Gibson RN (eds) Oceanogr Mar Biol Ann Rev. CRC Press, London, pp 161–236
Anthony KRN, Ridd PV, Orpin AR, Larcombe P, Lough J (2004) Temporal variation in light
availability in coastal benthic habitats: effects of clouds, turbidity, and tides. Limnol Oceanogr
49:2201–2211. https://doi.org/10.4319/lo.2004.49.6.2201
Barnes DKA, Conlan KE (2007) Disturbance, colonization and development of Antarctic ben-
thic communities. Philos Trans R Soc Lond Ser B Biol Sci 362(1477):11e38. https://doi.
org/10.1098/rstb.2006.1951
Barnes DKA, Souster T (2011) Reduced survival of Antarctic benthos linked to climate-induced
iceberg scouring. Nat Clim Chang 1:1–4. https://doi.org/10.1038/nclimate1232
Barnes DKA, Fenton M, Cordingley A (2014) Climate-linked iceberg activity massively reduces
spatial competition in Antarctic shallow waters. Curr Biol 24(12):R553–R554. https://doi.
org/10.1016/j.cub.2014.04.040
Barnes DKA, Fleming A, Sands CJ, Quartino ML, Deregibus D (2018) Icebergs, blue carbon and
Antarctic climate feedbacks. Philos Trans A Math Phys Eng Sci 376(2122):20170176. https://
doi.org/10.1098/rsta.2017.0176
Bartsch I, Paar M, Fredriksen S, Schwanitz M, Daniel C, Hop H, Wiencke C (2016) Changes in
kelp forest biomass and depth distribution in Kongsfjorden, Svalbard, between 1996–1998
and 2012–2014 reflect Arctic warming. Polar Biol 39:2021–2036.  https://doi.org/10.1007/
s00300-015-1870-1
Becker S, Quartino ML, Campana GL, Bucolo P (2011) The biology of an Antarctic rhodophyte,
Palmaria decipiens: recent advances. Antarct Sci 23(5):419–430. https://doi.org/10.1017/
S0954102011000575
Bers AV, Momo F, Schloss IR, Abele D (2013) Analysis of trends and sudden changes in long-­
term environmental data from King George Island (Antarctica): relationships between global
climatic oscillations and local system response. Clim Chang 116(34):789–803
9  Carbon Balance Under a Changing Light Environment 187

Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML, Wiencke C (2009) Drivers
of colonization and succession in polar benthic macro- and microalgal communities. Bot Mar
52:655–667. https://doi.org/10.1515/BOT.2009.076
Campana GL, Zacher K, Deregibus D, Momo F, Wiencke C, Quartino ML (2018) Succession of
Antarctic benthic algae (Potter Cove, South Shetland Islands): structural patterns and glacial
impact over a four-year period. Polar Biol 41(2):377–396
Clark GF, Stark JS, Johnston EL (2013) Light-driven tipping points in polar ecosystems. Glob
Chang Biol 12:3749–3761. https://doi.org/10.1111/gcb.12337
Clark GF, Stark JS, Palmer AS, Riddle MJ, Johnston EL (2017) The roles of sea-ice, light and sedi-
mentation in structuring shallow Antarctic benthic communities. PLoS One 12(1):e0168391.
https://doi.org/10.1371/journal.pone.0168391
Constable AJ, Melbourne -Thomas J, Corney SP, Arrigo KR, Barbraud C, Barnes DK, Bindoff
NL et al (2014) Climate change and Southern Ocean ecosystems. I: How changes in physical
habitats directly affect marine biota. Glob Chang Biol 20:3004–3025. https://doi.org/10.1111/
gcb.12623
Cook AJ, Fox AJ, Vaughan DG, Ferrigno JG (2005) Retreating glacier fronts on the Antarctic
Peninsula over the past half-century. Science 308(5721):541–544
Cullen JJ (1990) On models of growth and photosynthesis in phytoplankton. Deep Sea Res Part A
37:667–683. https://doi.org/10.1016/0198-0149(90)90097-F
Dennison WC, Alberte RS (1985) Role of daily light period in the depth distribution of Zostera
manna (eelgrass). Mar Ecol Prog Ser 25:51–61
Deregibus D (2017) Efecto del retroceso glaciario inducido por el cambio climático sobre la comu-
nidad de macroalgas en nuevas áreas libres de hielo en un ecosistema costero antártico (Caleta
Potter, I.25 de Mayo, I. Shetland del Sur). PhD thesis, Universidad de Buenos Aires. http://hdl.
handle.net/20.500.12110/tesis_n6241_Deregibus
Deregibus D, Quartino ML, Campana GL, Momo F, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in Potter
Cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166. https://doi.org/10.1007/
s00300-015-1679-y
Deregibus D, Quartino ML, Zacher K, Campana GL, Barnes D (2017) Understanding the link
between sea ice, ice scour and Antarctic benthic biodiversity; the need for cross station and
nation collaboration. Polar Rec 53:143–152. https://doi.org/10.1017/S0032247416000875
Drew EA, Hastings RM (1992) A year-round ecophysiological study of Himantothallus grandifo-
lius (Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31:262–277
Ducklow HW, Fraser WR, Meredith MP, Stammerjohn SE, Doney SC, Martinson DG, Sévrine F
et al (2013) West Antarctic Peninsula: an ice-dependent coastal marine ecosystem in transition.
Oceanography 26(3):190–203. https://doi.org/10.5670/oceanog.2013.62
Falk U, Sala H (2016) Winter melt conditions of the inland ice cap on King George Island, Antarctic
Peninsula. Erdkunde 341–363. https://doi.org/10.3112/erdkunde.2015.04.04
Falkowski PG, Raven JA (eds) (1997) Aquatic photosynthesis. Blackwell Scientific, Oxford
Ferron FA, Simões JC, Aquino FE, Setzer AW (2004) Air temperature time series for King George
Island, Antarctica. Pesqui Antarct Bras 4:155–169
Frenette J, Demers S, Legendre L, Dodson J (1993) Lack of agreement among models for esti-
mating photosynthetic parameters. Limnol Oceanogr 38(3):679–678. https://doi.org/10.4319/
lo.1993.38.3.0679
Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic
Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation? PLoS
One 10(8):e0134440
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, metabolic
carbon balance and zonation of sublittoral macroalgae from King George Island (Antarctica).
Mar Ecol Prog Ser 148:281–293
188 D. Deregibus et al.

Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U, Quartino ML, Dunton K, Wiencke C
(2009) Light and temperature demands of marine benthic microalgae and seaweeds in polar
regions. Bot Mar 52:593–608. https://doi.org/10.1515/BOT.2009.073
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27. https://doi.org/10.1016/j.pocean.2018.03.013
González PM, Deregibus D, Malanga G et  al (2017) Oxidative balance in macroalgae from
Antarctic waters. Possible role of Fe. J Exp Mar Biol Ecol 486:379–386
Graham MH (1996) Effect of high irradiance on recruitment of the giant kelp
Macrocystis (Phaeophyta) in shallow water. J Phycol 32:903–906. https://doi.
org/10.1111/j.0022-3646.1996.00903.x
Grange LJ, Smith CR (2013) Megafaunal communities in rapidly warming fjords along the West
Antarctic Peninsula: hotspots of abundance and beta diversity. PLoS One 8(12):e77917
Gutt J, Barratt I, Domack E, d’Udekem d’Acozd C, Dimmlere W, Grémare A, Heilmayer O, Isla E
et al (2011) Biodiversity change after climate-induced ice-shelf collapse in the Antarctic. Deep
Sea Res II 58(1):74–83
Gutt J, Bertler N, Bracegirdle TJ, Buschmann A, Comiso J, Hosie G, Isla E, Schloss IR et  al
(2015) The Southern Ocean ecosystem under multiple climate change stresses – an integrated
circumpolar assessment. Glob Chang Biol 21:1434–1453. https://doi.org/10.1111/gcb.12794
Hanelt D (1996) Photoinhibition of photosynthesis in marine macroalgae. Sci Mar 60:243–248
Hanelt D, Figueroa FL (2012) Physiological and photomorphogenic effects of light on marine
macrophytes. In: Wiencke C, Bischof K (eds) Seaweed biology: novel insights into ecophysiol-
ogy, ecology and utilization. Springer, Berlin, pp 3–23
Hanelt D, Melchersmann B, Wiencke C, Nultsch W (1997) Effects of high light stress on photo-
synthesis of polar macroalgae in relation to depth distribution. Mar Ecol Prog Ser 149:255–266
Hendry KR, Meredith MP, Ducklow HW (2018) The marine system of the West Antarctic
Peninsula: status and strategy for progress. Phil Trans R Soc A 376:20170179. https://doi.
org/10.1098/rsta.2017.0179
Henley WJ (1993) Measurements and interpretation of photosynthetic light-response curves in
algae in the context of photoinhibition and diel changes. J Phycol 29:729–739. https://doi.
org/10.1111/j.0022-3646.1993.00729.x
Hill R, Bellgrove A, Macreadie PI, Petrou K, Beardall J, Steven A, Ralph PJ (2011) Can macroal-
gae contribute to blue carbon? An Australian perspective. Limnol Oceanogr 60(5):1689–1706.
https://doi.org/10.1002/lno.10128
Hurd CL, Harrison PJ, Bischof K, Lobban CS (eds) (2014) Seaweed ecology and physiology.
Cambridge University Press, Cambridge
IPCC (Intergovernmental Panel on Climate Change) (2014) Climate change 2014: impacts, adap-
tation, and vulnerability. Summaries, frequently asked questions, and cross-chapter boxes. A
Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental
Panel on Climate Change. World Meteorological Organization, Geneva. 190 pp
IPCC (Intergovernmental Panel on Climate Change) (2019) Special report on the ocean and cryo-
sphere in a changing climate. https://www.ipcc.ch/srocc/download-report/
Jassby AD, Platt T (1976) Mathematical formulation of the relationship photosynthesis and light
for phytoplankton. Limnol Oceanogr 21:540–547. https://doi.org/10.4319/lo.1976.21.4.0540
Jerosch K, Scharf FK, Deregibus D, Campana GL, Zacher K, Pehlke H, Falk U, Hass HC, Quartino
ML, Abele D (2019) Ensemble modeling of Antarctic macroalgal habitats exposed to glacial
melt in a polar fjord. Front Ecol Evol 7:207. https://doi.org/10.3389/fevo.2019.00207
Johnston EL, Connell SD, Irving AD, Pile AJ, Gillanders BM (2007) Antarctic patterns of shallow
subtidal habitat and inhabitants in Wilke’s Land. Polar Biol 30:781–788
Jones CT, Craig SE, Barnett AB, MacIntyre HL, Cullen JJ (2014) Curvature in models of the
photosynthesis-­irradiance response. J Phycol 50(2):341–355. https://doi.org/10.1111/jpy.12164
Kain JM (1989) The seasons in the subtidal. Br Phycol J 24:203–215
9  Carbon Balance Under a Changing Light Environment 189

Kortsch S, Primicerio R, Beuchel F, Renaud PE, Rodrigues J, Lønne OJ, Gulliksen B (2012)
Climate-driven regime shifts in Arctic marine benthos. PNAS 109(35):14052–14057. https://
doi.org/10.1073/pnas.1207509109
Krause-Jensen D, Duarte CM (2014) Expansion of vegetated coastal ecosystems in the future
Arctic. Front Mar Sci 1:77. https://doi.org/10.3389/fmars.2014.00077
Krause-Jensen D, Duarte CM (2016) Substantial role of macroalgae in marine carbon sequestra-
tion. Nat Geosci 9(10):737
Krause-Jensen D, Marbà N, Olesen B, Sejr MK, Christensen PB, Rodrigues J, Renaud PE et al
(2012) Seasonal sea ice cover as principal driver of spatial and temporal variation in depth
extension and annual production of kelp in Greenland. Glob Chang Biol 18:2981–2994
Krüger M (2016) Photosynthese Lichtkurven ausgewählter makroalgenarten des Kongsfjords
(Spitsbergen, Norwegen) als Grundlage für Abschätzungnezur Produktivitäät des arktischen
Kelpwaldes. Diploma Thesis, Technical University Bergakademie Freiberg, 194 pp
Kühl M, Glud R, Borum J, Roberts R, Rysgaard S (2001) Photosynthetic performance of surface-­
associated algae below sea ice as measured with a pulse-amplitude-modulated (PAM) fluorom-
eter and O2 microsensors. Mar Ecol Prog Ser 223:1–14. https://doi.org/10.3354/meps223001
Lagger C, Servetto N, Torre L, Sahade R (2017) Benthic colonization in newly ice-free soft-­bottom
areas in an Antarctic fjord. PLoS One 12:e0186756
Lagger C, Nime M, Torre L, Servetto N, Tatián M, Sahade R (2018) Climate change, glacier retreat
and a new ice-free island offer new insights on Antarctic benthic responses. Ecography 40:1–12
Lüning K (ed) (1990) Seaweeds, their environment biogeography and ecophysiology. Wiley, New
York. 527 pp
Marcías ML, Deregibus D, Saravia L, Campana GL, Quartino ML (2017) Life between tides: spa-
tial and temporal variations of an intertidal macroalgal community at Potter Peninsula, South
Shetland Islands, Antarctica. Estuar Coast Shelf Sci 187:193–203
Marina TI, Salinas V, Cordone G, Campana GL, Moreira E, Deregibus D, Torre L, al e (2018) The
food web of Potter Cove (Antarctica): complexity, structure and function. Estuar Coast Shelf
Sci 200:141–151
Matta JL, Chapman DJ (1991) Photosynthetic responses and daily carbon balance of Colpomenia
peregrina: seasonal variations and differences between intertidal and subtidal populations. Mar
Biol 108:303–313. https://doi.org/10.1007/BF01344345
Meredith MP, King JC (2005) Rapid climate change in the ocean west of the Antarctic Peninsula
during the second half of the 20th century. Geophys Res Lett 32:L19604
Meredith MP, Falk U, Bers AV, Mackensen A, Schloss IR, Ruiz Barlett E, Jerosch K, Silva Busso
A, Abele D (2018) Anatomy of a glacial meltwater discharge event in an Antarctic cove. Philos
Trans R Soc A 376:20170163. https://doi.org/10.1098/rsta.2017.0163
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48. https://doi.org/10.1093/icb/31.1.35
Moon HW, Hussin WMRW, Kim HC, Ahn I-Y (2015) The impacts of climate change on Antarctic
nearshore megaepifaunal benthic assemblages in a glacial fjord on King George Island: responses
and implications. Ecol Indic 57:280–292. https://doi.org/10.1016/j.ecolind.2015.04.031
Navarro NP, Huovinen P, Gómez I (2019) Photosynthetic characteristics of geographically disjunct
seaweeds: a case study on the early life stages of Antarctic and Subantarctic species. Prog
Oceanogr 174:28–36. https://doi.org/10.1016/j.pocean.2018.11.001
Nelson S, Siegrist AW (1987) Comparison of mathematical expressions describing light-saturation
curves for photosynthesis by tropical marine macroalgae. Bull Mar Sci 41:617–622
Paar M, Voronkov A, Hop H, Brey T, Bartsch I, Schwanitz M et al (2016) Temporal shift in bio-
mass and production of macrozoobenthos in the macroalgal belt at Hansneset, Kongsfjorden,
after 15 years. Polar Biol 39:2065. https://doi.org/10.1007/s00300-015-1760-6
Pasotti FE, Manini D, Giovannelli D, Wolfl A-K, Monien D, Verleyen E, Braeckman U, Abele D,
Vanreusel A (2014) Antarctic shallow water benthos under glacier retreat forcing. Mar Ecol
36:716–733. https://doi.org/10.1111/maec.12179
190 D. Deregibus et al.

Pavlov AK, Leu E, Hanelt D, Bartsch I, Karsten U, Hudson SR, Gallet J-C, Cottier F et al (2019)
The underwater light climate in Kongsfjorden and its ecological implications. In: Hop H,
Wiencke C (eds) The ecosystem of Kongsfjorden, Svalbard, Advances in Polar Ecology, vol 2.
Springer, Cham, pp 137–172
Pritchard DW, Hurd CL, Beardall J, Hepburn CD (2013) Survival in low light: photosynthesis and
growth of a red alga in relation to measured in situ irradiance. J Phycol 49(5):867–879. https://
doi.org/10.1111/jpy.12093
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223
Roleda MY, Campana GL, Wiencke C, Hanelt D, Quartino ML, Wulff A (2009) Sensitivity of
Antarctic Urospora penicilliformis (Ulotrichales, Chlorophyta) to ultraviolet radiation is life-­
stage dependent. J Phycol 45:600–609. https://doi.org/10.1111/j.1529-8817.2009.00691.x
Rückamp M, Braun M, Suckro S, Blindow N (2011) Observed glacial changes on the King George
Island ice cap, Antarctica, in the last decade. Glob Planet Change 79:99–109. https://doi.
org/10.1016/j.gloplacha.2011.06.009
Runcie KW, Riddle MJ (2011) Distinguishing downregulation from other non-photochemical
quenching of an Antarctic benthic macroalga using in situ fluorometry. Eur J Phycol 46:171–180
Runcie JW, Riddle MJ (2012) Estimating primary productivity of marine macroalgae in East
Antarctica using in situ fluorometry. Eur J Phycol 47(4):449–460
Runcie JW, Paulo D, Santos R, Sharon Y, Beer S, Silva J (2009) Photosynthetic responses of
Halophila stipulacea to a light gradient: I – In situ energy partitioning of non photochemical
quenching. Aquat Biol 7:143–152
Sahade R, Lagger C, Torre L, Momo FR, Monien P, Schloss I et al (2015) Climate change and
glacier retreat drive shifts in an Antarctic benthic ecosystem. Sci Adv 1:e1500050. https://doi.
org/10.1126/sciadv.1500050
Scherrer KJN, Kortsch S, Varpe Ø, Weyhenmeyer GA, Gulliksen B, Primicerio R (2018)
Mechanistic model identifies increasing light availability due to sea ice reductions as cause
for increasing macroalgae cover in the Arctic. Limnol Oceanogr 64(1):330–341. https://doi.
org/10.1002/lno.11043
Schloss IR, Abele D, Moreau S, Norkko A, Cummings V, Thrush S (2012) Response of phyto-
plankton dynamics to 19 year (1991–2009) climate trends in Potter Cove (Antarctica). J Mar
Syst 92:53–66. https://doi.org/10.1016/j.Jmarsys.2011.10.006
Schofield O, Ducklow HW, Martinson DG, Meredith MP, Moline MA, Fraser WR (2010) How do
polar marine ecosystems respond to rapid climate change? Science 328:1520–1523. https://doi.
org/10.1126/science.1185779
Schwarz AM, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photo-
synthesis near the southern global limit for growth; Cape Evans, Ross Sea, Antarctica. Polar
Biol 26:789–799. https://doi.org/10.1007/s00300-003-0556-2
Smale DA, Barnes DKA (2008) Likely responses of the Antarctic benthos to climate-­
related changes in physical disturbance during the 21st century, based primarily on evi-
dence from the West Antarctic Peninsula region. Ecography 31:289–305. https://doi.
org/10.1111/j.0906-7590.2008.05456.x
Smith EL (1936) Photosynthesis in relation to light and carbon dioxide. Proc Natl Acad Sci
22:504–511
Spurkland T, Iken K (2011) Salinity and irradiance effects on growth and maximum photosyn-
thetic quantum yield in subarctic Saccharina latissima (Laminariales, Laminariaceae). Bot
Mar 54:355–365. https://doi.org/10.1515/BOT.2011.042
Stammerjohn SE, Martinson DG, Smith RC, Ianuzzi RA (2008) Sea ice in the western Antarctic
Peninsula region: spatio-temporal variability from ecological and climate change perspectives.
Deep-Sea Res II 55:2041–2058. https://doi.org/10.1016/j.dsr2.2008.04.026
Steele J (1962) Environmental control of photosynthesis in the sea. Limnol Oceanogr 7:137–150.
https://doi.org/10.4319/lo.1962.7.2.0137
9  Carbon Balance Under a Changing Light Environment 191

Torre L, Servetto N, Eöry ML, Momo F, Tatián M, Abele D, Sahade R (2012) Respiratory responses
of three Antarctic ascidians and a sea pen to increased sediment concentrations. Polar Biol
35:1743–1748. https://doi.org/10.1007/s00300-012-1208-1
Turner J, Bindschadler RA, Convey P, Di Prisco G, Fahrbach E, Gutt J, Hodgson DA et al (2009)
Antarctic climate change and the environment. SCAR, Cambridge
Turner J, Barrand NE, Bracegirdle TJ, Convey P, Hodgson DA, Jarvis M et al (2013) Antarctic
climate change and the environment  - an update. Polar Rec 50(3):237–259. https://doi.
org/10.1017/S0032247413000296
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental gradients in a marine macroalgal-dominated subtidal community on the western
Antarctic Peninsula. PLoS One 10(9):e0138582
Vaughan DG, Marshall GJ, Connolley WM, Parkinson C, Mulvaney R, Hodgson DA et al (2003)
Recent rapid regional climate warming on the Antarctic Peninsula. Clim Chang 60:243–274.
https://doi.org/10.1023/A:1026021217991
Webb WL, Newton M, Starr D (1974) Carbon dioxide exchange of Alnus rubra: a mathematical
model. Oecologia 17:281–291. https://doi.org/10.1007/BF00345747
Wiencke C (1990) Seasonality of brown macroalgae from Antarctica-a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600. https://doi.org/10.1007/
BF00239370
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Wiencke
C, Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and uti-
lization. Ecological studies, vol 219. Springer, Heidelberg, pp  265–292. https://doi.
org/10.1007/978-3-642-28451-9
Wiencke C, Clayton MN, Gómez I, Iken K, Lüder UH, Amsler CD et  al (2006) Life strategy,
ecophysiology and ecology of seaweeds in polar waters. Rev Environ Sci Biotechnol 6:95–126
Wiencke C, Gómez I, Dunton K (2011) Phenology and seasonal physiological performance
of polar seaweeds. In: Wiencke C (ed) Biology of polar benthic algae. De Gruyter, Berlin,
pp 181–194. https://doi.org/10.1007/978-1-4020-6285-8_13
Zacher K, Rautenberger R, Hanelt D, Wulff A, Wiencke C (2009) The abiotic environment of polar
marine benthic algae. Bot Mar 52:483–490. https://doi.org/10.1515/BOT.2009.082
Chapter 10
Life History Strategies, Photosynthesis,
and Stress Tolerance in Propagules
of Antarctic Seaweeds

Nelso Navarro, Pirjo Huovinen, and Iván Gómez

Abstract  Reproduction is one of the most important processes to maintain sea-


weed populations. In general, growth and reproduction of seaweeds depend on envi-
ronmental cues, such as change in temperature, light, and nutrients. However, the
fact that Antarctic waters show a small variation in temperature and nutrient levels
over the year, these biological processes depend mainly on variables related to light
conditions, especially daylength. This seems to be more obvious in the eulittoral
and shallow sublittoral species, because the reproduction and growth coincides with
the spring season. However, in species inhabiting the deeper sublittoral zone, repro-
duction seems to be controlled by a free-running endogenous clock synchronized by
the seasonal variation of daylength or by photoperiodisms. Whatever the case, the
Antarctic environment imposes physiological constraints to reproductive output,
settlement and development of propagules, recruitment, and growth of seaweeds.
Early life stages (e.g., spores, gametes, propagules, and plantlets) are extremely
shade-adapted and susceptible to environmental stress, such as exposure to UV
radiation; however, they are thermally well adapted, at least for short periods of
time, allowing them to develop in a highly variable environment. In this chapter, we
review the main reproduction strategies that Antarctic seaweeds display to cope
with the extreme environment. Additionally, we review recent studies on stress tol-
erance of early developmental stages from selected species. In scenarios of the
changing Antarctic environment due to warming, UV radiation, freshening, and

N. Navarro (*)
Laboratorio de Ecofisiología y Biotecnología de Algas (LEBA), Universidad de Magallanes,
Punta Arenas, Chile
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
e-mail: nelso.navarro@umag.cl
P. Huovinen · I. Gómez
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems IDEAL, Valdivia, Chile
e-mail: pirjo.huovinen@uach.cl; igomezo@uach.cl

© Springer Nature Switzerland AG 2020 193


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_10
194 N. Navarro et al.

other emergent stressors, the knowledge on adaptive life strategies of early develop-
mental phases can allow us better predicting the fate of seaweed communities.

Keywords  Life history stages · Photosynthetic light requirements · Reproduction


· Seaweed propagules · Seasonal development

10.1  Seasonal Strategies and Life History Cycles

In their environment, seaweeds are exposed to a complex suite of abiotic variables,


whose interaction may affect reproduction synergistically or antagonistically. In the
case of Antarctic seaweeds, the life strategy of the individual species is regulated by
the strong seasonal variation in light conditions (Wiencke et al. 2009). Two different
growth (and reproduction) strategies have been identified: the season anticipator
and season responder strategy (sensu Kain 1989; Wiencke and Clayton 2002). These
strategies have been corroborated through long-term laboratory culture experiments
in which temperature and nutrient levels were kept constant and only light and day-
length were modified in order to simulate the seasonally fluctuating Antarctic irradi-
ances (Wiencke 1990a, b; Dummermuth and Wiencke 2003). Moreover, other
phenological events such as seasonal induction of propagules, their release, and the
growth of early developmental stages have been examined in the field (Roleda et al.
2007, 2008; Zacher et al. 2009; Navarro et al. 2016).

10.1.1  Season Anticipators

This group of algae grows and reproduces in winter under short-day and low-light
conditions (Fig. 10.1). Thus, physiological and reproduction processes seem to be
controlled by a free-running endogenous annual rhythm synchronized by the sea-
sonal changes of daylength or by photoperiodisms and not by environmental condi-
tions (such as levels of light or temperature) as demonstrated by Lüning and tom
Dieck (1989), tom Dieck (1989), and Lüning and Kadel (1993) in several species
from temperate regions. Likewise, the growth of Antarctic season anticipators has
been related to increasing daylength during the late winter and early spring (Wiencke
et al. 2007, 2009). Many endemic Antarctic seaweeds with sublittoral distribution
are regarded as season anticipators, e.g., the brown seaweeds Himantothallus gran-
difolius (A.Gepp and E.S.Gepp) Zinova, Desmarestia anceps Montagne, D. antarc-
tica Moe et Silva, Phaeurus antarcticus Skottsberg (Wiencke 1990a), Ascoseira
mirabilis Skottsberg (Gómez et al. 1995, 1996; Wiencke 1990a), and D. menziesii
J.  Agardh (Gómez and Wiencke 1996) and the red seaweeds Palmaria decipiens
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 195

Fig. 10.1  Life history development of some conspicuous Antarctic seaweeds. The green line
shows the period when growth starts. In the case of season anticipators, growth take place in late
winter onwards, while in responders, growth occur during spring onwards
196 N. Navarro et al.

(Reinsch) Ricker (Wiencke 1990b), Paraglossum salicifolia (Reinsch) Schowe


M.  Lin, Fredericq, and Hommersand (formerly Delesseria salicifolia Reinsch),
Gymnogongrus antarcticus Skottsberg; G. turquetii Hariot, Hymenocladiopsis pro-
lifera (Reinsch) M. J. Wynne (formerly H. crustigena R.L. Moe), Trematocarpus
antarcticus (Hariot) Fredericq and R. Moe (formerly Kallymenia antarctica Hariot),
and Phyllophora ahnfeltioides Skottsberg (Dummermuth and Wiencke 2003).
Reproductive responses to the environment are particularly evident in season
anticipators with strongly heteromorphic phase expression, such as in members of
the genus Desmarestia (Wiencke et al. 1991, 1995, 1996), Himantothallus grandi-
folius (Wiencke and Clayton 1990), and Phaeurus antarcticus (Clayton and Wiencke
1990). The heteromorphic life history of large brown algae is characterized by the
development of large perennial sporophytes and a marked reduction of the gameto-
phytic generation (Clayton 1988). In these species, microscopic gametophytes and
early stages of sporophytes grow under limited light conditions during winter,
whereas adult stages of macroscopic sporophytes grow in late winter–spring
(Fig. 10.1). In the case of Desmarestia anceps, one of the most important seaweeds
in terms of biomass in the Antarctic region, the microstage of male and female
gametophytes becomes fertile between July and September under a daylength of 5
and 7 h day−1 at photon fluence rates <3 μmol photon m−2 s−1 (Wiencke et al. 1996).
The induction of fertility is a photoperiodic short-day response as revealed by the
effect of a night-break regime (Wiencke 1990b; Wiencke et al. 1996), while in con-
tinuous darkness gamete formation was inhibited (Wiencke et  al. 1996).
Gametogenesis under short daylengths was also demonstrated in other Desmarestiales
members, e.g., Himantothallus grandifolius (Wiencke and Clayton 1990) and
Desmarestia menziesii (Wiencke et al. 1995; Gómez and Wiencke 1997), whereas
no daylength dependence of gamete formation has been found in Desmarestia ant-
arctica (Wiencke et  al. 1991). In this latter species and in Phaeurus antarcticus,
gametogenesis occurs both in short and long days. According to Wiencke et  al.
(2009) the phenology in these species is controlled by the sporophytic stage, which
becomes fertile at daylengths between 6 and 8 h day−1, while gametophytes form
gametangia soon after germination (Clayton and Wiencke 1990; Wiencke 1990a;
Wiencke et al. 1991). A typical feature of Antarctic Desmarestiales is the fact that
they exhibit in situ fecundation, and the juvenile sporophytes remain attached to the
female gametophytes (Wiencke et al. 1995, 1996). This feature could have ecologi-
cal significance for the sporophytes recruitment and dominance of this group in
Antarctic environment (Wiencke et al. 2006).
The brown alga Ascoseira mirabilis, another season anticipator, exhibits maxi-
mum growth rates in late winter–spring, while the minimum growth rates were
recorded in May–June (Wiencke 1990a). However, a further, much smaller growth
optimum became evident between January and March. On the other hand, unlike
typical season anticipators, fertile fronds in A. mirabilis are present all year round,
and growing and reproducing when environmental conditions are favorable (see
below). A. mirabilis is the only member of the order Ascoseirales, and the Antarctic
environmental constraints might have exerted an evolutionary pressure to develop a
unique life history and reproductive biology when compared with other
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 197

Phaeophyceae (Roleda et  al. 2007). The species is monoecious with sexual (iso-
gamia) reproduction. There is one, free-living diploid generation, and zygotes
develop into new individuals (Wiencke and Clayton 2002). Conceptacles are scat-
tered all over the blades and the extrusion of gametangial masses through the osti-
oles precedes the release of heterokont gametes (Müller et  al. 1990). Zygote
formation follows immediately after fusion of gametes.
For red seaweeds, the phenology of six season anticipator species from the
Antarctic, Paraglossum salicifolium, Gymnogongrus antarcticus, Gymnogongrus
turquetii, Hymenocladiopsis prolifera, Trematocarpus antarcticus, and Phyllophora
ahnfeltioides were investigated by Dummermuth and Wiencke (2003) in a two-year
culture study under fluctuating daylengths simulating the Antarctic conditions. The
period of highest growth rate in these species was registered between September
and November (late winter–spring) and the formation of new blades occurred from
January/February onwards. Before the summer solstice, growth ceased. However, in
Hymenocladiopsis prolifera, the seasonal growth peak was observed in August
when the light conditions increased from 3 to 25 μmol photon m−2 s−1. This suggests
that the phenology of season anticipators could not only be controlled by daylength
but also by photon fluence rates. Thus, in species distributed along wide ranges of
depth (e.g., 5–30 m), the seasonal growth peak could be later in the season at deeper
water depths and earlier in shallower waters. Reproductive fronds were not observed,
except in Trematocarpus antarcticus, which completed its life cycle with carpo-
spore formation between June and August, but with the first cystocarps found in
March (Dummermuth and Wiencke 2003). These results agree with the findings
reported by Lamb and Zimmermann (1977), who reported cystocarps in thalli of
T. antarcticus in January. Similarly, in Gymnogongrus antarcticus, cystocarps are
formed in the summer (Skottsberg 1953; Cormaci et al. 1992). Cystocarps and tet-
rasporangia of Paraglossum salicifolium have been observed in late winter (Wynne
1982). Likewise, spermatangia, cystocarps, and tetrasporangia in Delesseria san-
guinea, a comparable species in the same family, are formed during winter
(Kornmann and Sahling 1977). In the case of G. turquetii (Kylin and Skottsberg
1919; Skottsberg 1953) and Phyllophora ahnfeltioides (Kylin and Skottsberg 1919),
cystocarpic fronds have been reported between May and June (autumn).
The red seaweed Palmaria decipiens, one of the dominating species in terms of
biomass, is considered as season anticipator (Wiencke 1990b; Weykam and Wiencke
1996) and displays a heteromorphic life history perfectly adapted to the Antarctic
conditions. In this species female gametophytes represent the microscopic phase,
while the male gametophyte develops into a macro-thallus similar in morphology to
the tetrasporophytes (Fig. 10.1). The male and tetrasporophytic blades are formed
in winter (Wiencke 1990b; Weykam et al. 1997) and the optimum growth period and
high rates of net photosynthesis and photosynthetic efficiency coincide with increas-
ing light intensities in spring (Wiencke 1990b; Weykam and Wiencke 1996).
Tetrasporophytes become fertile in February and tetraspores develop in May into
semiglobular to discoid gametophytes. The females become fertile only once from
May to June. After fertilization, the female gametophyte is overgrown by the devel-
oping sporophyte, which matures and releases tetraspores in the next summer.
198 N. Navarro et al.

Interestingly, it takes about a year until male gametophytes become fertile; thus
fertilization of females is only possible by mature males of the previous season,
indicating a life span of the species of several years (Wiencke 1990b).

10.1.2  Season Responders

In these organisms growth and reproduction coincide with favorable light condi-
tions in spring and summer. Thus, these species react directly to the primary factors
in their environment (such as light availability) and show an opportunistic life strat-
egy (Wiencke 1990a, b). Most of the season responder species are distributed in the
eulittoral and upper sublittoral zone, and they can have temperate or cold-temperate
affinities (Wiencke et  al. 2007; Navarro et  al. 2019; see Chap. 12 by Campana
et  al.). Well-known members of this group are the red seaweeds Iridaea cordata
(Turner) Bory (Weykam et  al. 1997) and Gigartina skottsbergii Setchell et
N.L.  Gardner (Wiencke 1990b), the brown alga Adenocystis utricularis (Bory)
Skottsberg (Wiencke 1990a), and the green seaweeds Ulva hookeriana (Kützing)
H.  S. Hayden, Blomster, Maggs, P.  C. Silva, Stanhope, and Walland (formerly
Enteromorpha bulbosa (Suhr) Montagne and Acrosiphonia arcta (Dillwyn)
J. Agardh (Wiencke 1990b).
The pseudoperennial Gigartina skottsbergii and Iridaea cordata have a triphasic
life history with isomorphic haploid male and female gametophytes and a diploid
tetrasporophyte. They occur normally in eulittoral pools and in the upper sublittoral,
but also can be found down to 30 m (Wiencke and Clayton 2002; Navarro et al.
2016). Both species show the maximum growth rate during the spring-summer sea-
son (e.g., December), while the minimum growth rates were recorded from May to
July (Wiencke 1990b). Mature tetrasporophytes and gametophytes of Iridaea cor-
data were observed during spring-summer (Roleda et al. 2008; Navarro et al. 2016).
Tetraspores and carpospores of this species germinate normally forming a discoid
germling from which new plantlets arise from July onwards. The plantlets show a
growth optimum between September and November and large blades are formed in
summer (Wiencke 1990b). Regrowth from the perennial basal parts of the blades is
possible (Wiencke and Clayton 2002), which could explain its dominance at the
eulittoral (Marcías et  al. 2017). In the case of G. skottsbergii, Wiencke (1990b)
reported the induction of sporangium formation in tetrasporophytes in the labora-
tory by the end of September, when irradiances were between 27 and 46 μmol pho-
ton  m−2  s−1, but spores were not released before June. In contrast, in the field,
reproductive fronds with viable propagules have been collected in October (Roleda
et  al. 2008) and January (Navarro et  al. 2016). This discrepancy in reproductive
periods might be related to differences related to the experimental setup of labora-
tory cultures by Wiencke (1990b). As suggested for the season anticipator
Hymenocladiopsis prolifera, photon fluence rates might also control the seasonal
phenology of this species. Thus, these algae apparently have the capacity to repro-
duce during a prolonged time span under changing environmental conditions.
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 199

Adenocystis utricularis has a heteromorphic life cycle with a sporophytic macro-


thallus and gametophytic microthallus. Spores from the macrothalli develop into
microscopic filamentous, dioecious gametophytes (Wiencke and Clayton 2002), the
dominant life phase under winter conditions in laboratory culture (Wiencke 1990a).
Macrothalli start to develop asexually on crustose parts of the microthalli from June
onwards. Between October and December, growth rates of macrothalli are optimal.
Reproductive macrothalli are present in January–February, while release of spores
occurs in February, after which the thalli disintegrate.
As suggested for other species with heteromorphic phase expression, the micro-
thallus is probably an important over-wintering stage (Wiencke and Clayton 2002).
However, eventually all developmental stages can be present at the same time
depending where they occur. De Reviers and Délépine (1981) reported that macro-
thalli are present throughout the year with juveniles being most abundant in October
in the eulittoral zone, while in the sublittoral zone, small macrothalli are present
only from November to June. Laboratory experiments where the photon fluence
rates varied from 2 to 46  μmol  photon  m−2  s−1 confirmed this field observation
(Wiencke 1990a). Thus, in the eulittoral zone, A. utricularis occurs as an aseasonal
annual, while in upper sublittoral zone, the species probably occurs as a seasonal
annual due to less available light. The species has been reported to occur down to
20 m (Wiencke and Clayton 2002), and at these depths the alga possibly is biannual
as suggested by Wiencke (1990a).

10.2  Photosynthetic Light Requirements of Early Stages

Irrespective of the life history strategy, the Antarctic environment imposes physio-
logical constraints to the reproductive output (propagules), settlement, recruitment,
and growth of seaweeds. However, seaweeds have adapted their biological pro-
cesses by evolving different functional mechanisms to cope with the Antarctic light
and temperature conditions. In general, Antarctic seaweeds are very low light
adapted, adult phases being able to photosynthesize at irradiances as low as
10 μmol photon m−2 s−1, while propagules can photosynthesize at even lower irradi-
ances (Gómez et al. 2009). In Table 10.1 the saturating irradiances of photosynthe-
sis (Ek values) of different Antarctic seaweed propagules are summarized, and in
Fig. 10.2 mean values of Ek for propagules from eulittoral and sublittoral algae are
plotted. With the exception of some eulittoral species, most of the studied Antarctic
seaweeds exhibit Ek values lower than 60 (μmol photon m−2 s−1). Although differ-
ences in saturation irradiance between eulittoral (65 ± 20 μmol photon m−2 s−1) and
sublittoral (40 ± 15 μmol photon m−2 s−1) exist, propagules are able to adapt to dif-
ferent light conditions (quantity and quality) mainly during the winter–spring tran-
sition. This is particularly evident in propagules of species that colonize a wide
range of vertical distribution (e.g., Desmarestia anceps). After sea ice breakup in
King George Island (South Shetland Islands), light can penetrate down to 30  m
200

Table 10.1  Overview of light requirement for saturation (Ek), photosynthetic efficiency (αETR), inhibition by PAR and PAR + UV radiation, and subsequent
recovery in different life history stages of Antarctic seaweeds collected from different depths. Propagule size and date of collection are also indicated.
Irradiance values represent conditions during incubation in the laboratory. PAR and PAR + UV inhibition were calculated after 4 h of exposition according to
information provided in the papers)
Cell Irradiance
Date of diameter (μmol PAR PAR + UV
Species Depth collection (μm) m−2 s−1) Ek αETR Fv/Fm inhibition inhibition Recovery Reference
Urospora Upper Oct 2004 6 22 87 0.16 ~0.50 No ~37% (PA; 100% Roleda
penicilliformis eulittoral 4 h); ~39% (24 h) et al.
zoospores, (0 m) (PAB; 4 h) (2009)
gametes
Urospora Upper Oct 2004 22 252 0.18 ~0.51 No ~20% (PA; 94% Roleda
penicilliformis eulittoral 4 h); ~23% (1 h) et al.
gametophytes (0 m) (PAB; 4 h) (2009)
Pyropia Eulittoral Jan–Feb 12 ± 1 13 101 ± 19 0.14 ± 0.01 0.35 ± 0.01 No ~16% (PAB, 100% Navarro
endiviifolia (0 m) 2015 4 h) (4 h) et al. (2016,
carpospores 2019)
Pyropia Upper Jan–Mar 15 ± 2 23 33 0.12 0.49 ± 0.04 46% (4 h) ~56% (PA; 100% Zacher
endiviifolia eulittoral 2005 4 h); ~60% (48 h) et al.
monospores (0 m) (PAB; 4 h); (2007)
Adenocystis Eulittoral Jan–Feb 6 ± 1 13 59 ± 6 0.45 ± 0.01 0.64 ± 0.02 No ~2% (PAB, 100% Navarro
utricularis (0–1 m) 2015 4 h) (4 h) et al.
zoospores (2016)
Adenocystis Eulittoral Jan–Mar 4 23 64 0.14 0.46 ± 0.11 No ~20% (PA; 100% Zacher
utricularis (0–1 m) 2005 4 h); ~57% (48 h) et al.
zoospores (PAB; 4 h); (2007)
N. Navarro et al.
Cell Irradiance
Date of diameter (μmol PAR PAR + UV
Species Depth collection (μm) m−2 s−1) Ek αETR Fv/Fm inhibition inhibition Recovery Reference
Monostroma Eulittoral Jan–Feb 5 ± 0.5 13 295 ± 84 0.09 ± 0.01 0.26 ± 0.01 23% (4 h) ~28% (PAB; 100% Navarro
hariotii gametes (0–1 m) 2015 4 h) (4 h) et al. (2016,
2019)
Monostroma Eulittoral Jan–Mar 7 23 83 0.065 0.29 ± 0.04 50% (4 h) ~66% (PA; 100% Zacher
hariotii gametes (0 m) 2005 4 h); ~71% (48 h) et al.
(PAB; 4 h); (2007)
Iridaea cordata Eulittoral Jan–Feb 18 ± 2 13 46 ± 6 0.17 ± 0.03 0.39 ± 0.06 No ~20% (PAB; 100% Navarro
tetraspores (0–1 m) 2015 4 h) (4 h) et al. (2016,
2019)
Iridaea cordata Upper Oct 2004 20 ± 2 22 57 0.12 0.47 ± 0.04 53% (4 h) ~78% (PA; 100% (2 Zacher
tetraspores sublittoral 4 h); ~82% d) et al.
(0 m) (PAB; 4 h) (2009)
Gigartina Sublittoral Jan–Feb 25 ± 2 13 27 ± 9 0.27 ± 0.05 0.42 ± 0.32 52% (4 h) ~80% (PAB, 35–60% Navarro
skottsbergii (5–8 m) 2015 4 h) (4 h) et al.
carpospores (2016)
Gigartina Sublittoral Oct 2004 27 ± 2 22 54 ± 2 0.14 0.40 ± 0.03 73% ~87% (PA; 100% (2 Roleda
skottsbergii (3 m) 4 h);89% d) et al.
carpospores (PAB; 4 h) (2008)
Gigartina Sublittoral Oct 2004 23 ± 1.5 22 44 ± 21 0.14 0.31 ± 0.07 69% (4 h) ~86% (PA; 100% (2 Roleda
skottsbergii (3 m) 4 h);90% d) et al.
tetraspores (PAB; 4 h) (2008)
Ascoseira Deep Jan–Feb 22 ± 4 13 17 ± 5 0.21 ± 0.03 0.48 ± 0.02 16% (4 h) ~30% (PAB; 85% Navarro
mirabilis sublittoral 2015 4 h) (4 h) et al.
gametangia (17–30 m) (2016)
(continued)
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules…
201
Table 10.1 (continued)
202

Cell Irradiance
Date of diameter (μmol PAR PAR + UV
Species Depth collection (μm) m−2 s−1) Ek αETR Fv/Fm inhibition inhibition Recovery Reference
Ascoseira Upper Oct 2004 8–10 22 52 0.09 0.40 ± 0.06 85% (4 h) ~91% (PA; 100% (2 Roleda
mirabilis sublittoral 4 h); 95% d) et al.
gametangia (2 m) (PAB; 4 h) (2007)
Ascoseira Shallow Jan 2014 13 52 ± 14 0.20 ± 0.04 ~0.6 ~0% (6 h) 100% Huovinen
mirabilis sublittoral (14 h) and Gómez
conceptacles (1 m) (2015)
Cystosphaera Deep Jan–Feb 13 13 ± 6 0.28 ± 0.01 ~15–21% ~ 90% Huovinen
jacquinotii sublittoral 2015 (2 h) (4 h) and Gómez
receptacles (2015)
N. Navarro et al.
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 203

Fig. 10.2  Summary of light requirement for saturation (Ek) and inhibition by PAR and PAR + UV
radiation in propagules of eulittoral and sublittoral Antarctic seaweeds. PAR and PAR + UV inhibi-
tion were calculated after 4-h exposure according to the information provided in each of the studies
consulted. For references, see Table 10.1

depth reaching an average of photon fluence rates of 70 μmol photon m−2 s−1 (Gómez


et  al. 1997). This level of irradiance can also be strongly attenuated in terms of
spectral characteristics under the canopy of large brown algae (Huovinen et  al.
2016; Gómez et al. 2019). Below the canopy the spectrum is enriched in green and
in far red light, probably affecting photosynthesis as well as the photomorphoge-
netic development of the understory species (Salles et  al. 1996) (see Chap. 7 by
Huovinen and Gómez).
204 N. Navarro et al.

Fig. 10.3  Rapid light curves (PAR vs rETR) and rETR/ rETRmax ratio of Antarctic and sub-­
Antarctic populations of Adenocystis utricularis and Iridaea cordata measured using chlorophyll
fluorescence. rETRmax represents the maximum value for each curve. (Modified from Navarro
et al. 2019)

10.2.1  E
 stimating Photosynthetic Parameters
from Chlorophyll Fluorescence

Photosynthetic characteristics of propagules of Antarctic seaweeds are normally


determined estimating photosynthetic parameters (ETRmax, alpha and saturation
irradiance, Ek) calculated from P-E curves (summarized in Gómez et  al. 2009).
When P-E curves are based on chlorophyll fluorescence measurements, the electron
transport rates (ETR) are commonly used as a parameter (Fig. 10.3). Considering
the limitations of the fluorescence method, as well as various factors that can affect
light requirements, e.g., form, season, size, number of cells, chlorophyll concentra-
tion, etc., the P-E curve-derived light requirements for photosynthesis (Ek) represent
the best measures to estimate shade adaptation in adult (Huovinen and Gómez
2013) and early stages (Zacher et al. 2007; Roleda et al. 2008, 2009; Navarro et al.
2016, 2019). However, two important aspects for calculation of electron transport
rates are sometimes not considered: (1) the proportional rates of chlorophyll a at
each photosystem (FII factor), which is different between red, green, and brown
algae (Grzymski et  al. 1997), and (2) the amount of light absorbed by the algal
samples (absorptance). Due to the difficulty of measuring absorptance in a propa-
gule suspension, the use of the relative ETR has been proposed (Beer et al. 2001).
This parameter provides useful information for the description of relative changes
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 205

in photosynthetic activity if the experimental approach uses standardized


measurements.
Thus, the comparison of photosynthetic parameters of propagules using the
ETR-based P-E curves must be made with caution. Alternatively, the rETR can be
normalized to rETRmax (with rETRmax as the maximum value for each curve) express-
ing the rETR curve between 0 and 1 (relative units), which allows comparing propa-
gules from species with very different ETR values. For example, Navarro et  al.
(2019) showed differences in photosynthetic performance of propagules from con-
generic and conspecific Antarctic/sub-Antarctic seaweeds using the rETR and
rETR/rETRmax ratio curves. While the rETR curve demonstrates that tetraspores of
I. cordata from Antarctic populations exhibit very low rETR values when compared
to the sub-Antarctic population, the rETR/rETRmax ratio allows detecting differ-
ences in the shape of the curves (Fig. 10.3).

10.3  E
 ffects of Environmental Factors on the Biology
of Propagules

10.3.1  High Solar Radiation

Environmental shifts will affect recruitment, and consequently, the whole fate of the
seaweed population and their maintenance through time. Once Antarctic seaweed
spores or gametes are released, they face a completely different physical environ-
ment than what existed when they were housed in the parental reproductive struc-
tures (Amsler et  al. 1992; Zacher 2014). During the last decade, various studies
have examined the effects of stress factors (e.g., temperature and UV radiation) on
microscopic stages, e.g., propagules and plantlets, of some selected Antarctic sea-
weeds (Gómez et al. 2009; Roleda et al. 2009; Zacher et al. 2009; Navarro et al.
2016). There is a consensus that propagules are the most susceptible life stage of
seaweeds in terms of their responses to environmental perturbations. However, the
effect of a given factor on the propagule physiology is highly variable, depending on
a suite of environmental and endogenous factors, which can act synergistically or
antagonistically.
It is well known that UV wavelengths cause direct and indirect effects on algal
cells (e.g., Karsten et al. 2009). The direct effects are normally mediated by absorp-
tion of UV by important biomolecules, in particular the DNA, enzymes, and mem-
brane components (Vass 1997). In the case of propagules, which attain small size,
translucent cytosol and an incipient development of the cell wall, UV radiation can
easily reach the DNA where diverse injuries are produced, e.g., formation of
cyclobutane pyrimidine dimers (CPDs) (Wiencke et al. 2000). This results in the
inability of RNA and DNA polymerases to recognize the damaged sectors, causing
the interruption of gene transcription and DNA replication (Britt 1995).
206 N. Navarro et al.

Consequently, modifications in the metabolism, cellular division, and germination


of unicellular propagules can occur (Huovinen et al. 2000).
DNA damage has been reported in propagules of eulittoral Antarctic seaweeds
Adenocystis utricularis, Monostroma hariotii, and Iridaea cordata after exposure to
different doses of UV radiation (Roleda et al. 2007, 2008; Zacher et al. 2007, 2009).
In general, the amount of CPDs increases with increasing UV-B dose; however,
lesions can be effectively repaired after 48  h under photoreactivation processes
(Zacher et al. 2007, 2009). Besides, contrasting patterns have been detected in spe-
cies from different depths: propagules of Pyropia endiviifolia from upper eulittoral
did not exhibit CPDs under different UV-B doses (Zacher et al. 2007), while propa-
gules of Gigartina skottsbergii and Ascoseira mirabilis from the deep sublittoral
were more affected and not able to repair their damaged DNA completely after 8-h
UV-B exposure (0.4 Wm−2) (Roleda et al. 2007, 2008). Interestingly, in the case of
Gigartina skottsbergii, the accumulation of DNA damage was related to the ploidy
level of the propagules: DNA damage was lower in diploid carpospores (2n) com-
pared to haploid tetraspores (n) suggesting that diploid carpospores are more toler-
ant to UV radiation in terms of UV-B-induced DNA damage (Roleda et al. 2008).
These authors suggested that higher UV-stress tolerance of diploid carpospores than
haploid tetraspores could be related to the genetic buffering hypothesis, which says
that diploid organisms are more vigorous and tolerant to stress than haploid ones,
i.e., the two copies of every gen confer them advantages to withstand the effects of
deleterious recessive mutations (Raper and Flexer 1970; Gerstein et  al. 2010).
However, diverse studies have stated many important genetic advantages of hap-
loidy such as lower mutation load and more rapid spread of beneficial alleles and of
diploidy, e.g., protection from somatic mutation and heterozygote advantage (Otto
and Gerstein 2008). In fact, in spite of the higher DNA damage, tetraspores of
G. skottsbergii exhibited a higher DNA damage repair rate than carpospores when
the UVR was excluded. It must be noted that DNA damage in spores exposed to
high UV-B dose was not repaired completely after 2 days of post-cultivation, and
the remaining DNA damage was lower in carpospores than in tetraspores (Roleda
et al. 2008).
UV radiation affects also photochemical processes, especially inhibiting the
energy transfer within the PSII reaction center by blocking the electron flow. UV-B
radiation affects the D1/D2 protein complex (Richter et al. 1990) mainly by frag-
menting the D1 protein (Vass 1997; Bischof et al. 2006) through UV-active chromo-
phores on both the donor and acceptor side of this protein (Bouchard et al. 2006).
On the oxidizing side, the oxygen evolving system (water splitting complex) is
another sensitive target of UV-B (Renger et al. 1986). Moreover, it has been sug-
gested that UV-B can affect the antenna complex through the functional shutdown
of the photosystem, resulting in a failure in the transfer of energy to the reaction
center (Renger et al. 1986; Lorenz et al. 1997; Bischof et al. 2006). In propagules of
Antarctic seaweeds, UV radiation has also been pointed out as responsible for the
decrease in photosynthetic activity, measured as decreases in optimum quantum
yield-Fv/Fm. For example, Navarro et al. (2016) reported that propagules of species
from the eulittoral (e.g., Iridaea cordata, Pyropia endiviifolia, Adenocystis
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 207

utricularis) showed <20% inhibition in Fv/Fm from UV (1.5 and 0.26  Wm−2 of
UV-A and UV-B, respectively) after 4 h of exposure, while propagules of the red
alga G. skottsbergii collected in the sublittoral were more sensitive exceeding 30%
inhibition in Fv/Fm in the same condition. It is important to emphasize, however, that
photochemical reactions of propagules from Antarctic seaweeds can also be strongly
photoinhibited by PAR (Fig. 10.2). For example, 1-h exposure under 22 μmol pho-
tons m−2 s−1 of PAR decreased Fv/Fm in propagules of the sublittoral G. skottsbergii
(53–58%) (Roleda et al. 2008) and A. mirabilis (62%) (Roleda et al. 2007) and in
the eulittoral M. hariotii (62%), P. endiviifolia (81%) (Zacher et  al. 2007), and
I. cordata (~25%) (Zacher et al. 2009). Increasing exposure time further reduced
Fv/Fm in all these species, with exception of M. hariotii (Zacher et  al. 2007). In
contrast, in the case of propagules of the eulittoral Adenocystis utricularis (Zacher
et al. 2007) and Urospora penicilliformis (Roleda et al. 2009), the photosynthetic
activity was not affected by PAR.  PAR supplemented with UV-A (~4.3 Wm−2)
decreased photosynthetic efficiency significantly compared to only PAR treatment
in all mentioned species during 1-h exposure. However, additional UV-B
(~0.35  Wm−2) revealed a further decrease of Fv/Fm only in sublittoral Ascoseira
mirabilis (25%) and G. skottsbergii (3–7%) (Roleda et al. 2007, 2008). Although
UV radiation further decreased photosynthetic efficiency in these species, all propa-
gules recovered completely after 48 h (Table 10.1).
Additionally, the UV susceptibility has been related to propagule size as cell path
length affects various bio-optical processes such as scattering and spectral extinc-
tion (Swanson and Druehl 2000; Roleda et al. 2008; Navarro et al. 2016). However,
at the cellular level, UV tolerance does not seem to respond to complex biochemical
and bio-optical processes. For example, tetraspores of I. cordata from Antarctica
exhibit a smaller size but very high UV tolerance compared to tetraspores of the
same species from sub-Antarctic  region (Navarro et  al. 2019). UV tolerance can
also be related to the presence and/or the capacity to induce formation of
UV-absorbing compounds, what could result in a more effective UV photoprotec-
tion, still in small propagules (Roleda et  al. 2008). To our knowledge, only few
studies have described absorption of UV in Antarctic seaweed propagules under UV
stress. Higher concentration of palythine (λmax  =  320  nm) than shinorine
(λmax = 334 nm) has been reported in freshly released tetraspores of G. skottsbergii
(Roleda et al. 2008) and I. cordata (Zacher et al. 2009). However, contrasting pat-
terns in MAA content were observed after 8 h under PAR or PAR + UV treatments,
while the total content of MAAs in tetraspores of G. skottsbergii was not signifi-
cantly different between control (freshly released spores) and treatment. In contrast,
MAA concentration in spores of I. cordata decreased in treated compared to freshly
released spores. Based on these findings, it could be suggested that (1) freshly
released propagules could have a basal level of UV-absorbing substances due to the
higher in situ incident solar radiation in the field and (2) the level of UV-absorbing
substances can acclimate depending on environmental conditions. In the first case,
the synthesis of UV-absorbing substances would take place when the spores are still
protected by the thick tissue of the parental thalli (tetrasporangial tissue in the case
208 N. Navarro et al.

of tetraspores of G. skottsbergii and I. cordata). For I. cordata, Karsten et al. (2000)


reported a higher amount of MAAs in tetrasporangial tissue than in vegetative parts
of the thalli. Similarly, Huovinen and Gómez (2015) reported that reproductive tis-
sue of Ascoseira mirabilis and Cystosphaera jacquinotii contain higher amounts of
soluble phlorotannins, a type of UV-absorbing phenols found in brown algae. The
presence of these compounds in reproductive tissues could ensure the maturation,
survival, and germination of released propagules when they are exposed to UV
radiation in the water column. Although photoprotection was only partial in labora-
tory experiments, propagules of I. cordata and G. skottsbergii tetraspores exposed
to UV-B radiation showed the higher total MAAs in comparison with those incu-
bated under only PAR (Roleda et al. 2008; Zacher et al. 2009).

10.3.2  Temperature

Antarctic seaweed propagules are adapted to low temperature. Cold adaptation was
confirmed by the high photosynthetic efficiency (in terms of maximum quantum
yield of fluorescence – Fv/Fm) at 0 °C in six Antarctic distributed species (Navarro
et al. 2016). This low temperature requirement for photosynthesis is certainly the
result of the long Antarctic cold-water history of at least 14  Ma (Crame 1993).
However, it is well known that photosynthesis increases progressively with increas-
ing temperature and then rapidly declines near upper critical temperature (Davison
1991). In the case of Antarctic species, the optimum temperature for photosynthesis
is between 10 and 20 °C (Eggert and Wiencke 2000; Eggert 2012), lower than that
reported for algae from other geographic regions (reviewed in Gómez et al. 2009).
In propagules of eulittoral species such as Adenocystis utricularis, Monostroma
hariotii, and Pyropia endiviifolia and shallow sublittoral Ascoseira mirabilis, the
highest photosynthetic efficiency was observed at 25 °C (Navarro et al. 2016). This
suggests that propagules of these species are thermally well adapted (eurythermal
species), allowing them to develop in a highly variable environment or in different
biogeographic regions. For example, A. utricularis and M. hariotii are widely dis-
tributed in sub-Antarctic and temperate coasts of South America (Huovinen and
Gómez 2012, see Chap. 2 by Oliveira et al. and Chap. 4 Macaya et al.). In contrast,
the high photosynthetic efficiency exhibited by propagules of Antarctic endemic
Ascoseira mirabilis at 25 °C could be explained by the upper vertical distribution of
the parental sporophytes or could be a conserved trait related to the fact that the spe-
cies is probably a relic of Mesozoic (Gondwana) marine flora, which was highly
diverse when the average water temperatures were close to 12 °C (Clayton 1994).
Temperature is a factor modifying the susceptibility/tolerance to UV radiation.
The influence of this factor apparently depends on the position of parental thalli on
the shore. In this context, a recent study provided evidence that propagules of
Antarctic seaweeds are relatively tolerant to enhanced temperature, which can fur-
thermore modulate UV tolerance at least under laboratory conditions (Navarro et al.
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 209

2016). These authors observed that the exposure of propagules to a combination of


UV radiation and temperature stress inhibits the photosynthetic capacity of propa-
gules of six Antarctic seaweed species from the eulittoral (Pyropia endiviifolia,
Iridaea cordata, Adenocystis utricularis, and Monostroma hariotii) and the sublit-
toral (Ascoseira mirabilis, and Gigartina skottsbergii), the former group being more
tolerant to UV and enhanced temperature than the sublittoral group. Additionally,
propagules of eulittoral species P. endiviifolia, I. cordata, and A. utricularis exhibit
negative UV effects at 2  °C compared to 7 and 12  °C, suggesting that enhanced
temperature improves UV tolerance. On the contrary, this positive interaction was
not observed in propagules of the shallow sublittoral A. mirabilis, where an increase
in temperature exacerbates the reduction of photosynthetic efficiency (Navarro et al.
2016). It is known that various processes related to photoprotection, e.g., D1 protein
turnover, enzyme repair mechanisms, and dissipative quenching, operate more effi-
ciently at higher temperatures (Wünschmann and Brand 1992; Becker et al. 2010).
Thus, the lower inhibition of photosynthesis observed at 12 °C compared to 2 and
7 °C can be regarded as an efficient acclimation of photosynthesis in these cells.
Even though photosynthesis was inhibited by UV radiation, propagules from eulit-
toral species recover completely after 4 h under dim visible light, whereas sublitto-
ral ones do not. A fast turnover of D1 protein may be responsible for the fast
reversible photoinhibition of photosynthesis in eulittoral macroalgae as suggested
for Urospora penicilliformis propagules (Roleda et al. 2009). However, the recovery
is not influenced by a temperature increase in the studied species (Navarro
et al. 2019).
Antarctic propagules can retain their capacity to tolerate elevated temperatures,
which is evident when they are compared with their sub-Antarctic counterparts. For
example, Fv/Fm measured in I. cordata tetraspores from Antarctica was not inhibited
by UV radiation at 2 °C or 8 °C, while propagules from sub-Antarctic populations
exhibited a decrease after a 4-h exposure, mainly at 2  °C in PAR (30%) and
PAR + UV (67%). Considering only the effects of temperature, Fv/Fm decreased by
14% in tetraspores from sub-Antarctic population exposed at 2 °C when compared
to the control (8  °C). Surprisingly, photosynthetic activity in tetraspores from
Antarctic increased by 2% relative to control. These results suggest that low tem-
peratures may exacerbate UV stress to photosynthesis in spores from the sub-­
Antarctic population, whereas Antarctic spores would be adapted to low temperature
and UV. The results also confirm previous evidence obtained in adult thalli of Ulva
spp. from Antarctic and sub-Antarctic region by Rautenberger and Bischof (2006).
At 10 °C the inhibition of Ulva hookeriana (known as Enteromorpha bulbosa (Suhr)
Montagne) from Antarctica was comparable to its sub-Antarctic counterpart Ulva
clathrata (10% of control). However, at 0  °C, inhibition was of 50% in the sub-­
Antarctic Ulva clathrata and 37% in U. hookeriana (Rautenberger and Bischof
2006). Overall, the results indicate that in cold-adapted species, stress tolerance can
be efficient, which allow many shallow sublittoral, and especially eulittoral species,
to thrive under extremely changing thermal conditions.
210 N. Navarro et al.

10.3.3  Other Environmental Stressors

In the Antarctic environment, seaweeds are also facing fluctuations of other envi-
ronmental factors such as salinity, influenced by local meltwater influx and calving
glaciers as well as desiccation when algae are exposed to air during low tides.
Although the effects of salinity on seaweeds are relatively well known (reviewed in
Kirst 1990 and Karsten 2012), few studies have been conducted on Antarctic sea-
weeds (e.g., Jacob et al. 1991, 1992a, b; Karsten et al. 1991a, b). In general, it has
been reported that seaweeds respond to external salinity changes with osmotic accli-
mation processes involving the control of internal organic (e.g., proline, sucrose,
β-dimethylsulphoniopropionate) and inorganic (K+, Na+, Mg2+, Cl−, SO42−, and
PO43−) ions (Karsten et al. 1991a, b; Kirst 1990). Antarctic seaweeds inhabiting the
eulittoral and supralittoral zone can be characterized as euryhaline organisms,
which can survive salinities between 7 and 102 PSU with a low rate of mortality.
Most taxa grow, photosynthesize, and respire optimally under normal seawater con-
ditions with rather broad tolerances between 7 and 68 PSU. Hitherto, there is no
information of the effect of salinity on Antarctic seaweed propagules. On the other
hand, emergent stressors in Antarctic environment, e.g., ocean acidification (Hurd
et al. 2009) and marine pollution (Goutte et al. 2013), can pose risks to adult and
early phases of Antarctic seaweeds. Ocean acidification can affect the physiology of
seaweeds; however, practically no data exist on their effects on early phases of mac-
roalgae. In the giant kelp Macrocystis pyrifera, pH between 7.59 and 7.60 reduced
meiospore germination, which was ameliorated when CO2 was added (Roleda et al.
2011). Hitherto there is no information on the effects of these compounds on the
biology of Antarctic propagules.
It has been suggested that metals may inhibit reproduction in brown algae by
interfering with the ability of sperm to find eggs, perhaps via interference of the
pheromone attractant (Maier 1993). However, the effect of trace metals is expected
to be detrimental to propagules (spores, gametes, and zygotes) due to poor develop-
ment of the protective cell wall. Moreover, cell walls of brown seaweeds composed
of alginate and fucoidan can bind cations and have a high affinity for copper (Lignell
et al. 1982), affecting the settlement and germination of propagules. For example, in
Lessonia, copper drastically affected spore release by mature sporophytes as well as
spore settlement. The highest copper concentration applied interrupted the develop-
ment of the spores totally after settlement (Contreras et al. 2007). In all, the impor-
tance of studying the effects of metals and other pollutants (hydrocarbons, pesticides,
other persistent pollutants, and so on) on Antarctic algae propagules lies in the
recent increase of contaminant concentration in Antarctic due to human activities
(Bargagli 2008). On the other hand, although the harmful effects of metals, e.g.,
copper toxicity, have been analyzed in brown species (reviewed by Coelho et al.
2000; Contreras et al. 2007), the effects of these new, emergent stressors on the biol-
ogy of Antarctic seaweeds have to be examined in a context of the combined action
of multiple factors (see Chap. 7 by Huovinen and Gómez).
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 211

10.4  C
 oncluding Remarks: Biology of Propagules under
Climate Change

Despite having a crucial importance in the biology of seaweeds, propagules have


not been sufficiently studied in relation to their physiological requisites to respond
to climate change. The importance of understanding the effects of global climate
change on reproductive stages lies in the fact that early stages of development are
essential for recruitment, especially for those species that rely their dominance
entirely on reproductive abilities.
The predicted increase of temperature and the prevalence of episodes of depleted
ozone around the Antarctic Peninsula region and adjacent islands will impose phys-
iological constraints to reproductive output, settlement, and recruitment of different
species of seaweeds. Increase in seawater temperature could also influence the phe-
nology and the formation of propagules (spores and/or gametes) and consequently,
the timing and formation of juvenile thalli, especially in species inhabiting the eulit-
toral and shallow sublittoral zone (Zacher et  al. 2007; Campana et  al. 2009; see
Chap. 12 by Campana et  al.). Furthermore, as a consequence of temperature
increase, glaciers can retreat opening new free space for recruitment of benthic
organisms, including macroalgae (Quartino et  al. 2013; see Chap. 8 by Quartino
et al.). In these new open areas, however, alteration in light, salinity, sedimentation,
and disturbance processes can occur, limiting settlement of established communi-
ties and even favoring the arrival of cold-temperate species (see also Chap. 9 by
Deregibus et al.). Increased turbidity can have, however, contrasting implications
for the biology of reproductive cells, which can become favored by a minimized
impact of UV radiation, but decreasing available irradiance for photosynthesis.
Undoubtedly, Antarctic seaweeds have developed life strategies to colonize and
form a complex structure in the coastal ecosystems. In scenarios of climate change
and warming in the Antarctic, dispersal and colonization of Antarctic coastal zones
via efficient adaptations of early developmental phases of seaweeds are central to
envision the future seaweed diversity in Antarctica (see Chap. 2 by Oliveira et al.
and Chap. 5 by Pellizzari et al.).

Acknowledgments This work was supported by  CONICYT-Chile: Fondecyt Iniciación


N°11160520, Anillo ART1101, and Centro Fondap-IDEAL 15150003.

References

Amsler C, Reed D, Neushul M (1992) The microclimate inhabited by macroalgal propagules. Eur
J Phycol 27:253–270
Bargagli R (2008) Environmental contamination in Antarctic ecosystems. Sci Total Environ
400:2012–2226
Becker S, Graeve M, Bischof K (2010) Photosynthesis and lipid composition in the Antarctic
rhodophyte Palmaria decipiens: effects of changing light and temperature levels. Polar Biol
33:945–955
212 N. Navarro et al.

Beer S, Björk M, Gademan R, Ralph P (2001) Measurements of photosynthesis rate in seagrass.


In: Short FT, Cole R (eds) Global seagrass research methods. Elsevier Publishing, Amsterdam,
pp 183–198
Bischof K, Gómez J, Molis M et  al (2006) UV radiation shapes seaweed communities. Rev
Environ Sci Biotechnol 5:141–166
Bouchard JN, Roy S, Campbell DA (2006) UVB effects on the photosystem II-D1 protein of phy-
toplankton and natural phytoplankton communities. Photochem Photobiol 82:936–951
Britt AB (1995) DNA damage and repair in plants. Annu Rev Plant Physiol Plant Mol Biol
47:75–100
Campana GL, Zacher K, Fricke A et al (2009) Drivers of radiation colonization and succession in
polar benthic macro- and microalgal communities. Bot Mar 52:655–667
Clayton MN (1988) Evolution and life histories of brown algae. Bot Mar 31:379–387
Clayton MN (1994) The evolution of the Antarctic marine benthic algal flora. J Phycol 30:897–904
Clayton MN, Wiencke C (1990) The anatomy, life history and development of the Antarctic brown
alga Phaeurus antarcticus (Desmarestiales, Phaeophyceae). Phycologia 29:303–315
Coelho SM, Rijstenbil JW, Brown MT (2000) Impacts of anthropogenic stresses on the early
development stages of seaweeds. J Aquat Ecosyst Stress Recover 7:317–333
Contreras L, Medina MH, Andrade S et al (2007) Effects of copper on early developmental stages
of Lessonia nigrescens Bory (Phaeophyceae). Environ Pollut 145:75–83
Cormaci M, Furnari G, Scammacca B (1992) The benthic alga flora of Terra Nova Bay (Ross Sea,
Antarctica). Bot Mar 35:541–552
Crame JA (1993) Latitudinal range fluctuations in the marine realm through geological time.
Trends Ecol Evol 8:162–266
Davison IR (1991) Environmental effects on algal photosynthesis: temperature. J Phycol 27:2–8
de Reviers B, Délépine R (1981) Biological data on a southern Phaeophyceae Adenocystis utricu-
laris (Bory) Skottsberg, possible material for alginate industry. Proc 10th Int Seaweed Symp,
p 345–350
Dummermuth AL, Wiencke C (2003) Experimental investigation of seasonal development in six
Antarctic red macroalgae. Antarct Sci 15:449–457
Eggert A (2012) Seaweed responses to temperature. In: Wiencke C, Bischof K (eds) Seaweed biol-
ogy: novel insights into ecophysiology, ecology and utilization. Ecological studies, vol 219.
Springer, Berlin, pp 47–66
Eggert A, Wiencke C (2000) Adaptation and acclimation of growth and photosynthesis of five
Antarctic red algae to low temperatures. Polar Biol 23:609–618. https://doi.org/10.1007/
s003000000130
Gerstein AC, Cleathero LA, Mandegar MA, Otto SP (2010) Haploids adapt faster than diploids
across a range of environments. J Evol Biol 24:531–540
Gómez I, Wiencke C (1996) Photosynthesis, dark respiration and pigment contents of gametophytes
and sporophytes of the Antarctic brown alga Desmarestia menziesii. Bot Mar 39:149–157
Gómez I, Wiencke C (1997) Seasonal growth and photosynthetic performance of the Antarctic
macroalga Desmarestia menziesii (Phaeophyceae) cultured under fluctuating Antarctic day-
lengths. Bot Acta 110:25–31
Gómez I, Thomas DN, Wiencke C (1995) Longitudinal profiles of growth, photosynthesis and
light independent carbon fixation in the Antarctic brown alga Ascoseira mirabilis. Bot Mar
38:157–164
Gómez I, Wiencke C, Thomas DN (1996) Variations in photosynthetic characteristics of the
Antarctic marine brown alga Ascoseira mirabilis in relation to thallus age and size. Eur J
Phycol 31:167–172
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, metabolic
carbon balance and zonation of sublittoral macroalgae from King George Island (Antarctica).
Mar Ecol Prog Ser 148:281–229
Gómez I, Wulff A, Roleda M et al (2009) Light and temperature demands of marine benthic micro-
algae and seaweeds in polar regions. Bot Mar 52:593–608
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 213

Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27
Goutte A, Chevreuil M, Alliot F et al (2013) Persistent organic pollutants in benthic and pelagic
organisms off Adélie land, Antarctica. Mar Pollut Bull 77:82–89
Grzymski J, Johnsen G, Sakshaug E (1997) The significance of intracellular self-shading on the
bio-optical properties of brown, red and green macroalgae. J Phycol 33:408–414
Huovinen P, Gómez I (2012) Cold-temperate seaweed communities of the Southern Hemisphere.
In: Wiencke C, Bischof K (eds) Seaweed biology: Novel insights into ecophysiology, ecology
and utilization. Ecological studies, vol 219. Springer, Berlin, pp 293–313
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
Huovinen P, Gómez I (2015) UV sensitivity of vegetative and reproductive tissues of three
Antarctic macroalgae is related to differential allocation of phenolic substances. Photochem
Photobiol 91:1382–1388
Huovinen PS, Oikari AOJ, Soimasuo MR, Cherr GN (2000) Impact of UV radiation on the early
development of the giant kelp (Macrocystis pyrifera) gametophytes. Photochem Photobiol
72:308–313
Huovinen P, Ramírez J, Gómez I (2016) Underwater optics in sub-Antarctic and Antarctic coastal
ecosystems. PLoS One 11(5):e0154887
Hurd CL, Hepburn CD, Currie KI et al (2009) Testing the effects of ocean acidification on algal
metabolism: considerations for experimental designs. J Phycol 45:1236–1251
Jacob A, Kirst GO, Wiencke C, Lehmann H (1991) Physiological responses of the Antarctic green
alga Prasiola crispa ssp. antarctica to salinity stress. J Plant Physiol 139:57–62
Jacob A, Lehmann H, Kirst GO, Wiencke C (1992a) Changes in the ultrastructure of Prasiola
crispa ssp. antarctica under salinity stress. Plant Biol 105:41–46
Jacob A, Wiencke C, Lehmann H, Kirst GO (1992b) Physiology and ultrastructure of desiccation
in the green alga Prasiola crispa from Antarctica. Bot Mar 35:297–303
Kain JM (1989) The seasons in the subtidal. Eur J Phycol 24:203–215
Karsten U (2012) Seaweed acclimation to salinity and desiccation stress. In: Wiencke C, Bischof K
(eds) Seaweed biology: Novel insights into ecophysiology, ecology and utilization. Ecological
studies, vol 219. Springer, Berlin, Heidelberg, pp 87–107
Karsten U, Wiencke C, Kirst GO (1991a) The effect of salinity changes upon physiology of eulit-
toral green macroalgae from Antarctica and southern Chile. I. Cell viability, growth, photosyn-
thesis and dark respiration. J Plant Physiol 138:667–673
Karsten U, Wiencke C, Kirst GO (1991b) The effect of salinity changes upon physiology of eulit-
toral green macroalgae from Antarctica and southern Chile. II. Inorganic ions and organic com-
pounds. J Exp Bot 42:1533–1539
Karsten U, Sawall T, West J, Wiencke C (2000) Ultraviolet sunscreen compounds in epiphytic red
algae from mangroves. Hydrobiologia 432:159–171
Karsten U, Wulff A, Roleda MY, Müller R, Steinhoff F, Fredersdorf J, Wiencke C (2009)
Physiological responses of polar benthic micro- and macroalgae to ultraviolet radiation. Bot
Mar 52:639–654
Kirst GO (1990) Salinity tolerance of eukaryotic marine algae. Annu Rev Plant Physiol Plant Mol
Biol 41:21–53
Kornmann P, Sahling PH (1977) Meeresalgen von Helgoland. Benthische Grün-, Braun-und
Rotalgen. Helgolander Meeresun 29:1–289
Kylin H, Skottsberg C (1919) Zur Kenntnis der Subantarktischen und Antarktischen Meeresalgen,
II.  Rhodophyceen. In: Nordenskjöld O (ed) Wissenschaftliche Ergebnisse der schwedischen
Südpolarexpedition 1901–1903, Band IV, Lieferung 15. Stockholm, Lithographisches Institut
des Generalstabs, 88 pp
Lamb IM, Zimmerman MH (1977) Benthic marine algae of the Antarctic Peninsula. Antarct Res
Ser 23:129–229
214 N. Navarro et al.

Lignell A, Roomans GM, Pedersen M (1982) Localization of absorbed cadmium in Fucus vesicu-
losus by X-ray microanalysis. J Plant Physiol 105:103–109
Lorenz M, Schubert H, Forster RM (1997) In vitro- and in vivo-effects of ultraviolet-B radiation on
the energy transfer in phycobilisomes. Photosynthetica 33:517–527
Lüning K, Kadel P (1993) Daylength range for circannual rhythmicity in Pterygophora californica
(Alariaceae, Phaeophyta) and synchronization of seasonal growth by daylength cycles in sev-
eral other brown algae. Phycologia 32:379–387
Lüning K, tom Dieck I (1989) Environmental triggers in algal seasonality. Bot Mar 32:389–397
Maier I (1993) Gamete orientation and induction of gametogenesis by pheromones in algae and
plants. Plant Cell Environ 16:891–907
Marcías ML, Deregibus D, Saravia LA et  al (2017) Life between tides: spatial and temporal
variations of an intertidal macroalgal community at Potter Peninsula, South Shetland Islands,
Antarctica. Estuar Coast Shelf Sci 187:193–203
Müller DG, Westermeier R, Peters A, Boland W (1990) Sexual reproduction of the Antarctic
brown alga Ascoseira mirabilis (Ascoseirales, Phaeophyceae). Bot Mar 33:251–255
Navarro NP, Huovinen P, Gómez I (2016) Stress tolerance of Antarctic macroalgae in the early life
stage. Rev Chil Hist Nat 89:5
Navarro NP, Huovinen P, Gómez I (2019) Photosynthetic characteristics of geographically disjunct
seaweeds: a case study on the early life stages of Antarctic and Subantarctic species. Prog
Oceanogr 174:28–36
Otto SP, Gerstein AC (2008) The evolution of haploidy and diploidy. Curr Biol 18:R1121–R1124
Quartino ML, Deregibus D, Campana GL et  al (2013) Evidence of macroalgal colonization
on newly ice-free areas following glacial retreat in Potter Cove (South Shetland Islands),
Antarctica. PLoS One 8(3):e58223
Raper JR, Flexer AS (1970) The road to diploidy with emphasis on a detour. Symposium of the
Society of General Microbiology 20:401–432
Rautenberger R, Bischof K (2006) Impact of temperature on UV-susceptibility of two Ulva
(Chlorophyta) species from Antarctic and Subantarctic regions. Polar Biol 28:988–996
Renger G, Voss M, Graber P, Schulze A (1986) Effect of UV irradiation on differential partial
reactions of the primary processes of photosynthesis. In: Worrest RC, Caldwell MM (eds)
Stratospheric ozone reduction, solar ultraviolet radiation and plant life. NATO ASI series, vol
G8. Springer, Heidelberg, pp 171–184
Richter M, Rühle W, Wild A (1990) Studies on the mechanism of photosystem II photoinhibition
I. a two-step degradation of D1-protein. Photosynth Res 24:229–235
Roleda M, Zacher K, Wulff A et al (2007) Photosynthetic performance, DNA damage and repair in
gametes of the endemic Antarctic brown alga Ascoseira mirabilis exposed to ultraviolet radia-
tion. Austral Ecol 32:917–926
Roleda MY, Zacher K, Wulff A et  al (2008) Susceptibility of spores of different ploidy levels
from Antarctic Gigartina skottsbergii (Gigartinales, Rhodophyta) to ultraviolet radiation.
Phycologia 47:361–370
Roleda MY, Campana G, Wiencke C et al (2009) Sensitivity of Antarctic Urospora penicilliformis
(Ulotrichales, Chlorophyta) to ultraviolet radiation is life stage dependent. J Phycol 45:600–609
Roleda MY, Morris JN, McGraw CM, Hurd CL (2011) Ocean acidification and seaweed repro-
duction: increased CO2 ameliorates the negative effect of lowered pH on meiospore germina-
tion in the giant kelp Macrocystis pyrifera (Laminariales, Phaeophyceae). Glob Chang Biol
18:854–864
Salles S, Aguilera J, Figueroa FL (1996) Light field in algal canopies: changes in spectral light
ratios and growth of Porphyra leucosticta. Thur. In Le Jol. Sci Mar 60:29–38
Skottsberg C (1953) On two collections of Antarctic marine algae. Arkiv Botanik Serie 2 2:531–566
Swanson AK, Druehl LD (2000) Differential meiospore size and tolerance of ultraviolet light
stress within and among kelp species along a depth gradient. Mar Biol 136:657–664
Tom Dieck I (1989) Vergleichende Untersuchungen zur Ökophysiologie und Kreuzbarkeit
innerhalb der digitaten Sektion der Gattung Laminaria. PhD thesis, University of Hamburg,
Hamburg, Germany, 168 pp
10  Life History Strategies, Photosynthesis, and Stress Tolerance in Propagules… 215

Vass I (1997) Adverse effects of UV-B light on the structure and function of the photosynthetic
apparatus. In: Pessarakli M (ed) Handbook of photosynthesis. Marcel Dekker Inc., New York,
pp 931–949
Weykam G, Wiencke C (1996) Seasonal photosynthetic performance of the endemic Antarctic alga
Palmaria decipiens (Reinsch) Ricker. Polar Biol 16:357–361
Weykam G, Thomas DN, Wiencke C (1997) Growth and photosynthesis of the Antarctic red algae
Palmaria decipiens (Palmariales) and Iridaea cordata (Gigartinales) during and following
extended periods of darkness. Phycologia 36:395–405
Wiencke C (1990a) Seasonality of brown macroalgae from Antarctica – a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600
Wiencke C (1990b) Seasonality of red and green macroalgae from Antarctica – a long-term culture
study under fluctuating Antarctic daylengths. Polar Biol 10:601–607
Wiencke C, Clayton MN (1990) Sexual reproduction, life history, and early development in cul-
ture of the Antarctic brown alga Himantothallus grandifolius (Desmarestiales, Phaeophyceae).
Phycologia 29:9–18
Wiencke C, Clayton MN (2002) In: Wägele JW, Sieg J (eds) Antarctic seaweeds. Synopses of the
Antarctic benthos, vol 9. Gantner, Ruggell, Lichtenstein. 239 pp
Wiencke C, Stolpe U, Lehmann H (1991) Morphogenesis of the brown alga Desmarestia antarc-
tica cultivated under seasonally fluctuating Antarctic daylengths. Ser Cient INACH 41:65–78
Wiencke C, Clayton MN, Schulz D (1995) Life history, reproductive morphology and develop-
ment of the Antarctic brown alga Desmarestia menziesii J. Agardh. Plant Biol 108:201–208
Wiencke C, Clayton MN, Langreder C (1996) Life history and seasonal morphogenesis of the
endemic Antarctic brown alga Desmarestia anceps Montagne. Bot Mar 39:435–444
Wiencke C, Gómez I, Pakker H, Flores-Moya A, Altamirano M, Hanelt D, Bischof K, Figueroa
F-L (2000) Impact of UV radiation on viability, photosynthetic characteristics and DNA of
brown algal zoospores: implications for depth zonation. Mar Ecol Prog Ser 197:217–229
Wiencke C, Roleda MY, Gruber A et al (2006) Susceptibility of zoospores to UV radiation deter-
mines upper depth distribution limit of Arctic kelps: evidence through field experiments. J Ecol
94:455–463
Wiencke C, Clayton MN, Gómez I et al (2007) Life strategy, ecophysiology and ecology of sea-
weeds in polar waters. Rev Environ Sci Biotechnol 6:95–126
Wiencke C, Gómez I, Dunton K (2009) Phenology and seasonal physiological performance of
polar seaweeds. Bot Mar 52:585–559
Wünschmann G, Brand JJ (1992) Rapid turnover of a component required for photosynthe-
sis explains temperature dependence and kinetics of photoinhibition in a cyanobacterium,
Synechococcus 6301. Planta 186:426–433
Wynne MJ (1982) Observations on four species of Delesseriaceae (Rhodophyta) from the South
Sandwich Islands, the Antarctic. Contr Univ Mich Herb 15:325–337
Zacher K (2014) The susceptibility of spores and propagules of Antarctic seaweeds to UV and
photosynthetically active radiation  — field versus laboratory experiments. J Exp Mar Biol
Ecol 458:57–63
Zacher K, Roleda MY, Hanelt D, Wiencke C (2007) UV effects on photosynthesis and DNA in
propagules of three Antarctic seaweeds (Adenocystis utricularis, Monostroma hariotii and
Porphyra endiviifolium). Planta 225:1505–1516
Zacher K, Roleda MY, Wulff A et al (2009) Responses of Antarctic Iridaea cordata (Rhodophyta)
tetraspores exposed to ultraviolet radiation. Phycol Res 57:186–193
Chapter 11
Form and Function in Antarctic Seaweeds:
Photobiological Adaptations, Zonation
Patterns, and Ecosystem Feedbacks

Iván Gómez and Pirjo Huovinen

Abstract  Morpho-functional traits of Antarctic seaweeds are modeled by different


physical and biological factors. Due to the extreme seasonality, which imposes light
limitation for extended periods, Antarctic seaweeds are shade-adapted organisms
that are physiologically able to thrive at considerable depths down to 40 m. This
vertical distribution is defined by a suite of bio-optical and morphological features
that allow algae occupying habitats with different environmental conditions in the
water column. However, various species can also colonize the highly perturbed
intertidal zone where environmental setting, e.g. ice scouring, high solar radiation,
extremely variable temperature, limit growth, and reproduction. In the maritime
Antarctic region, large endemic brown algae attaining a massive (leathery) mor-
phology and perennial life history dominate at depths below 10 m or less. Here, they
coexist with perennial highly shade-adapted coarsely branched rhodophytes, which
show understory characteristics. At shallower locations, various annual species with
very rapid growth can be found. The intertidal zone, characterized by a depauperate
diversity, is populated mostly by ephemeral and delicate green algae. In the present
chapter, form and function of seaweeds is revisited in the context of a changing
Antarctic environment. Here, the functional groups display different acclimation
mechanisms, which can operate at different temporal scales and consequently with
variable impact on the biogeochemical coastal processes. The role of canopy-­
forming algae, whose “bioengineering” processes alleviate the impact of environ-
mental variability, is fundamental in determining the fate of the benthic communities
in the coastal system.

Keywords  Canopy-forming algae · Life strategy · Light absorption ·


Morpho-­functional traits · Vertical zonation

I. Gómez (*) · P. Huovinen


Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: igomezo@uach.cl; pirjo.huovinen@uach.cl

© Springer Nature Switzerland AG 2020 217


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_11
218 I. Gómez and P. Huovinen

11.1  Brief Overview of Form and Function in Seaweeds

In contrast to unicellular algae, organisms bearing multiple cells (from simple plu-
ricellular colonies to more advanced forms attaining, e.g. parenchyma) have to syn-
chronize a more complex structural organization, characterized by different
function-specific elements that follow a morpho-genetic-based program. Thus, mul-
ticellular algae adjust form and function through an intricate molecular network that
allows them to interact with their physical, chemical, and biotic environment
(Grosberg and Strathmann 2007). A suite of morphological models found in benthic
organisms (e.g. seaweeds, sponges, corals, bryozoa) coexist in a physical environ-
ment, which raises a question of how similar convergent forms have evolved even in
species phylogenetically very distant or, alternatively, how related organisms dis-
play completely different growth patterns and shape (Kaandorp and Kübler 2001).
A striking characteristic of almost all seaweeds is their morphological plasticity,
i.e. although the basic thallus plan is based on a determined morphogenetic design,
body shape can change through the life span of an organism or within the life his-
tory sequence. These variations in the morphological traits within a genotype can be
subtle or drastic depending on the intensity of the endogenous and environmentally
driven shifts (Innes 1984; Taylor and Hay 1984). Especially in sites where the phys-
ical perturbations are extreme, seaweeds display complex mechanisms to adjust
form and function to the prevailing environmental condition (Hay 1986). Phenotypic
plasticity, one of the most well-known types of intrinsic morphological variability,
is normally prompted by environmental conditions and thereby complicates efforts
to identify the routes of morpho-functional responses within a multi-specific assem-
blage: intrinsic properties at an organismal level mask the morpho-functional differ-
ences at community scale (Steneck and Dethier 1994). Changes in morphology due
to ontogenetic development and heteromorphic phase expression within of life
cycle are also important to characterize form and function in seaweeds. For exam-
ple, in the brown alga Himantothallus grandifolius, the largest Antarctic seaweed,
the thallus undergoes considerable changes with development: while juvenile indi-
viduals are characterized by partial cortication and coarsely branched morphology,
adult plants are characterized by a thick leathery strap-like anatomical structure,
where lateral branches are absent (Moe and Silva 1981; Wiencke and Clayton 1990).
Traditionally, functional groups of seaweeds (which could also be applied to
other groups of benthic marine organisms) are defined by their thallus architecture
(also called life form). This concept implies intuitively a series of intrinsic proper-
ties of an organism, which can or not be shared by other unrelated organisms.
Clearly this gives a high value to the anatomical features (form) and less emphasis
on the function. Such conceptual framework represented the basis on, which Littler
and Littler (1980) and Steneck and Watling (1982) developed their general func-
tional form models, where functionality of algae, e.g. resistance to biotic distur-
bance, was inferred from gross morphology. Although the general applicability of
these models has been questioned since similar morphologies often show different
functional responses to, e.g. disturbance gradients (Phillips et al. 1997; Ingólfsson
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 219

2005; Padilla and Allen 2000), functionality, which depends on extrinsic factors
(e.g. resource utilization, disturbance, biotic interactions, etc.), has an unavoidable
expression in the morphology. The rationale to understand in what ways form and
function is modeled by the environment, known as “the holy grail framework”,
considers necessarily different areas of knowledge, e.g. genomics, physiology, ecol-
ogy, demography, etc., and has been used with different emphasis to explain the
structure of different types of vegetation, both in terrestrial and aquatic realms
(Littler and Littler 1980; Grime 1981; Steneck and Dethier 1994; Lavorel and
Garnier 2002). Thus, if one assumes that different habitats have a different environ-
mental setting, it is possible to argue that they host assemblages of organisms with
similar morphology but different functional attributes. For example, in littoral
stress, tolerance of macroalgae depends strongly on fine photochemical adjust-
ments, which are related to their position on the shore and less with functional form
groups (Aguilera et al. 1999; Gómez and Huovinen 2011; Balata et al. 2011). This
has been commonly found in different types of terrestrial vegetation where environ-
mental tolerance, and not gross morphology, defines functional groups (Grime
1981; Ackerly and Reich 1999; Poorter and Bongers 2006).

11.2  Functional Groups of Seaweeds in the Antarctic

The coastal systems around Antarctica may be regarded as highly inhospitable for
life, where physical disturbance associated with ice scouring, extreme light limita-
tion, and low temperatures imposes severe restrictions for marine organisms.
However, benthic algae thrive in these habitats displaying different functional strat-
egies and morphologies (Fig. 11.1). Based on various surveys, 131 species of sea-
weeds (Fig.  11.2) are distributed  among different  types of functional groups:
filamentous and finely branched (45); foliose (9); coarsely branched, including cor-
ticated species (48); thick leathery, including terete forms (11); and postrate species
(18) (Fig. 11.2). However, when the different functional forms are grouped accord-
ing to the major phylogenetic categories, it is possible to observe that 64% of green
algae are filamentous, while practically the totality of thick leathery forms belong to
brown algae. In red algae, 58% of the known species can be recognized as finely and
coarsely branched morphs. In the case of Chlorophyta, most of the species attain del-
icate filamentous or sheet-like morphs, and with the exception of the endemic
Lambia antarctica and Monostoma hariotii (Wiencke and Clayton 2002; see also
Chap. 2 by Oliveira et al. and Chap. 5 by Pellizzari et al.), all are restricted to inter-
tidal zones. However, it should be emphasized that many species cannot be easily
assigned to these major functional categories. For example, the brown algae
Adenocystis utricularis and Utriculidium durvillaei are the only species with a sac-
cate morphology (and thus were not included in this analysis). Moreover, the num-
ber of postrate species, which may include crustose, calcareous, and endophytic life
forms, is largely underrepresented. In fact, these algae have been very little studied,
220 I. Gómez and P. Huovinen

Fig. 11.1  Diversity of gross morphologies in Antarctic seaweeds. (a) Crustose Rhodophyta; (b)
Saccate morphology (Adenocystis utricularis); (c) Filamentous tubular (Ulva intestinalis); (d)
finely branched Plocamium cartilagineum; (e) coarsely branched (Trematocarpus antarcticus); (f)
Thick leathery (Ascoseira mirabilis). (Photos a, d, e and f by Ignacio Garrido; b and c by
Iván Gómez)

due to mostly that they are not ubiquitous (e.g. encrusting morphs) or can inhabit
deeper locations (e.g. calcified coralline algae) (Alongi et al. 2002).
The absence of a marked dominance of a given seaweed gross morphology in the
Antarctic can only be explained in terms of the distribution of these different func-
tional forms in the mosaic of benthic habitats. In fact, the arrangement of species
across different environmental gradients implies also an ordination of organismal
traits that can be classified in different functional entities (e.g. gross morphology,
size, life forms, physiological responses). Thus, it is possible to understand why
similar morpho-functional “solutions” are exploited by different taxa, many of them
phylogenetically unrelated. For example, of the 25 species of green algae recorded
in the Antarctic, 56% correspond to widely distributed taxa, which normally display
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 221

Fig. 11.2  Major functional-form groups in different divisions of Antarctic seaweeds based on dif-
ferent surveys (Clayton et al. 1997; Wiencke and Clayton 2002; Hommersand et al. 2009; Ramírez
2010; Charles Amsler, personal communication)
222 I. Gómez and P. Huovinen

Fig. 11.3  Antarctic seaweeds and their organization in relation to form and function, size, geo-
graphic affinity, and taxonomy. Number of taxa extracted from Fig. 11.2

filamentous or sheet-like forms (Fig. 11.3). According to Gómez et al. (2019), form


and function and biogeographic affinity are highly correlated in Antarctic green
algae. In contrast, only 12% of Rhodophyta can be regarded as widely distributed,
contrasting with the high prevalence of endemic (42%) and Antarctic/sub-Antarctic
taxa (46%). Similar pattern can be observed in brown algae where endemic and
Antarctic/sub-Antarctic species (characterized by thick leathery and saccate
morphs) account by 88% of the total numbers of recorded taxa (Fig. 11.2).

11.3  T
 he Vertical Zonation of Antarctic Seaweeds:
A Paradigm of Spatial Distribution of Different
Morpho-functional Traits

Knowledge on zonation and in general the structure of the submarine landscape in


the shallow sublittoral in the Antarctic began to increase in the 1960s and 1970s,
along with scuba diving–based surveys (Neushul 1965; Delépine et  al. 1966;
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 223

Zaneveld 1966; DeLaca and Lipps 1976; Lamb and Zimmerman 1977). These stud-
ies across different geographical zones demonstrated that the vertical distribution of
macroalgae could be defined in terms of functional groups, which apparently do not
follow uniform patterns, principally due to differences in latitude, substrate, influ-
ence of ice, associated fauna, etc. The vertical distribution of Antarctic seaweeds
has been much more studied in the Western Antarctic Peninsula and adjacent
islands, an eco-region known as the Maritime Antarctic. Due to the relatively milder
climatic conditions, seaweed assemblages reach their maximal development in
terms of abundance and diversity in the north-western part of the Antarctic Peninsula,
decreasing the macroalgal biodiversity towards the southern regions (Wiencke et al.
2014; Mystikou et al. 2014). The zonation in the Maritime Antarctic can be charac-
terized by a dominance of large canopy-forming endemic species of the order
Desmarestiales (Desmarestia menziesii, D. anceps, and Himantothallus grandifo-
lius) between 10 and 40 m or greater depth (Fig. 11.4a). These three species have
thick leathery and terete gross morphology and can alternate their dominance
depending on the substrate characteristics, whose consolidation can vary consider-
ably depending on closeness to glaciers, slope of the vertical profile, terrestrial run-
off, etc. (Klöser et al. 1994). Coexisting at this level, it is possible to found delicate
understory red algae, e.g. Myriogramme, Gymnogongrus, and Georgiella (Amsler
et al. 1995). Between 0 and 5 m depth, a zone marked by ice abrasion and waves,
the substrate is colonized by fast-growing species, algae with an ability for re-sprout
from basal shoots and crustose forms. In contrast, the intertidal rocky shores are
dominated by ephemeral, turf species, mainly filamentous Chlorophyceans (e.g.
Urospora, Ulva, Ulothrix) and the saccate brown alga Adenocystis utricularis
(Huovinen and Gómez 2013; Marcías et al. 2017).
In areas outside the Western Antarctic Peninsula, e.g. around the Ross Sea and
some sites along the East Antarctica, the diversity and abundance of seaweeds
decreases and their vertical distribution is much more constrained by available sub-
strata and the longer permanence of sea ice cover (Zaneveld 1966; Miller and Pearse
1991; Gambi et  al. 1994; Johnston et  al. 2007; Clark et al. 2011). In these sites,
although some large Desmarestiales (e.g. H. grandifolius and D. menziesii) can be
found at deeper locations, in general the coarsely branched red algae Iridaea cor-
data and Phyllophora antarctica are the dominant assemblages, especially at inter-
mediate depths (between 2 and 20 m) (Cormaci et al. 2000) (Fig. 11.4b). Another
particular feature of these ecosystems is the massive presence of crustose coralline
red algae at deeper locations, especially of Phymatolithon foecundum (Hommersand
et al. 2009), which can cover >70% of the available substrate under the canopy of
red and brown algae (Irving et al. 2005). Remarkably, algae have to adapt to very
low light conditions for primary productivity, irrespective of their functional form
architecture. In fact, due to their extreme shade adaptation, these species can reach
considerable depths and live with <2% of surface irradiances (Schwarz et al. 2003,
2005; see also Chap. 7 by Huovinen and Gómez).
224 I. Gómez and P. Huovinen

Fig. 11.4  Patterns of vertical distribution of seaweeds in Western Antarctic Peninsula (a) and
Eastern Antarctica (b) indicating major functional groups and photobiological processes. Synthetic
schemes from observations reported in Gambi et al. (1994), Johnston et al. (2007), and Huovinen
and Gómez (2013)
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 225

11.4  L
 ight Use Characteristics as a Major Factor
Delineating Physiological Thallus Anatomy of Seaweeds

The arrangement of different functional forms along the depth gradient is strongly
determined by different physico-chemical and biological factors. However, spatial
and temporal availability of light is probably the most relevant factor by which
zonation of Antarctic seaweeds can be explained. Because light governs the primary
processes of photosynthesis, and, hence, primary productivity and biomass forma-
tion, Antarctic seaweeds, irrespective of their morphological organization, display
efficient mechanisms of light harvesting. In fact, in the Antarctic, the marked sea-
sonality in light availability defines strongly an intrinsic shade adaptation of mac-
roalgae. Here the whole phenology of algae is closely tuned with the Antarctic light
regime, which exposes organisms to darkness in winter (Wiencke et  al. 2009).
However, although in summer algae can be exposed to very high doses of solar
radiation due to extended daylengths of up to 24 h at the highest latitudes, they do
not acclimate and retain the capacity for very low light requirements for metabolism
(Gómez and Huovinen 2015). This ability to use very low irradiances for photosyn-
thesis and an intrinsic positive metabolic carbon balance (an indicator of compensa-
tion of carbon losses due to respiration) (Gómez et al. 1997; Deregibus et al. 2016)
has important implications for the spatial dimension of the algal zonation: it allows
Antarctic algae to colonize shaded locations, especially deeper sites. As a conse-
quence, many Antarctic species can occupy extended ranges of depth and hence
different light fields (Gómez et al. 1997). This situation contrasts with zonation pat-
terns of various cold and temperate coasts, where the different algal groups are
arranged in well-defined “belts” (Lüning 1990). Light trapping, especially under
very limited conditions of irradiance, requires not only a specific pigment configu-
ration but also morphological features such as thickness and thallus translucency,
which are important in terms of absorptance of the different wavelengths (Gómez
and Huovinen 2011). Algae increase light trapping through their thallus architec-
ture, which can result in different in vivo spectral absorptance (Lüning and Dring
1985; Gómez et  al. 2019). In Fig.  11.5, the spectral characteristics of several
Antarctic macroalgae with different functional form and thickness are exemplified.
In the case of thick leathery and coarsely branched morphs (e.g. Himantothallus,
Desmarestia, Iridaea), attaining thallus thickness >500 μm, show high absorptance
practically along the whole spectrum. In contrast, delicate morphs with thick-
ness <100 μm, mostly foliose and finely branched such as Monostroma, Pantoneura,
Myriogramme, and Pyropia, show decreased absorptance between 550 and 650 nm.
Interestingly, some thin filamentous algae, e.g. Ulva intestinalis, Acrosiphonia, and
Urospora, can exhibit high absorptance at these wavelengths, which is related to
their turf arrangement, i.e. the overlapping of different filaments equals the several
cell layers of thicker algae. Overall, these patterns are related to algal taxonomy and
distribution in the zonation profile. In fact, thick leathery forms commonly belong
to the brown algae, and their efficient light absorption over an extended range of
wavelengths allows using the impoverished light field at higher depths (Gómez and
226 I. Gómez and P. Huovinen

Fig. 11.5  Patterns of spectral in vivo absorptance of several Antarctic seaweeds with different
morpho-functional organization. Examples of thallus gross morphology, cross section, and thick-
ness are indicated for selected species. Cross-sectional microphotographs courtesy of Nelso Navarro
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 227

Huovinen 2015). Similarly, understory coarsely branched and finely branched red
algae, inhabiting deep sites with high absorption between 400 and 500 nm, are well
suited to live at these depths (Gómez et al. 2019). Considering that, irrespective of
their thallus shape and taxonomical affinity, Antarctic seaweeds are shade-adapted
organisms, whose morphological and optical traits (e.g. thickness), pigment compo-
sition, and intrinsic photochemical capacities are superimposed in their vertical dis-
tribution (Huovinen and Gómez 2013; Gómez et al. 2019).

11.5  F
 orm and Function in the Context of Life Strategies
and Stress Tolerance

Since the publication of MacArthur and Wilson in 1967 on r and k selection, which
put into a context the evolutionary divergences of organisms in relation to the pro-
duction and care of offsprings, many studies have tried to expand these concepts to
other adaptive traits (concept discussed in Pianka 1970). Because r and k strategies
involve normally differences in energy allocation and, hence, body size, the form
and function concept could be easily integrated in the theory (Grime 1981). For the
case of marine macroalgae, in the 1980s, Joanna Kain used the terms “season
responder” and “season anticipator” to describe the different phenological responses
of seaweeds to seasonal changes in the environment (Kain 1987). Accordingly,
“season responders” correspond to organisms that grow and reproduce under
favourable environmental conditions, which could be analogous to r strategists,
while “season anticipators” are organisms whose development is triggered by envi-
ronmental factors at which they anticipate. The latter classification resembles
k-selection strategy. Although not a strict rule, most of the season responders identi-
fied in the literature seem to correspond to delicate, small-sized forms, which nor-
mally exhibit an opportunistic strategy. In contrast a number of season anticipators
can be associated with long-lived (perennial) seaweeds attaining normally large
thalli (Kain 1989). However, although intuitively one may argue that differences in
thallus complexity (including size) are correlated with differential responses to sea-
son, in general they are complex and depend on different environmental factors,
type of biological indicator (growth, photosynthesis, reproduction, etc.), age and
life history phase, and endogenous rhythms, all which can show complementary or
divergent patterns (Kain 1986; Lüning and Kadel 1993; see also Chap. 10 by
Navarro et al.).
In the Antarctic benthos, seaweeds are exposed to a marked seasonality, and thus,
the concepts of “responders” and “anticipators” could explain well the different
phenological patterns found in Antarctic seaweeds. In fact, various Antarctic spe-
cies, e.g. Iridaea cordata, Ulva intestinalis, Acrosiphonia arcta, and Adenocystis
utricularis, have been regarded as “season anticipators”, while Antarctic
Desmarestiales, Ascoseira mirabilis and Palmaria decipiens, can be considered as
“season anticipators” (Wiencke 1990a, b). Although these classifications were
228 I. Gómez and P. Huovinen

based mostly on growth responses to the Antarctic light regime, it has been shown
that photosynthetic light use characteristics can respond in the same seasonal man-
ner. For example, the brown alga Adenocystis utricularis and the red alga Iridaea
cordata, two species regarded as responders, maintain high photosynthetic func-
tionality still in winter, when light is very limited (Gutkowski and Maleszewski
1989; Weykam et  al. 1997). This strategy is completely different in large
Desmarestiales and Ascoseira mirabilis, as well as the red alga Palmaria decipiens,
which activate their photosynthetic apparatus during early spring to optimize the
available irradiance after the ice break-up (Gómez et al. 1995a, b; Wiencke et al.
2009). For large brown algae whose thalli can have length of various meters, these
responses have important morpho-functional implications: Firstly, photosynthesis
and growth during early spring are strongly synchronized to potentiate the use of
newly fixed and stored carbon. Secondly, there is spatial separation between carbon
production and sink zones with different metabolic activity, which can also be
exposed to very contrasting light fields. Because these massive thick leathery spe-
cies require compensation for the enhanced carbon burning due to high rates of dark
respiration during the rapid biomass formation, the lamina elongation is powered by
carbon stored in the previous season (Gómez and Wiencke 1998), similarly as in
high-latitude kelps (e.g. Laminaria, Saccharina) (Dunton and Schell 1986). A well-­
studied case is Ascoseira mirabilis, which grows through the action of an intercalary
meristem and presents “conducting channels” in medullary cell regions (Clayton
and Ashburner 1990; Gómez et al. 1995b). In this species, during the growth phase,
carbon stored as laminarin in distal parts is remobilized through the conducting
cells (normally as mannitol and some amino acids) towards the meristem to replen-
ish carbon substrates in the so-called light-independent carbon fixation (LICF) reac-
tions (Kremer 1981; Gómez and Huovinen 2012). Such morpho-functional strategies
have not only been demonstrated in large brown algae: in the perennial, coarsely
branched red alga Palmaria decipiens, LICF reactions accounting 9% of the total
fixed carbon have been reported (Weykam et al. 1997), suggesting that this type of
mechanisms are operating in algae with complex thallus anatomy and season antici-
pation phenology, which allow them thriving at high depths and under extreme sea-
sonality in the Antarctic.
Form and function in the context of stress tolerance have been revisited in the last
years. Interestingly, several anatomical traits related to resistance to physical distur-
bance, e.g. multilayered architecture, thickness, and large size, are also functional to
increase light trapping, e.g. efficient absorptance (Gómez and Huovinen 2011).
Recently it was claimed that populations of three species of Desmarestiales
(D. anceps, D. menziesii, and H. grandifolius) and Ascoseira mirabilis extending
between 5 and 30  m depth show similar photosynthetic characteristics along the
depth profile (Gómez and Huovinen 2015). However, not only the efficient and
highly conserved light use across different irradiances but also an intrinsic capacity
for UV stress tolerance was shown in these algae (see Chap. 7 by Huovinen and
Gómez). Although all these traits conferring UV shielding show a strong overlap-
ping with other factors, e.g. competence for space and overgrowth, scape from her-
bivores, there appears to exist a trade-off between photoprotection against enhanced
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 229

solar UV radiation, mostly due to an increased thallus cross section (low area/
weight ratio) and ultrastructural compounds, and highly efficient shade adaptation
(Gómez and Huovinen 2015). A key element explaining this feature in algae rarely
exposed to UV radiation is their constitutively high levels of phenolics (phlorotan-
nins) (Flores-Molina et al. 2016). These secondary metabolites in Antarctic brown
algae represent multifunctional compounds with putative roles in, e.g. resistance to
grazing, temperature, and UV radiation (Amsler et  al. 2005; Iken et  al. 2009;
Huovinen and Gómez 2013; Rautenberger et al. 2015) (for a description of func-
tional roles of phlorotannins in Antarctic seaweeds, see Chap. 17 by Amsler et al.
and Chap. 18 by Gómez and Huovinen).
In the case of delicate morphs, mostly filamentous and finely branched green
algae, the opportunistic life strategy of these organisms allows them to respond
rapidly to environmental stressors and, in virtue of their high metabolic rates per
weight, to restore the homeostasis at short term (Holzinger and Karsten 2013).
Albeit stress tolerance of seaweeds living in the intertidal zone would rely on highly
efficient metabolic adjustments (Holzinger and Lutz 2006; Karsten et  al. 2009;
Gómez and Huovinen 2011), some structural adaptations have been described. For
example, in Urospora penicilliformis a dense cell wall, presence of mucilage and
external mineral deposition provide efficient shielding from high solar radiation and
desiccation (Roleda et al. 2010). In many cases, filamentous green algae can form
mats or turf-like structures that are effective to minimize the harmful effects of
changing environment (Bischof et al. 2006). In all, in terms of photosynthetic char-
acteristics and physiological responses to stress, form and function of some Antarctic
seaweed assemblages have been related to biogeographic affinity and depth. Based
on 31 species from King George Island, three major groups of species were defined:
(a) coarsely branched Rhodophyta are mostly found at shallow subtidal sites and
have an Antarctic-sub-Antarctic origin; (b) endemic Antarctic brown algae are dom-
inant at depths between 10 and 30 m and practically all attain thick leathery mor-
phology; and (c) filamentous and sheet-like green algae, mostly intertidal species,
normally can be categorized as algae with wide geographic distribution (Gómez
et al. 2019).

11.6  F
 unctional Traits of Seaweeds and Properties
of Benthic Communities

Seaweeds in practically all cold-temperate and polar coastal ecosystems represent


foundational organisms, whose processes and fate determine key community indi-
cators, such as structure and functional and taxonomic richness (Chapman 1987;
Lüning 1990). In King George Island, the distribution and composition of different
functional groups in both intertidal and subtidal sites are regulated by different fac-
tors, which are defined by some species that account by 90% of the dissimilarities
between depth strata (Valdivia et al. 2014). When representative taxa are analysed,
230 I. Gómez and P. Huovinen

effects are scale dependent: variance components increase at the finer scale of varia-
tion (from centimeters  to meters) compared to shore level (hundreds of meters)
(Valdivia et al. 2014). In the intertidal system dominated by filamentous and finely
branched morphs, the grazing by the limpet Nacella concinna is probably one of the
most important biological interactions (Kim 2001; Segovia-Rivera and Valdivia
2016). Apart from green algae, N. concinna exerts control on periphyton, thus deter-
mining far-reaching ecological processes, e.g. the fate of re-colonization and suc-
cession in these systems (Campana et al. 2009; Valdivia et al. 2019; see also Chap.
12 by Campana et al. and Chap. 13 by Valdivia).
At the subtidal zone, facilitative interactions held by large brown algae through
bioengineering seem relevant for the structure and maintenance of the benthic com-
munities (Valdivia et al. 2015). These canopy-forming seaweeds are important as
they shelter other species of algae and invertebrates from harmful environmental
conditions and thus have an important effect on the community biomass of the
whole ecosystem (Valdivia et al. 2015; Ortiz et al. 2016; see also Chap. 15 by Momo
et al. and Chap. 16 by Ortiz et al.). However, in locations exposed to severe impact
of physical disturbance, small organisms can be favoured while canopy-forming
algae would be more sensitive (Smale 2007). For example, in eastern Antarctica
where ice cover can be considerably extended through spring, canopy-forming mac-
roalgae were only abundant at sites where sea-ice cover break-up occurs during
spring, but absent at sites that retained ice cover until summer (Johnston et al. 2007).
Thus, these organisms appear to respond slowly to the changing environment due,
for example, to enhanced warming. For example, in new ice-free areas originated
from glacier retreat where enhanced sediment input limits light penetration, estab-
lishment of large brown algae is highly constrained (Quartino et al. 2013). In these
highly perturbed sites, ice scouring and unconsolidated substrate affect consider-
ably the presence of canopy-forming algae and hence the taxonomic richness
(Klöser et al. 1994; Smale 2007; Smale et al. 2008; Valdivia et al. 2015). On the
other hand, environmental shifts driven by climate change can affect the morpho-­
functional responses of Antarctic species. For example, physiology of canopy-­
forming algae (e.g. Desmarestia spp.) may have consequences for the whole benthic
community (Schoenrock et al. 2015). In the case of crustose species, fleshy encrust-
ing forms (Hildenbrandia) could be favoured in scenarios of changing pH and tem-
perature compared to calcified Coralline species (Clathromorphum) (Schoenrock
et al. 2016). In general, morpho-functional and anti-stress mechanisms of macroal-
gae to cope with sharp physical gradients percolate towards upper hierarchies
through insurance of functional richness in the community, which set high degree of
resilience to physical perturbation (Ortiz et  al. 2016, 2017) or to minimize the
impact of alien species (Arenas et al. 2006; see Chap. 16 by Ortiz et al.).
Considering some functional form models for marine seaweeds, similitudes and
analogies with terrestrial vegetation strategies can be identified. For example,
according to the functional groups described by Grime (1981) for terrestrial vegeta-
tion, opportunistic green algae growing at the intertidal zone could correspond to
the “ruderal” species, permanently subjected to strongly physical perturbation. In
contrast, large endemic brown algae, which thrive in sites with lower physical
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 231

Fig. 11.6  CSR Grime’s triangle describing the main strategies and environmental trade-off of the
major functional groups of Antarctic seaweeds. (Photos Myriogramme and Desmarestia by Ignacio
Garrido; Urospora by Iván Gómez)

perturbation, can be analogue to “competitive species” in virtue of their exuberant


canopy and perennial characteristics. For many temperate ecosystems, the life his-
tory traits conferring advantages under high levels of disturbance are convergent in
different types of algal assemblages, suggesting that some patterns could be gener-
alized (Steneck and Dethier 1994). However, in Antarctic communities some fac-
tors associated with disturbance and stress require adjustment to the extreme
Antarctic conditions. In the conceptual framework in Fig. 11.6, three major func-
tional groups of Antarctic seaweeds (turf algae, dominated by intertidal green algae;
canopy-forming algae, especially large brown algae; and corticated red algae,
grouping many understory species) can be oriented through the three axes following
a Grime’s CSR triangle schema. Here, the extreme action of ice (perturbation), light
limitation (stress), and biomass (competition) dimensions determine the separation
among algal groups. Corticate red algae in virtue of their extreme shade adaptation
represent the stress tolerant group. Here, many crustose species growing at very low
light conditions in the eastern Antarctic can also be added to this group. In the per-
turbation axis, filamentous and finely branched green algae and some little saccate
brown algae (Adenocystis utricularis) exemplify the colonizers, well adapted to
occupy sites highly perturbed by ice, terrestrial run-off, and high solar radiation.
Under these conditions abundance and species richness are less influenced by bio-
logical interactions (Valdivia et  al. 2014; Segovia-Rivera and Valdivia 2016).
232 I. Gómez and P. Huovinen

Finally, the canopy-forming algae, represented by species of the order Desmarestiales,


and Ascoseira and Cystosphaera that exhibit high biomass production, are dominat-
ing at sites with lower physical perturbation. However, they show competitive abili-
ties for light and substrate (Gómez et al. 1997; Valdivia et al. 2015).

11.7  Concluding Remarks

The main ecological expression of the morpho-functional adaptation of Antarctic


seaweeds is the macroalgal zonation, which is not only a vertical arrangement of
species but also represents an ordination of organismal traits that can be classified
in different functional entities (e.g. gross morphology, life forms, physiological
responses). These attributes can be scaled up to community structure and ecosystem
functioning. The concept, well studied in plants, has been revitalized in the last
decade in the context of the contemporary climate change.
Due to the seasonally changing light conditions, characterized in the highest lati-
tudes by several months of very dim light, Antarctic seaweeds are adapted to very
low light levels. In contrast, after the ice break-up in spring, they suddenly can be
exposed to strong solar radiation. Thus the adaptations of Antarctic algae are finely
tuned with the daylength, changes in water turbidity, and ice perturbations. This
environmental variability is fully exploited by seaweeds in virtue of their efficient
morpho-functional adaptations. However, due to climate change, the environmental
settings in which Antarctic seaweeds have evolved for millions of years are chang-
ing. In these new scenarios, the adaptive capacities of these organisms as well as the
ecosystem functions they provide will be challenged (Constable et al. 2014; Gutt
et al. 2015). Although one can recognize that polar seaweeds are particularly sus-
ceptible to these changes with unpredictable consequences for the whole coastal
ecosystem, we have still a limited understanding on how physiological and morpho-
logical traits respond and how they will be integrated in, for example, molecular
mechanisms of environmental tolerance and stress resilience.

Acknowledgements The authors acknowledge the financial support from Conicyt Chile


(FONDAP 15150003, Anillo-PIA ART1101, Fondecyt 1161129) and INACH T-20-09 from the
Instituto Antártico Chileno. The helpful assistance and collaboration of the members of our labo-
ratories at Universidad Austral de Chile as well as the staff of the Instituto Antártico Chileno for
during various Antarctic expeditions are acknowledged.

References

Ackerly DD, Reich PB (1999) Convergence and correlations among leaf size and function in seed
plants: a comparative test using independent contrasts. Am J Bot 86:1272–1128
Aguilera J, Karsten U, Lippert H, Vögele B, Philipp E, Hanelt D, Wiencke C (1999) Effects of solar
radiation on growth, photosynthesis and respiration of marine macroalgae from the Arctic. Mar
Ecol Prog Ser 191:109–119
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 233

Alongi G, Cormaci M, Furnari G (2002) The Corallinaceae (Rhodophyta) from the Ross Sea
(Antarctica): a taxonomic revision rejects all records except Phymatolithon foecundum.
Phycologia 41:140–146
Amsler CD, Rowley RJ, Laur DR, Quetin LB, Ross RM (1995) Vertical distribution of Antarctic
peninsular macroalgae: cover, biomass and species composition. Phycologia 34:424–430
Amsler CD, Iken K, McClintock JB, Amsler MO, Peters KJ, Hubbard JM, Furrow FB, Baker JB
(2005) Comprehensive evaluation of the palatability and chemical defenses of subtidal mac-
roalgae from the Antarctic Peninsula. Mar Ecol Prog Ser 294:141–159
Arenas F, Sánchez I, Hawkins SJ, Jenkins SRJ (2006) The invasibility of marine algal assem-
blages: role of functional diversity and identity. Ecology 87(11):2851–2861
Balata D, Piazzi L, Rindi F (2011) Testing a new classification of morphological functional groups
of marine macroalgae for the detection of responses to stress. Mar Biol 158:2459–2469
Bischof K, Rautenberger R, Brey L, Pérez-Lloréns JL (2006) Physiological acclimation to gradi-
ents of solar irradiance within mats of the filamentous green macroalga Chaetomorpha linum
from southern Spain. Mar Ecol Prog Ser 306:165–175
Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML, Wiencke C (2009) Drivers
of colonization and succession in polar benthic macro- and microalgal communities. Bot Mar
52:655–667
Chapman ARO (1987) Population and community ecology of seaweeds. Adv Mar Biol 23:1–161
Clark GF, Stark JS, Perrett LA, Hill NA, Johnston EL (2011) Algal canopy as a proxy for the dis-
turbance history of understorey communities in East Antarctica. Polar Biol 34:781–790
Clayton MN, Ashburner CM (1990) The anatomy and ultrastructure of “conducting channels” in
Ascoseira mirabilis (Ascoseirales, Phaeophyceae). Bot Mar 33:63–70
Clayton M, Wiencke C, Klöser H (1997) New records of temperate and sub Antarctic marine ben-
thic macroalgae from Antarctica. Polar Biol 17:141–179
Constable AJ, Melbourne-Thomas J, Corney SP, Arrigo KR, Barbraud C, Barnes DK, Bindoff
NL et al (2014) Climate change and Southern Ocean ecosystems. I: how changes in physical
habitats directly affect marine biota. Glob Chang Biol 20:3004–3025. https://doi.org/10.1111/
gcb.12623
Cormaci M, Furnari G, Scammacca B (2000) The macrophytobenthos of Terra Nova Bay. In:
Faranda FM et al (eds) Ross Sea ecology. Italian Antarctic expeditions (1987–1995). Springer,
Berlin, pp 493–502
DeLaca TE, Lipps JH (1976) Shallow-water marine associations, Antarctic Peninsula. Antarct J
US 11:12–20
Delépine R, Lamb JM, Zimmermann MH (1966) Preliminary report on the vegetation of the
Antarctic Peninsula. Proc Int Seaweed Symp 5:107–116
Deregibus D, Quartino ML, Campana GL, Momo FR, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in Potter
Cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166
Dunton KH, Schell DM (1986) Seasonal carbon budget and growth of Laminaria solidungula in
the Alaskan high Arctic. Mar Ecol Prog Ser 31:57–66
Flores-Molina MR, Muñoz P, Rautenberger R, Huovinen P, Gómez I (2016) Stress tolerance to UV
radiation and temperature of the endemic Antarctic brown alga Desmarestia anceps is medi-
ated by high concentrations of phlorotannins. Photochem Photobiol 92:455–466
Gambi M, Lorent IM, Russo G, Scipione M (1994) Benthic associations of the shallow hard bot-
toms off Terra Nova Bay, Ross Sea: zonation, biomass and population structure. Antarct Sci
6:449–462
Gómez I, Huovinen P (2011) Morpho-functional patterns and zonation of south Chilean seaweeds:
the importance of photosynthetic and bio-optical traits. Mar Ecol Prog Ser 422:77–91
Gómez I, Huovinen P (2012) Morpho-functionality of carbon metabolism in seaweeds. In:
Wiencke C, Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and
utilization. Springer, Berlin, pp 25–46
234 I. Gómez and P. Huovinen

Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic


Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation? PLoS
One 10(8):e0134440
Gómez I, Wiencke C (1998) Seasonal changes in C, N and major organic compounds, and their
significance to morpho-functional processes in the endemic Antarctic brown alga Ascoseira
mirabilis. Polar Biol 19:115–124
Gómez I, Thomas DN, Wiencke C (1995a) Longitudinal profiles of growth, photosynthesis and
light independent carbon fixation in the Antarctic brown alga Ascoseira mirabilis. Bot Mar
38:157–164
Gómez I, Wiencke C, Weykam G (1995b) Seasonal photosynthetic characteristics of Ascoseira
mirabilis (Ascoseirales, Phaeophyceae) from King George Island, Antarctica. Mar Biol
123:167–172
Gómez I, Weykam G, Klöser H, Wiencke C (1997) Photosynthetic light requirements, daily carbon
balance and zonation of sublittoral macroalgae from King George Island (Antarctica). Mar
Ecol Prog Ser 148:281–293
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27
Grime JP (ed) (1981) Plant strategies and vegetation processes. Wiley, Chichester, p 222
Grosberg RK, Strathmann RR (2007) The evolution of multicellularity: a minor major transition?
AREES 38:621–654. https://doi.org/10.1146/annurev.ecolsys.36.102403.114735
Gutkowski R, Maleszewski S (1989) Seasonal changes of the photosynthetic capacity of the
Antarctic macroalga Adenocystis utricularis (Bory) Skottsberg. Polar Biol 10:145–148
Gutt J, Bertler N, Bracegirdle TJ, Buschmann A, Comiso J, Hosie G, Isla E, Schloss IR et  al
(2015) The Southern Ocean ecosystem under multiple climate change stresses – an integrated
circumpolar assessment. Glob Chang Biol 21:1434–1453. https://doi.org/10.1111/gcb.12794
Hay ME (1986) Functional geometry of seaweeds: ecological consequences of thallus layering
and shape in contrasting light environments. In: Givnish TJ (ed) On the economy of plant
form and function. Proceedings of the Sixth Maria Moors Cabot Symposium, “Evolutionary
Constraints on Primary Productivity: Adaptive Patterns of Energy Capture in Plants”, Harvard
Forest, August 1983. Cambridge University Press, Cambridge
Holzinger A, Karsten U (2013) Desiccation stress and tolerance in green algae: consequences
for ultrastructure, physiological, and molecular mechanisms. Plant Sci 22:327. https://doi.
org/10.3389/fpls.2013.00327
Holzinger A, Lütz C (2006) Algae and UV irradiation: effects on ultrastructure and related meta-
bolic functions. Micron 37:190–207. https://doi.org/10.1016/j.micron.2005.10.015
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52:509–534
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
Iken K, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009) Field studies on deterrent roles
of phlorotannins in Antarctic brown algae. Bot Mar 52:547–557
Ingólfsson A (2005) Community structure and zonation patterns of rocky shores at high latitudes:
an interocean comparison. J Biogeogr 32:169–182
Innes DJ (1984) Genetic differentiation among populations of marine algae. Helgoländer
Meeresunters 38:401–417
Irving A, Connell S, Johnston E, Pile A, Gillanders B (2005) The response of encrusting coral-
line algae to canopy loss: an independent test of predictions on an Antarctic coast. Mar Biol
147:1075–1083
Johnston EL, Connell SD, Irving AD, Pile AJ, Gillanders BM (2007) Antarctic patterns of shallow
subtidal habitat and inhabitants in Wilke’s land. Polar Biol 30:781–788
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 235

Kaandorp JA, Kübler JE (eds) (2001) The algorithmic beauty of seaweeds, sponges and corals.
Springer, Heidelberg, p 194
Kain JM (1986) Plant size and reproductive phenology of six species of Rhodophyta in subtidal
Isle of Man. Br Phycol J 21:129–138
Kain JM (1987) Seasonal growth and photoinhibition in Plocamium cartilagineum (Rhodophyta)
off the Isle of Man. Phycologia 26:88–99
Kain JM (1989) The seasons in the subtidal. Br Phycol J 24:203–215
Karsten U, Wulff A, Roleda MY, Müller R, Steinhoff FS, Fredersdorf J, Wiencke C (2009)
Physiological responses of polar benthic algae to ultraviolet radiation. Bot Mar 52:639–654
Kim D (2001) Seasonality of marine algae and grazers of an Antarctic rocky intertidal, with empha-
sis on the role of the limpet Nacella concinna Strebel (Gastropoda: Patellidae). Fachbereich
Biologie/Chemie, Ph.D. thesis, University of Bremen, Germany
Klöser H, Mercuri G, Laturnus F, Quartino ML, Wiencke C (1994) On the competitive balance of
macroalgae at Potter Cove (King George Island, South Shetlands). Polar Biol 14:11–16
Kremer BP (1981) Aspects of carbon metabolism in marine macroalgae. Oceanogr Mar Biol Annu
Rev 19:41–94
Lamb IM, Zimmermann MH (1977) Benthic marine algae of the Antarctic Peninsula. Antarct Res
Ser 23:130–229
Lavorel S, Garnier E (2002) Predicting changes in community composition and ecosystem func-
tioning from plant traits: revisiting the holy grail. Funct Ecol 16:545–556
Littler MM, Littler DS (1980) The evolution of thallus form and survival strategies in benthic
marine macroalgae: field and laboratory tests of a functional form model. Am Nat 116:25–44
Lüning K (ed) (1990) Seaweeds: their environment, biogeography and ecophysiology. Wiley, New
York, p 527
Luning K, Dring MJ (1985) Action spectra and spectral quantum yield of photosynthesis in marine
macroalgal with thin and thick thalli. Mar Biol 87:119–129
Lüning K, Kadel P (1993) Daylength range for circannual rhythmicity in Pterygophora californica
(Alariaceae, Phaeophyta) and synchronization of seasonal growth by daylength cycles in sev-
eral other brown algae. Phycologia 32:379–387
Marcías ML, Deregibus D, Saravia LA, Campana GL, Quartino ML (2017) Life between tides:
spatial and temporal variations of an intertidal macroalgal community at Potter Peninsula,
South Shetland Islands, Antarctica. Estuar Coast Shelf Sci 187:193–203
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48
Moe RL, Silva PC (1981) Morphology and taxonomy of Himantothallus (including Phaeoglossum
and Phyllogigas), an Antarctic member of the Desmarestiales (Phaeophyceae). J Phycol
17:15–29. https://doi.org/10.1111/j.1529-8817.1981.tb00814.x
Mystikou A, Peters AF, Asensi AO, Fletcher KF, Brickle P, van West P, Convey P, Küpper FC
(2014) Seaweed biodiversity in the south-western Antarctic Peninsula: surveying macroalgal
community composition in the Adelaide Island/Marguerite Bay region over a 35-year time
span. Polar Biol 37(11):1607–1619
Neushul M (1965) Diving observations of subtidal Antarctic marine vegetation. Bot Mar 8:234–243
Ortiz M, Berríos F, González J, Rodríguez-Zaragoza FF, Gómez I (2016) Macroscopic network
properties and short-term dynamic simulations in coastal ecological systems at Fildes Bay
(King George Island, Antarctica). Ecol Complex 28:145–157. https://doi.org/10.1016/j.
ecocom.2016.06.003
Ortiz M, Hermosillo-Nuñez B, González J, Rodríguez-Zaragoza F, Gómez I, Jordán F (2017)
Quantifying keystone species complexes: ecosystem-based conservation management in the
King George Island (Antarctic Peninsula). Ecol Indic 81:453–460. https://doi.org/10.1016/j.
ecolind.2017.06.016
Padilla DK, Allen BJ (2000) Paradigm lost: reconsidering functional form and group hypotheses
in marine ecology. J Exp Mar Biol Ecol 250:207–221
236 I. Gómez and P. Huovinen

Phillips JC, Kendrick GA, Lavery PS (1997) A test of a functional group approach to detecting
shifts in macroalgal communities along a disturbance gradient. Mar Ecol Prog Ser 153:125–138
Pianka ER (1970) On r and K selection. Am Nat 104(940):592–597
Poorter L, Bongers F (2006) Leaf traits are good predictors of plant performance across 53 rain
forest species. Ecology 87:1733–1743
Quartino ML, Deregibus D, Campana GL, Edgar G, Latorre J, Momo FR et al (2013) Evidence of
macroalgal colonization on newly ice-free areas following glacial retreat in Potter Cove (South
Shetland Islands), Antarctica. PLoS One 8:e58223
Ramírez ME (2010) Flora marina bentónica de la región austral de Sudamérica y la Antártica. An
Inst Patagon 38(1):57–71
Rautenberger R, Huovinen P, Gómez I (2015) Effects of increased seawater temperature on
UV-tolerance of Antarctic marine macroalgae. Mar Biol 162:1087–1109
Roleda MY, Lütz-Meindl U, Wiencke C, Lütz C (2010) Physiological, biochemical, and ultrastruc-
tural responses of the green macroalga Urospora penicilliformis from Arctic Spitsbergen to UV
radiation. Protoplasma 243:105–116
Schoenrock KM, Schram JB, Amsler CD, McClintock JB, Angus RA (2015) Climate change
impacts on overstory Desmarestia spp. from the western Antarctic Peninsula. Mar Biol
162:377–389. https://doi.org/10.1007/s00227-014-2582-8
Schoenrock KM, Schram JB, Amsler CD, McClintock JB, Angus RA, Vohra YK (2016) Climate
change confers a potential advantage to fleshy Antarctic crustose macroalgae over calcified
species. J Exp Mar Biol Ecol 474:58–66
Schwarz A, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photosyn-
thesis near the southern global limit for growth, Cape Evans, Ross Sea, Antarctica. Polar Biol
26:789–799
Schwarz A-M, Hawes I, Andrew N, Mercer S, Cummings V, Thrush S (2005) Primary production
potential of non-geniculate coralline algae at Cape Evans, Ross Sea, Antarctica. Mar Ecol Prog
Ser 294:131–140
Segovia-Rivera V, Valdivia N (2016) Independent effects of grazing and tide pool habitats on the
early colonisation of an intertidal community on western Antarctic Peninsula. Rev Chil Hist
Nat 89:3
Smale DA (2007) Ice disturbance intensity structures benthic communities in nearshore Antarctic
waters. Mar Ecol Prog Ser 349:89–102
Smale DA, Brown KM, Barnes DKA, Fraser KPP, Clarke A (2008) Ice scour disturbance in
Antarctic waters. Science 321:371
Steneck RS, Dethier MN (1994) A functional group approach to the structure of algal-dominated
communities. Oikos 69:476–498
Steneck RS, Watling L (1982) Feeding capabilities and limitations of herbivorous molluscs: a
functional group approach. Mar Biol 68:299–319
Taylor PR, Hay ME (1984) Functional morphology of intertidal seaweeds: adaptive significance of
aggregate vs. solitary forms. Mar Ecol Prog Ser 18:295–302
Valdivia N, Díaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9(6):e100714. https://doi.org/10.1371/journal.
pone.0100714
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental stress gradients in a marine macroalgal-dominated subtidal community on the
Western Antarctic Peninsula. PLoS One 10(9):e0138582
Valdivia N, Huovinen P, Gómez I, Macaya E, Pardo LM (2019) Different ecological mechanisms
lead to similar grazer controls on the functioning of periphyton Antarctic and sub-Antarctic
communities. Prog Oceanogr 174:7–16. https://doi.org/10.1016/j.pocean.2018.01.008
Weykam G, Thomas DN, Wiencke C (1997) Growth and photosynthesis of the Antarctic red algae
Palmaria decipiens (Palmariales) and Iridaea cordata (Gigartinales) during and following
extended periods of darkness. Phycologia 36:395–405
11  Form and Function in Antarctic Seaweeds: Photobiological Adaptations, Zonati… 237

Wiencke C (1990a) Seasonality of brown macroalgae from Antarctica – a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600
Wiencke C (1990b) Seasonality of red and green macroalgae from Antarctica – a long-term culture
study under fluctuating Antarctic daylengths. Polar Biol 10:601–607
Wiencke C, Clayton MN (1990) Sexual reproduction, life history, and early development in cul-
ture of the Antarctic brown alga Himantothallus grandifolius (Desmarestiales, Phaeophyceae).
Phycologia 29:9–18
Wiencke C, Clayton MN (2002) Antarctic seaweeds. In: Wägele JW (ed) Synopses of the Antarctic
benthos, vol 9. A.R.G Gantner Verlag KG, Ruggell, p 239
Wiencke C, Gómez I, Dunton K (2009) Phenology and seasonal physiological performance of
polar seaweeds. Bot Mar 52:585–592
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: De Broyer C, Koubbi
P, Griffiths H, Raymond B, Udekem d’Acoz C et al (eds) Biogeographic atlas of the Southern
Ocean, vol 8. Scientific Committee on Antarctic Research, Cambridge, pp 66–73
Zaneveld JS (1966) Vertical zonation of Antarctic and subantarctic benthic marine algae. Antarct
J US 1(5):211–213
Part IV
Biological Interactions and Ecosystem
Processes
Chapter 12
Successional Processes in Antarctic
Benthic Algae

Gabriela L. Campana, Katharina Zacher, Fernando R. Momo,


Dolores Deregibus, Juan Ignacio Debandi, Gustavo A. Ferreyra,
Martha E. Ferrario, Christian Wiencke, and María L. Quartino

Abstract  Despite the importance of benthic algal communities to Antarctic coastal


ecosystems, much information about their dynamics is still needed. Primary succes-
sion processes in the Antarctic benthos are frequently initiated by ice-mediated dis-
turbance and by the creation of denuded substrate following glacier retreat, both
expected to increase in the future. Primary succession of benthic algae starts with
rapid colonization by diatoms, ephemeral green algal filaments and propagules of
annual and pseudoperennial macroalgae. Early stages of macroalgae can be particu-
larly vulnerable to environmental stress factors, being critical for the structure of
mature communities. The Antarctic environment is severely affected by global
change, and successional patterns can change due to species-specific susceptibilities
to abiotic and biotic drivers, introducing changes in the matter and energy flow in
the coastal food webs.
This chapter summarizes new advances in our knowledge on the successional
dynamics of benthic primary producers in the Antarctic hard-bottom benthos.
Manipulative experiments on the effects of grazing and ultraviolet (UV) radiation as
drivers of the succession at early stages and long-term experiments carried out at
sites with different environmental conditions are compiled. The gathered informa-
tion can contribute to achieve a deeper knowledge of these key communities and
their structure and functioning in a changing environment.

G. L. Campana (*)
Departamento de Biología Costera, Instituto Antártico Argentino (IAA) - Dirección Nacional
del Antártico (DNA), Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: gcampana@dna.gov.ar
K. Zacher · C. Wiencke
Alfred-Wegener-Institute, Helmholtz Centre for Polar and Marine Research (AWI),
Bremerhaven, Germany
e-mail: katharina.zacher@awi.de; christian.wiencke@awi.de

© Springer Nature Switzerland AG 2020 241


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_12
242 G. L. Campana et al.

Keywords  Early colonizers · Ecological succession · Grazing ·


Microphytobenthos · Rocky substrate · UV radiation

12.1  Introduction

The Antarctic benthos is frequently exposed to ice-mediated disturbance leading to


successional processes and to the coexistence of patches of different developmental
stages (Gutt 2001; Quartino et  al. 2005; Barnes and Conlan 2007; Teixidó et  al.
2007). Furthermore, climate change-related phenomena have accelerated the retreat
of glaciers opening newly available substrate formerly covered by ice and, hence,
initiating primary succession processes (Quartino et al. 2013; Sahade et al. 2015;
Lagger et al. 2017, 2018). Even though the successional process in marine rocky
coasts has been subject to a great number of studies worldwide, they were con-
ceived mainly on temperate habitats (Noël et al. 2009; Benincà et al. 2015), whereas
successional processes in polar regions are still less known.
Studies on the successional patterns on the Antarctic benthos have mainly been
focussed on sessile faunal assemblages (Barnes and Conlan 2007; Dayton et  al.
2016; Lagger et  al. 2017, 2018, and references therein). These studies showed a

F. R. Momo
Instituto de Ciencias, Universidad Nacional de General Sarmiento (UNGS),
Los Polvorines, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: fmomo@campus.ungs.edu.ar
D. Deregibus
Departamento de Biología Costera, Instituto Antártico Argentino (IAA) - Dirección Nacional
del Antártico (DNA), Buenos Aires, Argentina
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET),
Buenos Aires, Argentina
e-mail: dderegibus@dna.gov.ar
J. I. Debandi
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
G. A. Ferreyra
Centro Austral de Investigaciones Científicas (CADIC-CONICET), Ushuaia, Argentina
e-mail: gferreyra@cadic-conicet.gob.ar
M. E. Ferrario
Facultad de Ciencias Naturales y Museo (FCNyM), Universidad Nacional de La Plata
(UNLP), La Plata, Argentina
e-mail: meferra@fcnym.unlp.edu.ar
M. L. Quartino
Departamento de Biología Costera, Instituto Antártico Argentino (IAA) - Dirección Nacional
del Antártico (DNA), Buenos Aires, Argentina
e-mail: lquartino@dna.gov.ar
12  Successional Processes in Antarctic Benthic Algae 243

general low cover of the substrate even after several years, postulating predictable
successional processes as a result of slow growth and high seasonality (Bowden
2005; Bowden et al. 2006). However, the occurrence of high interannual changes as
a result of episodic events of massive colonization was also reported (Dayton 1989;
Dayton et al. 2016). More recent studies revealed a rapid colonization by a few spe-
cies after ice shelves collapse (Raes et  al. 2010; Gutt et  al. 2011; Fillinger et  al.
2013) and the discovery of complex communities in newly ice-free areas (Lagger
et al. 2017, 2018).
For Antarctic hard-bottom coasts dominated by macroalgae, the successional
stages of the assemblages were frequently indirectly assessed, by relating it to the
identity and relative abundance of the dominant taxa and to the disturbance regime
of the site (Klöser et  al. 1996; Quartino et  al. 2005; Valdivia et  al. 2014). For
instance, the presence of the green seaweed Monostroma hariotii Gain and colonial
diatoms in habitats severely affected by ice scouring was signalled as indicators of
early stages of succession (Klöser et al. 1996). In the study carried out by Quartino
et al. (2005), the cause of higher macroalgal diversity recorded at intermediate depth
ranges (5–10 m) was attributed to the coexistence of patches in different succes-
sional stages. More recent studies performed in newly ice-free hard-bottom areas
demonstrated the capability of macroalgae to colonize them rapidly (Quartino et al.
2013). The complexity of these communities, in terms of cover and richness, was
positively correlated to the elapsed time since the generation of the space and the
lower level of stress and disturbance imposed by a retreating glacier (Quartino et al.
2013; see also Chap. 8 by Quartino et al.).
In Potter Cove, a small fjord at the western Antarctic Peninsula (South Shetland
Islands), quite a number of studies were focussed to unravel the successional pat-
terns of benthic algal communities at the inter- and subtidal and over different time
spans (months to years) (Fig. 12.1). This chapter summarizes these works and the
advances in our knowledge on the successional process of primary producers in the
hard-bottom Antarctic benthos. We review in situ primary succession studies that
have been performed at this site and that have allowed for the description of the
structure of benthic primary producer communities during succession. Early stages
of succession are characterized through studies performed over the first 2–3 months
of colonization in the rocky substrate in the intertidal (Zacher et al. 2007a, b) and
the upper subtidal, at approximately 2 m depth (Campana et al. 2008a, b; Zacher
and Campana 2008; Campana 2018) (Fig. 12.1). These studies included a multifac-
torial design to test the effects of ultraviolet radiation and grazing on the developing
communities. Furthermore, experiments performed over a 4-year period with
monthly surveys at the subtidal (3–5 m depth) allowed for the characterization of
communities over a longer time span, at a site with no glacial influence, located at
the outer part of the cove and at an inner cove site, in close proximity to a retreating
glacier (Campana et al. 2018) (Fig. 12.1). We also discuss the importance of abiotic
and biotic drivers on community succession, the importance of their interactions
and the experimental approaches applied so far, with an emphasis on the effects of
global change-related phenomena.
244 G. L. Campana et al.

Fig. 12.1  Map of Potter Cove (South Shetland Islands, western Antarctic Peninsula), showing the
location of successional experiments performed at the upper subtidal (at Peñón de Pesca, situated
in the outer part of the cove and at a site located in the inner cove, in close proximity to a retreating
glacier) and at an intertidal platform (Peñón Uno, located on Potter Peninsula). The effects of
ultraviolet radiation and grazing on developing communities were studied at the subtidal of Peñón
de Pesca and the intertidal of Peñón Uno (yellow square); long-term successional studies were
performed at Peñón de Pesca and at the inner cove over 4 years (green square). (Satellite image of
Potter Cove: Google Earth 2016).  (Photos: left  by Gabriela L.  Campana; right by Katharina
Zacher)  

12.2  S
 tructural Patterns and Changes in Algal Community
Composition during Succession

Along the Antarctic hard-bottom coastlines studied in Potter Cove, the primary suc-
cession of benthic algae begins with a rapid colonization by diatoms, filamentous
green algae and propagules of annual and pseudoperennial macroalgae (Zacher
et al. 2007a, b; Campana et al. 2011; Campana 2018) (Fig. 12.2). This general taxo-
nomic composition is similar to studies carried out at the Arctic for early stages of
succession (Fricke et al. 2008, 2011).
As early colonizers in the marine benthos, diatoms may enhance or inhibit mac-
roalgal recruitment and growth (Huang and Boney 1985; Noël et al. 2009). As an
example, it was shown for the Arctic that diatoms apparently facilitate the establish-
ment of macroalgal propagules (Fricke et al. 2008). In the Antarctic benthos, few
investigations have focussed on the simultaneous study of microalgae and macroal-
gae in the successional process, particularly including detailed taxonomic analyses
(Wahl et al. 2004; Zacher et al. 2007a, b; Zacher and Campana 2008; Campana et al.
2008a, b; Campana 2018). Furthermore, the ecology and taxonomy of Antarctic
benthic diatoms are generally scarce when compared to macroalgae studies (Wulff
et al. 2011).
Over the first months of colonization, diatom assemblages were mainly domi-
nated by pennate, typically benthic diatom species (Zacher et al. 2007b; Campana
12  Successional Processes in Antarctic Benthic Algae 245

0 - 2,5 months 1 year 2 years 3 years 4 years

colonial diatoms annual and pseudoperennial macroalgae


green algal filaments
red algal propagules

Colonial diatoms Palmaria decipiens Phaeurus antarcticus

Green algal filaments Monostroma hariotii Iridaea cordata

Red algal propagules Adenocystis utricularis Corallinaceae

Fig. 12.2  The composition of the community of benthic primary producers over succession in an
Antarctic upper subtidal rocky coast. The scheme compiles the information obtained by in situ
studies focussed on the effects of ultraviolet radiation and grazing at early succession (0–2.5 months)
and long-term successional studies carried out from 1 to 4 years (Campana et al. 2008a, b, 2018;
Campana 2018). Both studies were performed at the same subtidal site (Peñón de Pesca, Fig. 12.1),
at low depths (less than 3 m), and were started during summer (December and February, respec-
tively). The succession of the bacterial community during biofilm development is not described
(see Lee et al. 2016). The new substrate is initially colonized by pioneer groups that occupy rapidly
the space; the colonization starts with dominance of diatoms, and a gradual higher space occupa-
tion by green algal filaments occurs. Propagules of red algae are established since the beginning of
the colonization, P. decipiens being the more abundant. These young sporelings probably develop
the sporophyte during the following spring, which coincides with the appearance of individuals in
communities of 1 year of colonization, during summer. At longer time scales, annual and pseudo-
perennial macroalgae dominate the substrate. Monostroma hariotii and P. decipiens are identified
in these stages, after 1 and 2 years of colonization, whereas P. antarcticus and I. cordata were more
abundant after 3 and 4 years, respectively. The dominance of A. utricularis was evident since the
first year, and a higher space occupation by red crustose algae occurs over time. The 4-year-old
communities were similar to the surrounding communities at low depths (less than 5  m), with
absence of large thalli of Desmarestiales that usually form dense forests at deeper depths

et al. 2008b; Campana 2018). The small naviculoid Navicula perminuta Grunow
was conspicuous in the intertidal and the subtidal assemblage (Zacher et al. 2007b;
Campana 2018). The assemblages further included colonial forms, some of which
reached macroscopic dimensions in the subtidal, such as Fragilaria striatula
Lyngbye and Fragilaria islandica Grunow var. adeliae Manguin; pedunculated
forms, such as Pseudogomphonema kamtschaticum (Grunow) Medlin and the spe-
cies of the genus Licmophora; and adnate forms, which have a firm adhesion to the
substrate through most of the valvar surface, such as the species of the genus
Cocconeis (Zacher et al. 2007b; Campana et al. 2008b; Campana 2018).
These benthic diatoms are probably more relevant to the early stages of succes-
sion and were shown to contribute less to the differences between years over longer
time scales, when macroalgae dominate the substrate (Campana et  al. 2018)
(Fig. 12.2). During these early stages, diatom abundance can differ drastically when
comparing intertidal and subtidal assemblages, being much higher in the subtidal
(Zacher and Campana 2008). This pattern could be related to the more stable
246 G. L. Campana et al.

prevailing conditions in the subtidal, with lower influence of stress and disturbance
factors and/or lower grazing pressure when compared to the intertidal (Zacher and
Campana 2008; Zacher et al. 2011).
Regarding seaweeds, within the first 2–3 months of development, no adult thalli
were observed in the intertidal, whereas only filamentous chlorophytes can be mac-
roscopically visible in the subtidal (Zacher et al. 2007a; Zacher and Campana 2008;
Campana 2018). Even though they gradually increased in space occupation in the
subtidal, they were never dominant (Campana 2018). In contrast, seaweeds can
dominate the space on artificial tiles within 2–3 months in temperate and tropical
habitats, a fact that may support the idea that succession proceeds more rapidly than
in the polar benthos (Wahl et al. 2004).
During early stages of succession (less than 10 months of development), young
algal recruits of M. hariotii and Palmaria decipiens (Reinsch) Ricker were regis-
tered in intertidal (Zacher et al. 2007a) and subtidal assemblages (Campana et al.
2011; Debandi et  al. 2015) (Fig.  12.2). Both species were the only macroalgae
inhabiting a newly ice-free area severely affected by a retreating glacier, showing
their ability not only to colonize newly open space but also to grow in areas sub-
jected to high sedimentation, low light penetration and high ice disturbance
(Quartino et al. 2013). Monostroma hariotii is an annual species that can be very
abundant in intertidal habitats, which are subject to high stress and disturbance
intensity (Wiencke and Clayton 2002; Kim 2001; Marcías et al. 2017).
It has also been proposed as a pioneer species in the subtidal, as it has been
recorded in sites repeatedly affected by ice scouring (Klöser et al. 1996). Similarly,
P. decipiens can cope with marked changes in salinity and light availability and can
grow rapidly in intertidal sites subjected to ice scouring (Becker et  al. 2011;
Deregibus et al. 2016; Marcías et al. 2017). For these reasons, it was signalled as
having a high competitive capacity and considered to be an opportunistic species in
the colonization process (Becker et al. 2011). The recruits of this species probably
develop the sporophyte during the next spring, which coincides with their appear-
ance in the communities after 1  year of colonization (Campana et  al. 2018)
(Fig. 12.2).
The brown seaweed Adenocystis utricularis (Bory) Skottsberg is another species
that can also be present at early stages of succession, as small recruits were recorded
after only 1  month of colonization in the intertidal (Zacher et  al. 2007a) and
10 months in the subtidal (Debandi et al. 2015). However, it was dominant over the
whole experimental exposure at a four-year successional study performed at the
upper subtidal, at 3 m depth (Campana et al. 2018) (Fig. 12.2). Adenocystis utricu-
laris is an aseasonal annual species in Antarctic intertidal habitats that can be very
abundant (de Reviers and Délépine 1981; Wiencke and Clayton 2002). It has a crus-
tose phase that can survive winter, and it is biannual in deeper subtidal habitats
(Wiencke 1990). The studies performed so far revealed that this species tends to
monopolize the substrate and, thus, may impede the recruitment—or resist the inva-
sion—by other species (Sousa 1979).
The brown seaweed Phaeurus antarcticus Skottsberg and the red Iridaea cordata
(Turner) Bory de Saint-Vincent were found to reach their maximum cover after 3
12  Successional Processes in Antarctic Benthic Algae 247

and 4 years of colonization, respectively (Fig. 12.2). A dense canopy of P. antarcti-


cus was found after 3 years of colonization, when the highest macroalgal cover and
space occupation were attained (Campana et al. 2018). The branched thalli of this
species generated several layers and a general more complex three-dimensional
structure of the communities (Campana et  al. 2018). Between the third and the
fourth year, competitive interactions may result in the liberation of space, and the
red alga I. cordata reached its maximum cover, during the following summer. These
four-year-old communities were also composed of the crustose phase of A. utricu-
laris and red coralline algae, which showed a higher space occupation over time
(Campana et al. 2018). The same study has also shown that coralline algae have the
potential to establish and grow on the crusts of A. utricularis (Campana 2018;
Campana et al. 2018).
A great space occupation by crustose red algae can be a characteristic trait of
mature communities at the same study site, at these low depths (Klöser et al. 1994).
Besides, they can be indicators of low disturbance in the Antarctic subtidal (Barnes
et al. 1996), where they are frequently found under the canopies of Desmarestiales
(Irving et al. 2005; Clark et al. 2011). Long-term studies could reveal the role of
these algae in successional processes (Barner et al. 2016) as they have been pro-
posed to favour the recruitment of late colonizers by providing structurally simple
habitats, with higher rugosity than the natural substrate (Maggi et al. 2011).
It has been suggested that communities in the upper subtidal of polar regions are
permanently in early successional stages, with low species diversity (Witman and
Dayton 2001). In the long-term experiment carried out at Potter Cove, the experi-
mental tiles were colonized by fast-growing macroalgae, with absence of large thalli
of perennial species, and after 4 years of colonization, these communities resembled
the surrounding community at low depths (Quartino et  al. 2005). Even though
recruits of perennial species were found in the colonization tiles, they did not grow
until maturity and reached only a few centimetres (Campana et al. 2018). Repeated
ice abrasion, interspecific competition with fast-growing species and/or high irradi-
ance conditions in spring if no canopies are present were signalled as possible driv-
ers of these communities that deflect the successional trajectory and maintain the
communities in early stages of succession at these low depths (less than 5  m)
(Campana et al. 2018).
In macroalgal dominated Antarctic subtidal sites, climax communities can be
identified by the abundance of large Desmarestiales that form extensive and persis-
tent patches that seem to be stable over time (Campana et al. 2011). Among these
species, some representatives can be the season anticipators Desmarestia menziesii
J.  Agardh, D. anceps Montagne and Himantothallus grandifolius (A.  Gepp and
E.S. Gepp) Zinova, as they are terete to leathery, perennial macroalgae with high
biomass and a complex three-dimensional structure and are habitat providers and
chemically defended (Wiencke and Clayton 2002; Amsler et  al. 2005; Quartino
et  al. 2005, 2008; Huang et  al. 2006, 2007; see also Chap. 11 by Gomez and
Huovinen for a description of form and function in Antarctic seaweeds). These spe-
cies gather many of the attributes of species recorded in mature communities in
other environments (sensu Littler and Littler 1980; Noël et al. 2009; Wiencke and
248 G. L. Campana et al.

Amsler 2012; Gómez et al. 2019). They can also be considered as foundation spe-
cies (sensu Dayton 1972) that provide structurally complex habitats favouring the
presence of smaller organisms that live associated to them, such as amphipods and
gastropods (Huang et  al. 2007; Amsler et  al. 2015). Whereas certain macroalgal
species are found as epiphytes on these species, it is noteworthy the low abundance
of epiphytic algae in this Antarctic subtidal forests (Wiencke and Clayton 2002;
Peters 2003). In contrast, there is a high occurrence of filamentous endophytes
(Peters 2003) that were shown to be highly palatable to sympatric amphipods
(Amsler et al. 2009; see also Chap. 17 by Amsler et al.).

12.3  E
 cological Factors Influencing Antarctic
Algal Succession

12.3.1  Ultraviolet Radiation

Ultraviolet-B (UV-B) is an environmental stress factor that can limit the develop-
ment of benthic primary producers causing inhibition of photosynthesis and damage
to biomolecules such as DNA, proteins and lipids, among other damaging effects
(Villafañe et al. 2003; Bischof et al. 2006; Häder et al. 2011; Karsten et al. 2011). In
fact, it has been well established that Antarctic seaweed assemblages show species-­
specific sensitivity towards UV, which determine major aspects of the ecology of
these organisms (Bischof et al. 1998, 2006; Wiencke et al. 2007). Thus, coloniza-
tion, establishment and further development of the benthic algae are affected by UV
radiation. As an example, green algae that inhabit the intertidal have a rapid and
high acclimation potential to UV, whereas some red algae that are found under the
canopy provided by other species are very UV-sensitive (Bischof et  al. 2006).
Besides, certain species can show high phenotypic plasticity revealing a correspon-
dence between UV-B tolerance to their vertical distribution or growth sites
(Rautenberger et al. 2013; see Chap. 7 by Huovinen and Gómez).
The experiments carried out in the intertidal and the subtidal of Potter Cove
revealed that UV is an important structuring driver of the benthic primary producers
community in Antarctica at early stages of succession (Zacher et  al. 2007a, b;
Campana et al. 2008a, b; Zacher and Campana 2008; Campana 2018). Ultraviolet
radiation was shown to affect algal groups differently: whereas diatoms were mostly
unaffected, the establishment and/or growth of green and red algal germlings was
limited by ambient UV radiation (Zacher et  al. 2007a, b; Campana et  al. 2008b;
Campana 2018). Antarctic benthic diatoms from soft-bottom habitats have shown a
high resistance to UV, with low levels of photoinhibition and efficient repair mecha-
nisms (Wulff et al. 2008). In these habitats, vertical migration can be a mechanism
of UV avoidance (Karsten et al. 2011). However, as revealed by experiments per-
formed in rocky coasts, a high resistance of these algae was also shown where this
mechanism of avoidance is not possible (Zacher et al. 2007b; Campana et al. 2008b).
12  Successional Processes in Antarctic Benthic Algae 249

Particularly for the subtidal, this group dominates the substrate at early stages of
succession, and when grazers were excluded (Fig.  12.3a), they provided a
UV-resistant canopy that could positively influence macroalgal recruitment and/or
growth (Molis and Wahl 2004).
Early life stages of macroalgae in polar environments can be more vulnerable to
UV radiation than their adult thalli (Bischof et al. 2006; Wiencke et al. 2007; Roleda
et al. 2009; Karsten et al. 2011). For instance, filaments of Antarctic specimens of
Urospora penicilliformis (Roth) Areschoug have high light requirements, but their
spores are more UV-B sensitive than adult stages (Roleda et al. 2009), being their
settlement possibly limited by UV-B radiation (Campana 2018). Besides, the nega-
tive UV-B effects on the photosynthesis and DNA of propagules can be correlated
to the depth collection of the adult thalli (Wiencke et al. 2000). In Antarctica, propa-
gules of macroalgae dwelling in the intertidal have a high resistance to UV, with
high recovery capacity and scarce or even null DNA damage (Zacher et al. 2007c).
In contrast, the germination of spores of species that inhabit the subtidal can be
severely affected by ambient UV, an effect that can be potentiated by UV-B in labo-
ratory conditions (Zacher 2014). Besides, UV was also shown to cause higher pho-
tosynthesis inhibition in spores of subtidal species compared to intertidal ones
(Navarro et al. 2016; see also Chap. 10 by Navarro et al.).
The available information suggest that at the beginning of the successional pro-
cess, during the colonization of the substrate, the differential sensitivity to UV radi-
ation among different groups of benthic algae (viz. diatoms and macroalgae), among
different seaweed species and/or different developmental stages of a species, can
shape the structure and functioning of the communities (Lotze et al. 2002; Villafañe
et al. 2003; Zacher et al. 2007a, b; Campana et al. 2008a; Zacher 2014), which seem
to be more UV-resistant over time (Lotze et al. 2002; Wahl et al. 2004; Molis and
Wahl 2004, 2009).

12.3.2  Grazing

According to the available information, grazing effects in the Antarctic benthos are
expected to be intense in early stages of algal succession, when communities are
composed of more vulnerable life forms such as diatoms and certain early life stages
of macroalgae (Brêthes et al. 1994; Kim 2001; Zacher et al. 2007a, b), a pattern that
has been recorded in other environments as well (Sousa 1979; Farrell 1991;
Lubchenco 1983; Sousa and Connell 1992). For instance, the conspicuous limpet
Nacella concinna Strebel exerts a high influence in structuring the assemblages of
primary producers in the intertidal, particularly at low shore levels (Kim 2001;
Segovia-Rivera and Valdivia 2016; Valdivia et  al. 2019; see also Chap. 13 by
Valdivia). On their upwards migration from the subtidal to the intertidal, N. con-
cinna feeds on microphytobenthos and green filamentous algae at low shore levels
(Brêthes et  al. 1994; Kim 2001), more severely affecting the early life stages of
macroalgae compared to adults (Kim 2001). In fact, grazing was identified as a
250 G. L. Campana et al.

Fig. 12.3 (a) Experimental unit used to determine the effects of UV and grazing on developing
algal communities; the same experiment was performed at the subtidal of Peñón de Pesca and the
intertidal site Peñón Uno. Note tiles (100 cm2) obtained from a (i) caged (grazer-excluded treat-
ment) and (ii) an uncaged treatment, at the subtidal. (b) Underwater view of a set-up used to allow
colonization over 10  months; tiles (25  cm2) were subsequently transported to the laboratory to
perform further experiments. Note a detail of the obtained communities, composed of young germ-
lings of the red seaweed P. decipiens of different age (i and ii), brown A. utricularis (iii) and green
M. hariotii (iv). (c) Underwater view of a long-term successional study performed in Peñón de
Pesca and at an inner cove site exposed to the influence of a retreating glacier (Fig. 12.1); note
bigger-sized colonization tiles (500 cm2) applied to perform the study over 4 years. (Photos: (a) left
panel by Katharina Zacher, (b) left panel by Argentine Army diving crew, (c) left panel by Marcelo
Mammana; (a) and (c) right panel by Gabriela L. Campana, (b) right panel by Gabriela L. Campana
and Juan I. Debandi)
12  Successional Processes in Antarctic Benthic Algae 251

Fig. 12.3 (continued)

strong driver of the structure of developing algal communities in the studies per-
formed in Potter Cove. Biomass reduction was mainly caused by the gastropods
N. concinna and Laevilacunaria antarctica Martens (Zacher et  al. 2007a, b;
Campana et al. 2008a, b; Campana 2018). At both depth ranges (intertidal and sub-
tidal), grazers significantly consumed diatoms, and particularly for the subtidal site,
feeding on colonial diatoms exerted a drastic change in the physiognomy of the
assemblages (Campana 2018) (Fig. 12.3a). This is a trait that was also observed for
freshwater (Steinman 1996) and other marine habitats (Nicotri 1977; Sommer
1999a, b; Hillebrand et al. 2000), where the consumption of canopy algae results in
an increase in the relative abundance of smaller taxa and prostrate forms (Nicotri
1977; Hillebrand et al. 2000).
A differential susceptibility towards grazing was detected among different mac-
roalgal species, which resulted in changes in species composition (Zacher et  al.
2007a; Campana 2018). Among green algae, Urospora penicilliformis (Roth)
Areschoug was not affected by grazing, but early stages of M. hariotii could be
severely affected (Zacher et al. 2007a, Campana 2018). Besides, the reduction of the
abundance of the dominant taxa (diatoms) was shown to be beneficial for the estab-
lishment of certain red algal species whose early stages are firmly attached to the
substrate, such as P. decipiens (Zacher et al. 2007a; Campana 2018). The diminish-
ing space pre-emption and/or the reduction of shading may result in a higher abun-
dance of red algae, causing increased evenness and, thus, diversity of the community
(Campana et al. 2011; Campana 2018). A similar effect was observed at the intertidal
site at Potter Cove where a higher macroalgal diversity was attributed to the increased
spatial heterogeneity (Sommer 2000) caused by the simultaneous presence of
untouched biofilms and areas visibly consumed by gastropods (Zacher et al. 2007a).
In contrast to vulnerable early successional stages, chemically defended mac-
roalgae can be characteristic at later successional stages. Dominant brown
252 G. L. Campana et al.

macroalgae (except for Desmarestia antarctica R.L. Moe and P.C. Silva) are chemi-
cally defended against the three most common sympatric consumers: the sea star
Odontaster validus Koehler, the fish Notothenia coriiceps J.  Richardson and the
amphipod Gondogeneia antarctica Chevreux (Amsler et al. 2005). Macroalgal pal-
atability on the sea urchin Sterechinus neumayeri Meissner was not tested due to its
unsuitability for feeding bioassays (Amsler et al. 2005), but studies in McMurdo
revealed that both dominant macroalgae in that area are chemically protected against
this consumer (Amsler et al. 1998, 1999). These authors propose that macroalgae
are commonly unpalatable to sympatric consumers mainly as a result of chemical
defences, as nor physical properties such as the toughness of the thallus or the nutri-
tional content appeared to be related to the algal palatability (Amsler et al. 2005;
Peters et al. 2005; see also Chap. 17 by Amsler et al.). In this sense, several works
have stressed that there is a high macroalgal contribution to the food webs through
the detrital pathway (Fischer and Wiencke 1992; Amsler et al. 2005; Seefeldt et al.
2017; Braeckman et al. 2019). However, amphipod grazing can probably be consid-
ered a biological force that also shapes more mature seaweed communities (see
Huang et al. 2006; Amsler et al. 2009; Aumack et al. 2011; Bucolo et al. 2011).
Indeed, amphipod grazing is presumably responsible for the exclusion of subtidal
filamentous algae in the western Antarctic Peninsula (Peters 2003), and it is hypoth-
esized that they live in a mutualistic relationship with macroalgae, cleaning poten-
tially harmful epiphytes in a chemically defended habitat (Amsler et al. 2014; see
Chap. 13 by Valdivia and Chap. 17 by Amsler et al.).
It is important to point out that grazing effects are not unidirectional and interac-
tions with other biotic and abiotic stressors are expected to occur (e.g. Bothwell
et al. 1994). For instance, grazing by a limpet had a positive effect on Arctic mac-
roalgae germlings under intermediate levels of sedimentation (Zacher et al. 2016a).
In the subtidal experiment carried out at Potter Cove, some of the grazing effects
were more intense in UV or UV-B shielded communities: a direct effect on algae
causing lower palatability or a negative effect on grazers that reduce their activity or
density when this radiation is present was postulated as explanations for the observed
tendencies (Campana et al. 2008a).

12.3.3  Glacier Retreat

Climate change has already shown to have a strong influence on Antarctic benthic
communities (Smale and Barnes 2008; Pasotti et al. 2015; Sahade et al. 2015; Moon
et al. 2015). Antarctic macroalgae are cold-water-adapted organisms, and tempera-
ture stress may limit their development (Wiencke et  al. 2007). Furthermore, the
rapid glacier retreat observed over the western Antarctic Peninsula has opened
newly ice-free areas where reduced light penetration caused by increased sedimen-
tation are the prevailing conditions (Rückamp et  al. 2011; Quartino et  al. 2013;
Deregibus et al. 2016). Primary succession patterns may be affected by high sedi-
mentation rates, which may reduce macroalgal propagules survival or even prevent
12  Successional Processes in Antarctic Benthic Algae 253

the spore settlement on the rocky substrate (Airoldi 2003; Zacher et al. 2016a). In
fact, in Potter Cove, the cover and diversity of the assemblages of macroalgae colo-
nizing newly ice-free areas were inversely correlated to the level of stress and dis-
turbance imposed by the retreating glacier, being the lowest in sites close to the
glacier with high sedimentation rates, lowest light penetration and high ice distur-
bance (Quartino et al. 2013; see Chap. 8 by Quartino et al. and Chap. 9 by Deregibus
et al.).
Long-term successional patterns were also evaluated by performing a coloniza-
tion experiment in close proximity to this retreating glacier in Potter Cove (Campana
et al. 2018). On one hand, this study showed convergent patterns to the observed for
the site not affected by glacier influence (Campana et al. 2018). The assemblages
were also dominated by algae, in particular a few opportunistic species, with
A. utricularis reaching a similar cover—approximately of 70%—after 2  years.
Besides, there were significant interannual changes in the assemblages and an
increase in cover over time until the third year, followed by a significant decline
between the third and fourth year. As both experiments were carried out with a year
of difference in their starting points, these convergent patterns can point to a predict-
able successional process for subtidal macroalgal communities in the Antarctic ben-
thos (Campana et al. 2018).
On the other hand, the communities located close to the retreating glacier showed
lower macroalgal richness and a decreased diversity trend over time, which was
attributed to (i) a lower spore availability in more simple communities established
in newly ice-free areas, (ii) a higher sedimentation causing direct abrasion or burial
of propagules and reduced light penetration and/or (iii) a higher ice disturbance
caused by ice block landslides from the glacier (Quartino et  al. 2013; Deregibus
et al. 2016; Campana et al. 2018). More recent colonization studies performed in
newly ice-free areas with different glacial influence showed similar patterns, with
an inverse relationship between the algal cover and diversity, and the degree of sedi-
mentation (Deregibus 2017; see Chap. 9 by Deregibus et al.).
The combined effects of increased temperature, sedimentation and grazing on
the early succession of benthic algae were also studied for subtidal communities
developed over 10  months in Potter Cove (Debandi et  al. 2015). These studies
revealed that increased sedimentation might favour the growth of P. decipiens and
M. hariottii and have neutral effects on brown algal early colonizers (Debandi 2019).

12.4  E
 xperimental Approaches to Study In Situ Succession
of Antarctic Benthic Algae

Different experimental approaches have been applied so far in order to study pri-
mary succession at the polar benthos (Barnes and Conlan 2007; Campana et  al.
2011; Dayton et al. 2016). In most of the cases, artificial substrates were used to
allow the colonization by benthic algae (Zacher et al. 2007a, b; Zacher and Campana
254 G. L. Campana et al.

2008; Campana et al. 2008a, b; Campana 2018) (Fig. 12.3). The use of settlement
tiles was shown to be a useful tool to tackle these studies, particularly for polar
regions, where relatively easy installation and retrieval are very much needed
(Stanwell-Smith and Barnes 1997; Campana et  al. 2011). This experimental
approach allows uniform settlement conditions and standardized replicates so that
several-factor designs and the possibility of deploying the same experiment simul-
taneously at different sites can be achieved (e.g. Wahl et  al. 2004; Zacher and
Campana 2008).
In order to assess the effects of abiotic and biotic drivers of communities at the
very early stages of succession in Antarctica, artificial substrates of relatively small
dimensions (25 to 100 cm2) were shown to be adequate (e.g. Zacher et al. 2007a, b;
Campana et al. 2008a, b; Debandi et al. 2015) (Fig 12.3a), as they were in other
polar regions (Fricke et al. 2008, 2011). Besides the mentioned advantages of their
use, these tiles can be transferred to the laboratory, where detailed analyses of dia-
toms and early stages of seaweeds, as well as quantifications of biomass, cell densi-
ties and percentage of cover of algae on known areas, can be assessed (Foster and
Sousa 1985; Stanwell-Smith and Barnes 1997; Campana et  al. 2011). Moreover,
these assemblages can be applied to physiological studies and further experimenta-
tion (Fig. 12.3b).
When studying successional patterns over a longer time scale (> a year), bigger
tiles, directly fixed to rocky substrate, were chosen (500 cm2) (Campana et al. 2018)
(Fig. 12.3c). These bigger dimensions still permitted an easy manipulation and rep-
lication of a somehow simpler design, and importantly, they allowed an adequate
assessment of community structure as organisms grew. For instance, the dense can-
opy formed by P. antarcticus included organisms reaching a maximum length of
1.2 m (Campana et al. 2018). In this experiment, tiles were monitored by two meth-
ods, as photographic samplings were performed on a monthly basis and detailed
laboratory analyses were done in spring and summer. Both methods revealed the
same patterns of succession and gave complementary information (Campana 2018).
Laboratory analyses applied the “point quadrat” method, which resulted to be the
most adequate to determine ecological indexes and community cover particularly at
later stages of succession, when several layers occur (Foster et al. 1991). This sam-
pling method gave information about the canopy and the understory algae and, thus,
about the three-dimensional characteristics of the analysed communities. On the
other hand, photographic samplings allowed for the detection of faster seasonal
changes that occurred over winter and autumn, achieving a higher time resolution of
the sampling.

12.5  Concluding Remarks and Perspectives

Even though preforming in situ experimental studies in the marine Antarctic ben-
thos can involve logistic constraints and difficulties, they are fundamental to eluci-
date mechanisms and to evaluate the influence of specific factors on biological
12  Successional Processes in Antarctic Benthic Algae 255

processes such as succession (Benedetti-Cecchi 2000; Meiners et al. 2015; Barner


et al. 2016). The studies performed so far reveal that continuous monitoring of envi-
ronmental conditions is essential to better explain the successional patterns of algae
in the Antarctic benthos. Light availability—including UV radiation—can be an
important driver of successional changes and is modified by season, ice cover, phy-
toplankton blooms and sediment input during the warmer seasons (Campana et al.
2011; Gómez et al. 2011). Furthermore, biological drivers such as grazers should
also be monitored; for instance, the increasing grazing pressure by N. concinna at
the end of winter can be signalled as an important factor in the control of the dynam-
ics of these communities in the rocky bottoms (Brêthes et  al. 1994; Kim 2001).
Non-selective grazing by this gastropod could be seasonal, exerting a higher pres-
sure during its migration from the deep subtidal to the intertidal at the end of winter
and early spring (Brêthes et al. 1994; Kim 2001; Zacher et al. 2007b; Campana 2018).
Early stages of succession were shown to be particularly vulnerable to UV radia-
tion and grazing and are probably controlled by complex interactions among abiotic
factors and biological interactions. Considering that UV radiation is an environmen-
tal stress factor for Antarctic ecosystems, it is important to explore the effects of
these wavelengths on algae-grazers interactions, such as the effects of UV on algal
palatability (e.g. Pavia et al. 1997; Macaya et al. 2005; Fairhead et al. 2006) and on
the physiology and behaviour of consumers (Sommaruga 2003; Obermüller et al.
2007). Furthermore, the interactions with climate change processes such as
increased temperature and acidification should also be considered (Rautenberger
et al. 2015; Flores-Molina et al. 2016; Häder 2018).
Antarctica, and particularly the Antarctic Peninsula, is one of the regions most
seriously affected by climate change (Turner et al. 2009), where a number of associ-
ated phenomena can exert a strong influence on the structure and functioning of
benthic communities (Quartino et al. 2013; Sahade et al. 2015; Moon et al. 2015;
Torre et al. 2017; Häder 2018; see Chap. 8 by Quartino et al.). Some of these phe-
nomena are bound to initiate colonization processes and affect the following stages
of succession. On one hand, glacier retreat on the western Antarctic Peninsula has
originated newly ice-free areas that were colonized by macroalgae (Quartino et al.
2013) and invertebrates (Sahade et al. 2015; Lagger et al. 2017, 2018). Furthermore,
recent blooms of benthic diatoms have been registered on the soft substrate at sites
exposed to glacier melting (Ahn et  al. 2016), pointing to a substantial change in
benthic communities. On the other hand, ice abrasion is postulated to increase due
to a lower formation of fast ice that allows higher movement of existing icebergs
and by the existence of new ice blocks and icebergs originated from retreating gla-
ciers (Barnes 2017; Deregibus et al. 2017).
Besides, the temporal and spatial reduction in fast ice formation may result in
higher light availability in the water column during winter and early spring, leading
to a higher primary production (Johnston et al. 2007; McClintock et al. 2008; Clark
et al. 2013; Deregibus et al. 2016) and favouring the colonization of deeper areas
(Miller and Pearse 1991; Gómez and Huovinen 2015). However, this will imply a
higher penetration and time exposure to damaging UV radiation (Gómez and
Huovinen 2015; see also Chap. 7 by Huovinen and Gómez). Besides, increased
256 G. L. Campana et al.

sedimentation during warm months may counteract negative UV effects but simul-
taneously reduce photosynthetically active radiation (PAR, 400–700 nm) availabil-
ity, leading to changes in community structure and affecting the carbon balance of
macroalgae (Quartino et al. 2013; Deregibus et al. 2016; see Chap. 9 by Deregibus
et al.). As already mentioned, higher sedimentation can also affect the survival or
interfere with the establishment of macroalgal propagules (Zacher et  al. 2016a),
benefitting the settlement of early space occupiers such as P. decipiens and M. hari-
otii (Quartino et al. 2013)
Overall, in this Antarctic environment seriously affected by global change phe-
nomena, it is important to assess the patterns of community development during
succession and how they can be modified due to a differential sensitivity of algae to
abiotic changes (i.e. a higher space availability but, simultaneously, higher tempera-
tures, changes in PAR availability, higher ice disturbance, higher sedimentation,
acidification) that may lead to changes in biological interactions, such as grazing
and competition (Schoenrock et al. 2015; Navarro et al. 2016; Zacher et al. 2016a,
b; Schram et al. 2017). Besides, studies using natural substrata, at greater depths,
and larger temporal and spatial scales are necessary to reveal the underlying mecha-
nisms and interactions (Campana et al. 2011). The gathered information can con-
tribute to achieve a deeper knowledge of these communities, whose structure and
functioning are key to the changing Antarctic coastal systems.

Acknowledgements  A significant part of the work reviewed here was performed within the
framework of the scientific collaboration between Instituto Antártico Argentino (IAA)/Dirección
Nacional del Antártico (DNA) and the Alfred Wegener Institute Helmholtz Centre for Polar and
Marine Research (AWI). We thank the support of Carlini Station and Dallmann Laboratory crews
over the several expeditions involved. Field assistance of Oscar González and Alejandro Ulrich
was always invaluable, as was the work of summer and overwinter expedition scientists and of the
Argentine Army and German diving crews. We acknowledge the financial support by grants from
CONICET, DNA-IAA, ANPCyT-DNA (PICTO 0116/2012–2015, PICT 2017-2691), AWI,
Deutsche Forschungsgemeinschaft (DFG, grant Za735/1-1) and MINCyT-BMBF Program
(AL/17/06-01DN18024). This chapter also presents an outcome of the International Research
Network IMCONet funded by the Marie Curie Action IRSES IMCONet (FP7 IRSES, Action No.
318718) and CoastCarb (Funding ID 872690, H2020-MSCA-RISE-2019-Research and Innovation
Staff Exchange).

References

Ahn I-Y, Moon H-W, Jeon M, Kang S-H (2016) First record of massive blooming of benthic
diatoms and their association with megabenthic filter feeders on the shallow seafloor of an
Antarctic fjord: does glacier melting fuel the bloom? Ocean Sci J 51:273–279. https://doi.
org/10.1007/s12601-016-0023-y
Airoldi L (2003) The effects of sedimentation on rocky coast assemblages. In: Gibson RN,
Atkinson RJA (eds) Oceanography and marine biology, an annual review, vol 41. Taylor &
Francis, Boca Raton, pp 161–236
Amsler CD, McClintock JB, Baker BJ (1998) Chemical defense against herbivory in the Antarctic
marine macroalgae Iridaea cordata and Phyllophora antarctica (Rhodophyceae). J Phycol
34:53–59. https://doi.org/10.1046/j.1529-8817.1998.340053.x
12  Successional Processes in Antarctic Benthic Algae 257

Amsler CD, McClintock JB, Baker BJ (1999) An Antarctic feeding triangle: defensive interactions
between macroalgae, sea urchins, and sea anemones. Mar Ecol Prog Ser 183:105–114. https://
doi.org/10.3354/meps183105
Amsler CD, Iken K, McClintock JB, Amsler MO, Peters KJ, Hubbard JM, Furrow FB, Baker BJ
(2005) Comprehensive evaluation of the palatability and chemical defenses of subtidal mac-
roalgae from the Antarctic Peninsula. Mar Ecol Prog Ser 294:141–159. https://doi.org/10.3354/
meps294141
Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009) Filamentous algal endophytes in mac-
rophytic Antarctic algae: prevalence in hosts and palatability to mesoherbivores. Phycologia
48:324–334. https://doi.org/10.2216/08-79.1
Amsler CD, McClintock JB, Baker BJ (2014) Chemical mediation of mutualistic interactions
between macroalgae and mesograzers structure unique coastal communities along the Western
Antarctic Peninsula. J Phycol 50:1–10. https://doi.org/10.1111/jpy.12137
Amsler MO, Huang YM, Engl W, McClintock MJ, Amsler CD (2015) Abundance and diver-
sity of gastropods associated with dominant subtidal macroalgae from the western Antarctic
Peninsula. Polar Biol 38:1171–1181. https://doi.org/10.1007/s00300-015-1681-4
Aumack CF, Amsler CD, McClintock JB, Baker BJ (2011) Impacts of mesograzers on epi-
phyte and endophyte growth associated with chemically defended macroalgae from the
Western Antarctic Peninsula: a mesocosm experiment. J Phycol 47:36–41. https://doi.
org/10.1111/j.1529-8817.2010.00927.x
Barner AK, Hacker SD, Menge BA, Nielse KJ (2016) The complex net effect of reciprocal interac-
tions and recruitment facilitation maintains an intertidal kelp community. J Ecol 104:33–43.
https://doi.org/10.1111/1365-2745.12495
Barnes DKA (2017) Iceberg killing fields limit huge potential for benthic blue carbon in Antarctic
shallows. Glob Chang Biol 23:2649–2659. https://doi.org/10.1111/gcb.13523
Barnes DKA, Conlan KE (2007) Disturbance, colonization and development of Antarctic ben-
thic communities. Philos Trans R Soc Lond B Biol Sci 362:11–38. https://doi.org/10.1098/
rstb.2006.1951
Barnes DKA, Rothery P, Clarke A (1996) Colonisation and development in encrusting communi-
ties from the Antarctic intertidal and sublittoral. J Exp Mar Biol Ecol 196:251–265. https://doi.
org/10.1016/0022-0981(95)00132-8
Becker SB, Quartino ML, Campana GL, Bucolo P, Wiencke C, Bischof K (2011) The biology of
an Antarctic rhodophyte, Palmaria decipiens: recent advances. Antarct Sci 23:419–430. https://
doi.org/10.1017/S0954102011000575
Benedetti-Cecchi L (2000) Predicting direct and indirect interactions during succession in a mid-­
littoral rocky shore assemblage. Ecol Monogr 70:45–72. https://doi.org/10.2307/2657167
Benincà E, Ballantine B, Ellner SP, Huisman J (2015) Species fluctuations sustained by a cyclic suc-
cession at the edge of chaos. PNAS 112:6389–6394. https://doi.org/10.1073/pnas.1421968112
Bischof K, Hanelt D, Wiencke C (1998) UV-radiation can affect depth zonation of Antarctic mac-
roalgae. Mar Biol 131:597–605. https://doi.org/10.1007/s002270050351
Bischof K, Gómez I, Molis M, Hanelt D, Karsten U, Lüder UH, Roleda MY, Zacher K, Wiencke
C (2006) Ultraviolet radiation shapes seaweeds communities. Rev Environ Sci Biotechnol
5:141–166. https://doi.org/10.1007/s11157-006-0002-3
Bothwell ML, Sherbot DMJ, Pollock CM (1994) Ecosystem response to solar ultraviolet-B radia-
tion: influence of trophic-level interactions. Science 265:97–100. https://doi.org/10.1126/
science.265.5168.97
Bowden DA (2005) Seasonality of recruitment in Antarctic sessile marine benthos. Mar Ecol Prog
Ser 297:101–118. https://doi.org/10.3354/meps297101
Bowden DA, Clarke A, Peck LS, Barnes DKA (2006) Antarctic sessile marine benthos: colonisa-
tion and growth on artificial substrata over three years. Mar Ecol Prog Ser 316:1–16. https://
doi.org/10.3354/meps316001
Braeckman U, Pasotti F, Vázquez S, Zacher K, Hoffmann R, Elvert M, Marchant H, Buckner C,
Quartino ML, Mac Cormack W, Soetaert K, Wenzhöfer F, Vanreusel A (2019) Degradation of
258 G. L. Campana et al.

macroalgal detritus in shallow coastal Antarctic sediments. Limnol Oceanogr 64:1423–1441.


https://doi.org/10.1002/lno.11125
Brêthes JC, Ferreyra GA, de la Vega S (1994) Distribution, growth and reproduction of the
limpet Nacella (Patinigera) concinna (Strebel 1908) in relation to potential food availabil-
ity, in Esperanza bay (Antarctic Peninsula). Polar Biol 14:161–170. https://doi.org/10.1007/
BF00240521
Bucolo P, Amsler CD, McClintock JB, Baker BJ (2011) Palatability of the Antarctic rhodophyte
Palmaria decipiens (Reinsch) RW Ricker and its endo/epiphyte Elachista antarctica Skottsberg
to sympatric amphipods. J Exp Mar Biol Ecol 396:202–206. https://doi.org/10.1016/j.
jembe.2010.10.023
Campana GL (2018) Microalgas y macroalgas bentónicas en un sistema costero Antártico (Caleta
Potter, Islas Shetland del Sur): Efectos de la radiación ultravioleta y el pastoreo sobre la colo-
nización y sucesión. PhD Thesis, Universidad Nacional de La Plata, Argentina
Campana GL, Momo FR, Quartino ML, Ferreyra GA (2008a) Effects of UVR and grazing on
biomass and primary production of subtidal benthic algae in Antarctica. Ber Polarforsch
Meeresforsch 571:278–286
Campana GL, Quartino ML, Yousif A, Wulff A (2008b) Effects of UV radiation and grazing on the
structure of a subtidal benthic diatom assemblage in Antarctica. Ber Polarforsch Meeresforsch
571:302–310
Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML, Wiencke C (2011) Drivers of
colonization and succession in polar benthic macro- and microalgal communities. In: Wiencke
C (ed) Biology of polar benthic algae. Walter de Gruyter GmbH & Co, Berlin, pp 399–320
Campana GL, Zacher K, Deregibus D, Momo FR, Wiencke C, Quartino ML (2018) Succession
of Antarctic benthic algae (Potter Cove, South Shetland Islands): structural patterns and
glacial impact over a four-year period. Polar Biol 41:377–396. https://doi.org/10.1007/
s00300-017-2197-x
Clark GF, Stark JS, Perrett LA, Hill NA, Johnston EL (2011) Algal canopy as a proxy for the
disturbance history of understorey communities in East Antarctica. Polar Biol 34:781–790.
https://doi.org/10.1007/s00300-010-0931-8
Clark GF, Stark JS, Johnston EL, Runcie JW, Goldsworthy PM, Raymond B, Riddle MJ (2013)
Light-driven tipping points in polar ecosystems. Glob Chang Biol 19:3749–3761. https://doi.
org/10.1111/gcb.12337
Dayton PK (1972) Toward an understanding of community resilience and the potential effects of
enrichments to the benthos at McMurdo Sound, Antarctica. In: Parker BC (ed) Proceedings of
the colloquium on conservation problems. Allen Press, Lawrence, pp 81–96
Dayton PK (1989) Interdecadal variation in an Antarctic sponge and its predators from oceano-
graphic climate shifts. Science 245:1484–1486. https://doi.org/10.1126/science.245.4925.1484
Dayton PK, Jarrell S, Kim S, Thrush S, Hammerstrom K, Slattery M, Parnell E (2016) Surprising
episodic recruitment and growth of Antarctic sponges: implications for ecological resilience. J
Exp Mar Biol Ecol 482:38–55. https://doi.org/10.1016/j.jembe.2016.05.001
de Reviers B, Délépine R (1981) Biological data on a southern Phaeophyceae Adenocystis utricu-
laris (Bory) Skottsberg, possible material for alginate industry. In: Levring T (ed) Proceedings
on the 10th International Seaweed Symposium. Walter de Gruyter, Berlin, pp 345–350
Debandi JI (2019) Efectos del calentamiento climático y el pastoreo sobre una comunidad de
algas marinas bentónicas antárticas, en etapas tempranas de la sucesión. Thesis, Universidad
Nacional de Luján, Argentina
Debandi JI, Zacher K, Campana GL, Deregibus D, Quartino ML (2015) Estudio experimen-
tal del efecto de la temperatura, la sedimentación y el pastoreo sobre comunidades de algas
marinas bentónicas antárticas. In: Riccialdelli L (ed) Abstracts of the IX Jornadas Nacionales
de Ciencias del Mar, Ushuaia, Tierra del Fuego. CADIC-CONICET-University of Tierra del
Fuego, Tierra del Fuego, p  187. https://jornadasdelmar2015.files.wordpress.com/2015/11/
libro-de-resc3bamenes-final.pdf
Deregibus D (2017) Efecto del retroceso glaciario inducido por el cambio climático sobre la comu-
nidad de algas marinas bentónicas en nuevas áreas libres de hielo en un ecosistema costero
12  Successional Processes in Antarctic Benthic Algae 259

antártico (Caleta Potter, I. 25 de Mayo, I.  Shetland del Sur). PhD Thesis, Universidad de
Buenos Aires, Argentina
Deregibus D, Quartino ML, Campana GL, Momo FR, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in potter
cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166. https://doi.org/10.1007/
s00300-015-1679-y
Deregibus D, Quartino ML, Zacher K, Campana GL, Barnes DKA (2017) Understanding the link
between sea ice, ice scour and Antarctic benthic biodiversity; the need for cross station and
nation collaboration. Polar Rec 53:143–152. 10.1017/S0032247416000875
Fairhead V, Amsler CD, McClintock JB, Baker BJ (2006) Lack of defense or phlorotannin induc-
tion by UV radiation or mesograzers in Desmarestia anceps and D. menziesii (Phaeophyceae).
J Phycol 42:1174–1183. https://doi.org/10.1111/j.1529-8817.2006.00283.x
Farrell TM (1991) Models and mechanisms of succession: an example from a rocky intertidal com-
munity. Ecol Monogr 61:95–113. https://doi.org/10.2307/1943001
Fillinger L, Janussen D, Lundälv T, Richter C (2013) Rapid glass sponge expansion after climate-­
induced Antarctic ice shelf collapse. Curr Biol 23:1330–1334. https://doi.org/10.1016/j.
cub.2013.05.051
Fischer G, Wiencke C (1992) Stable carbon isotope composition, depth distribution and fate
of macroalgae from the Antarctic peninsula region. Polar Biol 12:341–348. https://doi.
org/10.1007/BF00243105
Flores-Molina MR, Rautenberger R, Muñoz P, Huovinen P, Gómez I (2016) Stress tolerance of the
endemic Antarctic brown alga Desmarestia anceps to UV radiation and temperature is medi-
ated by high concentrations of phlorotannins. Photochem Photobiol 92:455–466. https://doi.
org/10.1111/php.12580
Foster MS, Sousa WP (1985) Succession. In: Littler MM, Littler DS (eds) Handbook of phycologi-
cal methods – ecological field methods: macroalgae. Cambridge University Press, Cambridge,
pp 269–290
Foster MS, Harrold C, Hardin DD (1991) Point vs. photo quadrat estimates of the
cover of sessile marine organisms. J Exp Mar Biol Ecol 146:193–203. https://doi.
org/10.1016/0022-0981(91)90025-R
Fricke A, Molis M, Wiencke C, Valdivia N, Chapman AS (2008) Natural succession of macroalgal-­
dominated epibenthic assemblages at different water depths and after transplantation from
deep to shallow water on Spitsbergen. Polar Biol 31:1191–1203. https://doi.org/10.1007/
s00300-008-0458-4
Fricke A, Molis M, Wiencke C, Valdivia N, Chapman AS (2011) Effects of UV radiation on
the structure of Arctic macrobenthic communities. Polar Biol 34:995–1009. https://doi.
org/10.1007/s00300-011-0959-4
Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic
Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation? PLoS
One 10(8):e0134440. https://doi.org/10.1371/journal.pone.0134440
Gómez I, Wulff A, Dunton K, Karsten U, Roleda M, Wiencke C (2011) Light and temperature
demands of benthic microalgae and seaweeds in polar regions. In: Wiencke C (ed) Biology of
polar benthic algae. Walter de Gruyter GmbH & Co, Berlin, pp 195–220
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27. https://doi.org/10.1016/j.pocean.2018.03.013
Gutt J (2001) On the direct impact of ice on marine benthic communities, a review. Polar Biol
24:553–564. https://doi.org/10.1007/s003000100262
Gutt J, Barratt I, Domack E, d’Udekem d’Acoz C, Dimmler W et al (2011) Biodiversity change
after climate-induced ice-shelf collapse in the Antarctic. Deep Sea Res Part II 58:74–83. https://
doi.org/10.1016/j.dsr2.2010.05.024
Häder D-P (2018) Polar Macroalgae. In: Häder D-P, Gao K (eds) Aquatic ecosystems in a chang-
ing climate. CRC Press, Princeton, pp 253–267
260 G. L. Campana et al.

Häder D-P, Helbling EW, Williamson CE, Worrest RC (2011) Effects of UV radiation on aquatic
ecosystems and interactions with climate change. Photochem Photobiol Sci 10:242–260.
https://doi.org/10.1039/c0pp90036b
Hillebrand H, Worm B, Lotze HK (2000) Marine microbenthic community structure regulated by
nitrogen loading and grazing pressure. Mar Ecol Prog Ser 204:27–38. https://doi.org/10.3354/
meps204027
Huang R, Boney AD (1985) Individual and combined interactions between litto-
ral diatoms and sporelings of red algae. J Exp Mar Biol Ecol 85:101–111. https://doi.
org/10.1016/0022-0981(85)90136-4
Huang Y, McClintock JB, Amsler CD, Peters KJ, Baker BJ (2006) Feeding rates of common
Antarctic gammarid amphipods on ecologically important sympatric macroalgae. J Exp Mar
Biol Ecol 329:55–65. https://doi.org/10.1016/j.jembe.2005.08.013
Huang Y, Amsler MO, McClintock JB, Amsler CD, Baker BJ (2007) Patterns of gammaridean
amphipod abundance and species composition associated with dominant subtidal macroal-
gae from the western Antarctic Peninsula. Polar Biol 30:1417–1430. https://doi.org/10.1007/
s00300-007-0303-1
Irving AD, Connell SD, Johnston EL, Pile AJ, Gillanders BM (2005) The response of encrusting
coralline algae to canopy loss: an independent test of predictions on an Antarctic coast. Mar
Biol 147:1075–1083. https://doi.org/10.1007/s00227-005-0007-4
Johnston EL, Connell SD, Irving AD, Pile AJ, Gillanders BM (2007) Antarctic patterns of shal-
low subtidal habitat and inhabitants in Wilke’s land. Polar Biol 30:781–788. https://doi.
org/10.1007/s00300-006-0237-z
Karsten U, Wulff A, Roleda MY, Müller R, Steinhoff FS, Fredersdorf J, Wiencke C (2011)
Physiological responses of polar benthic algae to ultraviolet radiation. In: Wiencke C (ed)
Biology of polar benthic algae. Walter de Gruyter GmbH & Co, Berlin, pp 195–220
Kim D (2001) Seasonality of marine algae and grazers of an Antarctic rocky intertidal, with
emphasis on the role of the limpet Nacella concinna Strebel (Gastropoda: Patellidae). Ber
Polarforsch Meeresforsch 397:1–120
Klöser H, Mercuri G, Laturnus F, Quartino ML, Wiencke C (1994) On the competitive balance of
macroalgae at Potter Cove (King George Island, South Shetlands). Polar Biol 14:11–16. https://
doi.org/10.1007/BF00240266
Klöser H, Quartino ML, Wiencke C (1996) Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in potter cove, King George Island, Antarctica.
Hydrobiologia 333:1–17. https://doi.org/10.1007/BF00020959
Lagger C, Servetto N, Torre L, Sahade R (2017) Benthic colonization in newly ice-free soft-­
bottom areas in an Antarctic fjord. PLoS One 12(11):e0186756. https://doi.org/10.1371/jour-
nal.pone.0186756
Lagger C, Nime M, Torre L, Servetto N, Tatián M, Sahade R (2018) Climate change, glacier
retreat and a new ice-free island offer new insights on Antarctic benthic responses. Ecography
40:1–12. https://doi.org/10.1111/ecog.03018
Lee YM, Cho KH, Hwang K, Kim EH, Kim M, Hong SG, Lee HK (2016) Succession of bacterial
community structure during the early stage of biofilm development in the Antarctic marine
environment. Korean J Microbiol 52:49–58. https://doi.org/10.7845/kjm.2016.6005
Littler MM, Littler DS (1980) The evolution of thallus form and survival strategies in benthic
marine macroalgae: field and laboratory tests of a functional form model. Am Nat 116:25–44.
https://doi.org/10.1086/283610
Lotze HK, Worm B, Molis M, Wahl M (2002) Effects of UV radiation and consumers on recruit-
ment and succession of a marine macrobenthic community. Mar Ecol Prog Ser 243:57–66.
https://doi.org/10.3354/meps243057
Lubchenco J (1983) Littorina and Fucus: effects of herbivores, substratum heterogeneity, and plant
escapes during succession. Ecology 64:1116–1123. https://doi.org/10.2307/1937822
Macaya EC, Rothäusler E, Thiel M, Molis M, Wahl M (2005) Induction of defenses and within-­
alga variation of palatability in two brown algae from the northern-central coast of Chile:
effects of mesograzers and UV radiation. J Exp Mar Biol Ecol 325:214–227
12  Successional Processes in Antarctic Benthic Algae 261

Maggi E, Bertocci I, Vaselli S, Benedetti-Cecchi L (2011) Connell and Slayter’s models of succes-
sion in the biodiversity era. Ecology 92:1399–1406. https://doi.org/10.1890/10-1323.1
Marcías ML, Deregibus D, Saravia LA, Campana GL, Quartino ML (2017) Life between tides:
spatial and temporal variations of an intertidal macroalgal community at Potter Peninsula, South
Shetland Islands, Antarctica. Estuar Coast Shelf Sci 187:193–203. https://doi.org/10.1016/j.
ecss.2016.12.023
McClintock JB, Ducklow H, Fraser W (2008) Ecological responses to climate change on the
Antarctic Peninsula. Am Sci 96:302–310. https://doi.org/10.1511/2008.73.302
Meiners SJ, Cadotte MW, Fridley JD, Pickett STA, Walker LR (2015) Is successional research
nearing its climax? New approaches for understanding dynamic communities. Funct Ecol
29:154–164. https://doi.org/10.1111/1365-2435.12391
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48. https://doi.org/10.1093/icb/31.1.35
Molis M, Wahl M (2004) Transient effects of solar ultraviolet radiation on the diversity and
structure of a field-grown epibenthic community at Lüderitz, Namibia. J Exp Mar Biol Ecol
302:51–62. https://doi.org/10.1016/j.jembe.2003.10.003
Molis M, Wahl M (2009) Comparison of the impacts of consumers, ambient UV, and future UVB
irradiance on mid-latitudinal macroepibenthic assemblages. Glob Chang Biol 15:1833–1845.
https://doi.org/10.1111/j.1365-2486.2008.01812.x
Moon H-W, Hussin WMRW, Kim H-C, Ahn I-Y (2015) The impacts of climate change on
Antarctic nearshore mega-epifaunal benthic assemblages in a glacial fjord on King George
Island: responses and implications. Ecol Indic 57:280–292. https://doi.org/10.1016/j.
ecolind.2015.04.031
Navarro NP, Huovinen P, Gómez I (2016) Stress tolerance of Antarctic macroalgae in the early life
stages. Rev Chil Hist Nat 89:5. https://doi.org/10.1186/s40693-016-0051-0
Nicotri ME (1977) Grazing effects of four marine intertidal herbivores on the microflora. Ecology
58:1020–1032. https://doi.org/10.2307/1936922
Noël LM-LJ, Griffin JN, Moschella PS, Jenkins SR, Thompson RC, Hawkins SJ (2009) Changes
in diversity and ecosystem functioning during succession. In: Wahl M (ed) Marine hard bottom
communities: patterns, dynamics, diversity, and change. Ecological studies, vol 206. Springer,
Heidelberg, pp 213–223
Obermüller B, Puntarulo S, Abele D (2007) UV-tolerance and instantaneous physiological stress
responses of two Antarctic amphipod species Gondogeneia antarctica and Djerboa furcipes
during exposure to UV radiation. Mar Environ Res 64:267–285. https://doi.org/10.1016/j.
marenvres.2007.02.001
Pasotti F, Manini E, Giovannelli D, Wölfl A-C, Monien D, Verleyen E, Braeckman U, Abele D,
Vanreusel A (2015) Antarctic shallow water benthos in an area of recent rapid glacier retreat.
Mar Ecol 36:716–733. https://doi.org/10.1111/maec.12179
Pavia H, Cervin G, Lindgren A, Åberg P (1997) Effects of UV-B radiation and simulated herbivory
on phlorotannins in the brown alga Ascophyllum nodosum. Mar Ecol Prog Ser 157:139–146.
https://doi.org/10.3354/meps157139
Peters AF (2003) Molecular identification, distribution and taxonomy of brown algal endo-
phytes, with emphasis on species from Antarctica. In: Chapman ARO, Anderson RJ, Vreeland
V, Davidson IR (eds) Proceedings on the 17th international seaweed symposium. Oxford
University Press, New York, pp 293–302
Peters KJ, Amsler CD, Amsler MO, McClintock JB, Dunbar RB, Baker BJ (2005) A comparative
analysis of the nutritional and elemental composition of macroalgae from the western Antarctic
Peninsula. Phycologia 44:453–463. https://doi.org/10.2216/0031-8884(2005)44[453:ACAOT
N]2.0.CO;2
Quartino ML, Zaixso H, Boraso de Zaixso A (2005) Biological and environmental characterization
of marine macroalgal assemblages in Potter Cove, South Shetland Islands, Antarctica. Bot Mar
48:187–197. https://doi.org/10.1515/BOT.2005.029
Quartino ML, Boraso de Zaixso A, Momo FR (2008) Macroalgal production and the energy cycle
of Potter Cove. Ber Polar Meeresforsch 571:68–74
262 G. L. Campana et al.

Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223. https://doi.org/10.1371/journal.pone.0058223
Raes M, Rose A, Vanreusel A (2010) Response of nematode communities after large-scale ice-­
shelf collapse events in the Antarctic Larsen area. Glob Chang Biol 16:1618–1631. https://doi.
org/10.1111/j.1365-2486.2009.02137.x
Rautenberger R, Wiencke C, Bischof K (2013) Acclimation to UV radiation and antioxidative
defence in the endemic Antarctic brown macroalga Desmarestia anceps along a depth gradient.
Polar Biol 36:1779–1789. https://doi.org/10.1007/s00300-013-1397-2
Rautenberger R, Huovinen P, Gómez I (2015) Effects of increased seawater temperature on UV
tolerance of Antarctic marine macroalgae. Mar Biol 162:1087–1097. https://doi.org/10.1007/
s00227-015-2651-7
Roleda MY, Campana GL, Wiencke C, Hanelt D, Quartino ML, Wulff A (2009) Sensitivity of
Antarctic Urospora penicilliformis (Ulotrichales, Chlorophyta) to ultraviolet radiation is life-­
stage dependent. J Phycol 45:600–609. https://doi.org/10.1111/j.1529-8817.2009.00691.x
Rückamp M, Braun M, Suckro S, Blindow N (2011) Observed glacial changes on the King George
Island ice cap, Antarctica, in the last decade. Glob Planet Change 79:99–109. https://doi.
org/10.1016/j.gloplacha.2011.06.009
Sahade R, Lagger C, Torre L, Momo FR, Monien P, Schloss IR, Barnes DKA, Servetto N, Tarantelli
S, Tatián M, Zamboni N, Abele D (2015) Climate change and glacier retreat drive shifts in an
Antarctic benthic ecosystem. Sci Adv 1(10):e1500050. https://doi.org/10.1126/sciadv.1500050
Schoenrock KM, Schram JB, Amsler CD, McClintock JB, Angus RA (2015) Climate change
impacts on overstory Desmarestia spp. from the western Antarctic Peninsula. Mar Biol
162:377–389. https://doi.org/10.1007/s00227-014-2582-8
Schram JB, Schoenrock KM, McClintock JB, Amsler CD, Angus RA (2017) Ocean warming and
acidification alter Antarctic macroalgal biochemical composition but not amphipod grazer
feeding preferences. Mar Ecol Prog Ser 581:45–56. https://doi.org/10.3354/meps1230
Seefeldt MA, Campana GL, Deregibus D, Quartino ML, Abele D, Tollrian R, Held C (2017)
Different feeding strategies in Antarctic scavenging amphipods and their implications for colo-
nisation success in times of retreating glaciers. Front Zool 14:59–74. https://doi.org/10.1186/
s12983-017-0248-3
Segovia-Rivera V, Valdivia N (2016) Independent effects of grazing and tide pool habitats on the
early colonisation of an intertidal community on western Antarctic peninsula. Rev Chil Hist
Nat 89:3. https://doi.org/10.1186/s40693-016-0053-y
Smale DA, Barnes DKA (2008) Likely responses of the Antarctic benthos to climate-­
related changes in physical disturbance during the 21st century, based primarily on evi-
dence from the West Antarctic Peninsula region. Ecography 31:289–305. https://doi.
org/10.1111/j.0906-7590.2008.05456.x
Sommaruga R (2003) UVR and its effects on species interactions. In: Helbling EW, Zagarese
H (eds) UV effects in aquatic organisms and ecosystems. The Royal Society of Chemistry,
Cambridge, pp 485–508. https://doi.org/10.1039/9781847552266-00485
Sommer U (1999a) The impact of herbivore type and grazing pressure on benthic microalgal diver-
sity. Ecol Lett 2:65–69. https://doi.org/10.1046/j.1461-0248.1999.22052.x
Sommer U (1999b) Periphyton architecture and susceptibility to grazing by periwinkles (Littorina
littorea, Gastropoda). Int Rev Hydrobiol 84:197–204. https://doi.org/10.1002/iroh.199900020
Sommer U (2000) Benthic microalgal diversity enhanced by spatial heterogeneity of grazing.
Oecologia 122:284–287. https://doi.org/10.1007/PL00008857
Sousa WP (1979) Experimental investigations of disturbance and ecological succession in a rocky
intertidal algal community. Ecol Monogr 49:227–254. https://doi.org/10.2307/1942484
Sousa WP, Connell JH (1992) Grazing and succession in marine algae. In: John DM, Hawkins
SJ, Price JH (eds) Plant-animal interactions in the marine benthos. Clarendon Press, Oxford,
pp 425–441
12  Successional Processes in Antarctic Benthic Algae 263

Stanwell-Smith D, Barnes DKA (1997) Benthic community development in Antarctica: recruit-


ment and growth on settlement panels at Signy Island. J Exp Mar Biol Ecol 212:61–79. https://
doi.org/10.1016/S0022-0981(96)02754-2
Steinman AD (1996) Effects of grazers on freshwater benthic algae. In: Stevenson RJ, Bothwell
ML, Lowe RL (eds) Algal ecology  – freshwater benthic ecosystems. Academic Press, San
Diego, pp 341–373
Teixidó N, Garrabou J, Gutt J, Arntz WE (2007) Iceberg disturbance and successional spatial pat-
terns: the case of the shelf Antarctic benthic communities. Ecosystems 10:143–158. https://doi.
org/10.1007/s10021-006-9012-9
Torre L, Carmona Tabares PC, Momo FR, Meyer JFCA, Sahade R (2017) Climate change effects
on Antarctic benthos: a spatially explicit model approach. Clim Chang 141:733–746. https://
doi.org/10.1007/s10584-017-1915-2
Turner J, Bindschadler R, Convey P, di Prisco G, Fahrbach E, Gutt J, Hodgson D, Mayewski P,
Summerhayes C (2009) Antarctic climate change and the environment. Scientific Committee
on Antarctic Research, Cambridge
Valdivia N, Díaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9(6):e100714. https://doi.org/10.1371/journal.
pone.0100714
Valdivia N, Pardo LM, Macaya EC, Huovinen P, Gómez I (2019) Different ecological mechanisms
lead to similar grazer controls on the functioning of periphyton Antarctic and sub-Antarctic
communities. Prog Oceanogr 174:7–16. https://doi.org/10.1016/j.pocean.2018.01.008
Villafañe VE, Sundbäck K, L Figueroa F, Helbling EW (2003) Photosynthesis in the aquatic envi-
ronment as affected by UVR. In: Helbling EW, Zagarese H (eds) UV effects in aquatic organ-
isms and ecosystems. The Royal Society of Chemistry, Cambridge, pp 357–397. https://doi.
org/10.1039/9781847552266-00357
Wahl M, Molis M, Davis A, Dobretsov S, Dürr S, Johansson J, Kinley J, Kirugara D, Langer M,
Lotze HK, Thiel M, Thomason JC, Worm B, Zeevi Ben-Yosef D (2004) UV effects that come
and go: a global comparison of marine benthic community level impacts. Glob Chang Biol
10:1962–1972. https://doi.org/10.1111/j.1365-2486.2004.00872.x
Wiencke C (1990) Seasonality of brown macroalgae from Antarctica – a long-term culture study
under fluctuating Antarctic daylengths. Polar Biol 10:589–600. https://doi.org/10.1007/
BF00239370
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Wiencke C,
Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and utilization.
Ecological studies, vol 219. Springer, Heidelberg, pp 265–291
Wiencke C, Clayton MN (eds) (2002) Antarctic seaweeds. A.R.G. Gantner Verlag KG, Ruggell
Wiencke C, Gómez I, Pakker H, Flores-Moya A, Altamirano M, Hanelt D, Bischof K, L Figueroa
F (2000) Impact of UV-radiation on viability photosynthetic characteristics and DNA of brown
algal zoospores: implications for depth zonation. Mar Ecol Prog Ser 197:217–229. https://doi.
org/10.3354/meps197217
Wiencke C, Clayton MN, Gómez I, Iken K, Lüder UH, Amsler CD, Karsten U, Hanelt D, Bischof
K, Dunton K (2007) Life strategy, ecophysiology and ecology of seaweeds in polar waters. Rev
Environ Sci Biotechnol 6:95–126. https://doi.org/10.1007/s11157-006-9106-z
Witman JD, Dayton PK (2001) Rocky subtidal communities. In: Bertness MD, Gaines SD, Hay
ME (eds) Marine community ecology. Sinauer Associates Inc, Sunderland, pp 339–366
Wulff A, Roleda MY, Zacher K, Wiencke C (2008) UV radiation effects on pigments, photo-
synthetic efficiency and DNA of an Antarctic marine benthic diatom community. Aquat Biol
3:167–177. https://doi.org/10.3354/ab00076
Wulff A, Iken K, Quartino ML, Al-Handal AY, Wiencke C, Clayton MN (2011) Biodiversity,
biogeography and zonation of benthic micro- and macroalgae in the Arctic and Antarctic.
In: Wiencke C (ed) Biology of polar benthic algae. Walter de Gruyter GmbH & Co, Berlin,
pp 23–52
264 G. L. Campana et al.

Zacher K (2014) The susceptibility of spores and propagules of Antarctic seaweeds to UV and
photosynthetically active radiation – field versus laboratory experiments. J Exp Mar Biol Ecol
458:57–63. https://doi.org/10.1016/j.jembe.2014.05.007
Zacher K, Campana GL (2008) UV and grazing effects on an intertidal and subtidal algal assem-
blage: a comparative study. Ber Polar Meeresforsch 571:287–294
Zacher K, Wulff A, Molis M, Hanelt D, Wiencke C (2007a) Ultraviolet radiation and consumer
effects on a field-grown intertidal macroalgal assemblage in Antarctica. Glob Change Biol
13:1201–1215. https://doi.org/10.1111/j.1365-2486.2007.01349.x
Zacher K, Hanelt D, Wiencke C, Wulff A (2007b) Grazing and UV radiation effects on an Antarctic
intertidal microalgal assemblage: a long-term field study. Polar Biol 30:1203–1212. https://doi.
org/10.1007/s00300-007-0278-y
Zacher K, Roleda MY, Hanelt D, Wiencke C (2007c) UV effects on photosynthesis and DNA
in propagules of three Antarctic seaweeds (Adenocystis utricularis, Monostroma hariotii and
Porphyra endiviifolium). Planta 225:1505–1516. https://doi.org/10.1007/s00425-006-0436-4
Zacher K, Rautenberger R, Hanelt D, Wulff A, Wiencke C (2011) The abiotic environment of polar
marine benthic algae. In: Wiencke C (ed) Biology of polar benthic algae. Walter de Gruyter
GmbH & Co, Berlin, pp 9–21
Zacher K, Bernard M, Bartsch I, Wiencke C (2016a) Survival of early life history stages of Arctic
kelps (Kongsfjorden, Svalbard) under multifactorial global change scenarios. Polar Biol
39:2009–2020. https://doi.org/10.1007/s00300-016-1906-1
Zacher K, Savaglia V, Bartsch I (2016b) Effects of temperature and interspecific competition
on growth and photosynthesis of two endemic Antarctic Desmarestia species. Algol Stud
151:103–122. https://doi.org/10.1127/algol_stud/2016/0269
Chapter 13
Seaweed-Herbivore Interactions: Grazing
as Biotic Filtering in Intertidal Antarctic
Ecosystems

Nelson Valdivia

Abstract  Consumers constitute a key component of the environmental filters that


restrict the establishment of colonists into local assemblages. Thus, the trophic
activity of consumers, particularly grazers, can be pivotal to control the develop-
ment of potential algal invaders in Antarctic coasts. Here, the consumptive effects of
coastal macrobenthic grazers on algal communities are reviewed to assess the
degree to which these consumers can mediate the introduction of seaweeds in inter-
tidal Antarctic communities. Gastropods and amphipods have strong consumptive
effects on algal communities. Yet, amphipods are sensitive to climate change factors
such as warming and acidification, which could hamper their ability to control
native and alien macroalgae. Alien macroalgae that modify the abiotic environment,
such as the gutweed Ulva intestinalis in tidepools, represent potential superior com-
petitors in Antarctic ecosystems. In this line, simulations based on a simple proba-
bilistic model showed that intermediate to high levels of frequency-dependent
consumption seem to be fundamental to allow for stable coexistence when the alien
species is competitively superior. With this work, I hope to stimulate further manip-
ulative research to assess the role of benthic consumers in mediating the coexistence
(or lack thereof) between alien and native seaweeds under multiple climate change
scenarios.

Keywords  Competition · Environmental filters · Grazing · Alien macroalgae

N. Valdivia (*)
Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral
de Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: nelson.valdivia@uach.cl

© Springer Nature Switzerland AG 2020 265


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_13
266 N. Valdivia

13.1  B
 iological Invasions and Their Impact on the Ecology
of Antarctic Coastal Systems

Biological invasions are one of the major anthropogenic threats to biodiversity


across multiple spatiotemporal scales (Simberloff et al. 2013). The negative conse-
quences of invasions range from local extinction of native species to the impairment
of ecosystem functioning and from alterations of interaction webs to the loss of
ecosystem services (Vitousek et  al. 1997; Bulleri et  al. 2010; White and Shurin
2011; Silva et al. 2019). In marine communities, for example, the introduction of
consumers into a bottom-up-controlled ecosystem can lead to severe compositional
shifts and significant declines in diversity (Kotta et al. 2018). Invasive macroalgae,
in addition, have been shown to reduce local diversity through interference competi-
tion that involves drastic environmental changes (Björk et al. 2005). The effects of
invasive species can be particularly severe on ecologically isolated ecosystems,
because of low functional redundancy and restricted niche coverage (Hughes and
Convey 2014). Antarctica is an iconic example for ecologically isolated ecosys-
tems—this isolation has influenced the evolution of a high proportion of endemic
species, which makes the Antarctic ecosystems a highly ranked scientific priority
(Kennicutt et al. 2014). Accordingly, the study of biological invasions, especially in
ecologically isolated ecosystems like Antarctica, is of widespread relevance for
conservation ecology.
Due to its well-described biological isolation and adverse abiotic conditions,
Antarctica provides a unique opportunity to improve our mechanistic understanding
of biological invasions (Chown et al. 2015; McCarthy et al. 2019). This is particu-
larly accurate for Antarctic coastal marine rocky communities, because of the mul-
tiple biotic and abiotic environmental changes that are currently increasing the risk
of non-native marine species in these ecosystems (McCarthy et  al. 2019). For
instance, increasing ship activity and high transportability have been suggested as
critical factors during the first stages of the invasion process of marine macroalgae
and invertebrates (Blackburn et al. 2011; McCarthy et al. 2019). Associated to the
transport of marine invaders, recent evidence indicates that rafting macroalgae are
able to cross the Antarctic Circumpolar Current and arrive to Antarctica (King
George Island) from Subantarctic and cold-temperate source populations (Fraser
et al. 2017; Fraser et al. 2018). In addition, the establishment of alien species in
Antarctic marine rocky communities could well be facilitated by the current sce-
nario of increasing water temperature and decreasing ice cover (Blunden et al. 2013;
Clark et  al. 2013; Quartino et  al. 2013; McCarthy et  al. 2019). These alterations
represent the weakening of dispersal limitations and also of abiotic filtering that
would benefit aliens over Antarctic natives (Duffy et al. 2017; Griffiths et al. 2017;
see also Chap. 3 by Fraser et al. and Chap. 4 by Macaya et al.).
In addition to dispersal and abiotic  environmental filtering, local community
assembly also depends upon chance (particularly when population numbers are
low) and biotic filters that restrict the establishment of some taxa (HilleRisLambers
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 267

et  al. 2012; Briski et  al. 2018). As such, both factors are relevant during the
­establishment of aliens. Once settlers have passed through the abiotic filtering (e.g.
environmental extremes in Antarctica), negative biotic interactions, such as compe-
tition, predation and herbivory, encompass local biotic filters that finally determine
the success of settled organisms (e.g. Guisan and Thuiller 2005; HilleRisLambers
et al. 2012; King and Howeth 2019). Although invasion theory predicts that general-
ist enemies (e.g. herbivores) have stronger impacts on native than alien competitors,
the analysis of local adaptation of consumers to native resources provides limited
evidence that generalist enemies are better adapted to attacking natives than exotics
(Keane and Crawley 2002). In addition, there are several examples of marine
Antarctic macroalgae that have developed chemical defences to local herbivores
(e.g. Amsler et al. 2005, 2009b; Aumack et al. 2010), which may allow them to have
fitness advantages over alien macroalgae (see Chap. 17  by Amsler et  al.). Thus,
negative biotic interactions in Antarctic coastal marine communities can have an
important role as local filters mediating the process of invasion in these ecosystems.

13.2  R
 ecent Introductions of Exotic Macroalgae
in Antarctica

Examples of alien macroalgae in Antarctica include green algae such as the gutweed
Ulva intestinalis, which was described as Enteromorpha intestinalis together with
the brown alga Petalonia fascia and the red alga Rhodymenia subantarctica in South
Shetland Islands (Clayton et al. 1997; see also Frenot et al. 2005; Campana et al.
2009). In addition, other green macroalgae have been suggested to be recent intro-
ductions to South Shetland Island shores, namely, the filamentous algae Urospora
penicilliformis and Ulothrix sp. (Gómez 2015). Considering the growing human
activity and ship traffic to South Shetland Islands (Frenot et al. 2005; Bender et al.
2016), the occurrence of exotic and potentially invasive species in these locations is
not surprising. Recent multivariate analyses suggest, moreover, a trend of biotic
homogenisation among seaweed assemblages of South Patagonia and both West
and East Antarctic Peninsula (Sanches et al. 2016). In this line, but not necessarily
related to recent biological invasions, a small subset of widely distributed macroal-
gae dominate the biogeographic structure of intertidal rocky shores in the Southern
Ocean, likely as a result of rafting across in the Antarctic Circumpolar Current
(Griffiths and Waller 2016; see also Pellizzari et  al. 2017). Although the recent
increases of Antarctic seaweed diversity could well be the result of improved and
more efficient techniques of sampling and molecular taxonomical methods (e.g.
Dubrasquet et al. 2018), the role of human-mediated transport and climate change
in modifying the biogeography of macroalgae in the Southern Ocean should not be
ruled out (Pellizzari et al. 2017; McCarthy et al. 2019; see Chap. 5 by Pellizzari
et al.).
268 N. Valdivia

13.3  Can Grazers Control Alien Macroalgae in Antarctica?

Worldwide, marine herbivores have profound and chiefly negative effects on the
abundance of primary producers—in particular, these effects are strongest in rocky
intertidal habitats (Lubchenco 1978; Hawkins and Hartnoll 1983; Poore et al. 2012).
In this line, manipulative and observational studies suggest that grazers can exert a
significant effect on the structure of algal communities in Antarctica. For example,
field-based manipulative experiments show a strong control of intertidal grazers,
namely, the limpet Nacella concinna (Fig.  13.1), on intertidal periphyton assem-
blages in Fildes Bay, King George Island (Segovia-Rivera and Valdivia 2016). An
important outcome of this work is that the overall negative effects of N. concinna on
the abundance of periphyton taxa were consistent across intertidal microhabitats
(i.e. emergent rocks and tidepools). Similarly, Zacher et  al. (2007b) demonstrate
that N. concinna’s effects on early-succession algal communities are consistent
across multiple levels of ultraviolet radiation (UVR) exposure. Moreover, the mag-
nitude and sign of the effects of this grazer on benthic periphyton in King George
Island can be similar to those of congeneric limpets in Chilean South Patagonia
(CSP), albeit channelled through different ecological mechanisms—while the
effects of Antarctic limpets appear to be frequency-dependent, the effects of
Magellan limpets seem to be related to niche complementarity in a rich community
of grazers (Aldea and Rosenfeld 2011; Valdivia et al. 2019). Generally speaking,
these results may indicate that the effects of N. concinna on algal abundance and
diversity may be consistent across abiotic environmental conditions related to emer-
sion time, desiccation, and photobiotic and osmotic stress. This conclusion is well
in line with a major meta-analysis that shows only a little influence of environmen-
tal conditions (i.e. latitude or mean annual water temperature) on the effects of graz-
ers on the abundance of primary producers (Poore et al. 2012).
Mesograzers, on the other hand, have also been suggested as strong top-down
controllers of Antarctic seaweed communities. Field experiments demonstrate that

Fig. 13.1  Aggregations of


individuals of Nacella
concinna in Antarctic
rocky tidepools. (Photo by
Nelson Valdivia, Proyecto
Anillo ART1101)
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 269

the abundant amphipod fauna rapidly consumes intertidal filamentous algae trans-
planted to the subtidal in West Antarctic Peninsula (WAP; Amsler et al. 2012). The
remarkable abundance of amphipods (e.g. up to 20 ind/g algal wet weight; Huang
et al. 2007) suggests that amphipod assemblages do have a significant population-­
level control on benthic algal communities. Indeed, amphipods and subtidal mac-
roalgae have developed mutualist  interactions in which the former benefit from
predator-sheltered habitats and the latter from reduced fouling (Amsler et al. 2014).
In this line, amphipods and other mesograzers are proposed to control the occur-
rence of filamentous macroalgae, which can explain the absence of these species in
Antarctic subtidal habitats (Peters 2003; Amsler et al. 2009a). Albeit less abundant
than amphipods, small-sized Antarctic gastropods can also play an important role as
top-down controls of epiphytic microalgae on large pseudo-kelps, as shown in
mesocosm experiments (Amsler et  al. 2015, 2019). Thus, it is highly likely that
mesograzers, through their consumptive activities, can be relevant for the assembly
of local macrobenthic communities in WAP (see Chap. 17 by Amsler et al.).
The results of the experiments described above are supported by early and recent
observational evidence. Kim (2001), for example, found a strong association
between the seasonal variation in the abundance of N. concinna and that of intertidal
filamentous algae in King George Island—although the role of ice scouring in medi-
ating this association cannot be ruled out. In addition, the analysis of stomach con-
tents and stable isotopes strongly supports the idea that amphipods are central in
WAP coastal food webs (Aumack et al. 2017; Zenteno et al. 2019). On the other
hand, observational evidence shows that notothenioid fish, like Notothenia rossi and
N. coriiceps, actively select for macroalgae as food (Casaux et al. 1990; Barrera-­
Oro et al. 2019). Despite correlation does not imply causality, the results of these
observational studies agree with field- and lab-based manipulative evidence of the
central role of herbivores in coastal Antarctic food webs.
Could these grazers prevent the establishment, or at least control the abundance,
of exotic seaweeds? As introduced above, theory predicts that parasites and general-
ist predators and herbivores (i.e. ‘enemies’) can have stronger impacts on native
than alien competitors, allowing the latter to expand their spatial distribution and
adopt an invasive behaviour (i.e. the Enemy Release Hypothesis, reviewed in Keane
and Crawley 2002). In addition, the current trend of seawater warming would also
reduce the physiological constrains imposed by Antarctic environmental extremes
to temperate seaweeds. This may picture a scenario of improving biotic and abiotic
environmental conditions for the establishment and spread of invasive seaweeds in
Antarctica. Yet, sophisticated anti-herbivory defences have evolved in several
Antarctic seaweeds (Amsler et  al. 1998, 2009b, 2019; Aumack et  al. 2010), and
today it is proposed that small-sized herbivores have actually positive effects on
macroalgae owing their antifouling consumptive activity (Amsler et al. 2014, 2019).
This could provide native seaweeds with consumer-mediated competitive advan-
tages over exotic seaweeds, provided that the latter are not equipped with anti-­
herbivory defences. Indeed, grazing has been shown to mediate the competitive
interaction between native and alien seaweeds elsewhere (Noè et  al. 2018). For
270 N. Valdivia

Antarctic communities, however, further empirical research is needed to assess the


role of consumers on competitive seaweed interactions.
How can climate change influence the potential effects of grazers on an invasion
process in Antarctica? The answer to this question seems to depend on the species
or functional type analysed. For instance, Antarctic grazing gastropods, including
N. concinna and Margarella antarctica, have been shown to resist the combined
effects of decreased pH and seawater warming (Schram et  al. 2014), hinting for
certain level of resistance of the grazing function to this stressor. However, the
authors warn that the slow growth rates and longevity of the analysed gastropods
could mask long-term sublethal effects of warming and acidification. On the other
hand, ocean acidification can significantly increase the mortality rates of amphi-
pods, while warming can have sublethal effects in terms of increased whole-body
protein content of those organisms (Schram et al. 2016). Acute warming, moreover,
is shown to modify the feeding preferences of amphipods (Schram et al. 2015). As
changes in consumer abundances and prey shift have been proposed as major causes
of food web variations (e.g. Lopez et al. 2017), ocean acidification and warming can
have profound indirect effects on the structure of coastal Antarctic communities. In
addition, recent evidence hints for indirect effects of climate change on grazer popu-
lations, as seawater warming and freshening can have significant and independent
effects on consumption rates of predatory fish (Navarro et al. 2019). Sedimentation,
which relates to warming-associated glacier melting, has been shown to mediate the
effects of grazers on the germination and development of young sporophytes of
Arctic kelps (Zacher et al. 2016). Finally, ocean warming and increased pCO2 could
mediate macroalgal competitive interactions, as it is demonstrated that both factors
in combination favour fleshy over crustose forms (Schoenrock et al. 2016). In this
way, the role of grazing and competition as biotic filtering in the assembly of local
Antarctic communities should be assessed in combination to climate change
stressors.

13.4  U
 lva intestinalis as a Case Study in a Simple,
Two-­Species Assembly Model

The gutweed Ulva intestinalis was described inhabiting Antarctic shores in 1997
(Clayton et  al. 1997). Elsewhere, this species is able to generate adverse abiotic
conditions for potential competitors in tidepools, which involve conditions of high
pH and low inorganic carbon concentrations—at the same time, U. intestinalis is
able to capture HCO3− under these conditions (Larsson et al. 1997). This ability is
suggested to provide U. intestinalis with competitive advantages over other mac-
roalgae in high-intertidal tidepools, explaining the dominance of this species
observed in Swedish Atlantic tidepools (Björk et  al. 2004, 2005). In Antarctica,
however, U. intestinalis is not a dominant species like in Sweden, and its spatial
distribution is usually confined to high-intertidal tidepools (Clayton et  al. 1997;
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 271

Gómez 2015). A first explanation for the restricted distribution of U. intestinalis


would be the climatic environmental conditions that prevail in WAP, particularly
average seawater temperatures that are below its optimum for somatic growth
(Bischoff and Wiencke 1993). According to the high consumption rates demon-
strated for Antarctic intertidal benthic grazers, including abundant populations of
amphipods (Huang et al. 2006) and the limpet N. concinna (e.g. Zacher et al. 2007a;
Segovia-Rivera and Valdivia 2016), a possible, non-exclusive explanation could be
that grazer activity defines the lower intertidal limit of this species. Intertidal herbi-
vores have been shown to control the abundance of fast-growing macroalgae in
tidepools of WAP (Segovia-Rivera and Valdivia 2016) and elsewhere (e.g. Noël
et al. 2009).
To improve our understanding of an invasion process in Antarctica, I used
U. intestinalis introduction into high-intertidal tidepools as a model system. The
scenario, therefore, includes two sites connected by dispersal (WAP and Chilean
South Patagonia, CSP), tidepools as local habitats, U. intestinalis as a competitively
superior alien (see previous paragraph) and the corticated red seaweed Iridaea cor-
data as a native competitor that is frequent in tidepools (Valdivia et al. 2014); for
simplicity, both species will be referred to as Ulva and Iridaea, respectively. To this
aim, I used a simple probabilistic model known as Moran model, which was origi-
nally generated to understand the temporal changes in allele frequencies in popula-
tions (Moran 1958). This model has been lately used and extended to simulate the
temporal dynamics of two-species communities (Hubbell 2001; Vellend 2016). The
basic structure of the model assumes a neutral, closed assemblage without specia-
tion, in which there are J individuals that belong to one of two species, either spe-
cies A or species B (either Ulva or Iridaea in this case). Since J is assumed to be
fixed (i.e. a zero-sum dynamic), there are j individuals of species A and J – j indi-
viduals of B. At each time, one individual of the community is selected at random to
die. At the next time, an individual is selected at random to produce one offspring
that replaces the dead individual. Each individual is chosen to be A or B with
probabilities

p j = jJ −1 , and

q j = ( J − j ) J −1 ,

for species A and B, respectively (Moran 1958). The community is then described
as a Markov chain in which the state (community structure at a given time) is defined
by the abundance of species (j) and a transition probabilities pjk from state j to k:

J 
p jk =   p kj q Jj − k .
k
In this work, and as previously done by Hubbell (2001) and Vellend (2016), I took
advantage of the simplicity of the Moran model to describe how differences in
272 N. Valdivia

c­ ompetitive abilities, reflected in differences between probabilities pj and qj, can


lead to differing community patterns. These differences were expressed as average
fitness ratios, in which larger ratios will indicate that species A (Ulva) generates a
larger offspring than species B (Iridaea); i.e. the former is competitively superior
than the latter (Vellend 2016). In addition, the simulations of this work were done in
a metacommunity context, in which the reproducing individual of the Moran model
was chosen at random from the entire set of habitat patches.
The rationale of these simulations was to resemble an invasion process in which
Ulva (species A) is an invasive and competitively superior species that already colo-
nised high-intertidal tidepools in the WAP from the nearest continental shore, i.e.
CSP. Iridaea, on the other hand, represents a native competitor in tidepools. The
simulations, thus, considered a simple metacommunity of two local sites that are
linked by dispersal (m in Fig. 13.2). Previous unpublished data was used to estimate
the initial abundances (in terms of frequency) of Ulva. Thus, the simulation included
empirical data only to set the initial abundances of Ulva. In addition, I assumed that
Ulva would be competitively dominant in high-intertidal tidepools in WAP but not
in CSP, because of the lower seaweed diversity observed in Antarctica that would
provide more niche opportunities to alien species. These differences were expressed
as fitness ratios >1 (1.3) in WAP and <1 (1/1.1) in CSP.
A central aspect of these simulations is the incorporation of negative density (or
frequency) dependence of competitive advantages as the result of grazing on the
abundance and fitness of competitors. If fitness ratio always is >1, that is, if species
A has consistently greater fitness than species B, then the former species will tend to
competitively exclude the latter, and no stable coexistence will take place (Chesson
2000; Letten et al. 2017). However, if the fitness advantages of species A over spe-
cies B are greater when species A is rare, and vice versa, then each species will have
relative advantage when rare, and there should be a stable coexistence (equilibrium)
point (Chesson 2000). This negative frequency dependence of growth rates can be
generated by, for instance, negative effects of consumers (predators and prey) that
become harsher with increasing prey abundances (i.e. enemy-mediated coexistence;
Holt et al. 1994). Therefore, the simulations in this work were done across a range
of strengths of negative frequency dependence (b in Fig. 13.2) to resemble potential
effects of grazers on the abundance of macroalgal competitors in Antarctic high-­
intertidal tidepools. These values were set to 0, −0.1, and −0.5, representing absence
and intermediate and high negative density dependence. All simulations were con-
ducted in R programming environment (R Core Team 2019).
The simulations showed that the complete absence of dispersal in the metacom-
munity resulted in extinction in both regions, even at intermediate levels of negative
frequency dependence (Fig. 13.2a,b). Stronger negative density dependence would
allow for coexistence at low abundances of Ulva only in CPS (segmented line in
Fig. 13.2c). Interestingly, dominance, but not competitive exclusion, in Antarctica
would be apparent even in the absence of significant negative frequency dependence
(i.e. grazing) and when dispersal is intermediate (Fig.  13.2d). As expected, high
levels of dispersal would lead to biotic homogenisation of the region (Fig. 13.2g–i).
Intermediate dispersal and increasing negative density dependence led to reduced
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 273

(A) m = 0, b = 0 (B) m = 0, b = −0.1 (C) m = 0, b = −0.5


1.0 1.0 1.0
Frequency of species A

Frequency of species A

Frequency of species A
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0.0 0.0 0.0
0 200 400 600 800 0 200 400 600 800 0 200 400 600 800
Time Time Time

(D) m = 0.1, b = 0 (E) m = 0.1, b = −0.1 (F) m = 0.1, b = −0.5


1.0 1.0 1.0
Frequency of species A

Frequency of species A

Frequency of species A
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0.0 0.0 0.0
0 200 400 600 800 0 200 400 600 800 0 20 40 60 80 100
Time Time Time

(G) m = 0.5, b = 0 (H) m = 0.5, b = −0.1 (I) m = 0.5, b = −0.5


1.0 1.0 1.0
Frequency of species A

Frequency of species A

Frequency of species A

0.8 0.8 0.8


0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0.0 0.0 0.0
0 200 400 600 800 0 200 400 600 800 0 20 40 60 80 100
Time Time Time

Fig. 13.2  Simulations of two-species communities in two sites connected by dispersal. The lines
represent the frequency of one of the species (species A) in each site. The solid and segmented lines
represent WAP and CSP, respectively. Since the number of individuals in each site is fixed (J), then
the frequency of species B is equal to 1 – freq. species A. The parameters m and b represent the
dispersal parameter and negative frequency dependency, respectively. Competitive exclusion is
evident when species A reaches a frequency of 1 or 0. It was assumed that Ulva is competitively
dominant in high-intertidal tidepools in WAP but not in CSP, which was expressed as fitness ratios
>1 (1.3) and <1 (1/1.1), respectively. The b parameter resembles the effects of grazers on the fre-
quency of competitors and was set to 0, −0.1, and −0.5, representing absence and intermediate and
high negative density dependence

dominance (Fig. 13.2d–f). Competitive exclusion by Ulva was expected under the


scenario of high dispersal and no or intermediate negative density dependence
(Fig. 13.2g, h). An important outcome of the simulation is that stable coexistence
can be reached when both dispersal between CPS and WAP and negative frequency
dependence are high (Fig. 13.2i). This could be caused by, on the one hand, ‘rescue
effect’ of immigrants in both sites (e.g. Altermatt et al. 2011) and, on the other hand,
by the controlling effect of consumers on the numerically dominant competitor.
274 N. Valdivia

13.5  Concluding Remarks

Benthic Antarctic grazers appear to have strong and deterministic effects on algal
communities across local environmental conditions, which can encompass a firm
biotic filter during the establishment stage of an alien seaweed. In addition, the evo-
lution of chemical anti-herbivory defences in Antarctic seaweed may provide them
with enemy-mediated competitive advantages over alien species. However, pro-
jected environmental conditions of warming and acidification can impair the ability
of amphipod grazers to control potential introductions in these ecosystems. The
results of a simple mathematical simulation, based on the introduction history of the
gutweed Ulva intestinalis, predict that intermediate to high levels of frequency-­
dependent consumption seem to be fundamental to allow for stable coexistence
when the alien species is competitively superior. This brief literature review and
simulations provide a benchmark to develop an experimental research agenda in the
WAP, in which competitive interactions between alien and native seaweeds could be
assessed as functions of consumption and abiotic climate change-related factors.
With this review, I hope to stimulate further empirical research on seaweed invasion
processes in Antarctica.

Acknowledgements  NV was financially supported by FONDAP 15150003 (IDEAL) and


FONDECYT 1190529 grants. Discussions with Viviana Segovia, Eliseo Fica, and Luis Miguel
Pardo contributed insightful ideas to the review. Gabriela Campana and Katharina Zacher provided
constructive criticism that greatly improved an early version of this manuscript.

References

Aldea C, Rosenfeld S (2011) Macromoluscos intermareales de sustratos rocosos de la playa Buque


Quemado, Estrecho de Magallanes, sur de Chile. Rev Biol Mar Oceanogr 46(2):115–124
Altermatt F, Bieger A, Carrara F, Rinaldo A, Holyoak M (2011) Effects of connectivity and recur-
rent local disturbances on community structure and population density in experimental meta-
communities. PLoS One 6(4):e19525
Amsler CD, McClintock JB, Baker BJ (1998) Chemical defense against herbivory in the Antarctic
marine macroalgae Iridaea cordata and Phyllophora antarctica (Rhodophyceae). J Phycol
34(1):53–59
Amsler CD, Katrin I, James BM, Margaret OA, Kevin JP, Joanna MH, Furrow FB, Bill JB (2005)
Comprehensive evaluation of the palatability and chemical defenses of subtidal macroalgae
from the Antarctic Peninsula. Mar Ecol Prog Ser 294:141–159
Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009a) Filamentous algal endophytes in mac-
rophytic Antarctic algae: prevalence in hosts and palatability to mesoherbivores. Phycologia
48(5):324–334
Amsler CD, Iken K, McClintock JB, Baker BJ (2009b) Defenses of polar macroalgae against her-
bivores and biofoulers. Bot Mar 52(6):535–545
Amsler CD, McClintock JB, Baker BJ (2012) Amphipods exclude filamentous algae from the
Western Antarctic Peninsula benthos: experimental evidence. Polar Biol 35(2):171–177
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 275

Amsler CD, McClintock JB, Baker BJ (2014) Chemical mediation of mutualistic interactions
between macroalgae and mesograzers structure unique coastal communities along the western
Antarctic Peninsula. J Phycol 50(1):1–10
Amsler MO, Huang YSM, Engl W, McClintock JB, Amsler CD (2015) Abundance and diver-
sity of gastropods associated with dominant subtidal macroalgae from the western Antarctic
Peninsula. Polar Biol 38(8):1171–1181
Amsler CD, Amsler MO, Curtis MD, McClintock JB, Baker BJ (2019) Impacts of gastropods
on epiphytic microalgae on the brown macroalga Himantothallus grandifolius. Antarct Sci
31(2):89–97
Aumack CF, Amsler CD, McClintock JB, Baker BJ (2010) Chemically mediated resistance to
mesoherbivory in finely branched macroalgae along the western Antarctic Peninsula. Eur J
Phycol 45(1):19–26
Aumack CF, Lowe AT, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2017) Gut content,
fatty acid, and stable isotope analyses reveal dietary sources of macroalgal-associated amphi-
pods along the western Antarctic Peninsula. Polar Biol 40(7):1371–1384
Barrera-Oro E, Moreira E, Seefeldt MA, Valli Francione M, Quartino ML (2019) The impor-
tance of macroalgae and associated amphipods in the selective benthic feeding of sister rock-
cod species Notothenia rossii and N. coriiceps (Nototheniidae) in West Antarctica. Polar Biol
42(2):317–334
Bender NA, Crosbie K, Lynch HJ (2016) Patterns of tourism in the Antarctic Peninsula region: a
20-year analysis. Antarct Sci 28(3):194–203
Bischoff B, Wiencke C (1993) Temperature requirements for growth and survival of macroalgae
from Disko Island (Greenland). Helgoländer Meeresun 47(2):167–191
Björk M, Axelsson L, Beer S (2004) Why is Ulva intestinalis the only macroalga inhabiting iso-
lated rockpools along the Swedish Atlantic coast? Mar Ecol Prog Ser 284:109–116
Björk M, Axelsson L, Beer S (2005) Photosynthetic traits of one alga can limit competition with
others: the case of Ulva Intestinalis growing alone in rockpools. Phycologia 44(4):7–7
Blackburn TM, Pyšek P, Bacher S, Carlton JT, Duncan RP, Jarošík V, Wilson JRU, Richardson DM
(2011) A proposed unified framework for biological invasions. Trends Ecol Evol 26(7):333–339
Blunden J, Arndt DS, Achberger C, Ackerman SA, Albanil A, Alexander P et al (2013) State of the
climate in 2012. Bull Am Meteorol Soc 94(8):S1–S258
Briski E, Chan FT, Darling JA, Lauringson V, MacIsaac HJ, Zhan A, Bailey SA (2018) Beyond
propagule pressure: importance of selection during the transport stage of biological invasions.
Front Ecol Environ 16(6):345–353
Bulleri F, Balata D, Bertocci I, Tamburello L, Benedetti-Cecchi L (2010) The seaweed Caulerpa
racemosa on Mediterranean rocky reefs: from passenger to driver of ecological change.
Ecology 91(8):2205–2212
Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML, Wiencke C (2009) Drivers
of colonization and succession in polar benthic macro- and microalgal communities. Bot Mar
52(6):655–667
Casaux RJ, Mazzotta AS, Barrera-Oro ER (1990) Seasonal aspects of the biology and diet of
nearshore Nototheniid fish at Potter Cove, South Shetland Islands, Antarctica. Polar Biol
11(1):63–72
Chesson P (2000) Mechanisms of maintenance of species diversity. Annu Rev Ecol Syst
31:343–366
Chown SL, Clarke A, Fraser CI, Cary SC, Moon KL, McGeoch MA (2015) The changing form of
Antarctic biodiversity. Nature 522(7557):431–438
Clark GF, Stark JS, Johnston EL, Runcie JW, Goldsworthy PM, Raymond B, Riddle MJ (2013)
Light-driven tipping points in polar ecosystems. Glob Chang Biol 19(12):3749–3761
Clayton MN, Wiencke C, Klöser H (1997) New records of temperate and sub Antarctic marine
benthic macroalgae from Antarctica. Polar Biol 17(2):141–149
276 N. Valdivia

Dubrasquet H, Reyes J, Sanchez RP, Valdivia N, Guillemin ML (2018) Molecular-assisted revision


of red macroalgal diversity and distribution along the Western Antarctic Peninsula and South
Shetland Islands. Cryptogam Algol 39(4):409–429
Duffy GA, Coetzee BWT, Latombe G, Akerman AH, McGeoch MA, Chown SL (2017) Barriers
to globally invasive species are weakening across the Antarctic. Divers Distrib 23(9):982–996
Fraser CI, Kay GM, Plessis M, Ryan PG (2017) Breaking down the barrier: dispersal across the
Antarctic polar front. Ecography 40(1):235–237
Fraser CI, Morrison AK, Hogg AM, Macaya EC, van Sebille E, Ryan PG et al (2018) Antarctica’s
ecological isolation will be broken by storm-driven dispersal and warming. Nat Clim Chang
8(8):704–708
Frenot Y, Chown SL, Whinam J, Selkirk PM, Convey P, Skotnicki M, Bergstrom DM (2005)
Biological invasions in the Antarctic: extent, impacts and implications. Biol Rev 80(1):45–72
Gómez I (ed) (2015) Flora marina antártica: patrimonio de biodiversidad. Ediciones Kultrún,
Valdivia
Griffiths HJ, Waller CL (2016) The first comprehensive description of the biodiversity and bioge-
ography of Antarctic and sub-Antarctic intertidal communities. J Biogeogr 43(6):1143–1155
Griffiths HJ, Meijers AJS, Bracegirdle TJ (2017) More losers than winners in a century of future
Southern Ocean seafloor warming. Nat Clim Chang 7(10):749–754
Guisan A, Thuiller W (2005) Predicting species distribution: offering more than simple habitat
models. Ecol Lett 8(9):993–1009
Hawkins SJ, Hartnoll RG (1983) Grazing of intertidal algae by marine invertebrates. Oceanogr
Mar Biol 21:195–282
HilleRisLambers J, Adler PB, Harpole WS, Levine JM, Mayfield MM (2012) Rethinking com-
munity assembly through the lens of coexistence theory. Annu Rev Ecol Evol Syst 43:227–248
Holt RD, Grover J, Tilman D (1994) Simple rules for interspecific dominance in systems with
exploitative and apparent competition. Am Nat 144(5):741–771
Huang YM, McClintock JB, Amsler CD, Peters KJ, Baker BJ (2006) Feeding rates of common
Antarctic gammarid amphipods on ecologically important sympatric macroalgae. J Exp Mar
Biol Ecol 329(1):55–65
Huang YM, Amsler MO, McClintock JB, Amsler CD, Baker BJ (2007) Patterns of gammaridean
amphipod abundance and species composition associated with dominant subtidal macroalgae
from the western Antarctic Peninsula. Polar Biol 30(11):1417–1430
Hubbell SP (2001) The unified neutral theory of biodiversity and biogeography. Princeton
University Press, Princeton
Hughes KA, Convey P (2014) Alien invasions in Antarctica-is anyone liable? Polar Res 33:1–13
Keane RM, Crawley MJ (2002) Exotic plant invasions and the enemy release hypothesis. Trends
Ecol Evol 17(4):164–170
Kennicutt MC, Chown SL, Cassano JJ, Liggett D, Massom R, Peck LS et al (2014) Six priorities
for Antarctic science. Nature 512(7512):23–25
Kim D (2001) Seasonality of marine algae and grazers of an Antarctic rocky intertidal, with
emphasis on the role of the limpet Nacella concinna Strebel (Gastropoda: Patellidae). PhD
Thesis, University of Bremen, Bremen
King GE, Howeth JG (2019) Propagule pressure and native community connectivity interact to
influence invasion success in metacommunities. Oikos 128:1549–1564
Kotta J, Wernberg T, Jänes H, Kotta I, Nurkse K, Pärnoja M, Orav-Kotta H (2018) Novel crab
predator causes marine ecosystem regime shift. Sci Rep 8(1):4956–4963
Larsson C, Axelsson L, Ryberg H, Beer S (1997) Photosynthetic carbon utilization by Enteromorpha
intestinalis (Chlorophyta) from a Swedish rockpool. Eur J Phycol 32(1):49–54
Letten AD, Ke PJ, Fukami T (2017) Linking modern coexistence theory and contemporary niche
theory. Ecol Monogr 87(2):161–177
Lopez DN, Camus PA, Valdivia N, Estay SA (2017) High temporal variability in the occurrence of
consumer-resource interactions in ecological networks. Oikos 126(12):1699–1707
13  Seaweed-Herbivore Interactions: Grazing as Biotic Filtering in Intertidal Antarctic… 277

Lubchenco J (1978) Plant species diversity in a marine intertidal community: importance of herbi-
vore food preference and algal competitive abilities. Am Nat 112(983):23–39
McCarthy AH, Peck LS, Hughes KA, Aldridge DC (2019) Antarctica: the final frontier for marine
biological invasions. Glob Chang Biol 25(7):2221–2241
Moran PAP (1958) Random processes in genetics. Math Proc Camb Philos Soc 54(1):60–71
Navarro JM, Paschke K, Ortiz A, Vargas-Chacoff L, Pardo LM, Valdivia N (2019) The Antarctic
fish Harpagifer antarcticus under current temperatures and salinities and future scenarios of
climate change. Prog Oceanogr 174:37–43
Noè S, Badalamenti F, Bonaviri C, Musco L, Fernández TV, Vizzini S, Gianguzza P (2018) Food
selection of a generalist herbivore exposed to native and alien seaweeds. Mar Pollut Bull
129(2):469–473
Noël L, Hawkins SJ, Jenkins SR, Thompson RC (2009) Grazing dynamics in intertidal rockpools:
connectivity of microhabitats. J Exp Mar Biol Ecol 370(1–2):9–17
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa LH,
Colepicolo P (2017) Diversity and spatial distribution of seaweeds in the South Shetland
Islands, Antarctica: an updated database for environmental monitoring under climate change
scenarios. Polar Biol 40(8):1671–1685
Peters AF (2003) Molecular identification, taxonomy and distribution of brown algal endophytes,
with emphasis on species Antarctica. In: Chapman ARO, Anderson RJ, Vreeland V, Davison
IR (eds) Proceedings of the 17th International Seaweed Symposium. Oxford University Press,
New York, pp 293–302
Poore AGB, Campbell AH, Coleman RA, Edgar GJ, Jormalainen V, Reynolds PL et  al (2012)
Global patterns in the impact of marine herbivores on benthic primary producers. Ecol Lett
15(8):912–922
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in Potter Cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223
R Core Team (2019) R: a language and environment for statistical computing, 3.0.3 edn. R
Foundation for Statistical Computing, Vienna
Sanches PF, Pellizzari F, Horta PA (2016) Multivariate analyses of Antarctic and sub-Antarctic
seaweed distribution patterns: an evaluation of the role of the Antarctic Circumpolar Current.
J Sea Res 110:29–38
Schoenrock KM, Schram JB, Amsler CD, McClintock JB, Angus RA, Vohra YK (2016) Climate
change confers a potential advantage to fleshy Antarctic crustose macroalgae over calcified
species. J Exp Mar Biol Ecol 474:58–66
Schram JB, Schoenrock KM, McClintock JB, Amsler CD, Angus RA (2014) Multiple stressor
effects of near-future elevated seawater temperature and decreased pH on righting and escape
behaviors of two common Antarctic gastropods. J Exp Mar Biol Ecol 457:90–96
Schram JB, McClintock JB, Amsler CD, Baker BJ (2015) Impacts of acute elevated seawater
temperature on the feeding preferences of an Antarctic amphipod toward chemically deterrent
macroalgae. Mar Biol 162(2):425–433
Schram JB, Schoenrock KM, McClintock JB, Amsler CD, Angus RA (2016) Seawater acidifica-
tion more than warming presents a challenge for two Antarctic macroalgal-associated amphi-
pods. Mar Ecol Prog Ser 554:81–97
Segovia-Rivera V, Valdivia N (2016) Independent effects of grazing and tide pool habitats on the
early colonisation of an intertidal community on western Antarctic Peninsula. Rev Chil Hist
Nat 89:1–9
Silva R, Vinagre C, Kitahara MV, Acorsi IV, Mizrahi D, Flores AAV (2019) Sun coral invasion of
shallow rocky reefs: effects on mobile invertebrate assemblages in southeastern Brazil. Biol
Invasions 21(4):1339–1350
Simberloff D, Martin JL, Genovesi P, Maris V, Wardle DA, Aronson J et al (2013) Impacts of bio-
logical invasions: what’s what and the way forward. Trends Ecol Evol 28(1):58–66
278 N. Valdivia

Valdivia N, Díaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9(6):e100714
Valdivia N, Pardo LM, Macaya EC, Huovinen P, Gómez I (2019) Different ecological mechanisms
lead to similar grazer controls on the functioning of periphyton Antarctic and sub-Antarctic
communities. Prog Oceanogr 174:7–16
Vellend M (2016) The theory of ecological communities (MPB-57). Princeton University Press,
Princeton
Vitousek PM, DAntonio CM, Loope LL, Rejmanek M, Westbrooks R (1997) Introduced species: a
significant component of human-caused global change. N Z J Ecol 21(1):1–16
White LF, Shurin JB (2011) Density dependent effects of an exotic marine macroalga on native
community diversity. J Exp Mar Biol Ecol 405(1–2):111–119
Zacher K, Hanelt D, Wiencke C, Wulff A (2007a) Grazing and UV radiation effects on an Antarctic
intertidal microalgal assemblage: a long-term field study. Polar Biol 30(9):1203–1212
Zacher K, Wulff A, Molis M, Hanelt D, Wiencke C (2007b) Ultraviolet radiation and consumer
effects on a field-grown intertidal macroalgal assemblage in Antarctica. Glob Chang Biol
13(6):1201–1215
Zacher K, Bernard M, Bartsch I, Wiencke C (2016) Survival of early life history stages of Arctic
kelps (Kongsfjorden, Svalbard) under multifactorial global change scenarios. Polar Biol
39(11):2009–2020
Zenteno L, Cárdenas L, Valdivia N, Gómez I, Höfer J, Garrido I, Pardo LM (2019) Unraveling
the multiple bottom-up supplies of an Antarctic nearshore benthic community. Prog Oceanogr
174:55–63
Chapter 14
Diversity and Functioning of Antarctic
Seaweed Microbiomes

Juan Diego Gaitan-Espitia and Matthias Schmid

Abstract Antarctic macroalgae are important primary producers and habitat-­


forming species that play fundamental roles in Antarctic coastal habitats and sustain
important communities of benthic organisms, including a not well-known microbi-
ota. Anthropogenic pressures, e.g., increasing ocean temperatures and extreme
events, have threatened the ecological integrity of several seaweed species and also
can modify the range shifts (e.g., introduction of Durvillaea antarctica in King
George Island), cause local extinctions, and alter the structure of these associations
in their natural habitats. However, understanding and prediction of the responses of
seaweeds to changing environment and rapid anthropogenic-driven change cannot
be done without considering the associated microbiome. These complex microbial
communities are intricately involved in the host health, defense, growth, and devel-
opment of seaweeds, thus with far-reaching implications for the ecology of the
whole coastal ecosystem. For most Antarctic seaweeds, the microbiome comprises
a stable core as well as microbes whose presence depends on local conditions and
transient microbial associates that are responsive to biotic and abiotic processes
across spatial and temporal scales. In this chapter, we will explore the ecological
and genetic diversity of microbiomes in Antarctic seaweeds and their functional
connections.

Keywords  Bacteria · Microbial communities · Genetic diversity · Holobionts ·


Microbiota

J. D. Gaitan-Espitia (*)
The Swire Institute of Marine Science and The School of Biological Sciences, The University
of Hong Kong, Hong Kong, SAR, China
e-mail: jdgaitan@hku.hk
M. Schmid
Institute for Marine and Antarctic Studies, University of Tasmania, Hobart, TAS, Australia
e-mail: matthias.schmid@utas.edu.au

© Springer Nature Switzerland AG 2020 279


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_14
280 J. D. Gaitan-Espitia and M. Schmid

14.1  I ntroduction: Environment and Antarctic Seaweed


Host-Microbiome

From very abundant and complex microbial communities to larger eukaryotes such
as seaweeds and marine mammals, Antarctic organisms have evolve a variety of
physiological, life history, and molecular adaptations that allow them to cope with
this challenging environment (Hoyer et al. 2002; Clark et al. 2004; Marx et al. 2007;
Rogers 2007). In Antarctic waters, the marine benthic flora is expected to be highly
susceptible to global changes (i.e., invasive species, ocean warming, ocean acidifi-
cation) due to its high degree of niche conservatism, reduced phenotypic plasticity,
and the low species richness (Clayton 1994; see also Chap. 11 by Gómez and
Huovinen). In order to survive in this harsh environment, Antarctic seaweeds have
evolved structural and functional adaptations such as high synthesis of photoprotec-
tive substances and antioxidant activity for mitigation of photodamage, molecular
adaptations of enzymes in order to maintain sufficient rates of enzyme-catalyzed
reactions of key metabolic processes, the evolution of cold shock and antifreeze
proteins, and different effective strategies for inorganic C acquisition and assimila-
tion (Morgan-Kiss et  al. 2006; Karsten et  al. 2009; Gómez et  al. 2009, 2019;
Huovinen and Gómez 2013; Hurd et al. 2014). While there has been a lot of interest
in documenting and understanding these functional adaptations, only scattered
information is available about other potential adaptive mechanisms that influence
functional regulation of Antarctic seaweeds. In fact, we know that seaweeds are not
independent biological units. They rely on tight relationships with their associated
microbiota for basic functions such as morphological development, growth, health,
defense, nutrient supply, and adaptation/acclimation to environmental stress (Egan
et al. 2013; Wichard 2015; Dittami et al. 2016; Singh and Reddy 2016). This sug-
gests that seaweeds and their microbiome interact as a unified functional entity or
holobiont (Egan et al. 2013). Therefore, it is essential to gain better understanding
of the adaptive role of seaweed host-microbiome interactions in changing/extreme
oceans and the mechanisms underlying their eco-evolutionary dynamics in
Antarctica.
The complex microbial communities (mutualistic symbionts and hazardous
pathogens) are intricately involved in health, defense, growth, and development of
seaweeds (Friedrich 2012; Egan et al. 2013; Martin et al. 2014; Egan and Gardiner
2016; Singh and Reddy 2016). For most eukaryotic hosts, the microbiome com-
prises a stable core as well as microbes whose presence depends on local conditions
and transient microbial associates that are responsive to biotic and abiotic processes
across spatial and temporal scales (Vandenkoornhuyse et  al. 2015; Hernandez-­
Agreda et al. 2016). Nevertheless, for seaweeds, the notion of host-specific micro-
bial taxa in the core microbiome is not a rule of thumb: some species have specific
microbial functional genes rather than taxonomic affiliations of microbial popula-
tions (Burke et al. 2011a, b). On the contrary, in other seaweeds, microbial taxo-
nomic diversity can be unique to each type of seaweed host but with high degree of
functional redundancy (Roth-Schulze et  al. 2016). Whether or not these patterns
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 281

(i.e., host-specific vs variable microbial diversity, functional redundancy) apply


more broadly to species from other seaweed phyla in extreme polar regions, and
considering temporal and spatial scales, is unknown.

14.2  F
 unctional Interactions of Antarctic Seaweeds
and Their Associated Microbiota

The seaweed microbiome represents a remarkably diverse array of microorganisms


that includes bacteria, archaea, fungi, and other eukaryotic unicellular organisms
and even viruses (Singh and Reddy 2016). Evidence suggests that the composition
of the seaweed microbiome is likely to be modulated, at least in part, by the host
because it is significantly different to the microbial community that is found in the
surrounding environment (Brodie et al. 2016). These microorganisms exhibit strong
seasonal and spatial changes in diversity and abundance (Tujula et al. 2010; Lachnit
et al. 2011; Campbell et al. 2015) that are also influenced by day length, nutrient
availability, and temperature (Gilbert et al. 2012; Moran 2015). Seaweeds can con-
trol and maintain communication with their associated microbes by producing sec-
ondary metabolites and exudates such as sugars and amino acids that serve as an
energy source, as well as for a variety of functions including antimicrobials, allo-
pathic molecules, and pathogen defenses (Friedrich 2012; Egan et al. 2013; Hollants
et al. 2013; Rout 2014). The seaweed microbiome, in turn, can exert influence on
trait expression by controlling growth and morphogenesis (Wichard 2015), acclima-
tion and physiological responses of the host to environmental gradients (Dittami
et al. 2016), survival and settlement of propagules (Morris et al. 2016), competition
among seaweeds by inhibiting the germination of algal spores (Egan et al. 2001),
and rapid defense adaptation (chemical controls) to the new bacterial epibionts and
pathogens during range shifts (Saha et al. 2016; Arnaud-Haond et al. 2017). All of
these effects influenced by associated microbial communities have important impli-
cations in the ecosystem services provided by the seaweed host (Rout 2014).
Although bacteria are by far the most abundant members of the seaweed microbi-
ome, our knowledge about epiphytic and endophytic bacterial communities living
in Antarctic seaweeds is quite limited. Very few studies have explored this compo-
nent of the seaweed microbiome. To date, a broad diversity of pigmented, Gram-­
positive epiphytic bacteria has been reported (mainly affiliated to Actinobacteria)
(Leiva et al. 2015), some of them with antibiotic activity that may influence micro-
bial dynamics (i.e., competition and colonization) of bacterial epibionts (Alvarado
et al. 2018). On the contrary, other components of the seaweed microbiome such as
fungi are very well documented. This group is characterized by the presence of
symbiont, saprobe, and parasitic species that form fungal assemblages on the sea-
weed host. The structure and dynamics of these assemblages are controlled by envi-
ronmental factors (e.g., availability of dissolved oxygen and organic matter), the
seaweed host, and the capacity of the fungi to tolerate or detoxify the array of
282 J. D. Gaitan-Espitia and M. Schmid

antifungal metabolites produced by the seaweed (Ogaki et  al. 2019). The typical
fungal community structure in Antarctic seaweeds is based on filamentous fungi
and yeasts belonging to the genera Geomyces, Antarctomyces, Oidiodendron,
Penicillium, Phaeosphaeria, Aureobasidium, Cryptococcus, Leucosporidium,
Metschnikowia, and Rhodotorula (Loque et al. 2010). These assemblages are domi-
nated by very few species (e.g., the filamentous fungus Pseudogymnoascus panno-
rum and the yeast Metschnikowia australis) (Loque et al. 2010; Godinho et al. 2013;
Furbino et al. 2014; Ogaki et al. 2019), some of which have the potential to degrade
algal biomass through agarolytic and carrageenolytic activities (Furbino et al. 2017).

14.3  D
 eciphering the Structure and Diversity
of Seaweed Microbiomes

Pioneering studies of microbial communities in Antarctic marine environments


employed cultivation-based methods, which are based on techniques aiming to
grow microorganisms under controlled laboratory conditions (Pham and Kim 2012).
This approach has several limitations as it only provides a reduced representation of
the microbial community diversity because most of the species in environmental
samples do not grow on standard media under laboratory conditions (Stewart 2012).
However, in the last few years, a major progress in microbial research has been
achieved, thanks to the development of several culture-independent molecular tech-
niques (e.g., polymerase chain reaction (PCR), hybridization, fingerprinting, Sanger
sequencing). These approaches have been routinely applied to study microbial com-
munities in Antarctic soils, water, ice, and biota (Loque et al. 2010; Godinho et al.
2013; Furbino et  al. 2014; Leiva et  al. 2015; Moreno-Pino et  al. 2016; Alvarado
et al. 2018; Castro-Sowinski 2019; Chua et al. 2018; Ogaki et al. 2019). From these,
the most powerful molecular approach to assess the structure and diversity of micro-
bial communities relies in the use of high-throughput sequencing (HTS) of house-
keeping genetic markers (amplicon sequencing) such as the internal transcribed
spacer region (ITS; fungi), the 18S gene (fungi and other eukaryotes), viral RNA
(viruses), and the 16 ribosomal RNA gene (16S rRNA; bacteria and archaea;
Fig. 14.1). The main drawback of microbial community profiling using these uni-
versal markers is that they cannot directly identify metabolic or other functional
capabilities of the microorganisms under study (Janda and Abbott 2007; Rausch
et al. 2019). This can be done using shotgun metagenomic sequencing (Fig. 14.1).
However, amplicon sequencing still offers many other advantages such as cost-­
efficiency, high precision, and fast speed characterization of taxonomic composition
and phylogenetic diversity of microbial communities (Langille et al. 2013; Rausch
et al. 2019). Although amplicon/targeted sequencing and shotgun sequencing strate-
gies differ in the type of information produced, phylogeny and biomolecular func-
tion are strongly correlated. For instance, phylogenetic trees based on 16S rRNA
closely resemble clusters obtained based on shared gene content, and researchers
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 283

Shotgun sequencing

Community DNA Bacteria genomes

Primers

Variable regions Constant regions


27F 357F 928F

V1 V2 V3 V4 V5 V6 V7 V8 V9
Targeted sequencing

518R 805R 926R 1492R

V1-V3
V3-V4

V3-V5

V6-V9

Fig. 14.1  Schematic representation of the workflow for bacterial community assessment. First,
epiphytes are collected from the seaweed host for posterior total DNA extraction. The DNA from
the microbiome is then used for targeted, amplicon sequencing (16S; down arrow) and/or shotgun
metagenome sequencing (right arrow). For the targeted approach, different combinations (red
lines) of variable (white boxes) and conserved (cyan boxes) 16S regions are sequenced via PCR
with traditional primers (blue, green, and pink arrows). For the shotgun sequencing approach,
DNA is fragmented and then sequenced using Illumina platforms, and the reads are finally sorted
and assembled into contigs and into circular genomes

often infer properties of uncultured organisms from cultured relatives (Langille


et al. 2013). Here, we will focus on targeted sequencing of bacterial 16S ribosomal
RNA gene (Fig.  14.1), as this approach has been frequently used to characterize
dynamics of microbial communities in Antarctic marine areas. Molecular approaches
used for characterizing other groups, such as fungi, are well described in Ogaki
et al. (2019).
The 16S rRNA gene is the most popular target gene in many molecular methods
(e.g., fluorescence in situ hybridization, FISH; quantitative PCR; terminal restric-
tion fragment length polymorphism, T-RFLP; denaturing-gradient gel electrophore-
sis, DGGE) of microbial studies (Bukin et al. 2019; Fuks et al. 2018; Janda and
Abbott 2007; Sambo et al. 2018). This gene offers several advantages for the study
of bacterial phylogeny and taxonomy. Such advantages are as follows:
(i) 16S rRNA is present in all prokaryotes.
(ii) Its function over time has not changed, suggesting that random sequence
changes are a more accurate measure of time (evolution).
(iii) It is approximately 1600 base pairs long and includes nine hypervariable, fast-­
evolving regions (V1–V9; Fig. 14.1), which can be used to classify organisms
at different taxonomic levels (genus or species).
284 J. D. Gaitan-Espitia and M. Schmid

(iv) The conserved, slow-evolving regions of this gene, which can be used for
determining higher-ranking taxa (Janda and Abbott 2007; Bukin et al. 2019).
These conserved regions have structural characteristics that allow to design broad-­
spectrum, degenerated primers for polymerase chain reaction (PCR) amplification,
which in turn can be used to isolate species-specific fast-evolving regions (Fig. 14.1)
(Sambo et al. 2018). The accuracy of 16S rRNA gene sequencing as a tool in micro-
bial identification (taxonomic assignment and phylogenetic placement) depends, in
a great extent, on the selection of the 16S region (Pootakham et al. 2017; Fuks et al.
2018; Sambo et al. 2018; Bukin et al. 2019). Nowadays, the majority of microbial
profiling studies utilizes the short-read V3–V4, V4–V5, or V5–V6 amplicons
instead of the full-length 16S rRNA sequences in environmental community sur-
veys (Pootakham et al. 2017). The advance in throughput has, however, come at the
cost of read length, and this trade-off has inevitably resulted in less accurate classi-
fication of partial 16S sequences, especially at the genus or species level (Bukin
et al. 2019; Pootakham et al. 2017). For seaweeds, there is currently no consensus
on the most appropriate hypervariable region(s) for profiling associated microbial
communities. Some of the studies using amplicon sequencing have targeted the 16S
hypervariable regions V1–V3 (Campbell et al. 2015), whereas others have used the
V3 (Lachnit et al. 2009), V4 (Lemay et al. 2018; Marzinelli et al. 2015), V3–V4
(Martin et al. 2015; Parrot et al. 2019), V6 (Brodie et al. 2016), or the full-length
16S (Leiva et al. 2015; Kumar et al. 2016; Alvarado et al. 2018; Serebryakova et al.
2018; Morrissey et al. 2019). Overall, and considering the percentage of sequences
that retrieve hits from public databases (e.g., Greengenes, NCBI, RDP, and SILVA),
it seems that hypervariable regions targeting V3–V4 (90%), V4 (91%), V4–V5
(88%), and V1–V9 (full-length; >99%) produce more reproducible results than V1–
V2 (30%) or V1–V3 (40%). These findings have been also documented in microbial
communities associated to other environmental and biological systems (Pootakham
et al. 2017; Almeida et al. 2018; Pollock et al. 2018).

14.4  V
 ariation of Bacterial Community Diversity
in Antarctic Seaweeds

In Antarctica, the diversity of marine prokaryotic (bacteria) communities in seawa-


ter correlates to both physical (Wilkins et  al. 2013) and chemical oceanographic
conditions (Giudice and Azzaro 2019). These factors can be highly variable in space
and time, shaping life in this polar region. As a result, prokaryotes have evolved a
wide range of pro-survival mechanisms such as habitat selection, life cycle strate-
gies, changes in cellular composition and enzyme activity, and the production of
extracellular polymeric substances (De Maayer et  al. 2014). These adaptations
might provide some functional benefits for the seaweed hosts. However, microbial
communities in seawater and sediments can differ in diversity and structure (as well
as in their functional properties) from those associated to seaweeds (Egan et  al.
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 285

2013). In coastal and marine sediments along the west part of the Antarctic Peninsula
and Antarctic islands, the predominant bacteria include representatives of the phyla
Proteobacteria, Actinobacteria, Chloroflexi, Bacteroidetes, Verrucomicrobia, and
Firmicutes, followed by Acidobacteria and Cyanobacteria (Chua et al. 2018; Flocco
et al. 2019). In seawater, on the other hand, the dominant phyla are Proteobacteria
(mainly Alphaproteobacteria and Gammaproteobacteria) and Bacteroidetes (Gentile
et al. 2006; Moreno-Pino et al. 2016). Although there are some similarities at the
higher taxonomic level, major differences in bacterial community diversity can be
found between sediments and seawater at the at genus level. In sediments, for exam-
ple, the most predominant genera is Sphingomonas, while in seawater the more
abundant bacteria belong to the family Rhodobacteraceae and the genera
Psychromonas, Pseudoalteromonas, and Balneatrix (Gentile et al. 2006; Moreno-­
Pino et al. 2016; Chua et al. 2018; Flocco et al. 2019; Giudice and Azzaro 2019; Lo
Giudice et al. 2019). This is also evidenced in bacterial communities associated to
Antarctic seaweeds (Fig.  14.2) (Leiva et  al. 2015; Alvarado et  al. 2018; Gaitan-­
Espitia et  al. unpublished). For instance, in a recent work, Gaitan-Espitia et  al.
(unpublished data) used the full-length 16S rRNA gene and PacBio SMRT sequenc-
ing in order to assess the diversity of microbiomes associated to some of the most
abundant seaweeds in the north of the Antarctic Peninsula. This study included rep-
resentative of the phylum Ochrophyta (Adenocystis utricularis, Geminocarpus
geminatus, Ascoseira mirabilis, Desmarestia antarctica, D. menziesii,

a b
i
rgi
be
amsleri

tts
ko

Seawater Sediments
as
Palm

ta
Paraglossum

da
rtin

r
co
aria

ga

ea
Po

da
Gi

Iri
rp

dec

ta
dicula
hy

appen
ra

phora
ipie

Phyllo
en
di

ns

Ulv
vi

a in
ifo

tes
tina
liu

lis
m

Lin
nae
us
Brown Red
Monostroma hariotii Seaweeds Seaweeds

Antarc
tosac
cion a
pplan
atum
-proteobacteria
-proteobacteria
Hi
s m -proteobacteria
lari an
icu -proteobacteria
ilis

to
utr
us

th -proteobacteria
tis all
irab

Green
at

ys us Bacteroidetes
De

noc
in

Desmarestia

gr Seaweeds
de Cyanobacteria
m

ira m

sm

A an
ge

Actinobacteria
di
are

fo
us

Verrucomicrobia
liu
ose
rp

sti

s Firmicutes
ca

Chlorobacteria
am
Asc
o
in

Acidobacteria
em

en
antarctica

Other bacteria
zie
G

sii

Fig. 14.2 (a) Phylogenetic relationships of Antarctic seaweeds; and (b) taxonomic distribution of
their associated microbiomes at the phylum level (Gaitan-Espitia unpublished)
286 J. D. Gaitan-Espitia and M. Schmid

Himantothallus grandifolius, Antarctosaccion applanatum), Chlorophyta


(Monostroma hariotii, Ulva intestinalis Linnaeus), and Rhodophyta (Porphyra
endiviifolium, Paraglossum amsleri, Phyllophora appendiculata, Iridaea cordata,
Gigartina skottsbergii, Palmaria decipiens) (Fig. 14.2). These species are character-
ized by different evolutionary histories, physiological and genomic architectures,
and distribution along the vertical zonation of Antarctic shores (Klöser et al. 1996;
Wiencke et al. 2007; Valdivia et al. 2014; Pellizzari et al. 2017). Overall, the find-
ings indicate that microbial community composition (at the family and genera level)
differs among species of seaweeds despite shared distribution in the vertical zona-
tion (Fig. 14.2). This suggests that for Antarctic seaweeds, the composition of epi-
phytic bacteria is, at least in part, likely regulated by the host (Gaitan-Espitia et al.
unpublished).
Studies analyzing the epiphytic and endophytic bacterial communities of tem-
perate and tropical seaweeds have shown different phylogenetic patterns across eco-
logical (e.g., temporal, spatial) and evolutionary (e.g., host) scales (Burke et  al.
2011b; Lachnit et al. 2011; Egan et al. 2013, 2017; Hollants et al. 2013; Campbell
et al. 2015; Marzinelli et al. 2015; Singh and Reddy 2016; Alvarado et al. 2018;
Serebryakova et al. 2018; Morrissey et al. 2019). This is consistent with findings
reported for Antarctic seaweed microbiomes (Alvarado et  al. 2018; Leiva et  al.
2015; Gaitan-Espitia et al. unpublished). However, in tropical and temperate sea-
weeds, Proteobacteria and Firmicutes are generally the most abundant bacteria
associated with seaweed hosts (Burke et al. 2011b; Lachnit et al. 2011; Egan et al.
2013, 2017; Hollants et al. 2013; Campbell et al. 2015; Marzinelli et al. 2015; Singh
and Reddy 2016; Serebryakova et al. 2018; Morrissey et al. 2019). In Antarctica, on
the other hand, members of phylum Actinobacteria are the most diverse and persis-
tent among different seaweed species, with a very low proportion of Firmicutes
(Leiva et  al. 2015; Alvarado et  al. 2018; Gaitan-Espitia et  al. unpublished). The
predominance of Actinobacteria is probably related to their essential roles in eco-
logical functions such as degradation of organic matter and maintenance of environ-
mental stability (Castro-Sowinski 2019). Within Actinobacteria, the most dominant
members belong to the family Micrococcaceae followed by Microbacteriaceae,
Dermabacteriaceae, Sanguibacteriaceae, and Nocardiaceae (Leiva et  al. 2015;
Gaitan-Espitia et  al. unpublished). Some of the genera related to these families
(Amycolatopsis, Arthrobacter, Agrococcus, Brevibacterium, Kocuria, Leucobacter,
Leifsonia, Microbacterium, Micrococcus, Micromonospora, Nocardiopsis,
Pseudarthrobacter, Pseudonocardia, Salinibacterium, and Streptomyces) are
known to possess antimicrobial activity (Hollants et al. 2013; Leiva et al. 2015),
likely controlling bacteria colonization and competition in the seaweed host (Leiva
et al. 2015; Busetti et al. 2017; Alvarado et al. 2018). Additionally, the high propor-
tion of some Actinobacteria, such as the pigmented, Gram-positive epiphytic bacte-
ria (Leiva et al. 2015), may confer some direct benefits to the seaweed host. These
non-photosynthetic bacteria have high concentration of carotenoids that protect the
cells against solar radiation and freeze–thaw cycles (Dieser et al. 2010), as well as
against oxidative damage (Glaeser and Klug 2005). Therefore, the regulation of
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 287

biofilm formation of these epiphytic bacteria may represent an adaptive mechanism


to enhance tolerances and survival of Antarctic seaweeds.

14.5  Conclusions and Future Perspectives

Species diversity is widely recognized as an essential adaptive ecological property


that increases the robustness and evolvability of biological systems (Kahilainen
et al. 2014). This applies to the framework of the seaweed holobiont, because the
diversity of microorganisms that live and interact with the seaweed host is known to
play a fundamental role in its health and resilience, particularly when faced with
environmental stress (Friedrich 2012; Egan et  al. 2013). Most studies exploring
seaweed microbiomes have determined the diversity (species and genes) of microbes
living in association with a seaweed host (mainly on surfaces = epiphytic communi-
ties) (Egan et al. 2001, 2013; Burke et al. 2011a, b; Friedrich 2012; Hollants et al.
2013; Rout 2014; Wichard 2015; Dittami et al. 2016; Morris et al. 2016; Saha et al.
2016; Arnaud-Haond et al. 2017). These studies reveal a seaweed microbiota that is
complex and very diverse and consists of a number of partners of different origins
and evolutionary trajectories (Friedrich 2012; Egan et al. 2013). However, diversity
without context provides limited insights into the mechanisms underpinning the
assembly of the microbiome, community patterns, and the benefits for the seaweed
host. High diversity is not necessarily “better” or “healthier,” whereas lower diver-
sity is not necessarily indicative of less stable or less “healthy” community (Shade
2016). In fact, microbial diversity can change across latitudinal–temporal scales
linked to biotic and abiotic factors rather than to the health or resilience of the host
(Gilbert et al. 2012; Koskella et al. 2017). Therefore, moving away from taxonomic
diversity toward functional diversity linked to the seaweed host phenotype and per-
formance under characterized environmental conditions could help us to understand
ecological drivers that shape the Antarctic seaweed microbiome and its functional
stability/plasticity in changing environments.

Acknowledgments  The work outlined in this review was partially supported by the Chilean
Antarctic Institute (INACH), the Commonwealth Scientific and Industrial Research Organisation
(CSIRO), the University of Hong Kong, and the Institute for Marine and Antarctic Studies (IMAS)
from the University of Tasmania. JDGE was supported by the Research Grants Council of Hong
Kong via the Early Career Scheme (Project ECS-27124318). MS was funded through the Deutsche
Forschungsgemeinschaft (DFG, grant ID: SCHM 3335/1-1).

References

Almeida A, Mitchell AL, Tarkowska A, Finn RD (2018) Benchmarking taxonomic assignments


based on 16S rRNA gene profiling of the microbiota from commonly sampled environments.
Gigascience 7:1–10
288 J. D. Gaitan-Espitia and M. Schmid

Alvarado P, Huang Y, Wang J, Garrido I, Leiva S (2018) Phylogeny and bioactivity of epiphytic
gram-positive bacteria isolated from three co-occurring Antarctic macroalgae. Antonie van
Leeuwenhoek Int J Gen Mol Microbiol 111:1543–1555
Arnaud-Haond S, Aires T, Candeias R, Teixeira S, Duarte C, Valero M, Serrão E (2017) Entangled
fates of holobiont genomes during invasion: nested bacterial and host diversities in Caulerpa
taxifolia. Mol Ecol 26:2379–2391
Brodie J, Williamson C, Barker GL, Walker RH, Briscoe A, Yallop M (2016) Characterising the
microbiome of Corallina officinalis, a dominant calcified intertidal red alga. FEMS Microbiol
Ecol 92:1–12
Bukin YS, Galachyants YP, Morozov IV, Bukin SV, Zakharenko AS, Zemskaya TI (2019)
The effect of 16s rRNA region choice on bacterial community metabarcoding results. Sci
Data 6:1–14
Burke C, Steinberg P, Rusch DB, Kjelleberg S, Thomas T (2011a) Bacterial community assembly
based on functional genes rather than species. Proc Natl Acad Sci U S A 108:14288–14293
Burke C, Thomas T, Lewis M, Steinberg P, Kjelleberg (2011b) Composition, uniqueness and vari-
ability of the epiphytic bacterial community of the green alga Ulva australis. ISME J 5:590–600
Busetti A, Maggs CA, Gilmore BF (2017) Marine macroalgae and their associated microbiomes as
a source of antimicrobial chemical diversity. Eur J Phycol 52:452–465
Campbell AH, Marzinelli EM, Gelber J, Steinberg PD (2015) Spatial variability of microbial
assemblages associated with a dominant habitat-forming seaweed. Front Microbiol 6:1–10
Castro-Sowinski S (ed) (2019) The ecological role of micro-organisms in the Antarctic environ-
ment. Springer Polar Sciences, Springer, Cham
Chua CY, Yong ST, González MA, Lavin P, Cheah YK, Tan GYA, Wong CMVL (2018) Analysis
of bacterial communities of King George and Deception Islands, Antarctica using high-­
throughput sequencing. Curr Sci 115:1701–1705
Clark MS, Clarke A, Cockell CS, Convey P, Detrich HW, Fraser KPP, Johnston IA, Methe BA et al
(2004) Antarctic genomics. Comp Funct Genomics 16:230–238
Clayton MN (1994) Evolution of the Antarctic marine benthic algal flora. J Phycol 30:897–904
De Maayer P, Anderson D, Cary C, Cowan DA (2014) Some like it cold: understanding the sur-
vival strategies of psychrophiles. EMBO Rep 15:508–517
Dieser M, Greenwood M, Foreman CM (2010) Carotenoid pigmentation in Antarctic heterotrophic
bacteria as a strategy to withstand environmental stresses. Arct Antarct Alp Res 42:396–405
Dittami SM, Duboscq-Bidot L, Perennou M, Gobet A, Corre E, Boyen C, Tonon T (2016) Host-­
microbe interactions as a driver of acclimation to salinity gradients in brown algal cultures.
ISME J 10:51–63
Egan S, Gardiner M (2016) Microbial dysbiosis: rethinking disease in marine ecosystems. Front
Microbiol 7:1–8
Egan S, James S, Holmström C, Kjelleberg S (2001) Inhibition of algal spore germination by the
marine bacterium Pseudoalteromonas tunicata. FEMS Microbiol Ecol 35:67–73
Egan S, Harder T, Burke C, Steinberg P, Kjelleberg S, Thomas T (2013) The seaweed holobiont:
understanding seaweed-bacteria interactions. FEMS Microbiol Rev 37:462–476
Egan S, Kumar V, Nappi J, Gardiner M (2017) Microbial diversity and symbiotic interactions with
macroalgae. In: Grube M, Seckbach J, Muggia L (eds) Algal and cyanobacteria symbioses.
World Scientific Publishing Europe Ltd, London/Singapore, pp 493–546
Flocco C, Cormack W, Smalla K (2019) Antarctic soil microbial communities in a changing
environment: their contributions to the sustainability of Antarctic ecosystems and the biore-
mediation of anthropogenic pollution. In: Castro-Sowinski S (ed) The ecological role of micro-­
organisms in the Antarctic environment. Springer Polar Sciences, Springer, Cham, pp 133–161
Friedrich MW (2012) Bacterial communities on macroalgae. In: Wiencke C, Bischof K (eds)
Seaweed biology: novel insights into ecophysiology, ecology and utilization. Springer, Berlin,
pp 189–201
Fuks G, Elgart M, Amir A, Zeisel A, Turnbaugh PJ, Soen Y, Shental N (2018) Combining
16S rRNA gene variable regions enables high-resolution microbial community profiling.
Microbiome 6:1–13
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 289

Furbino LE, Godinho VM, Santiago IF, Pellizari FM, Alves TMA, Zani CL, Junior PAS et  al
(2014) Diversity patterns, ecology and biological activities of fungal communities associated
with the endemic macroalgae across the Antarctic Peninsula. Microb Ecol 67:775–787
Furbino LE, Pellizzari FM, Neto PC, Rosa CA, Rosa LH (2017) Isolation of fungi associated with
macroalgae from maritime Antarctica and their production of agarolytic and carrageenolytic
activities. Polar Biol 41:527–535
Gentile G, Giuliano L, D’Auria G, Smedile F, Azzaro M, De Domenico M, Yakimov MM (2006)
Study of bacterial communities in Antarctic coastal waters by a combination of 16S rRNA and
16S rDNA sequencing. Environ Microbiol 8:2150–2161
Gilbert JA, Steele JA, Caporaso JG, Steinbrück L, Reeder J, Temperton B, Huse S, McHardy
AC, Knight R et al (2012) Defining seasonal marine microbial community dynamics. ISME J
6:298–308
Giudice A, Azzaro M (2019) Diversity and ecological roles of prokaryotes in the changing
Antarctic marine environment. In: Castro-Sowinski (ed) the ecological role of micro-organisms
in the Antarctic environment. Springer Polar Sciences, Springer, Cham, pp 109–131
Glaeser J, Klug G (2005) Photo-oxidative stress in Rhodobacter sphaeroides: protective role of
carotenoids and expression of selected genes. Microbiology 151:1927–1938
Godinho VM, Furbino LE, Santiago IF, Pellizzari FM, Yokoya NS, Pupo D, Alves TMA, Junior
PAS et al (2013) Diversity and bioprospecting of fungal communities associated with endemic
and cold-adapted macroalgae in Antarctica. ISME J 7:1434–1451
Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U, Quartino ML, Dunton K, Wiencke C
(2009) Light and temperature demands of marine benthic microalgae and seaweeds in polar
regions. Bot Mar 52:593–608
Gómez I, Navarro NP, Huovinen P (2019) Bio-optical and physiological patterns in Antarctic
seaweeds: a functional trait based approach to characterize vertical zonation. Prog Oceanogr
174:17–27
Hernandez-Agreda A, Leggat W, Bongaerts P, Ainsworth T (2016) The microbial signature pro-
vides insight into the mechanistic basis of coral success across reef habitats. MBio 7:1–10
Hollants J, Leliaert F, De Clerck O, Willems A (2013) What we can learn from sushi: a review on
seaweed-bacterial associations. FEMS Microbiol Ecol 83:1–16
Hoyer K, Karsten U, Wiencke C (2002) Induction of sunscreen compounds in Antarctic macroal-
gae by different radiation conditions. Mar Biol 141:619–627
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
Hurd C, Harrison P, Bischof K, Lobban C (eds) (2014) Seaweed ecology and physiology.
Cambridge University Press, Cambridge
Janda JM, Abbott SL (2007) 16S rRNA gene sequencing for bacterial identification in the diagnos-
tic laboratory: pluses, perils, and pitfalls. J Clin Microbiol 45:2761–2764
Kahilainen A, Puurtinen M, Kotiaho JS (2014) Conservation implications of species-genetic diver-
sity correlations. Glob Ecol Conserv 2:315–323
Karsten U, Wulff A, Roleda MY, Müller R, Steinhoff FS, Fredersdorf J, Wiencke C (2009)
Physiological responses of polar benthic algae to ultraviolet radiation. Bot Mar 52:639–654
Klöser H, Quartino ML, Wiencke C (1996). Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiologia 333: 1–17
Koskella B, Hall LJ, Metcalf CJE (2017) The microbiome beyond the horizon of ecological and
evolutionary theory. Nat Ecol Evol 1:1606–1615
Kumar V, Zozaya-Valdes E, Kjelleberg S, Thomas T, Egan S (2016) Multiple opportunistic patho-
gens can cause a bleaching disease in the red seaweed Delisea pulchra. Environ Microbiol
18:3962–3975
Lachnit T, Blümel M, Imhoff JF, Wahl M (2009) Specific epibacterial communities on macroalgae:
phylogeny matters more than habitat. Aquat Biol 5:181–186
290 J. D. Gaitan-Espitia and M. Schmid

Lachnit T, Meske D, Wahl M, Harder T, Schmitz R (2011) Epibacterial community patterns on


marine macroalgae are host-specific but temporally variable. Environ Microbiol 13:655–665
Langille M, Zaneveld J, Caporaso J, McDonald D, Knights D, Reyes J, Clemente J et al (2013)
Predictive functional profiling of microbial communities using 16S rRNA marker gene
sequences. Nat Biotechnol 31:814
Leiva S, Alvarado P, Huang Y, Wang J, Garrido I (2015) Diversity of pigmented gram-positive
bacteria associated with marine macroalgae from Antarctica. FEMS Microbiol Lett 362:1–6
Lemay MA, Martone PT, Hind KR, Lindstrom SC, Wegener-Parfrey L (2018) Alternate life his-
tory phases of a common seaweed have distinct microbial surface communities. Mol Ecol
27:3555–3568
Lo Giudice A, Caruso G, Rizzo C, Papale M, Azzaro M (2019) Bacterial communities versus
anthropogenic disturbances in the Antarctic coastal marine environment. Environ Sustain
2:297–310
Loque CP, Medeiros AO, Pellizzari FM, Oliveira EC, Rosa CA, Rosa LH (2010) Fungal commu-
nity associated with marine macroalgae from Antarctica. Polar Biol 33:641–648
Martin M, Portetelle D, Michel G, Vandenbol M (2014) Microorganisms living on macroal-
gae: diversity, interactions, and biotechnological applications. Appl Microbiol Biotechnol
98:2917–2935
Martin M, Barbeyron T, Martin R, Portetelle D, Michel G, Vandenbol M (2015) The cultiva-
ble surface microbiota of the brown alga Ascophyllum nodosum is enriched in macroalgal-­
polysaccharide-­degrading bacteria. Front Microbiol 6:1–14
Marx JC, Collins T, D’Amico S, Feller G, Gerday C (2007) Cold-adapted enzymes from marine
Antarctic microorganisms. Mar Biotechnol 9:293–304
Marzinelli E, Campbell A, Zozaya E, Vergés A, Nielsen S, Wernberg T, de Bettignies T et al (2015)
Continental-scale variation in seaweed host-associated bacterial communities is a function of
host condition, not geography. Environ Microbiol 17:4078–4088
Moran M (2015) The global ocean microbiome. Science 80:350, aac84551–aac84556
Moreno-Pino M, De la Iglesia R, Valdivia N, Henríquez-Castilo C, Galán A, Díez B, Trefault N
(2016) Variation in coastal Antarctic microbial community composition at sub-mesoscale: spa-
tial distance or environmental filtering? FEMS Microbiol Ecol 92:1–13
Morgan-Kiss RM, Priscu JC, Pocock T, Gudynaite-Savitch L, Huner NPA (2006) Adaptation and
acclimation of photosynthetic microorganisms to permanently cold environments microbiol.
Mol Biol Rev 70:222–252
Morris M, Haggerty J, Papudeshi B, Vega A, Edwards M, Dinsdale E (2016) Nearshore pelagic
microbial community abundance affects recruitment success of giant kelp, Macrocystis pyrif-
era. Front Microbiol 7:1–12
Morrissey KL, Çavas L, Willems A, De Clerck O (2019) Disentangling the influence of environ-
ment, host specificity and thallus differentiation on bacterial communities in siphonous green
seaweeds. Front Microbiol 10:1–12
Ogaki MB, de Paula MT, Ruas D, Pellizzari FM, García-Laviña CX, Rosa LH (2019) Marine fungi
associated with Antarctic macroalgae. In: Castro-Sowinski S (ed) The ecological role of micro-­
organisms in the Antarctic environment. Springer Polar Sciences, Springer, Cham, pp 239–255
Parrot D, Blümel M, Utermann C, Chianese G, Krause S, Kovalev A, Gorb SN, Tasdemir D (2019)
Mapping the surface microbiome and metabolome of brown seaweed Fucus vesiculosus by
amplicon sequencing, integrated metabolomics and imaging techniques. Sci Rep 9:1–17
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa LH,
Colepicolo P (2017) Diversity and spatial distribution of seaweeds in the South Shetland
Islands, Antarctica: an updated database for environmental monitoring under climate change
scenarios. Polar Biol 40:1671–1685
Pham VHT, Kim J (2012) Cultivation of unculturable soil bacteria. Trends Biotechnol 30:475–484
Pollock J, Glendinning L, Wisedchanwet T, Watson M (2018) The madness of microbiome:
attempting to find consensus. Appl Environ Microbiol 84:1–12
14  Diversity and Functioning of Antarctic Seaweed Microbiomes 291

Pootakham W, Mhuantong W, Yoocha T, Putchim L, Sonthirod C, Naktang C, Thongtham N,


Tangphatsornruang S (2017) High resolution profiling of coral-associated bacterial communi-
ties using full-length 16S rRNA sequence data from PacBio SMRT sequencing system. Sci
Rep 7:1–14
Rausch P, Hermes B, Doms S, Dagan T, Domin H, Fraune S, Humeida UH, Heinsen F, Jahn M,
Jaspers C et al (2019) Comparative analysis of amplicon and metagenomic sequencing meth-
ods reveals key features in the evolution of animal metaorganisms. Microbiome 7:1–19
Rogers AD (2007) Evolution and biodiversity of Antarctic organisms: a molecular perspective.
Philos Trans R Soc B Biol Sci 362:2191–2214
Roth-Schulze AJ, Zozaya-Valdés E, Steinberg PD, Thomas T (2016) Partitioning of functional
and taxonomic diversity in surface-associated microbial communities. Environ Microbiol
18:4391–4402
Rout ME (2014) The plant microbiome. In: Paterson AH (ed) Genomes of herbaceous land plants.
Advances in Botanical Research. Elsevier Ltd., London, pp 279–309
Saha M, Wiese J, Weinberger F, Wahl M (2016) Rapid adaptation to controlling new microbial
epibionts in the invaded range promotes invasiveness of an exotic seaweed. J Ecol 104:969–978
Sambo F, Finotello F, Lavezzo E, Baruzzo G, Masi G, Peta E, Falda M, Toppo S, Barzon L, Di
Camillo B (2018) Optimizing PCR primers targeting the bacterial 16S ribosomal RNA gene.
BMC Bioinformatics 19:1–10
Serebryakova A, Aires T, Viard F, Serrão EA, Engelen AH (2018) Summer shifts of bacterial com-
munities associated with the invasive brown seaweed Sargassum muticum are location and
tissue dependent. PLoS One 13:1–18
Shade A (2016) Diversity is the question, not the answer. ISME J 4:e2287v1
Singh R, Reddy C (2016) Unraveling the functions of the macroalgal microbiome. Front
Microbiol 6:1–8
Stewart EJ (2012) Growing unculturable bacteria. J Bacteriol 194:4151–4160
Tujula NA, Crocetti GR, Burke C, Thomas T, Holmström C, Kjelleberg S (2010) Variability and
abundance of the epiphytic bacterial community associated with a green marine ulvacean alga.
ISME J 4:301–311
Valdivia N, Díaz MJ, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all
around: scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula,
King George Island, Antarctica. PLoS One 9:1–12
Vandenkoornhuyse P, Quaiser A, Duhamel M, Le Van A, Dufresne A (2015) The importance of the
microbiome of the plant holobiont. New Phytol 206:1196–1206
Wichard T (2015) Exploring bacteria-induced growth and morphogenesis in the green macroalga
order Ulvales (Chlorophyta). Front Plant Sci 6:1–19
Wiencke C, Clayton MN, Gómez I, Iken K, Lüder UH, Amsler CD, Karsten U, Hanelt D, Bischof
K, Dunton K (2007) Life strategy, ecophysiology and ecology of seaweeds in polar waters. Rev
Environ Sci Biotechnol 6:95–126
Wilkins D, Yau S, Williams TJ, Allen MA, Brown MV, Demaere MZ, Lauro FM, Cavicchioli
R (2013) Key microbial drivers in Antarctic aquatic environments. FEMS Microbiol Rev
37:303–335
Chapter 15
Seaweeds in the Antarctic Marine Coastal
Food Web

Fernando R. Momo, Georgina Cordone, Tomás I. Marina, Vanesa Salinas,


Gabriela L. Campana, Mariano A. Valli, Santiago R. Doyle,
and Leonardo A. Saravia

Abstract  Antarctic macroalgae are the basis of marine food webs in most coastal
environments, especially the more confined ones such as bays and fjords. Whether
through direct consumption or via detritus, their role in maintaining biodiversity is

F. R. Momo (*)
Instituto de Ciencias, Universidad Nacional de General Sarmiento,
Los Polvorines, Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu), Luján, Buenos
Aires, Argentina
e-mail: fmomo@campus.ungs.edu.ar
G. Cordone
Centro Nacional Patagónico (CCT CONICET-CENPAT), Centro para el Estudio de Sistemas
Marinos (CESIMAR), Puerto Madryn, Chubut, Argentina
e-mail: gcordone@cenpat-conicet.gob.ar
T. I. Marina
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), Centro Austral de
Investigaciones Científicas (CADIC), Ushuaia, Tierra del Fuego, Argentina
e-mail: tomasimarina@cadic-conicet.gob.ar
V. Salinas · L. A. Saravia
Instituto de Ciencias, Universidad Nacional de General Sarmiento,
Los Polvorines, Buenos Aires, Argentina
e-mail: vsalinas@campus.ungs.edu.ar; lsaravia@campus.ungs.edu.ar
G. L. Campana
Departamento de Biología Costera, Instituto Antártico Argentino, Buenos Aires, Argentina
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
e-mail: gcampana@dna.gov.ar
M. A. Valli
Departamento de Ciencias Básicas, Universidad Nacional de Luján (UNLu),
Luján, Buenos Aires, Argentina
S. R. Doyle
Instituto de Ecología y Desarrollo Sustentable (INEDES), Universidad Nacional de Luján
(UNLu), Luján, Buenos Aires, Argentina
e-mail: sdoyle@campus.ungs.edu.ar

© Springer Nature Switzerland AG 2020 293


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_15
294 F. R. Momo et al.

essential. However, their relevance is due not only by direct trophic interactions but
also by indirect feedbacks, since macroalgae act as a habitat and refuge for multiple
benthic organisms and as a substrate for epiphytic microalgae. Macroalgae also
establish relations of exploitative competition, apparent competition, and mutual-
ism. The control over the biomass and diversity of the macroalgae itself does not
seem to be due to trophic interactions (top-down control) but rather to competition
and diverse abiotic factors such as substrata and light availability or physical distur-
bances (ice scouring). The extreme connectivity of trophic networks linked to algae
and their detritus determines that food webs are robust to local extinctions; however,
non-trophic interactions indicate that changes that affect the growth, biomass, and
distribution of macroalgae can have dramatic effects on the diversity of their associ-
ated fauna and, indirectly, on the networks of consumers of that fauna. In this chap-
ter, we present a detailed description of macroalgae relationship networks and
analyze the stability of the Antarctic community using food web theory.

Keywords  Biological interactions · Community structure · Ecological networks ·


Trophic level

15.1  Introduction

Macroalgae and their detritus or fragments are the basal energy source of Antarctic
coastal food webs. Approximately 150 species of macroalgae have been reported at
Antarctica (Wiencke et al. 2014; Pellizzari et al. 2017; see also Chap. 2 by Oliveira
et al.). Several local studies proved that macroalgae together with microphytoben-
thos constitute a direct pathway of energy and matter into organisms that feed on
them (Wiencke et al. 2014; Ahn et al. 2016) and indirectly through detritus pathway
(Iken et al. 1997; Tatián et al. 2004; Amsler et al. 2005, 2012; Quartino et al. 2008;
Campana et al. 2018). Besides this, benthic algae in some locations can have high
primary productivity similar to or higher than phytoplankton production (Fogg
1977). Furthermore, Antarctic macroalgae constitute structural and even chemical
refuges from predation (Amsler et al. 1999, 2009), habitat for a variety of associated
fauna (Huang et al. 2007) that provide a large fraction of the secondary production
to the benthos (Gómez et  al. 2009), and substrate for epiphytic communities
(Majewska et al. 2016). Thus, macroalgae play a fundamental role in the food web
being able to influence ecosystem dynamics and stability through propagation of
direct and indirect effects.
The network of interspecific relationships involving seaweeds is very complex
and includes not only consumption of algae but also several types of relations.
Direct interactions are, for instance, competition among different macroalgae spe-
cies for space (Smale and Barnes 2008; Quartino et al. 2013; Campana et al. 2009,
15  Seaweeds in the Antarctic Marine Coastal Food Web 295

2018), when hard substrate is a limiting resource, or for light (Klöser et al. 1994,
1996; Deregibus et al. 2016). On the other hand, seaweeds can establish relation-
ships with potentially harmful light-blocking epiphytes, whereas macroalgae pro-
vide a substrate for these organisms; however, seaweeds are chemically defended
against mesograzers (and other herbivores; see Chap. 18 by Amsler et al.) that, in
turn, use them as a refuge from predation: the macroalgae benefit in return because
the mesograzers remove epiphytic algae (Amsler et al. 2014), and in this way, com-
mensalism can become mutualism.
However, not only direct effects play an important role: several seaweed species
share predators and establish a complex dynamics that may be observed as much as
apparent competition or “apparent” mutualism depending on the density of preda-
tors and the closeness of macroalgae species involved. When a group of species
establish different types of interactions (i.e., trophic, commensalism, mutualism) in
an intricate manner, it is challenging to decide if the regulation of numbers or bio-
mass is mainly controlled by top-down, bottom-up, or wasp-waist effects or by dif-
ferent and nontrivial combinations of them. In consequence, it is necessary to take
into account the complexity of the ecosystem and all its ecological interactions (tro-
phic and non-trophic) to understand its stability in response to environmental
changes. In particular, macroalgal species are exposed to different environmental
disturbances such as sediment input or ice scouring (Sahade et al. 2015; Quartino
et al. 2013; Deregibus et al. 2016), resulting in a complex mosaic of effects. In this
sense, the understanding of the effects of climate change in the western Antarctic
Peninsula and a possible losses and gains in biodiversity urgently needs a deep
knowledge of the relations between species’ functional roles and the ecosystem
structure (Woodward et al. 2010).

15.2  Food Webs and Seaweeds

The complex network of interactions varies in space and time, and not all interac-
tions take place simultaneously (Poisot et al. 2015), but the interactions should not
be studied separately but be understood as a whole. For this purpose, we have a
powerful tool: the food web theory (Delmas et al. 2018).
After comparing 19 food web properties, Dunne et al. (2004) concluded that the
excessively low percentage of basal taxa in marine food webs compared to other
systems is clearly an artifact due to the poor resolution of primary producers and
consumer links to them. One of the methodological strengths of the food web stud-
ied here is the high taxonomic resolution of basal nodes. A good taxonomic resolu-
tion of the lower trophic levels, such as the macroalgal community, is essential to
understand ecosystem functioning, since there seems to be a species-specific selec-
tive consumption (Iken et  al. 1997). Implications for ecosystem functioning and
stability are only possible to elucidate in food webs where the species involved in
energy and matter transfer processes are well represented.
296 F. R. Momo et al.

A first step to identify the seaweeds embedded in the functional context of their
community is to visualize the food web in which they are integrated. For example,
Fig. 15.1 shows the food web of the Potter Cove ecosystem (Marina et al. 2018). We
can see that the base of the pyramid is filled with several macroalgal species accom-
panied by fresh and old detritus and microalgae. Arrows indicate matter and energy
fluxes through predation; the node size is proportional to its number of connections;
the placement of the node on the vertical coordinate is related to its weighted tro-
phic level. Also, macroalgae contribute to the detritivore pathway by decomposition
and accumulation in the cove, generating a link between macroalgae species and
detritus that is not represented in Fig. 15.1. However, we consider it in our studies
and conclusions.
An understanding of the relations between species functional roles and ecosystem
structure is an indispensable step toward the comprehension of change in the western
Antarctic Peninsula and subsequent biodiversity loss and gain (Woodward et  al.
2010). Several network properties are commonly used to describe food webs (Dunne
et  al. 2002): (1) number of species, S; (2) total number of interactions or trophic
links, L; (3) number of interactions per species or linkage density, L/S; (4) con-
nectance or trophic links divided by total number of possible interactions, C = L/S2;
(5) percentage of top species (species with preys, but without predators); (6) interme-
diate species (species having preys and predators); (7) basal species (species with
predators/consumers, but without preys); and (8) percentage of omnivores (species
eating prey from more than one trophic level). Additionally, the topology of the food
web can be studied by measuring three more properties: (9) characteristic path length
(ChPath), which is the average shortest path length between all pairs of species; (10)
clustering coefficient (CC), which is the average fraction of pairs of species one link
away from a species that are also linked to each other; and (11) distribution degree,
which is the fraction of trophic species P(k) that have k or more trophic links (both
predator and prey links) (Albert and Barabási 2002). These last three metrics give us
information related to the degree of self-organization shown by the web.
The trophic levels (TL) of species were calculated using the short-weighted TL
(Williams and Martinez 2004). Short-weighted trophic level is defined as the aver-
age of the shortest TL and prey-averaged TL. Shortest TL of a consumer in a food
web is equal to 1+ the shortest chain length from this consumer to any basal species.
Prey-averaged TL is equal to 1+ the mean TL of all consumer’s trophic resources,
calculated as follows:
S
TLi
TLj = 1 + ∑lij
nj.
i =1

where TLj is the trophic level of species j; S is the total number of species in the food
web; lij are the elements of the connection matrix with S rows and S columns; for
column j and row i, lij is 1 if species j consumes species i and 0 if not; and nj is the
number of prey species in the diet of species j. Therefore, short-weighted TL yields
a minimum estimate of TL and assumes a value of 1.0 for basal species (Williams
and Martinez 2004). We considered the mean TL of the web as the average of all
species’ TL.
15  Seaweeds in the Antarctic Marine Coastal Food Web 297

Fig. 15.1  Food web of Potter Cove. Vertical position is related to trophic level. Node size is pro-
portional to the total degree (in and out). Node colors are by functional group. Nodes inside the
boxes are the basal species. Network was plotted with Visone (version 2.9.2). 1Notothenia cori-
iceps, 2Notothenia rossii, 3Lepidonotothen nudifrons, 4Trematomus newnesi, 5Trematomus bernac-
chii, 6Harpagifer antarcticus, 7Parachaenichthys charcoti, 8Chaenocephalus aceratus,
9
Protomyctophum sp., 10Callophyllis atrosanguinea, 11Curdiea racovitzae, 12Georgiella confluens,
13
Gigartina skottsbergii, 14Iridaea cordata, 15Myriogramme manginii, 16Neuroglossum delesseriae,
17
Palmaria decipiens, 18Pantoneura plocamioides, 19Picconiella plumosa, 20Plocamium cartilag-
ineum, 21Porphyra plocamiestris, 22Trematocarpus antarcticus, 23Adenocystis utricularis,
24
Ascoseira mirabilis, 25Desmarestia anceps, 26Desmarestia antarctica, 27Desmarestia menziesii,
28
Geminocarpus geminatus, 29Phaeurus antarcticus, 30Lambia antarctica, 31Monostroma hariotii,
32
Urospora penicilliformis, 33Ulothrix sp., 34Epiphytic diatoms, 35Benthic diatoms, 36Phytoplankton,
37
Aged detritus, 38Nereidae, 39Margarella antarctica, 40Austrodoris kerguelensis, 41Eatoniella sp.,
42
Nacella concinna, 43Laevilacunaria antarctica, 44Dacrydyum sp., 45Laternula elliptica,
46
Neobuccinum eatoni, 47Euphausia superba, 48Paradexamine sp., 49Eurymera monticulosa,
50
Pontogeneiella sp., 51Gondogeneia antarctica, 52Hyperiids, 53Pariphimedia integricauda,
54
Bovallia gigantea, 55Cheirimedon femoratus, 56Gitanopsis antarctica, 57Prostebbingia gracilis,
58
Waldeckia obesa, 59Hippo-Orcho (Hippomedon kergueleni and Orchomene plebs collapsed),
60
Oradarea bidentata, 61Serolis sp., 62Glyptonotus antarcticus, 63Plakarthrium puncattissimum,
64
Hemiarthrum setulosum, 65Ophionotus victoriae, 66Odontaster validus, 67Diplasterias brucei,
68
Odontaster meridionalis, 69Perknaster fuscus antarticus, 70Perknaster aurorae, 71Sterechinus
neumayeri, 72squids, 73Copepods, 74Ascidians, 75Octopus sp., 76Oligochaetes, 77Hydrozoa,
78
Bryozoa, 79Priapulids, 80Parborlasia corrugatus, 81Salpidae, 82Mysida, 83fresh detritus, 84necro-
mass, 85zooplankton, 86Haliclonidae sp., 87Stylo-Myca (Stylocordila borealis and Mycale acerata
collapsed), 88Rosella sp., 89Dendrilla antarctica, 90Urticinopsis antarctica, 91Malacobelemnon day-
toni. (Modified from Marina et al. 2018)
298 F. R. Momo et al.

Fig. 15.2  Common-enemy graph for basal species of Potter Cove food web. Nodes represent spe-
cies and link indirect interactions (presence of shared predators). Only prey species are shown.
Node colors identify different preys (type of algae, or detritus, or microalgae). Node and link width
are proportional to the number of shared predators. (Modified from Marina et  al. (2018) and
Cordone et al. (2018))

The influence of the seaweed community in Potter Cove ecosystem is notorious


not only in the structure of the food web but also in the functioning. For instance,
the maximum trophic level for Potter Cove food web is 4.27 (Marina et al. 2018),
lower than most other food webs studied (Dunne et al. 2002, 2004). This implies
that top and basal species (dominated by macroalgae) are separated by few interme-
diate taxa. Therefore, the transfers of energy or nutrients from the base to the top of
Potter Cove food web is small, so that the number of times chemical energy is trans-
formed from a consumer’s diet into a consumer’s biomass along the food web is
also small. The mean trophic level for the mentioned food web is also low (2.10),
which is a consequence of the fact that most predators at intermediate levels (e.g.,
amphipods, isopods, bivalves, Notothenia coriiceps) feed predominantly on algae
species and/or detritus, being mainly the product of dead and decomposed macroal-
gae in Potter Cove (Iken et  al. 1998; Quartino et  al. 2008; see also Chap. 8 by
Quartino et al.). The macroalgal detritus decomposes and is eaten by detritivores
and suspensivores (e.g., sponges, ascidians, bryozoans, cnidarians), supporting an
important amount of the secondary production (Tatián et al. 2004). The obtained
low mean trophic level for Potter Cove food web clearly shows what species-­specific
and/or community studies have suggested. These characteristics of ecological com-
munities have a high impact on ecosystem functioning, such as nutrient and carbon
cycling, and trophic cascades (Post 2002).
However, our interest here is focused on macroalgae. To explore mainly their
interactions, we can use another useful information due to the secondary graph of
common-enemy graph (Fig. 15.2). This graph represents prey species linked with
each other according the existence of shared predators. In the representation of
15  Seaweeds in the Antarctic Marine Coastal Food Web 299

Fig. 15.2, the thickness of each link joining a couple of nodes is proportional to the
number of enemies shared by these nodes. The common-enemy graph is hypercon-
nected having a high-density value of interactions per node (Cordone et al. 2018);
this is an indicator of the existence of multiple alternative energy paths. This hyper-
connectivity makes the prediction of indirect relationships between species very
difficult, as each interaction involves positive and negative effects between species
abundances (Holt and Lawton 1994).

15.3  Network Dynamics and Robustness

An immediate question that we can try to answer is the following: can we predict
the behavior of the community against the (local) extinction of some algae? This
important question cannot be easily answered by field observations or experiments;
however, we can carry out in silico experiments using our food web (Cordone et al.
2018). We made a virtual experiment trying to determinate if successive extinctions
of base species in Potter Cove could produce phase transitions in the emergent prop-
erties of the food web. With this objective, we simulated extinctions by deleting
nodes (seaweeds, detritus, microalgae) according to different sequences established
by the degree of each macroalgae (total number of trophic interactions per node)
and their biomass in Potter Cove. Four sequences of extinctions were used: random
order, degree in ascending and descending order, and biomass in ascending order.
The method quantifies secondary extinctions taking into account the existence of
different extinction thresholds (Schleuning et al. 2016). That means that a species
suffers a secondary extinction when it loses a given percentage of its preys (Solé and
Montoya 2001). We define the threshold (v) as the minimum level of energy neces-
sary for species’ survival. After each node removal, the fraction of original incom-
ing energy e(i) is calculated for each species i, and when this fraction is equal or less
than the threshold, the species is secondarily lost. The classical topological approach
assumes that v is equal to 0, so only when the energy inflow is null a species goes
extinct (Bellingeri and Bodini 2013).
The results of the in silico experiment considering different thresholds for
extinction and the four sequences of primary elimination are indicated in Fig. 15.3.
We found that the Potter Cove food web is relatively stable to macroalgae loss, but
a significant number of secondary extinctions were obtained beyond a 50% thresh-
old (Cordone et al. 2018). The elimination of macroalgae species from the Potter
Cove food web does not seems to generate a catastrophic cascade of secondary
extinctions, suggesting that the Potter Cove food web could be more robust than
other similar networks (Allesina et al. 2006). Most connected macroalgae shared
more predators, which could indicates that these macroalgae species are function-
ally redundant (i.e., species with equivalent trophic interactions). Functional
redundancy has important consequences in potential cascade extinctions, since it
increases food web resistance by means of availability of alternative preys (Borrvall
et al. 2000; Petchey et al. 2008). However, these results are tied to the limitations
300 F. R. Momo et al.

Fig. 15.3  Secondary extinctions vs extinctions (total, primary plus secondary) when macroalgae
species and detritus are removed. The secondary extinctions after primary extinctions take place or
not depending on the numbers of preys that the predator needs to survive. This number is deter-
mined by the threshold of extinctions as the number of original preys multiplied by threshold. Each
box corresponds to a simulation in which the threshold of extinctions was constant and different
(0%, 25%, 50%, and 75%). The macroalgae loss was made following the four different sequences:
red, ascending degree; blue, descending degree; violet, random degree – mean and interval confi-
dence; and green, biomass – in ascending order

of our method. Typically, the topological approach underestimates the actual num-
ber of secondary extinctions and in consequence overestimates food web robust-
ness. The introduction of the threshold effect gives us a more accurate prediction
(Fig. 15.4).
Despite this, the topological approach enables the analysis of complex food
webs, since it only requires knowledge of network structure and could be used as
a proxy of food web stability (Eklöf et  al. 2013). Furthermore, we have to
15  Seaweeds in the Antarctic Marine Coastal Food Web 301

Fig. 15.4  This figure summarizes the effect of varying thresholds in the number of secondary
extinctions recorded. We tested thresholds from 5% to 95% by 5% when macroalgae are primary
lost (blue, 3; brown, 5; green, 10; red, 15; pink, 20). Points are means of secondary extinctions
when x number of primary extinctions took place at a particular threshold (by random simula-
tions). Shaded area represents confidence interval of the series. Secondary extinctions recorded at
a fixed threshold (0.25, 0.50, 0.75) from Fig. 15.3. (Modified from Cordone et al. 2018)

consider that this food web includes all recorded trophic interactions, though
these interactions do not always occur simultaneously in time and space. We can-
not ignore the potentially confounding effects of seasonality and spatial sampling
(Ings et al. 2009). In fact, the network is an average representation across seasons
and different habitats, and dynamic stability depends more on how interactions
are materialized in time and space. Some nodes are pulsatile, such as the massive
influx of Krill (Fuentes et al. 2016), which is difficult to reflect in a static descrip-
tion of the network topology. The consideration of all these factors might change
our estimation of the fragility of Potter Cove food web. Our current efforts are
focused on incorporating spatial and temporal variability into the modeling of
Potter Cove food web to achieve a more realistic picture of this ecosystem and its
macroalgae species.
302 F. R. Momo et al.

15.4  Non-Trophic Interactions

Ecological systems are complex and multifaceted in the nature of their interactions
and should be defined not only by lists of co-occurring species but also by the vari-
ety of interactions that take place between them. In order to satisfy different require-
ments, species interact in many ways with multiple partners and make associations
that imply more than a trophic relationship. Great advances and multiple practical
applications have been developed since the study of the structure, intensity, and
dynamics of trophic interactions. However, ecological interactions between co-­
existing species involve much more than simply feeding (Bascompte et al. 2003;
Kéfi et al. 2012; Pocock et al. 2012; Kéfi et al. 2015, 2016a).
Decades of empirical and theoretical studies have shown that specific non-­trophic
interactions, such as habitat modification, stress minimization, ecosystem engineer-
ing, and behavioral changes, can play important roles for community structure and
ecosystem functioning (Pocock et al. 2012; Kéfi et al. 2015, 2016a). Some works
have recognized this importance and have focused their study on isolated networks
of non-trophic interactions such as mutualistic networks (Jordano et al. 2009) and,
more recently, have incorporated different types of interaction (e.g., mutualistic,
competition, and trophic) in a single network usually named “multiplex” (Kéfi et al.
2012, 2015, 2016b; Pocock et al. 2012; García-Callejas et al. 2018).
To know how important non-trophic interactions are for the ecosystem, function-
ing triggers many interesting questions around these relationship types: How many
direct non-trophic interactions are there and how are they distributed among spe-
cies? Do the topological properties of trophic and non-trophic networks differ? Can
simple species attributes help predict the type of interaction between two species?
Can trophic characteristics help predict the non-trophic webs (Kéfi et  al. 2015)?
Here, we have focused on the first question to describe non-trophic interactions that
involve macroalgae species from the Potter Cove food web.
The number of non-trophic interactions identified in Potter Cove marine ecosys-
tem duplicates the number of trophic relationships: 1091 versus 454, respectively.
Here, non-trophic interactions include competition or negative interactions (−/−)
considered when a pair of species share at least one prey and represent the 74% of
the total number of non-trophic interactions; mutualism or positive interactions that
involve a benefit of both interacting species (+/+) that represent the 13.6% of the
total number of non-trophic interactions; and commensalism (+/0) and amensalism
(−/0) or neutral interactions in which one species benefit or not-benefit, respec-
tively, and the other species is not affected by the link, and these interactions repre-
sent the 12.4% of the total number of non-trophic interactions. The distribution of
these interactions (Fig. 15.5) depends on the trophic classification of the species in
basal, intermediate, or top species, e.g., competition, as we defined here, occurs
between intermediate and/or top species, which means that most non-trophic inter-
actions in Potter Cove ecosystem occur between species of high trophic levels. So,
we should wonder what the role of macroalgae in the non-trophic relationships is.
15  Seaweeds in the Antarctic Marine Coastal Food Web 303

Fig. 15.5  Number of non-trophic interactions for basal, intermediate, and top species. Each color
represents a type of interaction

Many works highlight the importance of studying the structure of non-trophic


networks versus the trophic network by suggesting that, in general, trophic net-
works tend to be compartmentalized, while positive non-trophic networks are more
nested (Bascompte et  al. 2003; Thébault and Fontaine 2010), which minimizes
competition, increases the number of co-existing species, and makes the community
more robust against random extinctions and diversity loss (Memmott et al. 2004;
Fortuna and Bascompte 2006).
In food webs, it is argued that modular patterns tend to increase the stability of
the network by retaining the impacts of disturbances within each module and mini-
mizing that impact on other modules in the network. In the particular case of Potter
Cove, it was observed that the food web has not modular characteristics, reason to
suppose that the network should be not very robust against random disturbances.
However, it has been shown that the trophic network is highly robust against differ-
ent types of disturbances, which makes it possible to assume that positive non-­
trophic interactions could have a stabilizing role in the Potter Cove ecosystem.
Then, we should not fail to recognize that macroalgae are a key component of the
system because they represent, directly or indirectly, a stabilizing species.
304 F. R. Momo et al.

15.5  Final Remarks

Network science offers a great opportunity to understand species in their context


and their relation to the community and the ecosystem. In our studies, the incorpora-
tion of a network perspective shows the relevance of macroalgae species to the
entire ecosystem of Potter Cove. We observed that Potter Cove food web is rela-
tively robust to local extinctions of macroalgal species under a topological approach
(considering extinction thresholds), but we also observed that a network collapse
could be reached by increasing the extinction threshold. This result demonstrates
that effects on single species could propagate to the entire network. Potter Cove
ecosystem is experiencing rapid climate-related changes in environmental factors
that have affected the benthic system, including the macroalgal community (Schloss
et al. 2012; Quartino et al. 2013; Sahade et al. 2015; Deregibus et al. 2016; Campana
et  al. 2018). In this sense, species-specific responses of macroalgae to changing
environmental factors should be explored as negative effects at the species level
could propagate through the food web, leading to changes in the structure and func-
tion of this system.

Acknowledgments  We wish to thanks the scientific and logistic assistance of Instituto Antártico
Argentino (IAA)/Dirección Nacional del Antártico (DNA). The academic interaction with the
Alfred Wegener Institute Helmholtz Centre for Polar and Marine Research (AWI) was essential for
this work. We acknowledge the financial support by grants from CONICET-UNGS (PIO
14420140100035CO). This chapter also presents an outcome of the international Research
Network IMCONet funded by the Marie Curie Action IRSES IMCONet (FP7 IRSES, Action No.
318718). Many thanks to C. Gardel and I. Corsini for their assistance.

References

Ahn I-Y, Moon H-W, Jeon M, Kang S-H (2016) First record of massive blooming of benthic
diatoms and their association with megabenthic filter feeders on the shallow seafloor of an
Antarctic fjord: does glacier melting fuel the bloom? Ocean Sci J 51:273–279. https://doi.
org/10.1007/s12601-016-0023-y
Albert R, Barabási A (2002) Statistical mechanics of complex networks. Rev Mod Phys 74:47–97.
https://doi.org/10.1103/RevModPhys.74.47
Allesina S, Bodini A, Bondavalli C (2006) Secondary extinctions in ecological networks: bottle-
necks unveiled. Ecol Model 194:150–161. https://doi.org/10.1016/j.ecolmodel.2005.10.016
Amsler CD, McClintock JB, Baker BJ (1999) An Antarctic feeding triangle: defensive interactions
between macroalgae, sea urchins, and sea anemones. Mar Ecol Prog Ser 183:105–114. https://
doi.org/10.3354/meps183105
Amsler CD, Iken K, McClintock JB, Amsler MO, Peters KJ, Hubbard JM, Furrow FB, Baker BJ
(2005) Comprehensive evaluation of the palatability and chemical defenses of subtidal mac-
roalgae from the Antarctic peninsula. Mar Ecol Prog Ser 294:141–159. https://doi.org/10.3354/
meps294141
Amsler CD, Iken K, McClintock JB, Baker BJ (2009) Defenses of polar macroalgae against herbi-
vores and biofoulers. Bot Mar 52:535–545. https://doi.org/10.1515/BOT.2009.070
15  Seaweeds in the Antarctic Marine Coastal Food Web 305

Amsler CD, McClintock JB, Baker BJ (2012) Palatability of living and dead detached Antarctic mac-
roalgae to consumers. Antarct Sci 24:589–590. https://doi.org/10.1017/S0954102012000624
Amsler CD, McClintock JB, Baker BJ (2014) Chemical mediation of mutualistic interactions
between macroalgae and mesograzers structure unique coastal communities along the Western
Antarctic Peninsula. J Phycol 50:1–10. https://doi.org/10.1111/jpy.12137
Bascompte J, Jordano P, Melián CJ, Olesen JM (2003) The nested assembly of plant–animal mutu-
alistic networks. PNAS 100:9383–9387. https://doi.org/10.1073/pnas.1633576100
Bellingeri M, Bodini A (2013) Threshold extinction in food webs. Theor Ecol 6:143–152. https://
doi.org/10.1007/s12080-012-0166-0
Borrvall C, Ebenman B, Jonsson T (2000) Biodiversity lessens the risk of cascading extinction
in model food webs. Ecol Lett 3:131–136. https://doi.org/10.1046/j.1461-0248.2000.00130.x
Campana GL, Zacher K, Fricke A, Molis M, Wulff A, Quartino ML (2009) Drivers of colonization
and succession in polar benthic macro- and microalgal communities. Bot Mar 52:655–667.
https://doi.org/10.1515/BOT.2009.076
Campana GL, Zacher K, Deregibus D, Momo FR, Wiencke C, Quartino ML (2018) Succession
of Antarctic benthic algae (Potter Cove, South Shetland Islands): structural patterns and
glacial impact over a four-year period. Polar Biol 41:377–396. https://doi.org/10.1007/
s00300-017-2197-x
Cordone G, Marina TI, Salinas V, Doyle SR, Saravia LA, Momo FR (2018) Effects of macroalgae
loss in an Antarctic marine food web: applying extinction thresholds to food web studies. Peer
J 6:e5531. https://doi.org/10.7717/peerj.5531
Delmas E, Besson M, Brice MH, Burkle LA, Dalla Riva GV, Fortin MJ, Gravel D, Guimarães PR
Jr, Hembry DH, Newman EA, Olesen JM, Pires MM, Yeakel JD, Poisot T (2018) Analysing
ecological networks of species interactions. Biol Rev 94:16–36. https://doi.org/10.1111/
brv.12433
Deregibus D, Quartino ML, Campana GL, Momo FR, Wiencke C, Zacher K (2016) Photosynthetic
light requirements and vertical distribution of macroalgae in newly ice-free areas in Potter
Cove, South Shetland Islands, Antarctica. Polar Biol 39:153–166. https://doi.org/10.1007/
s00300-015-1679-y
Dunne JA, Williams RJ, Martinez ND (2002) Food-web structure and network theory: the role
of connectance and size. PNAS 9920:12917–12922. https://doi.org/10.1073/pnas.192407699
Dunne JA, Williams RJ, Martinez ND (2004) Network structure and robustness of marine food
webs. Mar Ecol Prog Ser 273:291–302. https://doi.org/10.3354/meps273291
Eklöf A, Tang S, Allesina S (2013) Secondary extinctions in food webs: a Bayesian network
approach. Methods Ecol Evol 4:760–770. https://doi.org/10.1111/2041-210X.12062
Fogg GE (1977) Aquatic primary production in the Antarctic. Philos Trans R Soc Lond B Biol Sci
279:27–38. https://doi.org/10.1098/rstb.1977.0069
Fortuna MA, Bascompte J (2006) Habitat loss and the structure of plant–animal mutualistic net-
works. Ecol Lett 9:281–286. https://doi.org/10.1111/j.1461-0248.2005.00868.x
Fuentes V, Alurralde G, Meyer B, Aguirre GE, Canepa A, Wölfl A-C, Hass HC, Williams GN,
Schloss IR (2016) Glacial melting: an overlooked threat to Antarctic krill. Sci Rep 6:27234.
https://doi.org/10.1038/srep27234
García-Callejas D, Molowny-Horas R, Araújo MB (2018) Multiple interactions networks: towards
more realistic descriptions of the web of life. Oikos 127:5–22. https://doi.org/10.1111/
oik.04428
Gómez I, Wulff A, Roleda MY, Huovinen P, Karsten U, Quartino ML, Dunton K, Wiencke C
(2009) Light and temperature demands of marine benthic microalgae and seaweeds in polar
regions. Bot Mar 52:593–608. https://doi.org/10.1515/BOT.2009.073
Holt RD, Lawton JH (1994) The ecological consequences of shared natural enemies. Annu Rev
Ecol Evol Syst 25:495–520. https://doi.org/10.1146/annurev.es.25.110194.002431
Huang Y, Amsler MO, McClintock JB, Amsler CD, Baker BJ (2007) Patterns of gammaridean
amphipod abundance and species composition associated with dominant subtidal m ­ acroalgae
from the western Antarctic peninsula. Polar Biol 30:1417–1430. https://doi.org/10.1007/
s00300-007-0303-1
306 F. R. Momo et al.

Iken K, Barrera-Oro ER, Quartino ML, Casaux RJ, Brey T (1997) Grazing by the Antarctic fish
Notothenia coriiceps: evidence for selective feeding on macroalgae. Antarct Sci 9:386–391.
https://doi.org/10.1017/S0954102097000497
Iken K, Quartino ML, Barrera-Oro E, Palermo J, Wiencke C, Brey T (1998) Trophic relations
between macroalgae and herbivores. In: Wiencke C, Ferreyra G, Arntz W, Rinaldi C (eds) The
Potter Cove coastal ecosystem, Antarctica. Rep Polar Res, vol 299, pp 258–262
Ings TC, Montoya JM, Bascompte J, Blüthgen N, Brown L, Dormann CF, Edwards F, Figueroa
D, Jacob U, Jones JI, Lauridsen RB, Ledger ME, Lewis HM, Olesen JM, van Veen FJ, Warren
PH, Woodward G (2009) Ecological networks – beyond food webs. J Anim Ecol 78:253–269.
https://doi.org/10.1111/j.1365-2656.2008.01460.x
Jordano P, Vázquez D, Bascompte J (2009) Redes complejas de interacciones planta-animal. In:
Medel R, Aizen MA, Zamora R (eds) Ecología y evolución de interacciones planta-animal.
Editorial Universitaria, Santiago de Chile, pp 17–41
Kéfi S, Berlow EL, Wieters EA, Navarrete SA, Petchey OL, Wood SA, Boit A, Joppa LN, Lafferty
KD, Williams RJ, Martinez ND, Menge BA, Blanchette CA, Iles AC, Brose U (2012) More
than a meal…integrating non-feeding interactions into food webs. Ecol Lett 15:291–300.
https://doi.org/10.1111/j.1461-0248.2011.01732.x
Kéfi S, Berlow EL, Wieters EA, Joppa LN, Wood SA, Brose U, Navarrete SA (2015) Network
structure beyond food webs: mapping non-trophic and trophic interactions on Chilean rocky
shores. Ecology 96:291–303. https://doi.org/10.1890/13-1424.1
Kéfi S, Holmgren M, Scheffer M (2016a) When can positive interactions cause alternative stable
states in ecosystems? Funct Ecol 30:88–97. https://doi.org/10.1111/1365-2435.12601
Kéfi S, Miele V, Wieters EA, Navarrete SA, Berlow EL (2016b) How structured is the entan-
gled bank? The surprisingly simple organization of multiplex ecological networks leads to
increased persistence and resilience. PLoS Biol 14(8):e1002527. https://doi.org/10.1371/jour-
nal.pbio.1002527
Klöser H, Mercuri G, Laturnus F, Quartino ML, Wiencke C (1994) On the competitive balance
of macroalgae at Potter Cover (King George Island, South Shetlands). Polar Biol 14:11–16.
https://doi.org/10.1007/BF00240266
Klöser H, Quartino ML, Wiencke C (1996) Distribution of macroalgae and macroalgal com-
munities in gradients of physical conditions in Potter Cove, King George Island, Antarctica.
Hydrobiologia 333:1–17. https://doi.org/10.1007/BF00020959
Majewska R, Convey P, De Stefano M (2016) Summer epiphytic diatoms from Terra Nova Bay and
Cape Evans (Ross Sea, Antarctica) – a synthesis and final conclusions. PLoS One 11(4):1–30.
https://doi.org/10.1371/journal.pone.0153254
Marina TI, Salinas V, Cordone G, Campana GL, Moreira E, Deregibus D, Torre L, Sahade R,
Tatián M, Barrera Oro E, De Troch M, Doyle S, Quartino ML, Saravia LA, Momo FR (2018)
The food web of Potter Cove (Antarctica): complexity, structure and function. Estuar Coast
Shelf Sci 200:141–151. https://doi.org/10.1016/j.ecss.2017.10.015
Memmott J, Waser NM, Price MV (2004) Tolerance of pollination networks to species extinctions.
Proc R Soc Lond B 271:2605–2611. https://doi.org/10.1098/rspb.2004.2909
Pellizzari F, Silva MC, Silva EM, Medeiros A, Oliveira MC, Yokoya NS, Pupo D, Rosa LH,
Colepicolo P (2017) Diversity and spatial distribution of seaweeds in the South Shetland
Islands, Antarctica: an updated database for environmental monitoring under climate change
scenarios. Polar Biol 40:1671–1685. https://doi.org/10.1007/s00300-017-2092-5
Petchey OL, Eklöf A, Borrvall C, Ebenman B (2008) Trophically unique species are vulnerable to
cascading extinction. Am Nat 171:568–579. https://doi.org/10.1086/587068
Pocock MJ, Evans DM, Memmott J (2012) The robustness and restoration of a network of ecologi-
cal networks. Science 335:973–977. https://doi.org/10.1126/science.1214915
Poisot T, Stouffer DB, Gravel D (2015) Beyond species: why ecological interactions vary through
space and time. Oikos 124:243–251. https://doi.org/10.1111/oik.01719
Post DM (2002) Using stable isotopes to estimate trophic position: models, methods, and assump-
tions. Ecology 83(3):703–718
15  Seaweeds in the Antarctic Marine Coastal Food Web 307

Quartino ML, Boraso de Zaixso A, Momo FR (2008) Macroalgal production and the energy cycle
of Potter Cove. Rep Polar Mar Res 571:68–74
Quartino ML, Deregibus D, Campana GL, Latorre GEJ, Momo FR (2013) Evidence of macroalgal
colonization on newly ice-free areas following glacial retreat in potter cove (South Shetland
Islands), Antarctica. PLoS One 8(3):e58223. https://doi.org/10.1371/journal.pone.0058223
Sahade R, Lagger C, Torre L, Momo FR, Monien P, Schloss IR, Barnes DKA, Servetto N, Tarantelli
S, Tatián M, Zamboni N, Abele D (2015) Climate change and glacier retreat drive shifts in an
Antarctic benthic ecosystem. Sci Adv 1(10):e1500050. https://doi.org/10.1126/sciadv.1500050
Schleuning M, Fründ J, Schweiger O, Welk E, Albrecht J, Albrecht M, Beil M, Blüthgen N,
Bruelheide H, Böhning-Gaese K, Dehling DM, Dormann CF, Exeler N, Farwig N et al (2016)
Ecological networks are more sensitive to plant than to animal extinction under climate change.
Nat Commun 7:13965. https://doi.org/10.1038/ncomms13965
Schloss IR, Abele D, Moreau S, Demers S, Bers AV, González O, Ferreyra G (2012) Response of
phytoplankton dynamics to 19-year (1991–2009) climate trends in Potter Cove (Antarctica). J
Mar Syst 92:53–66. https://doi.org/10.1016/j.jmarsys.2011.10.006
Smale DA, Barnes DKA (2008) Likely responses of the Antarctic benthos to climate-­
related changes in physical disturbance during the 21st century, based primarily on evi-
dence from the West Antarctic peninsula region. Ecography 31:289–305. https://doi.
org/10.1111/j.0906-7590.2008.05456.x
Solé RV, Montoya M (2001) Complexity and fragility in ecological networks. Proc R Soc Lond B
268:2039–2045. https://doi.org/10.1098/rspb.2001.1767
Tatián M, Sahade R, Esnal GB (2004) Diet components in the food of Antarctic ascidians liv-
ing at low levels of primary production. Antarct Sci 16:123–128. https://doi.org/10.1017/
S0954102004001890
Thébault E, Fontaine C (2010) Stability of ecological communities and the architecture of mutu-
alistic and trophic networks. Science 329:853–856. https://doi.org/10.1126/science.1188321
Wiencke C, Amsler CD, Clayton MN (2014) Macroalgae. In: De Broyer C, Koubbi P, Griffiths HJ,
Raymond B, Udekem d’Acoz C et al (eds) Biogeographic atlas of the Southern Ocean. SCAR,
Cambridge, pp 66–73
Williams RJ, Martinez ND (2004) Limits to trophic levels and omnivory in complex food webs:
theory and data. Am Nat 163:458–468. https://doi.org/10.1086/381964
Woodward G, Benstead JP, Beveridge OS, Blanchard J, Brey T, Brown LE, Cross WF, Friberg
N, Ings TC, Jacob U, Jennings S et  al (2010) Ecological networks in a changing climate.
In: Woodward G (ed) Advances in ecological research, vol 42. Elsevier Ltd, Amsterdam,
pp 71–138. https://doi.org/10.1016/B978-0-12-381363-3.00002-2
Chapter 16
Trophic Networks and Ecosystem
Functioning

Marco Ortiz, Brenda B. Hermosillo-Núñez, and Ferenc Jordán

Abstract  The geographic isolation of the Antarctic continent offers an interesting


opportunity to quantify and qualify the actual ecological conditions and the most
sensitive components from an ecosystem perspective. Antarctic coastal ecosystems
are under severe stress as a consequence of climate change, which could facilitate
biological invasions, reduced growth of macroalgal species, and local extinctions.
The application of network analysis, representing the interactions among multiple
species, allows us to quantify macroscopic (emergent) system properties, to assess
overall health, to predict the propagation of direct and indirect effects, and to iden-
tify keystone species complexes within these complex ecological systems. Three
theoretical frameworks are used here for this analysis: (1) ecological network analy-
sis (ENA) considering thermodynamics and information theory (providing mea-
sures such as Ascendency), (2) semiquantitative (qualitative) mathematics based on
the structure and local stability of community matrices (Loop Analysis), and (3)
topological studies on interaction networks considering central node sets and defin-
ing keystone species complexes (KSCs).
Therefore, the integration of ecosystemic properties and keystone species com-
plexes could help us to facilitate the design and assessment of conservation and
monitoring measures, especially when the Antarctic coastal marine ecosystems are
being severely stressed. The protection of the Antarctic environment – as a whole –
not only should be focused on biological populations and communities but also

M. Ortiz (*)
Instituto de Ciencias Naturales Alexander von Humboldt, Facultad de Ciencias del Mar y
Recursos Biológicos, Universidad de Antofagasta, Antofagasta, Chile
Instituto Antofagasta, Universidad de Antofagasta, Antofagasta, Chile
e-mail: marco.ortiz@uantof.cl
B. B. Hermosillo-Núñez
Instituto Antofagasta, Universidad de Antofagasta, Antofagasta, Chile
e-mail: brenda.hermosillo@uantof.cl
F. Jordán
Balaton Limnological Institute, MTA Centre for Ecological Research, Tihany, Hungary
Stazione Zoologica, Napoli, Italy
e-mail: jordan.ferenc@okologia.mta.hu

© Springer Nature Switzerland AG 2020 309


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_16
310 M. Ortiz et al.

should consider changes in macroscopic properties, the propagation of direct and


indirect influences in the networks, and keystone species complexes, which emerge
using networks of interacting and coexisting species (or functional groups as system
components).

Keywords  Ascendency framework · Coastal benthic-pelagic ecosystem ·


Keystone species · Macroscopic network properties · Multispecies modelling

16.1  Introduction

Different scientific strategies – not necessarily mutually exclusive- can be used to


study, assess, and attempt to predict the transformations in natural systems caused
by direct and indirect human activities. Within of these strategies, we choose the
quantitative and semiquantitative (qualitative) simulations permitting us to deter-
mine whole-system properties (of structure and dynamics) of Antarctic coastal
benthic-­pelagic ecosystems (Pikitch et al. 2004). Antarctic coastal ecosystems face
several stressors as the increase in UV radiation, which reduces the growth of mac-
roalgae species (Richter et al. 2008), and the rise in temperature, which could mod-
ify the amount of the direct light to macroalgae and other organisms (King 1994;
Stark 1994). The impact of these stressors on the species and functional groups
could affect the structure of their trophic networks and their functioning of diverse
Antarctic ecosystems. Network analyses permit us to evaluate the macroscopic
properties at ecosystem scale, describing aspects related with growth, development,
complexity, and health (Odum 1969; Ulanowicz 1986; Levins 1998a, b; Costanza
and Mageau 1999). Likewise, another related analysis can be performed as the
propagation of direct and indirect effects (Hawkins 2004; Levins 1998a) and the
determination of species or functional groups that play key roles in these complex
ecological systems (Ortiz et al. 2013a).
The ecological concept of keystone species, introduced by Paine (1969) and
broadened by Power et al. (1996), has become a key issue in numerous research
programs in different communities and ecosystems around the planet (Mills et al.
1993; Power et al. 1996), especially given its use in the design and application of
conservation management and monitoring measures (Payton et  al. 2002; Barua
2011; Ortiz et al. 2013a, b). Power et al. (1996) named a key species as a species
whose effect is large, and disproportionately large relative to its abundance.
Although the concept seems to be sufficiently clear, its determination in communi-
ties and ecosystems is complex, requiring observations and experiments that include
different spatiotemporal scales, levels of organization, and taxonomic groups
(Power et al. 1996; Libralato et al. 2006). Similarly, some purely experimental stud-
ies (Pace et al. 1999) have omitted the propagation of direct and indirect effects,
16  Trophic Networks and Ecosystem Functioning 311

despite the ecological importance of these processes it has been recognized (Wootton
1994; Patten 1997; Yodzis 2001).
Various studies have determined the role that different species play in their eco-
logical systems by using different network indices (Jordán et al. 2007; Luczkovich
et al. 2003; Jordán and Scheuring 2004; Allesina and Bodini 2004; Libralato et al.
2006; Benedek et al. 2007; Ortiz et al. 2013a, 2017; Valls et al. 2015; Giacaman-­
Smith et al. 2016). Such analyses offer a complementary way to address some of the
limitations in the experimental identification of key groups. Quantitative trophic
analysis permits estimations of the strength of interactions between species or func-
tional groups by identifying the presence of keystone species, which occupy key
positions in the networks (Jordán and Scheuring 2004). At the same time, keystone-
ness can also be determined using semiquantitative or qualitative loop network
analysis. In this case, the key role of a species is defined as a dynamical conse-
quence of its changes, modifying the balance (prevalence) of positive and negative
feedbacks and, in turn, the local stability of the network (Ortiz et  al. 2013a).
Likewise, following field observations (Daily et al. 1993), two independent contri-
butions have proposed methodological extensions toward multispecies approaches
to keystones. One was given by Benedek et al. (2007), which is based on the central-
ity of node sets, and the other proposed by Ortiz et al. (2013a) based on quantitative
and semiquantitative multispecies trophic models. In both cases, the keystone spe-
cies complexes (KSCs) consist of a core of species and/or functional groups linked
by strong interspecific interactions. These more holistic concepts could facilitate the
design of conservation and monitoring programs in ecosystems since it is not guar-
anteed that always a single species plays the key role.
The aim of this chapter was to determine the macroscopic system properties
(e.g., level of development, organization, maturity, and health) derived from
Ascendency theoretical framework (sensu Ulanowicz 1986, 1997) of a model
benthic-­pelagic ecosystem at Fildes Bay, King George Island Peninsula (Fig. 16.1a).
The subtidal communities are dominated by the brown macroalgae species
Himmantothallus grandifolius and Desmarestia anceps; the red algae species
Gigartina skottsbergii, Trematocarpus antarcticus, and Plocamium cartilagineum;
the grazer species (gastropod) Nacella concinna; the asteroid predator species
Diplasterias brucei; and a variety of fish, sponge, bryozoan, and ascidian species
(Targett 1981; Huovinen and Gómez 2013; Valdivia et al. 2014, 2015). Considering
this network, dynamical simulations, local stability, and centrality of node sets
(KeyPlayer set) were performed to determine the keystone species complexes
(KSCs). Finally, the contribution of the components of KSCs on macroscopic sys-
tem properties will be also evaluated. For this purpose, quantitative and semiquan-
titative (qualitative) multispecies trophic network models will be used and analyzed.
The network model constructed represents exclusively transient states under short-­
term dynamics.
312 M. Ortiz et al.

Fig. 16.1  Study area and sampling sites at Fildes Bay (King Georg Island, Antarctica) (GC
Glacier Collins, IA Island Artigas, IS Island Shoa, HT Half Three, EN Estrecho Nelson, GN Glacier
Nelson) (a) and the trophic model for the coastal ecosystem of Fildes Bay (Note: vertical position
approximates trophic level, and the circle is proportional to the compartment biomass) (b). For
details, see Appendix 16.A. (Adapted from Fig. 2 in Ortiz et al. 2017)
16  Trophic Networks and Ecosystem Functioning 313

16.1.1  Q
 uantitative and Semiquantitative Multispecies Trophic
Modelling

Trophic mass balance models were constructed using the Ecopath with Ecosim
(v.5.0) (EwE) software, which was first developed by Polovina (1984) and subse-
quently extended by Christensen and Pauly (1992) and Walters et  al. (1997).
Ecopath allows the depiction of the flows of matter and/or energy in a stationary
state of an ecosystem within a given time, whereas Ecosim performs dynamic simu-
lations of the initial conditions (established with Ecopath) as a response to perturba-
tions (Christensen and Pauly 1992). The energy mass balance of a species or
functional group within a network is represented by linear differential equations for
each component in the model (Box 16.1). To employ Ecosim (see Walters and
Christensen 2007), an extension routine of Ecopath is included to define the con-
sumption by compartment which is represented by the control flow equation, allow-
ing to set if the flow control mechanism is top-down, bottom-up, or mixed (Box
16.1). We set the mixed control mechanism following the criteria of Ortiz (2008a,
b) and Ortiz et al. (2009, 2010, 2013a, b). We used Ecopath with Ecosim version
6.4, and details concerning this software package are given in Pauly et al. (2000).
The semiquantitative or qualitative trophic models were built using Loop
Analysis, which allows for the estimation of the local stability (as a measure of sus-
tainability) of an ecological system and the assessment of the propagation of both
direct and indirect effects as a response to external perturbations (Levins 1974,
1998a). This approach has been widely applied in different natural science fields
(Puccia and Levins 1991; Levins 1998b; Darmbacher et al. 2009; Ortiz and Levins
2011, 2017). The interactions are shown as signs that indicate the type of influence
each variable has upon another (positive, negative, or null). In ecological relation-
ships, the +1/−1 signs denote predator/prey or parasite/host interactions, +1/+1
signs express mutualism, +1/0 signs represent commensalism, and −1/0 signs show
amensalism. Loop Analysis is based on the relationships of differential equations
near equilibrium, Jacob-Levins’ community matrices, and their loop diagrams (Box
16.2). Local stability of the system is quantified using the Routh-Hurwitz criteria,
which require the following: (1) all feedbacks (on every level of complexity) must
be negative and (2) negative feedbacks on higher levels cannot be too great for com-
parison with the negative feedbacks on lower levels. Levins (1998a) proposes that
the system is more resistant (locally stable), whereas Fn (feedback corresponding to
higher level of complexity) is more negative (Box 16.2). The semiquantitative or
qualitative Jacob-Levins’ community matrices are based on the diet matrices used
for quantitative models (Ecopath with Ecosim). The polynomial equation required
for each community matrix was determined using the software LoopStability (Diaz-­
Ávalos and Ortiz, laboratory uses).
314 M. Ortiz et al.

Box 16.1 The Ecopath with Ecosim (EwE) Theoretical Framework

The mass balance of a species or functional group within a trophic network is


represented by the following mathematical expression:
P n
Q
Bi   EE i  B j   DC ji  Yi  BA i  Ei  0 (16.1)
 B  j 1  B j
i

where Bi and Bj are the biomass value of prey i and predator j; P/Bi is the pro-
ductivity of prey i, which is equivalent to total mortality (Allen 1971); EEi is
the ecotrophic efficiency or the fraction of the total production of a group or
species used in the system; Yi is the fishing production per unit of area and
time (Y = fishing mortality × biomass); Q/Bi is the food consumption per bio-
mass unit j; DCji is the fraction of prey i in the average diet of predator j; BAi
is the biomass accumulation rate for i; and Ei is the net migration of i (emigra-
tion minus immigration) (Christensen and Walters 2004).
Based on this equation, the input and output of matter (energy) in each
compartment of the system can be balanced. This energy balance is assured
for each group with this equation:
Q  P  R  UAF (16.2)
where Q is consumption, P is production, R is respiration, and UAF is unas-
similated food for each group or species of the system. Given the inclusion of
the factors BAi and Ei in Eq. (16.1), the focus of Ecopath is based on a steady
state. This situation allows changes in the network compartments when the
mathematical expression is expressed in a dynamic form.
To employ Ecosim (see Walters and Christensen 2007), an extension rou-
tine of Ecopath is included to define the consumption by compartment (Qij),
where Qij is represented by the following equation:
aij vij Bi B j
Qij  (16.3)
2 vij  aij B j

where aij represents the instantaneous mortality rate on prey i caused by a
single unit of predator j biomass. Similarly, aij can be understood as the rate of
effective search by predator j for prey i. Each aij is estimated directly from the
corresponding Ecopath models by aij = Qi/(BiBj), where Qi is the total con-
sumption of i. The vij represents the transfer rate between compartments i and
j. This parameter determines if the flow control mechanism is top-down, bot-
tom-­up, or mixed.
16  Trophic Networks and Ecosystem Functioning 315

Box 16.2 Loop Analysis Theoretical Framework

Loop Analysis estimates the local stability (as a component of sustainability)


of an ecological system and the assessment of the propagation of both direct
and indirect effects as a response to external perturbations (Levins 1974,
1998a). The effects are shown as signs that indicate the type of influence each
variable has upon another (positive, negative, or null). Interactions are
described according to pairwise sign combinations: +/− signs denote preda-
tor/prey or parasite/host interactions, +/+ signs express mutualism, +/0 signs
represent commensalism, and −/0 signs show amensalism or allelopathy.
Loop Analysis is based on the relationships of differential equations near equi-
librium, community matrices, and their loop diagrams. In a benthic system,
the element αij of the matrix and loop diagrams performs the effect of the
variable j in the growing of variable i, and this dynamic performs in the fol-
lowing way:
dXi
 fi  X1 , X 2 ,, X n ; C1 ,C2 ,,Cn  (16.4)
dt
where the change on the time of variable Xi is a function fi of the other vari-
ables – Xn and parameters Cn – which are interconnected. The link of Xj to Xi
is similarly to αij in Levins (1968):
fi  X  X 
 ij  (16.5)
X j

where x∗ indicates that it has been evaluated in equilibrium. The sign αij rep-
resents the link of j to i where the function of sign X is 1 when X > 0, 0 when
X = 0, and −1 when X < 0. Local stability of the system is quantified using the
Routh-Hurwitz criteria, which require the following: (1) all feedbacks  (on
every level of complexity) must be negative and (2) negative feedbacks on
higher levels cannot be too great for comparison with the negative feedbacks
on lower levels. Levins (1998a) proposes that the system is more resistant
(locally stable), whereas Fn (feedback corresponding to higher level of com-
plexity) is more negative:

F0  n  F1 n 1  F2  n  2    Fn 1  Fn  0 (16.6)


16.1.2  S
 election of Model Components, Sampling Programs,
and Data Sources

Three intensive field studies were conducted during the austral summers of 2013,
2014, and 2015 to identify the biological components (species or functional groups)
of the system model and to estimate the average biomass (B), average density, and
316 M. Ortiz et al.

food sources of the selected components (Valdivia et al. 2014). Sampling was per-
formed to directly estimate the average biomass and density of the macrobenthic
species (between 5 and 30 m depth) at six stations within Fildes Bay (Fig. 16.1a).
The production (P) and turnover rates (productivity) (P/B) were estimated using the
following allometric Eq. 16.1:

 Biomass 0.73 
Production      Density
 Density  

where 0.73 is the average exponent regression of annual production on body size for
macrobenthic invertebrates (for more details, see Warwick and Clarke 1993).
Food consumption rates were obtained from the literature (Cornejo-Donoso and
Antezana 2008; Ortiz 2008a; Ortiz et al. 2015; Pinkerton et al. 2010). To determine
the diets of N. concinna, Harpagifer antarcticus, Notothenia coriiceps, N. rossii,
and the asteroid species, the stomach and guts were revised, and the gut contents
were classified to the lowest possible taxonomic level; the frequency of occurrence
of each food item was then calculated. Several studies examining the trophic ecol-
ogy of several benthic and pelagic species were also used to determine the range of
food consumed (Cornejo-Donoso and Antezana 2008; Pinkerton et al. 2010; Kaehler
et al. 2000; Gili et al. 2001; Corbisier et al. 2004; Jacob et al. 2005; Norkko et al.
2007; Mincks et al. 2008).
A trophic model with 17 components was constructed for Fildes Bay. The com-
ponents represent the most abundant species or functional groups composed of mul-
tiple species. Seven components represented the following individual species: the
brown macroalgae H. grandifolius and D. anceps, the red algae G. skottsbergii, the
herbivores N. concinna and Margarella sp., the echinoid Sterechinus neumayeri,
and the asteroid D. brucei. The other components were functional groups that
included several species. These groups included multiple species of green (e.g.,
Monostroma hariotii), red (e.g., T. antarcticus and P. cartilagineum), and brown
(e.g., Ascoseira mirabilis and Halopteris obata) algae, respectively. The filter feeder
(FF) group was composed primarily of clams, hydrozoan, bryozoan, and sponge
species. The small epifauna (SE) component included gastropod, nematode, and
nemertean species. The group of other sea star species (SS) included principally
O. validus and Odontaster sp. The benthic fishes (BF) group was composed primar-
ily of H. antarcticus, N. coriiceps, and N. rossii. The three final groups were the
phytoplankton (Phy), zooplankton (Zoo), and detritus (Det) groups (Fig. 16.1b).
All the compartments are trophically linked by detritus – primarily as microbial
film and organic matter – because several studies have emphasized the importance
of bacteria as food for various species of molluscs (e.g., Epstein 1997; Grossmann
and Reichardt 1991; Plante and Mayer 1994; Plante and Shriver 1998), zooplankton
(Epstein 1997), and Echinodermata (Findlay and White 1983). The models were
constructed to depict the trophic relationships between the most important species
or functional groups in the benthic communities of Fildes Bay. Notably, the models
excluded the energy flows from epiphytes and the microphytobenthos, in addition to
16  Trophic Networks and Ecosystem Functioning 317

those leading to seals and birds, because insufficient scientific information was
available for these groups. Although these exclusions reduced the realism of the
model’s configuration, the most dominant interdependencies and energy flows are
reflected. Moreover, such a system-level error, if consistent, should not impede a
comparative analysis of ecological systems placed under similar limitations (e.g.,
kelp forest of SE pacific coast).

Box 16.3 Macroscopic Network Properties


The macroscopic descriptors are based on Ulanowicz’s Ascendency analysis
that enables quantification of the level of development and organization of
ecosystems (Ulanowicz 1986, 1997). This approach stems from informa-
tion theory:
• Total System Throughput (TST). Indicates the size of the system, that is, the
total number of flows in the system:
n 1 n  2
TST  Tij (16.7)
i j

where TST is the summation of all flows among compartments; i and j can
represent either an arbitrary system component or the environment.
• Average Mutual Information (AMI). Quantifies the organization of the sys-
tem in relation to the number and diversity of interactions between compo-
nents (complexity):
n2 n2  fij 
AMI   fij  Qi  log 2   (16.8)
Q

i j  j 
where fij is the fraction of flow in element ji in comparison to the TST; Qj
is the probability that species j is the host, calculated as the sum of flows in
the jth row divided by TST; Qi is the probability that i is the predator, cal-
culated as the sum of flows in the ith column divided by TST.
• Ascendency (A). Measures the growth and development of a system and
integrates TST and AMI of flows:
 Tij T0,0 
A  Tij log   (16.9)
T T

ij  i ,0 0, j 
where T is the summation of all flows among compartments (TST); i and j
are the prey and predator, respectively; 0 is the sum of flows of prey or
predators; with Ti,0 the flows from one prey to all their predators, T0,j is the
consumption of a predator over all its prey, and T0,0 is the total sum of flows
over prey and predators.

(continued)
318 M. Ortiz et al.

Box 16.3 (continued)
• Overhead (Ov). Quantifies the degrees of freedom preserved by the net-
work, can be used to estimate the ability of a network to withstand pertur-
bations, and can be estimated from the difference between the Development
Capacity and the Ascendency:
O  C  A (16.10)
• Development Capacity (C). Quantifies the upper limit of Ascendency:
C  A  O (16.11)
• Total Biomass/Total Throughput (TB/TST) ratio. Suggests different states
of system maturity (Christensen 1995).
• A/C and Ov/C ratios. Are used as indicators of ecosystem development and
the ability of the system to resist disturbances (Baird and Ulanowicz 1993;
Costanza and Mageau 1999; Kaufman and Borrett 2010).
• Relative Internal Ascendency (Ai/Ci ratio). Represent well-organized,
mature, and efficient systems that are therefore resistant against perturba-
tions (Baird et al. 1991; Baird and Ulanowicz 1993).

The balancing of models was performed on the basis of the following six criteria
proposed by Heymans et al. (2016): checking that (1) the ecotrophic efficiency (EE)
(for quantifying the proportion of production utilized by the next trophic level
through direct predation or fishing) of all compartments was <1.0 (Ricker 1968); (2)
the gross efficiency (GE) (gross food conversion efficiency) of all compartments
was <0.3 (Christensen and Pauly 1993); and when any inconsistency was detected,
the average biomass was modified within the confidence limits (± 1 standard devia-
tion); (3) the net efficiency (value for food conversion after accounting for unas-
similated food) of all compartments was >GE; (4) the respiration/assimilation
biomass (RA/AS) was <1.0; (5) respiration/biomass (RA/B) values for fish between
1 and 10 years−1 and for groups with higher turnover between 50 and 100 years−1;
and (6) the production/respiration (P/RA) was <1.0.

16.2  Macroscopic Ecosystem-Network Properties

The macrodescriptors based on Ulanowicz’s Ascendency (Ulanowicz 1986, 1997)


include the Total Biomass/Total Throughput (TB/TST) ratio, which suggests differ-
ent states of system maturity (Christensen 1995); Total System Throughput (TST),
which indicates the size of the system, that is, the total number of flows in the sys-
tem; Average Mutual Information (AMI), which quantifies the organization of the
system in relation to the number and diversity of interactions between components
16  Trophic Networks and Ecosystem Functioning 319

(complexity) (Ulanowicz 1986, 1997); Ascendency (A), which measures the growth
and development of a system and integrates TST and AMI of flows; and Overhead
(Ov), which quantifies the degrees of freedom preserved by the network, can be
used to estimate the ability of a network to withstand perturbations and can be esti-
mated from the difference between the Development Capacity (C) and the
Ascendency (A) (Ulanowicz 1986): Development Capacity (C), which quantifies
the upper limit of Ascendency, and A/C and Ov/C ratios, which are used as indica-
tors of ecosystem development and the ability of the system to resist disturbances
(Ulanowicz 1986, 1997; Baird and Ulanowicz 1993; Costanza and Mageau 1999;
Kaufman and Borrett 2010). High values of the Relative Internal Ascendency (Ai/Ci
ratio) represent efficient systems that are therefore resistant against perturbations
(Baird et al. 1991; Baird and Ulanowicz 1993). For more details about macrode-
scriptors, see Box 16.3.

16.2.1  M
 acroscopic Properties of Coastal Benthic-Pelagic
Ecosystem at Fildes Bay

The coastal benthic/pelagic ecological system of Fildes Bay dominated by brown


large macroalgae reached a Total System Throughput (TST) equals to 24,234.0 g
ww m−2 year−1; this magnitude was lower than that calculated for a kelp forest off
the Antofagasta Peninsula (Ortiz 2008a, 2010). However, compared to the benthic
communities of Tongoy Bay (Ortiz and Wolff 2002; Wolff 1994) and different estu-
aries around the world (Baird and Ulanowicz 1993; Patrício and Marques 2006;
Wolff et  al. 2000), TST was higher in the Fildes Bay system. The Development
Capacity (C) accounted for 110,354.4 flow bits, Ascendency (A) accounted for
32,953.9 flow bits, and the A/C and Ov/C ratios were 29.8% and 70.1%, respec-
tively (Table 16.1). Notably, the A/C value calculated for the Fildes Bay ecological
system was one of the lowest compared to those obtained for other coastal areas
along the Chilean coast and around the globe (Table 16.2). The difference between
the A/C and the Ai/Ci ratios for the Fildes Bay model may indicate a dependency of
this system on external connections (sensu Baird et al. 1991).
Ulanowicz (1997) proposed estimating Relative Ascendency by model compo-
nent to evaluate the contribution of each group to the overall structure and function
of the system. In this case, detritus accounted for ~33%, followed by phyto-­
zooplankton at ~26%, macroalgae at ~19%, filter feeders at ~7%, small epifauna at
~5%, and top predators at ~2%. Moreover, the sea star species groups, Chlorophyta,
and the red algae G. skottsbergii accounted for the complexity in the system; that is,
they exhibited the lowest percentage of Average Mutual Information (AMI) within
the Fildes model system (Table 16.2).
Macroscopic properties, such as the A/C ratio, Ov/C ratio, and Redundancy
values, indicate that Fildes Bay would be a less developed system but is more
resistant to disturbances than a kelp forest off the Mejillones Peninsula or the
320 M. Ortiz et al.

Table 16.1 Macroscopic Property Fildes Bay


properties of the trophic
Total System Throughput (TST) (g ww 21,432.00
network for the coastal
m−2 year−1)
benthic-pelagic ecosystem of
Fildes Bay (Antarctica) Total Biomass/Total System Throughput 0.1312
(TB/TST)
Ascendency (A) (g ww m−2 year−1 ∗ 29,758.60
bits)
Overhead (Ov) (g ww m−2 year−1 ∗ bits) 69,355.02
Development Capacity (C) (g ww m−2 99,114.87
year−1 ∗ bits)
Average Mutual Information (AMI) 1.39
A/C (%) 30.02
Ai/Ci (%) 14.00
Ov/C (%) 69.97

seagrass ­meadows of Tongoy Bay (Ortiz and Wolff 2002; Ortiz 2008a, 2010)
(Tables 16.1 and 16.2). This may be explained by the fact that Fildes Bay is nega-
tively affected by the Antarctic’s austral winters, which are characterized by lower
temperatures and freezing, leading to a reduction in the herbivore biomass and
thereby constraining the flow of energy/matter toward the upper trophic levels.
Additionally, as shown in Tables 16.1 and 16.2, the different estuaries and coral
reef systems appeared to be more developed (A/C and Ov/C) but less resistant to
perturbations compared to Fildes Bay and the benthic ecosystems studied along
the Chilean coast. This latter comparison should be taken with a degree of caution
because the trophic model constructed for Fildes Bay represents only a narrow
temporal window, and unknown system characteristics may emerge during the
rest of the year. The difference between the A/C and Ai/Ci ratios in the Fildes Bay
model system may primarily be a consequence of the omission of the flows to
birds and marine mammals from our analysis. The analysis of relative Ascendency
by component revealed that those groups that principally contributed to the over-
all structure and function of the Fildes Bay system (i.e., detritus, the phyto-zoo-
plankton complex, and macroalgae) differed from those that contributed to the
kelp forest (Mejillones Peninsula) (Ortiz 2008a, 2010) but were similar to those
within the benthic systems of Tongoy Bay and the La Rinconada Marine Reserve
(Antofagasta Bay) (Ortiz and Wolff 2002; Ortiz et al. 2010). This outcome indi-
cates that although a significant amount of the Antarctic system’s biomass is con-
centrated in macroalgae, these macroalgae would contribute fewer nutrients to the
coastal marine ecosystem than those within kelp forests (Duggins et  al. 1989;
Ortiz 2008a, 2010).
16  Trophic Networks and Ecosystem Functioning 321

Table 16.2  Macroscopic properties for system’s development and organization derived from
Ascendency network analysis for Fildes Bay (Antarctic) and other coastal ecosystems
Macroscopic properties of trophic networks
Ascendency theoretical framework (Ulanowicz 1986, 1997)
TST C A A/C Ov Ov/C
(g ww m−2 (flow (flow (flow
Coastal marine ecosystems year−1) bits) bits) % bits) %
Along the Chilean coast (SE Pacific)
Benthic/pelagic ecological 21,432.00 99,114.87 29,758.60 30.02 69,355.02 69.97
system of Fildes Bay1
Kelp ecological system 72,512.0 207,777.4 93,462.6 45.0 112,548.0 55.0
dominated by M. pyrifera,
Antofagasta Peninsula2
Kelp forest ecological system 50,105.0 200,609.4 77,613.5 38.7 117,678.9 61.3
dominated by L. trabeculata,
Antofagasta Peninsula2
Kelp forest ecological system, 85,217.0 332,041.6 117,939.7 35.5 211,848.3 64.5
Antofagasta Peninsula3
Seagrass habitat ecological 18,746.6 69,270.4 21,557.8 31.1 46,991.0 68.9
system of Tongoy Bay4
Mud habitat ecological system 17,451.3 59,139.0 19,354.8 32.7 39,433.4 67.3
of Tongoy Bay4
Benthic/pelagic ecological 20,834.9 80,689.8 26,312.6 32.6 54,377.2 67.4
system of Tongoy Bay5
La Rinconada Marine Reserve 20,124.0 80,321.0 24,375.1 30.3 55,945.9 69.7
coastal ecological system,
Antofagasta Bay6
Mejillones benthic/pelagic 29,429.8 142,897.9 34,395.1 24.1 108,353.1 75.9
ecological system of Mejillones
Bay7
Antofagasta benthic/pelagic 37,539.8 170,237.0 48,574.3 28.5 121,434.8 71.5
ecological system of
Antofagasta Bay7
Around the world
Coral reef ecosystem, 148,094.1 318,400.0 178,200.0 56.3 139,800.0 43.7
Chinchorro Bank, México8
Coral reef ecosystem, Mexican 194,758.4 890,301.6 308,428.7 34.64 581,872.9 65.35
Pacific Coast9
Mangrove estuary of Caeté, 10,558.6 44,741.4 12,261.6 27.0 31,129.8 63.0
Brazil10
Zostera meadows of Mondego 10,852.0 39,126.0 16,550.3 42.3 22,575.7 57.7
Estuary, Portugal11
Ems estuary in the 12,980.0 6085.0 2327.0 38.3 3758.0 61.7
Netherlands12
Benguela upwelling ecosystem, 8897.0 36,041.0 17,313.0 48.1 18,728.0 51.9
Namibia13
TST Total System Throughput, C System Capacity, A Ascendency
1
Current study, 2Ortiz (2008a), 3Ortiz (2010), 4Ortiz and Wolff (2002), 5Wolff (1994), 6Ortiz et al.
(2010), 7Ortiz et al. (2015), 8Rodriguez-Zaragoza et al. (pers. comm.), 9Hermosillo-Núñez et al.
(2018), 10Wolff et al. (2000), 11Patrício and Marques (2006), 12Baird and Ulanowicz (1993), and
13
Heymans and Baird (2000)
322 M. Ortiz et al.

16.3  Keystone Species Complex (KSC)

16.3.1  Functional Keystoneness Indices

Once the trophic model was balanced following the rules given by Heymans et al.
(2016), the functional keystone index (KSi) developed by Libralato et al. (2006) was
used. This index is an extension of the mixed trophic impacts (MTI, Ulanowicz and
Puccia 1990). Because every impact can be quantitatively positive or negative, a
new measure of the overall effect must be determined for each species or functional
group (εi) using the following mathematical Eq. 16.2:

n
i  m
j i
2
ij

where mij corresponds to the elements of the MTI matrix and quantifies the direct
and indirect effects that each (affecting) species or group i has on any (affected)
group j of the food web. However, the effect of the change in a group’s biomass on
the group itself (i.e., mii) is not included. The contribution of biomass from every
species or functional group with respect to the total biomass of the network was
estimated using the following Eq. 16.3:

Bi
pi  n

B k
i
where pi is the proportion of biomass of each species Bi with respect to the sum of
the total biomass Bk. Therefore, to balance the overall effect and the biomass, the
keystone index (KSi) for each species or functional group was established using the
Eq. 16.4, which integrates the Eqs. 16.2 and 16.3 as follows:

KSi  log  i  1  pi  



This index assigns high values of functional keystoneness to those variables (spe-
cies) or functional groups that have low biomass and a high overall effect.
The propagation of direct and indirect effects and system recovery time (SRT)
magnitudes estimated by Ecosim were treated in the same way as those obtained
with MTI in order to obtain two additional functional keystone indices. The Ecosim
simulations were used to evaluate the propagation of instantaneous direct and indi-
rect effects and the system recovery time (SRT) (as a system resilience measure) in
response to a steady increase in the total mortality (Z) of all compartments (see
Eqs. 16.5 and 16.6) which was set equivalent to 10%, 30%, and 50%. This proce-
dure was done between the first and second year of simulation for all components
considered in the model. These three magnitudes (scenarios) were set for prediction
16  Trophic Networks and Ecosystem Functioning 323

purposes as a measure of confidence. As the models studied represent only short-­


term (transient) dynamics, the propagation of instantaneous effects was determined
by evaluating the changes of biomass in the remainder variables in the third year of
simulation. All dynamic simulations by Ecosim were carried out using the following
vulnerabilities (flow control): (1) bottom-up (prey controls the flow), (2) top-down
(predator controls the flow), and (3) mixed (both prey and predator control the flow)
(Boxes 16.1 and 16.3):

Z  M  natural mortality   F  fishing mortality 



Production  P   Biomass B   Z

After that, Eqs. 16.2, 16.3, and 16.4 were used to obtain one keystone species index
related to the propagation of direct and indirect effects (KSiEcosim1), and Eqs. 16.3
and 16.4 were used to obtain another functional keystone species index related to
SRT magnitudes (KSiEcosim2). These indices show, similar as the KSi index (Libralato
et al. 2006), high values of keystoneness associated to compartments with low bio-
mass and a high overall effect.

16.3.2  Topological Keystone Index

The structural keystone index (Ki) developed by Jordán et  al. (1999) and Jordán
(2001), following Harary (1961), was also used in this work. This index is applica-
ble for trophic hierarchies (directed acyclic trophic networks) and considers direct
and indirect interactions in vertical directions (i.e., bottom-up and top-down). The
structural keystone index of the ith species or functional group (Ki) is calculated
using the following Eq. 16.7:
n n
1 1
Ki    1  K bc     1  K te 
c 1 dc e 1 fe

where n is the number of predator species eating species i, dc is the number of prey
of the cth predator, Kbc is the bottom-up keystone index of the cth predator, and sym-
metrically we have m as the number of prey species eaten by species i, fe as the
number of predators of its eth prey, and Kte as the top-down keystone index of the
eth prey. Within this index, the first and second components represent the bottom-up
(Kbc) and top-down (Kte) effects, respectively. Finally, the keystone index (Ki) cor-
responds to the highest value as a product of the addition of bottom-up (Kbc) and
top-down (Kte) components. For more details on this method, see Jordán (2001) and
Vasas et al. (2007). The Ki index has been shown to be one of the most robust cen-
trality indices (Fedor and Vasas 2009). It is important to indicate that only bottom-
­up and top-down components of Ki were used in the current work as a way to
324 M. Ortiz et al.

compare functional indices obtained using Ecosim simulations under comparable


flow control mechanisms.

16.3.3  Semiquantitative or Qualitative Keystone Index

A keystoneness index based on qualitative or semiquantitative loop models was also


calculated. Once the stabilized trophic matrix with Fn < 0 was obtained, the self-­
dynamics of each variable corresponding to the principal diagonal were modified to
estimate two new perturbed magnitudes of local stability Fp. Based on the distance
(△) between Fn and Fp as shown in Eq. 16.8

 Fn  Fp

it was possible to determine the change provoked by each variable on initial stability
(Fn), thereby obtaining a first qualitative keystone species index (KQiLA1) (selecting
only the largest change by variable). Because Loop Analysis does not consider the
abundance of the components, the difference (Δ) was treated in similar way to
Eq.  16.4 to obtain an additional keystone index (KQiLA2) in which high values of
keystoneness corresponded to variables with low biomass and a high overall effect.
Due to the qualitative-dialectic character of Loop Analysis, the prey-predator inter-
action is captured as a mixed control mechanism.

16.3.4  Centrality of Node Sets

Field studies suggest that in some situations, a small group of species behave as
keystones and they form a keystone species complex (Daily et al. 1993). The impor-
tance of this group is typically realized through their interspecific interaction net-
work, so a network approach to better understand multispecies keystone complexes
is reasonable. A particular approach was suggested by Borgatti (2003a) in order to
find the most central set of k nodes in a network. According to this, a topological
keystone species complex is defined as a solution of the KeyPlayer Problem (KPP)
(sensu Borgatti 2003a). The software KeyPlayer 1.44 (Borgatti 2003b) was used to
compute the importance of species combinations in maintaining the integrity of a
network. The importance of a set of nodes can be calculated by considering either
their fragmentation effect (KPP1) or their reachability effect (KPP2). In the first
case, we identify which k nodes should be deleted from the network of n nodes in
order to maximally increase its fragmentation. In the second case, we identify from
which k nodes the largest proportion of the other n–k nodes are reachable within a
certain distance. Based on fragmentation (F of KPP1), the best set of the deleted k
nodes can maximally increase the fragmentation of the network. This means an
16  Trophic Networks and Ecosystem Functioning 325

increase of the number of components and a larger average distance generated


within individual components. We used k = 1, 2, and 3 with 10,000 simulations for
each. We also consider the distance-based reachability approach (Rd of KPP2). We
simply count the number of nodes that are reachable within a given distance of m = l
step from a given set of k nodes. We have chosen m = 2 steps and increased the size
of the KP set from k = 1 to k = 3. We applied 10,000 runs for each simulation. The
outcome was three sets of nodes (for k = 1, 2, 3) for each network, containing spe-
cies codes. For each k, the software presents the percentage of nodes outside the KP
set but reachable from it in one step. If this percentage reaches 100%, then the
whole network is reachable from the KP set and we cannot create larger KP sets.

16.3.5  K
 eystone Species Complex in Benthic-Pelagic
Ecosystem at Fildes Bay

All individual indices integrated in the holistic keystone species complex (KSCi)
detected keystoneness for a variety of different species and functional groups, with
some agreement. In general terms, keystoneness properties were detected for spe-
cies of all different trophic levels, including primary producers, herbivores, and top
predators. Notably, the topological-structural (Ki) and the functional (KSi) indices
both identified sea stars (SS) as keystone species even though they are based on dif-
ferent algorithms. This outcome agrees partially with the field observations of
Gaymer and Himmelman (2008), who studied dominant sea star species in benthic
communities of northern Chile, establishing Meyenaster gelatinosus as a keystone
species in subtidal systems. Likewise, Ortiz et al. (2013a) determined that most of
the keystone species complexes identified by KSC indices in different ecological
systems along the Chilean coast include one asteroid species as the top predator.
The relevance of the sea star species determined in the present work also coincides
with the results described by Ortiz et al. (2009) regarding the longest system recov-
ery times (as a measure of “resilience”) being obtained in response to perturbations
on these species.
Similarly, both the qualitative and semiquantitative keystone indices (KQiLA1 and
KQiLA2) identified small epifauna (SE) to have keystoneness properties. This result
is very interesting because loop model predictions respond with a high degree of
certainty to external perturbations (Briand and McCauley 1978; Lane and Blouin
1985; Lane 1986; Hulot et al. 2000; Ortiz 2008b). Regarding the functional indices
based on Ecosim dynamical simulations (under three mortality levels and three
types of flow control mechanisms), KSiECOSIM1 and KSiECOSIM2 both identified the
group of Chlorophyta (Chloro), phytoplankton (Phyto), zooplankton (Zoo), the
small epifauna (SE), and the species S. neumayeri (Sn) (grazers) as having key-
stoneness. It is important to mention that it was not possible to determine species or
functional groups with keystoneness properties based on the functional KSiECOSIM2
index under a 30% increased mortality and using a top-down flow control mecha-
326 M. Ortiz et al.

Fig. 16.2  Keystone species complexes (KSC) determined for coastal benthic-pelagic ecosystem at
Fildes Bay (Antarctica) based on quantitative and semiquantitative networks (a) and centrality of
node sets (b). The share components are highlighted (Note: the circle and arrow mean negative and
positive effect, respectively). (Adapted from Fig. 3 in Ortiz et al. 2017)

nism because the model system does not return to initial steady-state conditions,
instead oscillating persistently. The core species as indicated by the keystone species
complex index (KSCi) for the coastal ecological system of Fildes Bay account by
24.2% of the total system biomass (Fig. 16.2a). Importantly, the KSC includes spe-
cies and functional groups that make up an ecological path clearly representing
three trophic levels.
The composition of the KeyPlayer sets is nested: For Fk = 1, the key group is phy-
toplankton (Phyto); for Fk = 2, the key groups are Phyto and benthic fishes (BF); and
for Fk = 3, the key groups are Phyto, BF, and the sea stars (SS). For Rkd=1 , the key
group is Phyto; for Rkd=2 , the key groups are Phyto and BF; and for Rkd=3 , the key
groups are Phyto, BF, and S. neumayeri (Sn). This means that the Phyto-Sn-SS
chain as well as BF together composes a core of species in this community
(Fig.  16.2b). The keystone species complex obtained by multinode centrality
­represents 24.4% of the total system biomass, a value quite similar to the previous
approach. The multinode approach based on the KP indices thus not only partly
reinforces the identity of some key players (e.g., S. neumayeri) but also suggests
new key organisms (e.g., zooplankton). This latter result is well supported by the
literature (e.g., Stibor et al. 2004; Murphy et al. 2007). That phytoplankton that was
identified as a component of the KSC could be a consequence of the higher level of
primary productivity in Antarctic waters (Smith et al. 2007; Cornejo-Donoso and
Antezana 2008).
16  Trophic Networks and Ecosystem Functioning 327

The keystone species complex indices (KSCi) determined for Fildes Bay inte-
grate fewer components than the one determined for the kelp forest of northern
Chile. However, both ecological systems share a sea urchin species (herbivore posi-
tioned at intermediate trophic level), a sea star species (top predator), and the
Chlorophyta (Chloro) (primary producers). After all, the outcomes obtained show
that the components with keystoneness properties in the benthic-pelagic system of
Fildes Bay are widely heterogeneous, coinciding with results reported for other
ecosystems (Power et  al. 1996; Piraino et  al. 2002; Libralato et  al. 2006).
Furthermore, Jordán et al. (2007, 2008) reported similar findings after comparing
several structural and functional keystone species indices. Hermosillo-Núñez et al.
(2018) showed that in spite the wide trophic heterogeneity of components with key-
stoneness properties, it is possible to observe that the core set of species and func-
tional groups are trophically linked. Okey (2004) arrived at similar results by
defining keystone guilds or clusters of species with keystoneness properties based
on a trophic model in Alaska. Thus, we believe that the keystone species complex
index (KSC) for coastal benthic/pelagic ecological systems of Fildes Bay would
facilitate the design and assessment of conservation and monitoring measures, espe-
cially when the Antarctic coastal marine ecosystems are being severely stressed by
the direct effects of the global warming and UV radiation (Richter et  al. 2008;
Pessoa 2012). The above notwithstanding, it is necessary also to recognize that the
use of the keystone species complex indices is still quite difficult because the tradi-
tional view of conservation and monitoring efforts is based principally on single
species such as keystone and/or niche-builder or bioengineer species. This mindset
undoubtedly imposes even greater challenges to understand how global changes,
covarying with variability of the natural system, act on networks of interacting
species.

16.4  C
 ontribution of Keystone Species Complex
to Macroscopic Network Properties

In global terms, the species and functional group belonging to KSCs showed differ-
ent degrees of contribution to the ecosystem’s emergent properties (i.e., growth,
organization, development, maturity, and health) (Table 16.3). KSC index and cen-
trality of node sets account to lower magnitudes for all macroscopic properties
(Table 16.3). The total biomass of KSCs accounted <25% of the total system bio-
mass, coinciding with the classical keystone concept (sensu Power et al. 1996). A
particular case is the functional group of phytoplankton, which concentrated a little
more biomass, contributing to higher values of TST, AMI, and Ascendency.
Likewise, the functional groups of Chlorophyta (Chloro) and sea stars (SS) pre-
sented the higher percentage magnitudes of Ov/C and TB/TST ratios, mainly con-
tributing – in term of flow – to the system’s resistance against perturbations and to
the whole system’s cycling (promoting efficiency). Different indices for the deter-
328 M. Ortiz et al.

Table 16.3  Contribution of species and functional groups of KSCs on macroscopic network
properties derived from Ascendency (Ulanowicz 1986, 1997) of the coastal benthic-pelagic
ecosystem of Fildes Bay (Antarctica)
TB % TST % AMI % A% Ov/C (%) A/C (%) TB/TST (%)
(a) KSC index
SS 0.658 0.382 0.357 0.357 6.551 3.614 5.953
SE 2.364 4.254 4.743 4.740 5.866 5.943 1.920
Sn 1.226 1.922 2.103 2.104 6.082 5.206 2.204
Chloro 1.369 0.381 0.340 0.340 6.675 3.184 12.402
Phy 10.665 22.283 18.897 18.872 5.398 7.526 1.654
Contribution KSC 16.283 29.223 26.440 26.413 30.572 25.473 24.132
(b) K node sets
BF 1.161 1.146 1.074 1.077 6.308 4.435 3.502
SS 0.658 0.382 0.357 0.357 6.551 3.614 5.953
Sn 1.226 1.922 2.103 2.104 6.082 5.206 2.204
FF 6.309 6.223 6.904 6.910 5.490 7.217 3.503
Zoo 4.355 7.836 7.154 7.155 5.868 5.931 1.920
Phy 10.665 22.283 18.897 18.872 5.398 7.526 1.654
Contribution KSC 24.375 39.792 36.490 36.474 35.698 33.929 18.735
(c) KSC integrated
SS 0.658 0.382 0.357 0.357 6.551 3.614 5.953
Sn 1.226 1.922 2.103 2.104 6.082 5.206 2.204
Phy 10.665 22.283 18.897 18.872 5.398 7.526 1.654
Contribution KSC 12.550 24.587 21.357 21.333 18.031 16.346 9.811
SS sea stars, SE small epifauna, Sn S. neumayeri, Chloro Chlorophyta, Phy phytoplankton, BF
benthic fishes, FF filter feeders, Zoo zooplankton

mination of keystone species complexes (KSCs) are suitable when biomass content
is considered, and it is consistent when using the total flows of matter and/or energy
(TST), the contribution to complexity (AMI), and system development and heath
(Ascendency).

16.5  Constrains and Perspectives

Although we were well aware that the quantitative trophic model built and analyzed
in this study is a partial representation of the overall trophic seascape and interac-
tions underlying the dynamics within Fildes Bay’s coastal benthic/pelagic ecologi-
cal system, such limitations, however, occur in any type of model and are independent
16  Trophic Networks and Ecosystem Functioning 329

of the model’s degree of complexity (Levins 1966; Ortiz and Levins 2011). In the
present study, the following limitations were identified: (1) The model only repre-
sented the austral summer condition, while the annual benthic/pelagic dynamics are
unknown; (2) system complexity was reduced in relation to the composition of sev-
eral functional groups, although the most abundant macroalgae, herbivore, and car-
nivore species were represented; and (3) regardless of the inherent, well-known
limitations and shortcomings of the Ecopath, Ecosim, Loop Analysis, and KeyPlayer
node sets theoretical frameworks, the constructed model and its dynamic simula-
tions represented underlying system processes based exclusively on short-term or
transient dynamics. In spite of these concerns, we claim that the most relevant tro-
phic relationships and energy/matter flows were well reflected in our model. Here,
the macroscopic properties and sensitive model compartments of the system were
quantified and compared adequately.
Likewise, the core species and functional groups that constituted the keystone
species complexes (KSCs) in the coastal ecological systems at Fildes Bay accounted
to lower magnitudes of total system biomass and macroscopic properties, coincid-
ing with those groups and species identified as keystones in experimental studies
using variations of the original keystone species concept (Menge et al. 1994; Estes
et al. 1998; Bond 2001). Several species or functional groups from different trophic
levels could have keystone properties. This result should not be considered as
ambiguous because populations inhabit heterogeneous and changing environments
(Levins 1968). This insight would support the design of putative conservation and
monitoring strategies in the Antarctic Peninsula, including a core of species or func-
tional groups linked trophically (Jordán et al. 2019), which could supplement the
unique species with keystone properties or those species considered as niche con-
structers (sensu Lewontin and Levins 2007) or bioengineers (sensu Jones et  al.
1994). We claim that the adequacy of the description of ecosystem properties based
on only one species or functional group is limited, especially when the task is to
conserve and monitor complex ecosystems (Jordán et al. 2019). In this sense, the
determination of the keystone species complexes (KSCs) could assist within an
ecosystem-­based conservation view under an ecological network context (sensu
Pikitch et al. 2004). Finally, it is relevant to assess the trajectory of the target spe-
cies, as they constitute compartments with a relevant role in the structure and tro-
phic functioning of the benthic-pelagic coastal ecosystem at Fildes Bay, Antarctica.

Acknowledgments This research was financed by the grant Proyecto Anillo ART1101


(CONICYT-PIA). We also thank the Instituto Antártico Chileno (INACH) for logistic support. The
work of Ferenc Jordán was supported by National Research, Development and Innovation Office –
NKFIH, grant OTKA K 116071.
330 M. Ortiz et al.

Appendix 16.A

Table 16.4  Parameter values entered (in bold), estimated (standard), and calculated (in italics) by
the Ecopath with Ecosim (EwE) software for the species and functional groups in the coastal
benthic-pelagic ecosystem of Fildes Bay (Antarctica)
Model compartments
Species/functional groups TL B P/B Q/B EE GE
(1) Benthic fishes (BF) 3.03 32.67 2.40 8.50 0.05 0.28
(2) Diplasterias brucei (Db) 2.98 37.83 1.40 5.00 0.18 0.28
(3) Sea stars (SS) 3.29 18.52 1.40 5.00 0.04 0.28
(4) Small epifauna (SE) 2.76 66.50 4.60 15.50 0.95 0.30
(5) Sterechinus neumayeri (Sn) 2.35 34.50 3.70 13.50 0.99 0.27
(6) Margarella sp. (Msp) 2.10 59.50 3.40 11.50 0.99 0.30
(7) Nacella concinna (Nc) 2.05 56.50 3.60 12.50 0.99 0.29
(8) Filter feeders (FF) 2.30 177.46 1.87 8.50 0.93 0.22
(9) Himmantothallus grandifolius (Hg) 1.00 597.67 2.10 – 0.35 –
(10) Desmarestia anceps (Da) 1.00 310.50 2.30 – 0.38 –
(11) Gigartina skottsbergii (Gs) 1.00 35.17 2.50 – 0.42 –
(12) Phaeophyceae (Phaeo) 1.00 650.50 2.80 – 0.39 –
(13) Chlorophyta (Chloro) 1.00 38.50 2.40 – 0.36 –
(14) Rhodophyta (Rhodo) 1.00 174.57 5.00 – 0.32 –
(15) Zooplankton (Zoo) 2.00 122.50 4.60 15.50 0.99 0.30
(16) Phytoplankton (Phy) 1.00 300.00 18.00 – 0.54 –
(17) Detritus (Det) 1.00 100.00 – – 0.07 –
TL trophic level, B biomass (g wet weight m−2), P/B productivity (year−1), Q/B consumption rate
(year−1), EE ecotrophic efficiency (dimensionless), GE gross efficiency (dimensionless)

References

Allen KR (1971) Ralation between production and biomass. J Fish Res Board Can 28:1573–1581
Allesina S, Bodini A (2004) Who dominates whom in the ecosystem? Energy flow bottlenecks and
cascading extinctions. J Theor Biol 230:351–358
Baird D, Ulanowicz RR (1993) Comparative study on the trophic structure, cycling and ecosystem
properties of four tidal estuaries. Mar Ecol Prog Ser 99:221–237
Baird D, McGlade JJ, Ulanowicz RE (1991) The comparative ecology of six marine ecosystems.
Philos Trans R Soc Biol Sci 333:15–29
Barua M (2011) Mobilizing metaphors: the popular use of keystone, flagship and umbrella species
concepts. Biodivers Conserv 20:1427–1440
Benedek Z, Jordán F, Báldi A (2007) Topological keystone species complexes in ecological inter-
action networks. Community Ecol 8:1–7
Bond W (2001) Keystone species – hunting the shark? Science 292:63–64
Borgatti SP (2003a) The key player problem. In: Breiger R, Carley Pattison P (eds) Dynamic
social network modeling and analysis: Workshop Summary and Papers. Committee on Human
Factors, National Research Council. National Academies Press, Washington, DC, pp 241–252
Borgatti SP (ed) (2003b) KeyPlayer. Analytic technologies. Harvard University Press, Boston
16  Trophic Networks and Ecosystem Functioning 331

Briand F, McCauley E (1978) Cybernetic mechanisms in lake plankton systems: how to control
undesirable algae. Nature 273:228–230
Christensen V (1995) Ecosystem maturity–towards quantification. Ecol Model 77:3–32
Christensen V, Pauly D (1992) Ecopath II: a software for balancing steady–state ecosystem models
and calculating network characteristics. Ecol Model 61:169–185
Christensen V, Pauly D (1993) Trophic models in aquatic ecosystem. ICLARM Conf Proc 26:390
Christensen V, Walters CJ (2004) Ecopath with Ecosim: methods, capabilities and limitations. Ecol
Model 172:109–139
Corbisier TN, Petti MAV, Skowronski RSP, Brito TAS (2004) Trophic relationships in the near-
shore zone of Martel Inlet (King George Island, Antarctica): δ13C stable-isotope analysis.
Polar Biol 27:75–82
Cornejo-Donoso J, Antezana T (2008) Preliminary trophic model of the Antarctic Peninsula eco-
system (sub-area CCAMLR 48.1). Ecol Model 218:1–17
Costanza R, Mageau M (1999) What is a healthy ecosystem? Aquat Ecol 33:105–115
Daily GC, Ehrlich PR, Haddad NM (1993) Double keystone bird in a keystone species complex.
PNAS 90:592–594
Darmbacher J, Guaghan D, Rochet M, Rossignol P, Trenkel V (2009) Qualitative modelling and
indicators of exploited ecosystems. Fish Fish 10:305–322
Duggins DO, Simenstad CA, Estes JA (1989) Magnification of secondary production by kelp
detritus in coastal marine ecosystems. Science 245:101–232
Epstein S (1997) Microbial food web in marine sediments: I. Trophic interactions and grazing rates
in two tidal flat communities. Microb Ecol 34:188–198
Estes JA, Tinker MT, Williams TM, Doak DF (1998) Killer whale predation on sea otters linking
oceanic and near shore ecosystems. Science 282:473–476
Fedor A, Vasas V (2009) The robustness of keystone indices in food webs. J Theor Biol
260:372–378
Findlay R, White D (1983) The effects of feeding by the sand dollar Mellita quinquiesperforata
(Leske) on the benthic microbial community. J Exp Mar Biol Ecol 72:25–41
Gaymer C, Himmelman JH (2008) A keystone predatory sea star in the intertidal zone is controlled
by a higher-order predatory sea star in subtidal zone. Mar Ecol Prog Ser 370:143–153
Giacaman-Smith J, Neira S, Arancibia H (2016) Community structure and trophic interactions in a
coastal management and exploitation area for benthic resources in central Chile. Ocean Coast
Manag 119:155–163
Gili JM, Coma R, Orejas C, López-González PJ, Zavala M (2001) Are Antarctic suspension feed-
ing communities different from those elsewhere in the world? Polar Biol 24:473–485
Grossmann S, Reichardt W (1991) Impact of Arenicola marina on bacteria in intertidal sediments.
Mar Ecol Prog Ser 77:85–93
Harary F (1961) Who eats whom? Gen Syst 6:41–44
Hawkins S (2004) Scaling up: the role of species and habitat patches in functioning of coastal
ecosystems. Aquat Conserv Mar Freshwat Ecosyst 14:217–219
Hermosillo-Núñez BB, Ortiz M, Rodríguez-Zaragoza FA (2018) Keystone species complexes in
kelp forest ecosystems along the northern Chilean coast (SE Pacific): improving multispecies
management strategies. Ecol Indic 93:1101–1111
Heymans J, Baird D (2000) A carbon flow model and network analysis of the northern Benguela
upwelling system, Namibia. Ecol Model 126:9–32
Heymans J, Coll M, Link JS, Mackinson S, Steenbeek J, Walters C, Christensen V (2016) Best
practice in Ecopath with Ecosim food–web models for ecosystem–based management. Ecol
Model 331:173–184
Hulot F, Lacroix G, Lescher-Moutoué F, Loreau M (2000) Functional diversity governs ecosystem
response to nutrient enrichment. Nature 405:340–344
Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic
seaweeds along the depth gradient. Polar Biol 36:1319–1332
332 M. Ortiz et al.

Jacob U, Mintenbeck K, Brey T, Knust R, Beyer K (2005) Stable isotope food web studies: a case
for standardized sample treatment. Mar Ecol Prog Ser 287:251–253
Jones C, Lawton J, Shachak M (1994) Organism as ecosystem engineers. Oikos 69:373–386
Jordán F (2001) Trophic fields. Community Ecol 2:181–185
Jordán F, Scheuring I (2004) Network ecology: topological constraints on ecosystem dynamics.
Phys Life Rev 1:139–172
Jordán F, Takács-Sánta A, Molnár I (1999) A reliability theoretical quest for keystones. Oikos
86:453–462
Jordán F, Benedek Z, Podani J (2007) Quantifying positional importance in food webs: a compari-
son of centrality indices. Ecol Model 205:270–275
Jordán F, Okey T, Bauer B, Libralato S (2008) Identifying important species: linking structure and
function in ecological networks. Ecol Model 216:75–80
Jordán F, Pereira J, Ortiz M (2019) Mesoscale network properties in ecological system models.
Curr Opin Syst Biol 13:122–128
Kaehler S, Pakhomov EA, McQuaid CD (2000) Trophic structure of the marine food web at the
Prince Edward Islands (Southern Ocean) determined by δ13C and δ15N analysis. Mar Ecol Prog
Ser 208:13–20
Kaufman AG, Borrett SR (2010) Ecosystem network analysis indicators are generally robust
to parameter uncertainty in a phosphorus model of Lake Sidney Lanier, USA.  Ecol Model
221:1230–1238
King JC (1994) Recent climate variability in the vicinity of the Antarctic Peninsula. Int J Climatol
14:357–369
Lane P (1986) Symmetry, change, perturbation, and observing mode in natural communities.
Ecology 67:223–239
Lane P, Blouin A (1985) Qualitative analysis of the pelagic food webs of three acid impacted lakes.
Int Rev Ges Hydrobiol 70:203–220
Levins R (1966) The strategy of model building in population biology. Am Sci 54:421–431
Levins R (ed) (1968) Evolution in changing environments. Princeton Monographs Series. Princeton
University Press, Princeton
Levins R (1974) The qualitative analysis of partially specified systems. Ann NY Acad Sci
231:123–138
Levins R (1998a) Qualitative mathematics for understanding, prediction, and intervention in com-
plex ecosystems. In: Raport D, Costanza R, Epstein P, Gaudet C, Levins R (eds) Ecosystem
health. Blackwell Science, MA, pp 178–204
Levins R (1998b) The internal and external in explanatory theories. Sci Cult 7:557–582
Lewontin R, Levins R (2007) Gene, organism, and environment. In: Lewontin R, Levins R (eds)
Biology under the influence: dialectical assays on ecology, agriculture, and health. Monthly
Review Press, NY, pp 221–234
Libralato S, Christensen V, Pauly D (2006) A method for identifying keystone species in food web
models. Ecol Model 195:153–171
Luczkovich J, Borgatti S, Johnson J, Everett M (2003) Defining and measuring trophic role simi-
larity in food webs using regular equivalence. J Theor Biol 220:303–321
Menge BA, Berlow EL, Balchette CA, Navarrete SA, Yamada SB (1994) The keystone species
concept: variation in interaction strength in a rocky intertidal habitat. Ecol Monogr 64:249–286
Mills L, Soulè M, Doak F (1993) The key-stone-species concept in ecology and conservation.
Bioscience 43:219–224
Mincks S, Smith CR, Jeffreys RM, Sumida PYG (2008) Trophic structure on the West Antarctic
Peninsula shelf: detritivory and benthic inertia revealed by δ13C and δ15N analysis. Deep-Sea
Res II 55:2502–2514
Murphy EJ, Watkins J, Trathan P, Reid K, Meredith M, Thorpe S, Johnston N, Clarke A, Tarling G
et al (2007) Spatial and temporal operation of the Scotia Sea ecosystem: a review of large-scale
links in a krill centred food web. Philos Trans R Soc Lond B Biol Sci 362:113–148
16  Trophic Networks and Ecosystem Functioning 333

Norkko A, Thrush SF, Cummings VJ, Gibbs MM, Andrew NL, Norkko J, Schwarz A-M (2007)
Trophic structure of coastal Antarctic food webs associated with changes in sea ice and food
supply. Ecology 88:2810–2820
Odum EP (1969) The strategy of ecosystem development. Science 164:262–270
Okey T (2004) Shifted community states in four marine ecosystems: some potential mechanisms.
PhD thesis, The University of British Columbia, Canada, 173 pp
Ortiz M (2008a) Mass balanced and dynamic simulations of trophic models of kelp ecosystems
near the Mejillones Peninsula of Northern Chile (SE Pacific): comparative network structure
and assessment of harvest strategies. Ecol Model 216:31–46
Ortiz M (2008b) The effect of a crab predator (Cancer porteri) on secondary producers versus
ecological model predictions in Tongoy Bay (SE Pacific coast): implications to management
and fisheries. Aquat Conserv 18:923–929
Ortiz M (2010) Dynamical and spatial models of kelp forest of Macrocystis integrifolia and
Lessonia trabeculata (SE Pacific) for assessment harvest scenarios: short-term responses.
Aquat Conserv 20:494–506
Ortiz M, Levins R (2011) Re–stocking practices and illegal fishing in northern Chile (SE Pacific
coast): a study case. Oikos 120:1402–1412
Ortiz M, Levins R (2017) Self–feedbacks determine the sustainability of human interventions
in eco–social complex systems: impacts on biodiversity and ecosystem health. PLoS One
12(4):e0176163. https://doi.org/10.1371/journal.pone.0176163
Ortiz M, Wolff M (2002) Trophic models of four benthic communities in Tongoy Bay (Chile):
comparative analysis and preliminary assessment of management strategies. J Exp Mar Biol
Ecol 268:205–235
Ortiz M, Avendaño M, Campos L, Berrios F (2009) Spatial and mass balanced trophic models of
La Rinconada Marine Reserve (SE Pacific coast), a protected benthic ecosystem: management
strategy assessment. Ecol Model 220:3413–3423
Ortiz M, Avendaño M, Cantillañez M, Berrios F, Campos L (2010) Trophic mass balanced mod-
els and dynamic simulations of benthic communities from La Rinconada Marine Reserve off
Northern Chile: network properties and multispecies harvest scenario assessments. Aquat
Conserv 20:58–73
Ortiz M, Levins R, Campos L, Berrios F, Campos F, Jordán F, Hermosillo B, González J,
Rodríguez-Zaragoza FA (2013a) Identifying keystone trophic groups in benthic ecosystems:
implications for fisheries management. Ecol Indic 25:133–140
Ortiz M, Campos L, Berrios F, Rodríguez-Zaragoza FA, Hermosillo-Núñez BB, González J
(2013b) Network properties and keystoneness assessment in different intertidal communities
dominated by two ecosystem engineer species (SE Pacific coast): a comparative analysis. Ecol
Model 250:307–318
Ortiz M, Berrios F, Campos L, Uribe R, Ramírez A, Hermosillo-Nuñez B, González J, Rodríguez-­
Zaragoza FA (2015) Mass balanced trophic models and short–term dynamical simulations for
benthic ecological systems of Mejillones and Antofagasta bays (SE Pacific): comparative net-
work structure and assessment of human impacts. Ecol Model 309–310:153–162
Ortiz M, Hermosillo-Núñez BB, González J, Rodríguez-Zaragoza FA, Gómez I, Jordán F (2017)
Quantifying keystone species complexes: ecosystem-based conservation management in the
King George Island (Antarctic Peninsula). Ecol Indic 81:453–460
Pace ML, Cole JJ, Carpenter SR, Kitchell JF (1999) Trophic cascades revealed in diverse ecosys-
tems. TREE 14:483–488
Paine RT (1969) A note of tropic complexity and community stability. Am Nat 103:91–93
Patrício J, Marques JC (2006) Mass balanced models of the food web in three areas along a gra-
dient of eutrophication symptoms in the south arm of the Mondego estuary (Portugal). Ecol
Model 197:21–34
Patten BC (1997) Synthesis of chaos and sustainability in a nonstationary linear dynamic model of
the American black bear (Ursus americanus Pallas) in the Adirondack Mountains of New York.
Ecol Model 100:11–42
334 M. Ortiz et al.

Pauly D, Christensen V, Walters C (2000) Ecopath, Ecosim, and Ecospace as tools for evaluating
ecosystem impact of fisheries. ICES J Mar Sci 57(3):697–706
Payton I, Fenner M, Lee W (2002) Keystone species: the concept and its relevance for conservation
management in New Zealand. Sci Conserv 203:5–29
Pessoa M (2012) Algae and aquatic macrophytes responses to cope to ultraviolet radiation- a
review. Emir J Food Agric 24:527–545
Pikitch EK, Santora C, Babcock EA, Bakun A, Bonfil R, Conover DO, Dayton P, Doukakis P,
Fluharty D et al (2004) Ecosystem-based fishery management. Science 305:346–347
Pinkerton MH, Bradford-Grieve JM, Hanchet SM (2010) A balanced model of the food web of the
Ross Sea, Antarctica. CCAMLR Sci 17:1–31
Piraino S, Fanelli G, Boero F (2002) Variability of species roles in marine communities: changes
of paradigms for conservation priorities. Mar Biol 140:1067–1074
Plante C, Mayer L (1994) Distribution and efficiency of bacteriolysis in the gut of Arenicola
marina and three additional deposit feeders. Mar Ecol Prog Ser 109:183–194
Plante C, Shriver A (1998) Patterns of differential digestion of bacteria in deposit feeders: a test of
resource partitioning. Mar Ecol Prog Ser 163:253–258
Polovina J (1984) Model of a coral reef ecosystem I.  ECOPATH model and its application to
French Frigate Shoals. Coral Reefs 3:1–11
Power ME, Tilman D, Estes JA, Menge BA, Bond WJ, Mills LS, Daily G, Castilla JC, Lubchengo
J, Paine RT (1996) Challenges in the Quest for Keystones. Bioscience 46:609–620
Puccia C, Levins R (1991) Qualitative modelling in ecology: loop analysis, signed digraphs, and
time averaging. In: Fishwick PA, Luker PA (eds) Qualitative simulation modeling and analysis.
Springer, New York, pp 119–143
Richter A, Wuttke S, Zacher K (2008) Two years of in situ UV measurements at Dallmann
Laboratory/Jubany Station. In: Wiencke C, Ferreyra G, Abele D, Marenssi S (eds) The Antarctic
ecosystem of Potter Cove, King George Island. Synopsis of research performed 1999–2006 at
Dallmann Laboratory and Jubany Station. Reports on Polar and Marine Research, vol 571.
AWI, Alfred-Wegener-Institut für Polar- und Meeresforschung, Bremerhaven, pp 12–19
Ricker WE (1968) Food from the sea. In: Committee on Resources and Man (ed) Resource and
man. US National Academies of Sciences, EH Freeman, San Francisco, Chap 5, pp 87–108
Smith W, Ainley D, Cattaneo-Vietti R (2007) Trophic interactions within the Ross Sea continental
shelf ecosystem. Philos Trans R Soc Lond B Biol Sci 362:95–111
Stark P (1994) Climate warning in the central Antarctic Peninsula area. Weather 49:215–220
Stibor H, Vadstein O, Diehl S, Gelzleichter A, Hansen T, Hantzsche F, Katechakis A, Lippert B,
Løseth K, Peters C et al (2004) Copepods act as a switch between alternative trophic cascades
in marine pelagic food webs. Ecol Lett 7:321–325
Targett TE (1981) Trophic ecology and structure of coastal Antarctic fish communities. Mar Ecol
Prog Ser 4:243–263
Ulanowicz RE (ed) (1986) Growth and development: ecosystems phenomenology. Springer,
New York
Ulanowicz R (1997) Ecology, the ascendent perspective. Complexity in ecological systems series.
Columbia University Press, New York
Ulanowicz R, Puccia CJ (1990) Mixed trophic impacts in ecosystems. Coenoses 5:7–16
Valdivia N, Díaz M, Holtheuer J, Garrido I, Huovinen P, Gómez I (2014) Up, down, and all around:
scale-dependent spatial variation in rocky-shore communities of Fildes Peninsula, King George
Island, Antarctica. PLoS One 9(6):e100714. https://doi.org/10.1371/journal.pone.0100714
Valdivia N, Díaz MJ, Garrido I, Gómez I (2015) Consistent richness-biomass relationship across
environmental stress gradients in a marine macroalgal-dominated subtidal community on the
Western Antarctic Peninsula. PLoS One 10(9):e0138582
Valls A, Coll M, Christensen V, Ellison AM (2015) Keystone species: toward an operational con-
cept for marine biodiversity conservation. Ecol Monogr 85:29–47
Vasas V, Lancelot C, Rousseau V, Jordán F (2007) Eutrophication and overfishing in temperate
nearshore pelagic food webs: a network perspective. Mar Ecol Prog Ser 336:1–14
16  Trophic Networks and Ecosystem Functioning 335

Walters C, Christensen V (2007) Adding realism to foraging arena predictions of trophic flow
rates in Ecosim ecosystem models: shared foraging arenas and bout feeding. Ecol Model
209:342–350
Walters C, Christensen V, Pauly D (1997) Structuring dynamic models of exploited ecosystems
from trophic mass–balance assessments. Rev Fish Biol Fish 7:139–172
Warwick RM, Clarke KR (1993) Comparing the severity of disturbance: a meta-analysis of marine
macrobenthic community data. Mar Ecol Prog Ser 92:221–231
Wolff M (1994) A trophic model for Tongoy Bay – a system exposed to suspended scallop culture
(Northern Chile). J Exp Mar Biol Ecol 182:149–168
Wolff M, Koch V, Isaac V (2000) A trophic flow model of the Caeté mangrove estuary (North
Brazil) with considerations for the sustainable use of its resources. Estuar Coast Shelf Sci
50:789–803
Wootton JT (1994) Predicting direct and indirect effects: an integrated approach using experiments
and path analysis. Ecology 75:151–165
Yodzis P (2001) Must top predators be culled for the sake of fisheries? TREE 16:78–84
Part V
Chemical Ecology
Chapter 17
Chemical Mediation of Antarctic
Macroalga-Grazer Interactions

Charles D. Amsler, James B. McClintock, and Bill J. Baker

Abstract  Macroalgal forests along the western Antarctic Peninsula (WAP) support
dense assemblages of small macroalgal-associated invertebrates, particularly
amphipods but also others including gastropods. Most of the macroalgal species,
including all the larger, ecologically dominant brown macroalgae, elaborate chemi-
cal defenses against herbivory to amphipods as well as fish and sea stars.
Consequently, the vast majority of the macroalgal biomass in these forests is unpal-
atable to potential consumers. A great deal of progress has been made on under-
standing these relationships during the past decade. Although the macroalgae are
seldom consumed by the associated invertebrates and fish, many of the inverte-
brates, particularly the amphipods, benefit from associating with the chemically
defended macroalgae because omnivorous fish avoid feeding on them. The amphi-
pods benefit their macroalgal hosts by greatly reducing biofouling by diatoms and
other epiphytic algae. This chapter reviews progress in understanding the chemical
defenses of Antarctic macroalgae. It also reviews the community-wide mutualistic
interaction between macroalgae and its associated amphipods as well as recent stud-
ies examining the extent to which this mutualistic interaction also occurs with
macroalgal-­associated gastropods.

Keywords  Amphipods · Biofouling · Chemical ecology · Herbivory · Mutualism ·


Trophic dynamics

C. D. Amsler (*) · J. B. McClintock


University of Alabama at Birmingham, Birmingham, AL, USA
e-mail: amsler@uab.edu; mcclinto@uab.edu
B. J. Baker
University of South Florida, Tampa, FL, USA
e-mail: bjbaker@usf.edu

© Springer Nature Switzerland AG 2020 339


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_17
340 C. D. Amsler et al.

17.1  Introduction

Chemistry has many, often central, roles in important biotic interactions of organ-
isms ranging from bacteria to most if not all groups of eukaryotes (e.g., Eisner and
Meinwald 1995; Waters and Bassler 2005; Dicke and Takken 2006; Wyatt 2014).
Chemical signals mediate a wide variety of forms of communication both within
and between species (e.g., Waters and Bassler 2005; Baldwin et al. 2006; Müller-­
Schwarze 2006; Wyatt 2014), and chemical compounds serve as defenses that help
organisms resist predation by herbivores or carnivores (e.g., Müller-Schwarze 2006;
Rosenthal and Berenbaum 2012), defenses against competitors (e.g., Inderjit and
Mallik 2002), and defenses against pathogens or biofoulers (e.g., Lane and Kubanek
2008). Such relationships are widespread in the marine environment (McClintock
and Baker 2001; Breithaupt and Thiel 2011; Brönmark and Hansson 2012; Puglisi
and Becerro 2019).
In marine macroalgae, some studies have examined chemical mediation of inter-
and intraspecific sensory ecology, but much more attention has been focused on
chemical mediation of defensive interactions, particularly defenses against herbiv-
ory (Amsler and Fairhead 2006; Amsler 2008, 2012). In the case of Antarctic mac-
roalgae, while chemical roles in sensory ecology have not been ignored (Zamzow
et al. 2010; Bucolo et al. 2012), the vast majority of studies to date have focused on
the chemical mediation of macroalga-herbivore interactions (Amsler et  al. 2008,
2009a, 2014; Avila et al. 2008; Núñez-Pons and Avila 2014; von Salm et al. 2019).
Macroalgae are especially important in shallow water, hard bottom communities
along the northern and north-central regions of the western Antarctic Peninsula
(WAP) where they form undersea forests covering much of the benthos, often with
very high standing biomass (Wiencke and Amsler 2012; Wiencke et  al. 2014).
However, their dominance decreases markedly toward the southern half of the WAP
(DeLaca and Lipps 1976; C.D. Amsler, personal observations), and while they are
present at many sites throughout the rest of coastal Antarctica, in these other areas
they are typically not present at the high levels of biomass or percent cover observed
in the northern WAP (Wiencke and Amsler 2012; Wiencke et al. 2014; Clark et al.
2017). To date, almost all studies of the chemical mediation of macroalga-­
invertebrate interactions have been performed in macroalgal-dominated communi-
ties along the northern portion of the WAP. Consequently, this chapter focuses on
the northern WAP but includes what is known from other parts of coastal Antarctica.

17.2  Feeding Bioassay Methodology

One of the most basic questions concerning macroalgal interactions with sympatric
invertebrate or vertebrate animals is “do the animals consume the algae?” The stan-
dard methodology to determine if a macroalga or any other potential food item is
palatable to a potential consumer is, conceptually, quite simple and straightforward.
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 341

One offers the potential food to a sympatric and otherwise ecologically relevant
potential consumer and asks if it is consumed (Hay et al. 1998). Depending on the
size of the consumer relative to the food item, the feeding bioassay can simply
determine if food item is eaten vs. not eaten or can measure how much is consumed
per unit time, in either case comparing the result to a control food item that is known
to be palatable to the animal.
If a macroalga or other food item is consumed by an animal, depending on the
overall questions being posed, one might ask a variety of specific questions about it
including its nutrient content or digestibility or its relative palatability or nutritive
value relative to other potential foods. If the macroalga is not palatable to the ani-
mal, a relevant question becomes “why not?” There are at least three reasons a
potential food could be unpalatable to a consumer. It might be physically defended
against consumption by being mechanically too tough to eat or by possessing spines
or other physical impediments to consumption (Littler and Littler 1980; Hay 1997).
Although not as common, it is also possible for a potential food item to be of such
low nutritive value that it is not energetically worthwhile for many animals to con-
sume (Bullard and Hay 2002). As further discussed throughout this chapter, a very
common reason that Antarctic macroalgae are unpalatable to sympatric potential
consumers is because of the production of chemical defenses that make them
unpalatable.
To establish a chemical role in feeding deterrence of a macroalga (or any other
possible food), the potential defensive chemicals, which typically are secondary
metabolites (Maschek and Baker 2008), need to be extracted from the macroalga
and used in further feeding bioassays (Hay et al. 1998). The first steps involve sol-
vents likely to extract a wide range of compounds, and the resulting mixtures are
referred to as crude extracts. There are many assumptions underlying extract bioas-
say methods as well as many variations on the extraction and bioassay protocols,
and it is likely that none are “perfect” at reproducing what occurs in nature. Bringing
attention to the various bioassay methods and organisms that have been used in
examining macroalga-consumer interactions in Antarctica and the inherent strengths
and limitations of each is a sub-goal of this chapter. Ideally, the protocol would
produce two or more crude extracts likely to capture all possible lipophilic and
hydrophilic secondary metabolites or other compounds that have a potential to be
responsible for feeding deterrence in intact macroalgal thalli. The assay procedure
would then present them to an ecologically relevant potential consumer in an artifi-
cial food that is as similar as possible to the chemical and nutritive composition of
the original, intact macroalgal thallus.

17.3  Antarctic Macroalgal Resistance to Herbivory

Faced with predation, a macroalga may continue to invest resources in growth and
reproduction at similar levels to those without predation, which is often referred to
as tolerating herbivory (Núñez-Farfán et  al. 2007). Alternately they may invest
342 C. D. Amsler et al.

resources that would otherwise have been used for growth and/or reproduction into
resisting herbivores through the production of physical and/or chemical defenses
(Herms and Mattson 1992). These strategies are not always mutually exclusive
(Arnold and Targett 2003; Núñez-Farfán et al. 2007).
Studies of the palatability of fresh thallus material and/or crude chemical extracts
of WAP macroalgae have utilized amphipods, sea stars, and fish as “taste tester”
animals (Table 17.1). Of these, the amphipods and fish are probably the most eco-
logically relevant potential consumers. The rationale for utilizing sea stars in feed-
ing bioassays is discussed further below (Sect. 17.3.3). Overall, almost all of the
macroalgae that have been used in fresh thallus bioassays with amphipods and/or
fish have been unpalatable to at least one of the animal species (Table 17.1). The
exceptions to this are the small, filamentous brown algae Geminocarpus geminatus
and Elachista antarctica and the red algae Palmaria decipiens and Porphyra ploca-
miestris (Table  17.1). Clearly the vast majority of the enormous live macroalgal
standing biomass on the WAP is unpalatable to these potential and numerous
consumers.

17.3.1  Macroalgal Palatability and Resistance to Amphipods

Amphipods are by far the most numerous and the most speciose group of animals
that associate with macroalgae on the WAP (Richardson 1971, 1977; Huang et al.
2007; Aumack et al. 2011a) although smaller numbers of gastropods, isopods, cope-
pods, and ostracods are also found in association with WAP macroalgae (e.g.,
Amsler et  al. 2015; Schram et  al. 2016). Currently there are at least 564 known
Antarctic amphipod species (De Broyer and Jażdżewska 2014), and amphipod
abundances are particularly high on finely branched species. Estimated amphipod
densities range up to 308,000 and 32,000 individuals m−2 benthos in stands of the
ecologically dominant, overstory brown macroalgae Desmarestia menziesii and
Desmarestia anceps, respectively (Amsler et al. 2008).
Gondogeneia antarctica (Fig. 17.1a) is a member of family Pontogeneiidae and
one of the best-studied shallow water Antarctic amphipods overall (e.g., Obermüller
et al. 2007; Doyle et al. 2012; Kang et al. 2015). As one of the most common amphi-
pods which associate with macroalgae (Richardson 1971, 1977; Huang et al. 2007;
Aumack et al. 2011a; Barrera-Oro et al. 2019), it has been used extensively in feed-
ing bioassays with WAP macroalgae (Table 17.1) and also with some benthic mac-
roinvertebrates (Ma et al. 2009; Amsler et al. 2009c; authors’ unpublished data). In
nature, the diet of G. antarctica consists largely of diatoms, but it also includes
macroalgae, particularly filamentous algae and Palmaria decipiens, as an important
diet component (Aumack et al. 2017; Zenteno et al. 2019). In feeding bioassays,
G. antarctica readily consumes filamentous macroalgae; artificial foods using
freeze-dried, ground, filamentous macroalgae as a feeding stimulant; and P. decipi-
ens (Table 17.1). One disadvantage of the use of this amphipod in feeding bioassays
with macroalgal extracts in artificial foods is that in some instances, it is probably
Table 17.1  Summary of fresh WAP macroalgal thalli and crude extract bioassays with amphipods, sea stars, and fish
Amphipods Sea star Fish
Gondogeneia antarctica Cheirimedon femoratus Other species Odontaster validus Notothenia coriiceps
Macroalgal species Thal LpEx HpEx LpEx Thal LpEx HpEx Thal LpEx HpEx Thal LpEx HpEx
Class Phaeophyceae
Adenocystis utricularis a3 r3 a5 r3 r3 a8 a3 r3 a3 a3
1
Ascoseira mirabilis r a3 a3a r5 a3 a3 r8 a3 r3 a3 r6
Chordaria linearis r3
Cystosphaera jacquinotii a3 r3 r3 r3 r3 r3 r6
1 5 7 3
Desmarestia anceps r a3 r3 r r r3 a3,8 r3 r3 a r3,6
Desmarestia antarctica r1 a3 r3 r3 r3 r3 r3 a3
Desmarestia menziesii r3 a3 r5 r7 r3 r3 a8 a3 r3 r3 a3 r6
4
Elachista antarctica a a4
Geminocarpus geminatus a1 a3
Halopteris obovata r2 a2,3 a2,3 r2 a2 a2 a3 r3
Himantothallus grandifolius r1 r3 r3 r3 r3 r3 r3 a3 r3,6
Phaeurus antarcticus r3  r3 a3 r3
Class Rhodophyceae
Austropugetia crassab a3 r3 r3 a3 a3 r3 r3
Callophyllis atrosanguinea a3 a3 r3 a3 r3 r3 a3
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions

Curdiaea racovitzae r1 a3 a3 r3 r3 r3 r3
Cystoclonium obtusangulum r2 r2 r2 r2,7 r2 r2
Delisea pulchra r3 a3 r7 r3 r3 a3 r3
5
Georgiella confluens r a3 r8
Gigartina skottsbergii r1 r5 a3 a8 a3
3 3
Gymnogongrus antarcticus a a a3 r3
Gymnogongrus turquetii r1 r3 r3 r3 a3 r3 r3 r3 a3
Iridaea cordata r1 a3 r3 r3 a3 r3
343

(continued)
Table 17.1 (continued)
344

Amphipods Sea star Fish


Gondogeneia antarctica Cheirimedon femoratus Other species Odontaster validus Notothenia coriiceps
Macroalgal species Thal LpEx HpEx LpEx Thal LpEx HpEx Thal LpEx HpEx Thal LpEx HpEx
Myriogramme mangini r1 a3 a3 r3 r3 a3 r3
1 3 3 3
Myriogramme smithii r r a3 r r r3 r3 r3 a3
3
Nereoginkgo adiantifolia a
Palmaria decipiens a1,2,3 a2 a2 r5 r2,4 a4 a2 a2 a3 a3
Pantoneura plocamioides r2 r2 r2 r2,7 r2 a2 a3 r3 a3 a3
Paraglossum amsleri b a3 r3 r3 a3 a3
3 3
Paraglossum salicifolium b r3 r3 r a r3 r3 r3 a3
3
Phycodrys austrogeorgica a
Picconiella plumosa r2 r2 r2 r2 a7 r2 r2 a3 r2 a3
Plocamium cartilagineum r2,7 r2,3 r2,3 r2,7 a7 r2 a2 r3 r3 r3 r3 r3 a3
3 3
Plumariopsis peninsularis a3 r3 r r a3 r3 a3
3
Porphyra plocamiestris a a3
Trematocarpus antarcticus r1 a3 a3 r3 a3 a3 r3 a3 a3
Varimenia macropustulosa b r1 a3 a3 r3 a3 a3 r3
Class Ulvophyceae
Lambia antarctica a3 a3 r3 a3 a3 a3
Thal fresh thallus bioassays, LpEx lipophilic extract bioassays, HpEx hydrophilic extract bioassays, a thallus or extract in artificial food accepted in bioassay,
r thallus or extract in artificial food rejected in bioassay. Superscript numbers = references (see table endnote).
1
Amsler et al. (2009b), 2Aumack et al. (2010), 3Amsler et al. (2005), 4Bucolo et al. (2011), 5Núñez-Pons et al. (2012), 6Iken et al. (2009), 7Amsler et al. (2013),
8
Núñez-Pons and Avila (2014)
Prior names in Amsler et al. (2005): Austropugetia crassa as undescribed sp. 1, Paraglossum amsleri as Delesseria lancifolia, Paraglossum salicifolium as
Delesseria salicifolia, Varimenia macropustulosa as Pachymenia orbicularis in Amsler et  al. (2005) and undescribed species A in Amsler et  al. (2009b).
Pachymenia orbicularis was a misidentification in Amsler et al. (2005). The other changes resulted from taxonomic revisions.
a
Water-soluble, apparently activated defense inhibits feeding (McDowell et al. 2014b).
b
All or some bioassay data published under other species name (see table endnote).
C. D. Amsler et al.
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 345

Fig. 17.1  Common potential consumers of macroalgae on the western Antarctic Peninsula: (a)
Gondogeneia antarctica; (b) Paradexamine fissicauda; (c) Odontaster validus; (d) Notothenia
coriiceps. (Photos (a) and (b) by M.O. Amsler, (c) and (d) by B.J. Baker)

conservative in identifying crude extracts with feeding deterrence. In these feeding


bioassays, the extracts by themselves – presumably because of primary metabolites
therein – are commonly preferred over control foods at statistically significant lev-
els (Amsler et  al. 2005; Schram et  al. 2015). Indeed, when either lipophilic or
hydrophilic crude extracts of the filamentous green alga Cladophora repens, which
is the feeding stimulant usually used in the artificial foods, are added to the artificial
foods, they are significantly preferred to the C. repens-based artificial foods without
extracts (Amsler et  al. 2005). Consequently, if there is no significant difference
between feeding rates on control and extract-containing artificial foods in a bioas-
say, could the extracts actually have been deterrent, but not enough to overcome the
inherent extra palatability of the extracts in general? Halopteris obovata and
Varimenia macropustulosa are species that were unpalatable to G. antarctica as
fresh thallus but in which feeding on one or both crude extracts was not significantly
different from the control food (Amsler et al. 2005; Aumack et al. 2010). This was
also true of Ascoseira mirabilis (Amsler et al. 2005), but there is an additional pos-
sible explanation for this as discussed below. Fresh thallus feeding bioassays using
G. antarctica have not been done with Callophyllis atrosanguinea, but it is unpalat-
able as fresh thallus to sea stars and fish (Table  17.1). There was no significant
preference of G. antarctica for either of its crude extracts (Amsler et al. 2005) sug-
gesting that there might be chemically based deterrence to amphipod herbivory.
There is a similar situation with Lambia antarctica although it is only unpalatable
as fresh thallus to sea stars (not to fish; Table 17.1) and only its hydrophilic crude
extract was not significantly different from the control food in feeding bioassay with
G. antarctica (Amsler et al. 2005).
346 C. D. Amsler et al.

Cheirimedon femoratus is the second most common amphipod that has been
used in feeding bioassays of Antarctic macroalgal extracts (Table 17.1). It is a mem-
ber of family Lysianassidae, which primarily consists of detritivores and necrovores
(Bousfield 1973). Consistent with that, it is commonly found in baited fish traps and
probably prefers carrion as a food when available (Bregazzi 1972). However, it also
feeds on detrital algal material, particularly as juveniles and ovigerous females
(Bregazzi 1972). Based on stable isotope analysis, Zenteno et al. (2019) classified
this species as a scavenger and reported a particularly high isotope signature appar-
ently originating from intertidal macroalgal sources. Cheirimedon femoratus has
been reported to be most common on sandy bottoms but also associated with rocky
substrata (Bregazzi 1972). Núñez-Pons et al. (2012) report that it associates with
live macroalgae at Deception Island (63°S latitude). Our group has examined liter-
ally hundreds of thousands if not more macroalgal-associated amphipods from
Anvers Island (64°S latitude) as well as smaller numbers from as far south as 68°S
and never found C. femoratus – indeed only very few members of its family (Huang
et al. 2007) – associated with macroalgae. However, we have sometimes found it
from Anvers Is. (but not further south) in airlift samples that included substratum
material from macroalgal communities as well as occasionally in baited fish traps
(authors, unpublished) suggesting that it is present in the macroalgal communities,
probably associated with the substratum, but not on the upright portions of the mac-
roalgae. Richardson (1977) reported very small numbers of C. femoratus relative to
other species associated with the overstory brown alga D. anceps at Signy Island
(60°S latitude), but his samples included holdfasts which ours usually did not. At
King George Island (62°S latitude), C. femoratus has been reported in low numbers
from samples including both macroalgae and surrounding substratum but com-
monly in the guts of fish from these communities (Barrera-Oro et al. 2019).
Even though Cheirimedon femoratus does not associate with intact macroalgae
in many locations along the WAP and as such may not be a primary factor in select-
ing for chemical defenses in the algae, the palatability of detached and degrading
macroalgal material to this amphipod is likely to be relevant to the flow of carbon
into the surrounding communities. Macroalgae, probably as detritus, are an impor-
tant carbon source to a wide variety of animals in these communities (Dunton 2001;
Corbisier et al. 2004; Aumack et al. 2017; Braeckman et al. 2019; Zenteno et al.
2019). Once dead, chemically defended brown macroalgae can begin to become
palatable to G. antarctica and probably other amphipods after a period of weeks
(Reichardt and Dieckmann 1985; Amsler et al. 2012a). However, they can survive a
very long time after being detached (Brouwer 1996; Amsler et al. 2012a), and drift
thalli probably need to be buried in sediments and begin microbial decomposition
to become palatable to many animals (Braeckman et al. 2019).
Cheirimedon femoratus has only been used in feeding bioassays with lipophilic
macroalgal extracts, but the results of those bioassays are generally comparable to
lipophilic extract feeding bioassays using Gondogeneia antarctica (Table  17.1).
One difference is that feeding deterrence of Desmarestia anceps to G. antarctica
was only in the hydrophilic extract while a lipophilic extract was unpalatable to
C. femoratus. It is possible that this is attributable to different extraction procedures
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 347

in the two studies (Amsler et al. 2005; Núñez-Pons et al. 2012) which could have
resulted in the compounds responsible for the response partitioning differently. A
second difference is that extracts of A. mirabilis were unpalatable to C. femoratus
but not to G. antarctica (Table 17.1), but as noted above, the conservative nature of
the G. antarctica extract bioassay could have resulted in a false negative. A more
perplexing difference is that the extract of P. decipiens was unpalatable to C. femo-
ratus (Núñez-Pons et al. 2012) since in our experience it is one of the few generally
palatable macroalgae in the community. It is readily consumed as fresh thallus by
G. antarctica (Amsler et al. 2009b, 2013; Aumack et al. 2010; Bucolo et al. 2011),
by the amphipod Oradarea bidentata (Bucolo et al. 2011), and by sea stars and fish
(Table  17.1). It is not palatable as fresh thallus to the amphipods Prostebbingia
gracilis and Paraphimedia integricauda (Aumack et al. 2010; Bucolo et al. 2011),
but this is almost certainly because of biophysical limitations since P. gracilis pref-
erentially consumes extracts of P. decipiens (Aumack et al. 2010) and P. integri-
cauda, which consumes primarily diatoms, has a mandible morphology unsuited for
grinding macroalgal thalli (Aumack et al. 2017).
Although the vast majority of WAP macroalgae are unpalatable to amphipods,
usually because of chemical defenses (Table 17.1), there is one amphipod species
known to consume otherwise chemically defended macroalgae. Plocamium
“cartilagineum”1 is unpalatable as thallus to the amphipods Gondogeneia antarc-
tica, Prostebbingia gracilis, Paraphimedia integricauda, Oradarea bidentata, and
Schraderia gracilis but readily consumed by Paradexamine fissicauda (Fig. 17.1b)
(Amsler et al. 2013). The unpalatability to most species is clearly due to defensive
chemistry as its extracts are highly deterrent in feeding bioassays with G. antarctica
and P. gracilis (Amsler et al. 2005; Aumack et al. 2010) as well as with sea stars and
fish (Table 17.1). Paradexamine fissicauda is also able to consume Picconiella plu-
mosa (Amsler et al. 2013) which is unpalatable both as fresh thallus and as crude
extracts to G. antarctica and P. gracilis (Aumack et  al. 2010). However, P. fissi-
cauda does not consume the chemically defended red algae Pantoneura plocamioi-
des, Cystoclonium obtusangulum, or Delisea pulchra or the chemically defended
brown algae Desmarestia menziesii or Desmarestia anceps (Amsler et  al. 2013).
Predators that are able to consume chemically defended organisms including mac-
roalgae are generally only able to do so in single species or in closely related taxa
with similar or identical defensive metabolites because the biochemical pathways
they use to detoxify and/or excrete the defensive metabolites are relatively com-
pound specific (Sotka and Whalen 2008; Sotka et al. 2009). Plocamium “cartilag-
ineum” and P. plumosa are in separate taxonomic orders (Plocamiales and
Ceramiales, respectively; Hommersand et al. 2009) which are not closely related
and which were recently estimated to have been divergent for close to 400 million
years (Yang et al. 2016). Although separate or convergent evolution of secondary
metabolites is not common, it does occur (e.g., Daly 2004; Cutignano et al. 2012),

1
 This entity is genetically distinct from P. cartilagineum found elsewhere and certainly represents
a separate species, but to date, no formal alternative name has been proposed (Hommersand
et al. 2009).
348 C. D. Amsler et al.

and it is possible that defensive metabolites of P. “cartilagineum” and P. plumosa


are similar enough to allow both to be eaten by P. fissicauda. Paradexamine fissi-
cauda does not simply detoxify the halogenated secondary metabolite defenses of
P. “cartilagineum” but also incorporates at least some of them into its tissues where
they appear to deter predation on the amphipods by fish (Amsler et  al. 2013).
Although such sequestration of chemical defenses from the diet for use by the con-
sumer as a defense against its own predators is known in terrestrial arthropods
(Eisner and Meinwald 1995) and in marine systems from some opisthobranch mol-
luscs (reviewed by Jormalainen and Honkanen 2008; Cimino and Ghiselin 2009;
Hay 2009), we are not aware of other examples of this phenomenon in marine
organisms.
While chemical defenses against amphipod herbivory are clearly very common
in WAP macroalgae, we know specific compounds involved in that unpalatability in
only a few cases. Purified phlorotannins from both Desmarestia anceps and
Cystosphaera jacquinotii (but not several other brown algae) are deterrent in feed-
ing bioassays with Gondogeneia antarctica (Iken et al. 2009) which is consistent
with the deterrence observed in the hydrophilic crude extracts of these species
(Table 17.1). Two halogenated monoterpenes, anverene and epi-plocamene D, from
Plocamium “cartilagineum” have been shown to deter feeding by G. antarctica
(Ankisetty et  al. 2004). Antarctic P. “cartilagineum” produces a large number of
halogenated secondary metabolites (Young et al. 2013), and it is likely that addi-
tional compounds contribute to the strong unpalatability of this species in most
feeding bioassays. Similarly, it is possible that more than phlorotannins are respon-
sible for feeding deterrence in D. anceps or C. jacquinotii, particularly if differential
extraction is responsible for the differences discussed above between responses of
Cheirimedon femoratus and G. antarctica to crude extracts of D. anceps.
Secondary metabolites are not the only molecules that can be involved in chemi-
cal defense. For example, reactive oxygen species (ROS) are known to be involved
in defense against pathogens in a variety of macroalgae (Potin 2008; Thomas et al.
2014). Recent evidence suggests that this is probably true in Antarctic macroalgae
and also that ROS has the potential to be involved in deterring predation by small
grazers such as amphipods. Nine of 13 WAP macroalgae accumulated ROS within
their tissues within 70  min of wounding, and four of five species tested released
strong oxidants into the surrounding seawater within 1 min of wounding (McDowell
et al. 2014a). Palmaria decipiens, the only macroalgal species tested that is eaten by
Gondogeneia antarctica, increased ROS production at the site of grazing by G. ant-
arctica (McDowell et al. 2014a). Although P. decipiens is consumed as fresh thallus
by G. antarctica, as already noted, Ascoseira mirabilis is not (Table 17.1). Ascoseira
mirabilis produces ROS upon wounding in the light but not in the dark (McDowell
et al. 2016), and wounded A. mirabilis inhibits feeding of G. antarctica on P. decipi-
ens disks within the same experimental container (McDowell et al. 2014b). Although
hydrogen peroxide does not appear to be part of the ROS response in A. mirabilis,
it is in other Antarctic macroalgae (McDowell et al. 2014a), and addition of hydro-
gen peroxide at concentrations comparable to those released by macroalgae also
inhibits feeding by G. antarctica (McDowell et al. 2014b). ROS production in the
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 349

light has been shown to be dependent on photosynthetic electron transfer in a tem-


perate kelp species, and this is probably true in P. decipiens and A. mirabilis as well
(McDowell et al. 2015). The light dependence of this potential defense against her-
bivores likely has ecological consequences as amphipods are active and probably
feeding at night (authors’ personal observations). This includes species such as
G. antarctica moving off of chemically defended hosts to the palatable species
P. decipiens at night, presumably to feed (Aumack et al. 2011a).

17.3.2  Macroalgal Palatability and Resistance to Fish

The species of fish most commonly observed in WAP macroalgal forests, Notothenia
coriiceps (Fig. 17.1d), commonly has macroalgae in its guts when collected from
macroalgal communities (e.g., Iken et al. 1997; Zamzow et al. 2011) and apparently
selectively feeds on some macroalgal species including Palmaria decipiens (Iken
et al. 1997). In feeding bioassays with fresh thallus material, P. decipiens is one of
only four of 28 species tested that was consumed by N. coriiceps (Table 17.1). We
have observed N. coriiceps taking bites out of P. decipiens in laboratory aquaria and
observed P. decipiens individuals in nature with similar, apparent bites missing from
the thallus (authors, unpublished). Of the other three subtidal WAP macroalgal con-
sumed by the fish as fresh thallus, Porphyra plocamiestris is in a genus that is com-
monly palatable to animals, including humans, and is very thin and easily torn. It is
not surprising that it is palatable to the fish. Lambia antarctica is palatable as extract
to both amphipods and sea stars (Table  17.1) indicating no apparent chemical
defenses although it was rejected as thallus by sea stars (Table 17.1). Lambia ant-
arctica is not as physically tough as P. decipiens (Amsler et al. 2005) suggesting
that N. coriiceps should be able to bite off pieces of its thallus as it does with
P. decipiens. The fourth species the fish consumed as thallus, Gigartina skottsbergii
(Table  17.1), is different. It was the physically toughest of 30 species tested for
puncture resistance by Amsler et al. (2005). Although the fish swallowed the small
pieces used in the feeding bioassays, we believe that it is very unlikely that they
could remove pieces of thallus from an intact G. skottsbergii blade in nature.
Fewer extract feeding bioassays have been performed with Notothenia coriiceps
than with the amphipod Gondogeneia antarctica or sea star Odontaster validus
(Table 17.1). Of the 24 macroalgal species known to be unpalatable to N. coriiceps,
extract bioassays have not been performed with seven. Of the other 17 species,
extract bioassays indicate that chemical defenses can explain the unpalatability in
10. In four others, only one crude extract type has been bioassayed (Table 17.1), so
it is possible that defensive compounds could have been present in the other extract.
Only three species that were rejected as thallus have had both crude extract types
tested and shown to be palatable to the fish including Trematocarpus antarcticus
which similarly is unpalatable as thallus but not as either crude extract to amphipods
and sea stars (Table 17.1).
350 C. D. Amsler et al.

The only purified compounds that have been shown to deter feeding of Notothenia
coriiceps to date are phlorotannins. The strongest deterrent response was seen with
phlorotannins from Desmarestia menziesii, but pure phlorotannins from
Cystosphaera jacquinotii were also significantly rejected by the fish (Iken et  al.
2009). It is important to note that these assays were done with isolated phlorotan-
nins, not simply methanolic extracts that would be enriched in phlorotannins and
are commonly what is used in “phlorotannin” feeding bioassays even though they
also contain other compounds (Amsler and Fairhead 2006). Such methanolic
extracts of Desmarestia anceps and Ascoseira mirabilis deterred N. coriiceps, but
the deterrence was lost when the extracts were purified further with other solvents
and a microcellulose column (Iken et al. 2009).

17.3.3  Macroalgal Palatability and Resistance to Sea Stars

The sea star Odontaster validus (Fig. 17.1c) is one of the most common benthic
invertebrates throughout much of coastal Antarctica (Janosik and Halanych 2010)
and is typically present in WAP macroalgal communities (Dearborn and Fell 1974).
Although sea stars are usually not considered as macroalgal predators, O. validus
was one of two species out of 20 WAP sea stars surveyed that had macroalgae in its
guts (McClintock 1994). Although rare, we have personally observed O. validus
individuals with their cardiac stomachs extended over macroalgal thalli with epi-
phytes, but it seems most probable that macroalgae reported within their guts came
from detrital fragments. So although it appears unlikely that sea star predation
would have been important in selecting for the production of chemical defenses, the
fact that they do not bite into their prey as amphipods and fish do could mean that
thallus toughness is not as important a factor in their food choices. Fresh thallus
bioassays with them, therefore, could provide useful biophysical information rela-
tive to feeding deterrence observed in other predators. The feeding bioassay system
our group uses with O. validus (Amsler et al. 2005) is relatively quick and, for crude
extract bioassays, can use artificial foods leftover from preparation of foods for
amphipod or fish bioassays, providing comparative information on chemical deter-
rence in a separate predator.
To date, more fresh thallus bioassays with different macroalgal species (35) have
been done with Odontaster validus than with either amphipods or fish (Table 17.1).
Of those 35 species, fresh thallus was consumed by O. validus in 13 of the species,
a much higher acceptance rate than in fresh thallus assays with amphipods or fish
(Table 17.1). Five of the 13 species accepted by the sea star were rejected as thallus
by one or more of the amphipod species or fish and in several of these instances,
crude extract bioassays with one or more of the other species did not explain the
unpalatability (Table 17.1). The O. validus assay results in these instances would be
consistent with the thallus unpalatability to amphipods or fish being biophysi-
cally based.
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 351

The two studies that have examined the palatability of lipophilic macroalgal
extracts with Odontaster validus (Amsler et al. 2005; Núñez-Pons and Avila 2014)
used not only different extraction solvent systems, as discussed previously (Sect.
17.3.1) for studies on amphipods by the two different research groups, but also
markedly different bioassay procedures. It is probably not surprising that results
differed in three of the four macroalgal species assayed in both studies (Table 17.1).
One other difference between the two studies was in the composition and extract
loading of the artificial foods. Amsler et al. (2005) coated extracts onto freeze-dried
powder of a palatable, intertidal, filamentous Antarctic green alga (Cladophora
repens) that was then incorporated into alginate-based foods. The “natural concen-
tration” of the extract was calculated on a volumetric basis relative to the volume of
the final artificial food pellet. The pellets were 5% dry C. repens powder in 2%
alginate which based on Peters et al. (2005) should have contained approximately
2% protein. That protein content is within but at the lower end of the range observed
for 36 species of WAP macroalgae (Peters et  al. 2005). Núñez-Pons and Avila
(2014) used artificial foods composed of cubes of non-Antarctic shrimp which were
12.4% protein, a level at the very top of the range of protein contents in WAP mac-
roalgae (Peters et al. 2005). The lipophilic extracts were added on a dry weight basis
to the outside of the shrimp cubes, and while some would have soaked in, most
would have been retained near the surface which is the only place “tasted” by a
nonbiting sea star predator. So while the foods themselves were higher in nutritional
value – and, therefore, potentially more attractive foods – compared to those used
by Amsler et  al. (2005), the potentially unpalatable extracts added to them were
probably presented to the sea stars at higher concentrations. Another difference
between the studies is that the Amsler et al. (2005) assay placed food items onto the
sea stars’ chemosensory tube feet and recorded whether or not the food that had
been “tasted” was moved to and retained at the mouth. Núñez-Pons and Avila (2014)
placed the sea stars on top of the food cubes and recorded whether they were con-
sumed in 24 hours, which captures the full complement of feeding behaviors (“tast-
ing” followed by consumption) and is probably a more ecologically robust bioassay
approach.
Purified phlorotannins from Desmarestia menziesii and Himantothallus grandi-
folius but not from five other brown macroalgal species significantly deter feeding
by Odontaster validus (Iken et  al. 2009). Desmarestia menziesii also produces a
quinone derivative, menzoquinone, which deters O. validus feeding in bioassays
(Ankisetty et al. 2004).

17.3.4  M
 acroalgal Palatability and Resistance to Sea Urchins
in McMurdo Sound

As discussed in Sect. 17.1, macroalgae occur throughout coastal Antarctica although


their biomass, percent cover of the bottom, and species richness all become lower
south of the northern WAP (Wiencke and Amsler 2012; Wiencke et al. 2014). Only
352 C. D. Amsler et al.

two species of fleshy macroalgae, Phyllophora antarctica and Iridaea cordata,


occur throughout most of the year at the southernmost extent of macroalgal distribu-
tion (and seasonally open water) in McMurdo Sound (Ross Sea; Miller and Pearse
1991). The sea urchin Sterechinus neumayeri is a very common and important con-
sumer in McMurdo Sound as well as throughout much of the rest of Antarctica
(Pearse and Giese 1966; Pawson 1969; Saucède et al. 2014) and an obvious poten-
tial predator on P. antarctica and I. cordata in McMurdo Sound. Unfortunately,
S. neumayeri does not reliably or reproducibly consume anything we have tested it
with in laboratory aquaria, and so typical feeding bioassays are not practical with it
because of the high variability caused by many animals not eating at all. To get
around this limitation, Amsler et al. (1998) developed a “phagostimulation” bioas-
say which measured how long a small disk of algal thallus or crude extracts on
feeding-stimulant-containing filter paper disks were retained at the urchins’ mouths
when forcibly placed there. While far from an ideal assay, the responses were quite
strong which allowed for statistically robust outcomes. Thallus disks from both spe-
cies were retained at the mouth less than half as long as a paper disk without feeding
stimulants. Both lipophilic and hydrophilic extracts added to filter paper disks with
a feeding stimulant were rapidly rejected, while solvent controls were retained at
the mouth for several-fold longer times, indicating that the unpalatability of fresh
thalli is chemically based (Amsler et al. 1998).
Even though Sterechinus neumayeri does not consume Phyllophora antarctica
or Iridaea cordata, it prefers to cover itself with them over other potential cover
objects if the algae are present (Amsler et al. 1999). Either the macroalgae or other
objects used as cover help protect S. neumayeri from predatory sea anemones by
acting as a detachable “shield” that the anemones’ tentacles stick to, allowing the
urchins to release the cover material and escape (Dayton et al. 1970; Amsler et al.
1999). The algae also benefit from this relationship because they are the preferred
cover items for the urchins. Anchor ice tears macroalgal thalli from the rock substra-
tum, and most of the algal biomass in the communities are held by urchins rather
than being attached to the rock substratum (Dayton et al. 1969; Miller and Pearse
1991; Amsler et  al. 1999). The algae so retained within the photic zone remain
physiologically active (Schwarz et al. 2003) as well as reproductive (Amsler et al.
1999) and probably are responsible for most of the spore and gamete production
which maintains the populations.

17.4  M
 acroalga-Invertebrate Interactions on the Western
Antarctic Peninsula

As discussed in Sect 17.3.1, amphipods are the most abundant invertebrates associ-
ated with WAP macroalgae. As also discussed in Sect. 17.3.1, the vast majority of
macroalgal species deter predation by amphipods using chemical defenses. This
includes all of the large, ecologically dominant brown macroalgae as well as most
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 353

of the common red macroalgae. Consequently, the vast majority of the WAP mac-
roalgal biomass is not available or, at least, not at all preferable, as food to the dense
amphipod assemblage. Although the amphipods are not consuming most of the
large macroalgae, they do consume epiphytic microalgae, filamentous algal epi-
phytes, and emergent filaments from endophytic algae (Amsler et al. 2009b, 2012b;
Aumack et al. 2011b, 2017) and thereby clearly benefit their macroalgal hosts by
sustaining photosynthesis, gas exchange, and nutrient uptake capacities. In addition
to consuming the smaller algae, the amphipods also benefit from associating with
the chemically defended macroalgae by gaining a refuge from an important preda-
tor, the common benthic fish Notothenia coriiceps (Zamzow et  al. 2010, 2011).
Because macroalgae are the major structural component and primary producers in
these communities and because amphipods are the most abundant invertebrates, we
have previously described this association as a community-wide mutualism (Amsler
et  al. 2014). This relationship is presented diagrammatically in the upper half of
Fig. 17.2. None of the individual components of this mutualism are unique to the
WAP, and comparisons and contrasts of the WAP communities to specific lower-­
latitude communities are discussed at length by Amsler et al. (2014). However, the
overall importance of these two groups of organisms – the macroalgae which domi-
nate in terms of structure and primary productivity and the amphipods which are by
far the most abundant animals – to these nearshore communities makes the interac-
tion uniquely important compared to such interactions in lower-latitude communi-
ties (Amsler et al. 2014).
In addition to the “bottom-up” benefits to amphipods from gaining an associa-
tional refuge from fish predation, it is possible that the very high densities of amphi-
pods associated with macroalgae result in part from “top-down” influences of
predators on fish that prey on amphipods. Notothenia coriiceps is a benthic ambush
predator with relatively small home ranges (e.g., Daniels 1982; North 1996;
Campbell et al. 2008). Although it is common along the bottom beneath the mac-
roalgal canopy, it is rarely observed by divers up off the bottom within the canopy
(Casaux et al. 1990; authors’ personal observations). The smaller Harpagifer ant-
arcticus is found associated with small rocks in Antarctic macroalgal forests and is
also an amphipod predator with a cryptic lifestyle (Daniels 1983; Casaux 1998).
Several seal species including leopard seals have been observed with fish in their
gut contents or scat (Hall-Aspland and Rogers 2004; Casaux et al. 2009), and mem-
bers of our research group have observed and photographed leopard seals eating
N. coriiceps in shallow water (A. Shilling, personal communication). It is reason-
able to hypothesize that the cryptic behavior of the fish represents a nonconsumptive
effect of seal predators and thereby reduces predation on amphipods and other
macroalgal-­associated invertebrates.
Another feature of these communities that is at least somewhat unique is that
free-living filamentous algae are very uncommon throughout most of the year in the
subtidal zone but filamentous algae are very common growing as endophytes within
subtidal macroalgae (Peters 2003; Amsler et al. 2009b). Filamentous macroalgae
are frequently observed in the WAP intertidal (e.g., Lamb and Zimmerman 1977;
Marcías et  al. 2017; Valdivia et  al. 2019), indicating that there is nothing about
354 C. D. Amsler et al.

Fig. 17.2 Schematic
representation of species
interactions between algae
and amphipods on the
western Antarctic
Peninsula. See text for
details. + indicates a
positive effect, 0 indicates
no effect, and − indicates a
negative effect. (From
Amsler et al. (2014). Used
with permission)

Antarctica that excludes free-living filamentous algae. Peters (2003) noted that
escape from herbivory has long been hypothesized as a selective factor favoring
endophytism (e.g., Kylin 1937) and postulated that the dense assemblages of amphi-
pods on subtidal WAP macroalgae probably selected for this pattern of filamentous
algae rarely being free-living but commonly being endophytic in the WAP subtidal.
This is potentially problematic with respect to the idea of WAP macroalgae and
amphipods being mutualists. While not always true (e.g., Gauna et al. 2009), fila-
mentous algal endophytes are commonly pathogenic to their macroalgal hosts (e.g.,
Apt 1988; Correa and Sánchez 1996; Faugeron et al. 2000). If such pathogenicity is
common in WAP endophytes and endophytism is driven by amphipod herbivory,
could amphipods truly be viewed as beneficial to the larger macroalgae? Schoenrock
et  al. (2013) performed field experiments which showed that while endophytes
could reduce growth and/or survival of some species of macrophytes when present
in very high densities within the macroalgal thalli, this was not always true. In the
most heavily impacted species in terms of growth, Iridaea cordata, there was no
corresponding correlation between endophyte load and the density of either carpo-
sporangia or tetrasporangia, indicating no pathogenic impact on fecundity
(Schoenrock et al. 2015b). In laboratory experiments on nine macroalgal species,
almost no pathogenic impacts were seen on photosynthetic physiology, thallus
toughness, or susceptibility to grazers, and the few significant effects observed were
very mild (Schoenrock et al. 2015a). Consequently, although endophytes can some-
times be detrimental to their larger macroalgal hosts, overall, this almost certainly
does not negate the benefit macroalgae gain from having the dense assemblage of
associated amphipods (Fig. 17.2).
Although often not as abundant as amphipods, small gastropods are also com-
monly observed associated with WAP macroalgae (Richardson 1977; Picken 1979,
1980; Iken 1999; Amsler et al. 2015). Gastropods commonly consume microalgae
and filamentous macroalgae (Purchon 1977; Santhanam 2018) and consequently
could be analogous to amphipods in providing benefits to their macroalgal hosts by
controlling epiphytic algae (see also Chaps. 12 and 13). They may be particularly
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 355

important on the upper surfaces of large, blade-forming macroalgae such as


Himantothallus grandifolius and Gigartina skottsbergii where amphipods are not as
common as gastropods, at least during daylight hours (authors’ personal observa-
tions). Amsler et al. (2019) performed a mesocosm experiment which maintained
H. grandifolius blade sections either with no gastropods or with densities of gastro-
pods corresponding to natural gastropod densities on H. grandifolius. A very similar
mesocosm experiment had previously been done with or without natural densities of
amphipods by Aumack et al. (2011b). The gastropods controlled epiphytic diatom
densities on H. grandifolius blades (Amsler et al. 2019) but not to the extent that
amphipods were observed to do so for H. grandifolius (Aumack et al. 2011b). No
grazing marks from the gastropods were ever observed on the H. grandifolius thalli
either in the mesocosms or in an accompanying laboratory experiment, but almost
all of the gastropod species have radulae that are probably not able to rasp the tough
thalli of H. grandifolius (Amsler et al. 2019). That is unlikely to be the case for the
limpet Nacella concinna which was present in the experiment and has a very tough,
minerally hardened, chitinized radula. This limpet is presumably deterred from con-
suming its H. grandifolius hosts by the same chemical defenses observed to deter
feeding in other sympatric predators (Table 17.1).
The fish Notothenia coriiceps often consumes gastropods in addition to amphi-
pods in macroalgal communities (Zamzow et al. 2011). Although small gastropods
on chemically defended, branched macroalgae probably gain refuge from fish pred-
ators just as amphipods do, subjectively it is likely that a gastropod on a large bladed
macroalga such as Himantothallus grandifolius is just as available to a predatory
fish as it would be on bare rock. However, sea stars are also important predators of
Antarctic gastropods (McClintock 1994), and Odontaster validus is chemically
deterred from consuming H. grandifolius (Table 17.1). Because sea stars use their
tube feet both for locomotion and for chemically sensing their prey (Hennebert et al.
2013), Amsler et al. (2019) hypothesized that O. validus might be less likely to be
on H. grandifolius blades than on other possible substrata, giving gastropods on the
macroalga somewhat of an associational refuge from predation. This hypothesis
might be true, but was not supported in the one simple experiment performed
(Amsler et  al. 2019). Overall, compared to the macroalga-amphipod mutualism
illustrated in Fig. 17.2, there are similarities in macroalga-gastropod interactions but
they are not identical.

17.5  Overview

A very high percentage of Antarctic macroalgal species deter potential herbivores


by elaborating chemical defenses. We are not aware of examples elsewhere in the
world where such a high percentage of the macroalgal flora is chemically defended.
In North Carolina (USA), few palatable macroalgal species are apparent in the sum-
mer, but this is because grazing fish remove palatable species that are apparent in
the winter and spring (Hay 1986; Hay and Sutherland 1988; Duffy and Hay 1994).
356 C. D. Amsler et al.

Most of the Antarctic macroalgal flora are composed of perennial species that are
present throughout all seasons (Wiencke and Clayton 2002; Wiencke et al. 2014)
(see Chaps. 1, 11, 12). This high prevalence of chemical defenses against herbivory
appears to be a fairly unique feature of Antarctic macroalgal forests. Chemical
defenses against carnivory are also very common in Antarctic invertebrates (e.g.,
Amsler et al. 2001; Avila et al. 2008; McClintock et al. 2010; Moles et al. 2015),
making Antarctica a powerful natural laboratory for the study of chemically medi-
ated predator-prey relationships.
Macroalgal forests along the WAP are similar to temperate kelp forests in being
dominated by large, perennial brown macroalgae (Wiencke and Amsler 2012). All
of the overstory brown algal species that dominate these WAP forests are chemi-
cally defended from herbivory. Consequently, while being similar to kelp forests in
some ways, their trophic dynamics differ greatly. Temperate kelps are usually palat-
able to abundant kelp forest consumers such as sea urchins, and the macroalgal
biomass and numerical dominance are usually maintained by “top-down” factors
controlling the herbivore populations (Steneck et al. 2002). Although top-down fac-
tors may have a role in allowing amphipods to be so abundant in WAP forests, as
discussed above in the Sect. 17.4, the general unpalatability to herbivores of the
Antarctic macroalgae suggests that their persistence and community dominance
would exist even if only through this “bottom-up” factor.
As also discussed in Sect. 17.5, while the hypothesis of a community-wide mutu-
alism between amphipods and macroalgae is well supported, there are similarities
but also important differences in the interactions of macroalgal-associated gastro-
pods and their hosts compared to the macroalga-amphipod relationship. While per-
haps not as abundant, at least in terms of biomass, other crustaceans such as
copepods, isopods, and ostracods also associate with canopy-forming macroalgae
(Schram et  al. 2016), and polychaetes commonly associate with holdfasts of the
large macroalgae (Pabis and Sicinski 2010). The extent to which the relationships of
these other macroalgal-associated invertebrates and their hosts are similar to or dif-
ferent from the macroalga-amphipod mutualism illustrated in Fig. 17.2 has yet to be
examined.

Acknowledgments  We are grateful to all our Antarctic field team members over the years, in
particular to M. Amsler who also provided constructive comments on an earlier version of this
chapter. The original version of this chapter also benefited from the constructive comments of
Martin Thiel and an anonymous reviewer. Our Antarctic chemical ecology research has been
­generously supported by numerous awards from the NSF Antarctic Organisms and Ecosystem
program, most recently by PLR-1341333 (CDA, JBM) and PLR-1341339 (BJB).

References

Amsler CD (2008) Algal chemical ecology. Springer-Verlag, Berlin, p xviii. 313


Amsler CD (2012) Chemical ecology of seaweeds. In: Wiencke C, Bischof K (eds) Seaweed
biology: novel insights into ecophysiology, ecology and utilization. Springer-Verlag, Berlin,
pp 177–188
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 357

Amsler CD, Fairhead VA (2006) Defensive and sensory chemical ecology of brown algae. Adv
Bot Res 43:1–91
Amsler CD, McClintock JB, Baker BJ (1998) Chemical defense against herbivory in the Antarctic
marine macroalgae Iridaea cordata and Phyllophora antarctica (Rhodophyceae). J Phycol
34:53–59
Amsler CD, McClintock JB, Baker BJ (1999) An antarctic feeding triangle: defensive interactions
between macroalgae, sea urchins, and sea anemones. Mar Ecol Prog Ser 183:105–114
Amsler CD, Iken KB, McClintock JB, Baker BJ (2001) Secondary metabolites from Antarctic
marine organisms and their ecological implications. In: McClintock JB, Baker BJ (eds) Marine
chemical ecology. CRC, Boca Raton, pp 267–300
Amsler CD, Iken K, McClintock JB, Amsler MO, Peters KJ, Hubbard JM, Furrow FB, Baker BJ
(2005) Comprehensive evaluation of the palatability and chemical defenses of subtidal mac-
roalgae from the Antarctic Peninsula. Mar Ecol Prog Ser 294:141–159
Amsler CD, McClintock JB, Baker BJ (2008) Macroalgal chemical defenses in polar marine com-
munities. In: Amsler CD (ed) Algal chemical ecology. Springer-Verlag, Berlin, pp 91–103
Amsler CD, Iken K, McClintock JB, Baker BJ (2009a) Defenses of polar macroalgae against her-
bivores and biofoulers. Bot Mar 52:535–545
Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009b) Filamentous algal endophytes in mac-
rophytic Antarctic algae: prevalence in hosts and palatability to mesoherbivores. Phycologia
48:324–334
Amsler MO, McClintock JB, Amsler CD, Angus RA, Baker BJ (2009c) An evaluation of sponge-­
associated amphipods from the Antarctic Peninsula. Antarct Sci 21:579–589
Amsler CD, McClintock JB, Baker BJ (2012a) Palatability of living and dead detached Antarctic
macroalgae to consumers. Antarct Sci 24:589–590
Amsler CD, McClintock JB, Baker BJ (2012b) Amphipods exclude filamentous algae from the
Western Antarctic Peninsula benthos: experimental evidence. Polar Biol 35:171–177
Amsler MO, Amsler CD, von Salm JL, Aumack CF, McClintock JB, Young RM, Baker BJ (2013)
Tolerance and sequestration of macroalgal chemical defenses by an Antarctic amphipod: a
‘cheater’ among mutualists. Mar Ecol Prog Ser 490:79–90
Amsler CD, McClintock JB, Baker BJ (2014) Chemical mediation of mutualistic interactions
between macroalgae and mesograzers structure unique coastal communities along the western
Antarctic Peninsula. J Phycol 50:1–10
Amsler MO, Huang YM, Engl W, McClintock JB, Amsler CD (2015) Abundance and diver-
sity of gastropods associated with dominant subtidal macroalgae from the western Antarctic
Peninsula. Polar Biol 38:1171–1181
Amsler CD, Amsler MO, Curtis MD, McClintock JB, Baker BJ (2019) Impacts of gastropods
on epiphytic microalgae on the brown macroalga Himantothallus grandifolius. Antarct Sci
31:89–97
Ankisetty S, Nandiraju S, Win H, Park YC, Amsler CD, McClintock JB, Baker JA, Diyabalanage
TK, Pasaribu A, Singh MP, Maiese WM, Walsh RD, Zaworotko MJ, Baker BJ (2004) Chemical
investigation of predator-deterred macroalgae from the Antarctic Peninsula. J Nat Prod
67:1295–1302
Apt KE (1988) Etiology and development of hyperplasia induced by Streblonema sp. (Phaeophyta)
on members of the Laminariales (Phaeophyta). J Phycol 24:28–34
Arnold TM, Targett NM (2003) To grow and defend: lack of tradeoffs for brown algal phlorotan-
nins. Oikos 100:406–408
Aumack CF, Amsler CD, McClintock JB, Baker BJ (2010) Chemically mediated resistance to
mesoherbivory in finely branched macroalgae along the western Antarctic Peninsula. Eur J
Phycol 45:19–26
Aumack CF, Amsler CD, McClintock JB, Baker BJ (2011a) Changes in amphipod densities among
macroalgal habitats in day versus night collections along the Western Antarctic Peninsula. Mar
Biol 158:1879–1885
Aumack CF, Amsler CD, McClintock JB, Baker BJ (2011b) Impacts of mesograzers on epiphyte
and endophyte growth associated with chemically defended macroalgae from the western
Antarctic Peninsula: a mesocosm experiment. J Phycol 47:36–41
358 C. D. Amsler et al.

Aumack CF, Lowe AT, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2017) Gut content,
fatty acid, and stable isotope analyses reveal dietary sources of macroalgal-associated amphi-
pods along the western Antarctic Peninsula. Polar Biol 40:1371–1384
Avila C, Taboada S, Núñez-Pons L (2008) Antarctic marine chemical ecology: what is next? Mar
Ecol 29:1–71
Baldwin IT, Halitschke R, Paschold A, von Dahl CC, Preston CA (2006) Volatile signaling in
plant-plant interactions: “Talking trees” in the genomics era. Science 311:812–815
Barrera-Oro E, Moreira E, Seefeldt MA, Valli Francione M, Quartino ML (2019) The impor-
tance of macroalgae and associated amphipods in the selective benthic feeding of sister rock-
cod species Notothenia rossii and N. coriiceps (Nototheniidae) in West Antarctica. Polar Biol
42:317–334
Bousfield EL (ed) (1973) Shallow-water gammaridean amphipoda of New England.
Comstock, Ithaca
Braeckman U, Pasotti F, Vázquez S, Zacher K, Hoffmann R, Elvert M, Marchant H, Buckner C,
Quartino ML, Mác Cormack W, Soetaert K, Wenzhöfer F, Vanreusel A (2019) Degradation of
macroalgal detritus in shallow coastal Antarctic sediments. Limnol Oceanogr 64:1423–1441
Bregazzi PK (1972) Life cycle and seasonal movements of Cheirimedon femoratus (Pfeffer) and
Tryphosella kergueleni (Miers) (Crustacea: Amphipoda). Br Antarct Surv Bull 30:1–34
Breithaupt T, Thiel M (eds) (2011) Chemical communication in crustaceans. Springer-Verlag,
New York, p 565
Brönmark C, Hansson L-A (eds) (2012) Chemical ecology in aquatic systems. Oxford University
Press, Oxford
Brouwer PEM (1996) Decomposition in situ of the sublittoral Antarctic macroalga Desmarestia
anceps Montagne. Polar Biol 16:129–137
Bucolo P, Amsler CD, McClintock JB, Baker BJ (2011) Palatability of the Antarctic rhodophyte
Palmaria decipiens (Reinsch) RW Ricker and its endo/epiphyte Elachista antarctica Skottsberg
to sympatric amphipods. J Exp Mar Biol Ecol 396:202–206
Bucolo P, Amsler CD, McClintock JB, Baker BJ (2012) Effects of macroalgal chemical extracts
on spore behavior of the Antarctic epiphyte Elachista antarctica Phaeophyceae. J Phycol
48:1403–1410
Bullard SB, Hay ME (2002) Palatability of marine macro-holoplankton: nematocysts, nutritional
quality, and chemistry as defenses against consumers. Limnol Oceanogr 47:1456–1467
Campbell HA, Fraser KPP, Bishop CM, Peck LS, Egginton S (2008) Hibernation in an Antarctic
fish: on ice for winter. PLoS One 3:e1743
Casaux R (1998) The contrasting diet of Harpagifer antarcticus (Notothenioidei, Harpagiferidae)
at two localities of the South Shetland Islands, Antarctica. Polar Biol 19:283–285
Casaux R, Mazzotta A, Barrera-Oro E (1990) Seasonal aspects of the biology and diet of nearshore
nototheniid fish at Potter Cove, South Shetland Islands, Antarctica. Polar Biol 11:63–72
Casaux R, Baroni A, Ramon A, Carlini A, Bertolin M, DiPrinzio CY (2009) Diet of the leopard
seal Hydrurga leptonyx at the Danco Coast, Antarctic Peninsula. Polar Biol 32:307–310
Cimino G, Ghiselin MT (2009) Chemical defense and the evolution of opisthobranch gastropods.
Proc Calif Acad Sci 60:175–422
Clark GF, Stark JS, Palmer AS, Riddle MJ, Johnston EL (2017) The roles of sea-ice, light and
sedimentation in structuring shallow Antarctic benthic communities. PLoS One 12:e0168391
Corbisier TN, Petti MAV, Skowronski RSP, Brito TAS (2004) Trophic relationships in the near-
shore zone of Martel Inlet (King George Island, Antarctica): δ13C stable-isotope analysis. Polar
Biol 27:75–82
Correa JA, Sánchez PA (1996) Ecological aspects of algal infectious diseases. Hydrobiologia
326–327:89–95
Cutignano A, Villani G, Fontana A (2012) One metabolite, two pathways: convergence of polypro-
pionate biosynthesis in fungi and marine molluscs. Org Lett 14:992–995
Daly JW (2004) Marine toxins and nonmarine toxins: Convergence or symbiotic organisms? J Nat
Prod 67:1211–1215
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 359

Daniels RA (1982) Feeding ecology of some fishes of the Antarctic Peninsula. Fish Bull 80:575–585
Daniels RA (1983) Demographic characteristics of an Antarctic plunderfish, Harpagifer bispinis
antarcticus. Mar Ecol Prog Ser 13:181–183
Dayton PK, Robilliard GA, DeVries AL (1969) Anchor ice formation in McMurdo Sound,
Antarctica, and its biological effects. Science 163:273–275
Dayton PK, Robillard GA, Paine RT (1970) Benthic faunal zonation as a result of anchor ice
at McMurdo Sound, Antarctica. In: Holgate MW (ed) Antarctic ecology, vol 1. Academic,
New York, pp 244–258
De Broyer C, Jażdżewska A (2014) Chapter 5.17. Biogeographic patterns of Southern Ocean ben-
thic Amphipods. In: De Broyer C, Koubbi P, Griffiths HJ, Raymond B, d’Udekem d’Acoz C,
Van de Putte AP, Danis B, David B, Grant S, Gutt J, Held C, Hosie G, Huettmann F, Post A,
Ropert-Coudert Y (eds) Biogeographic atlas of the Southern Ocean. Scientific Committee on
Antarctic Research, Cambridge, 155–165
Dearborn JH, Fell FJ (1974) Ecology of echinoderms from the Antarctic Peninsula. Antarct J US
9:304–306
DeLaca TE, Lipps JH (1976) Shallow water marine associations, Antarctic Peninsula. Antarct J
US 11:12–20
Dicke M, Takken W (eds) (2006) Chemical ecology: from gene to ecosystem. Springer, New York
Doyle SR, Momo FR, Brêthes J-C, Ferreyra GA (2012) Metabolic rate and food availability of the
Antarctic amphipod Gondogeneia antarctica (Chevreux 1906): seasonal variation in allometric
scaling and temperature dependence. Polar Biol 35:413–424
Duffy JE, Hay ME (1994) Herbivore resistance to seaweed chemical defense: the roles of mobility
and predation risk. Ecology 75:1304–1319
Dunton K (2001) δ15N and δ13C measurements of Antarctic Peninsula fauna: trophic relationships
and assimilation of benthic seaweeds. Am Zool 41:99–112
Eisner T, Meinwald J (eds) (1995) Chemical ecology: the chemistry of biotic interaction. National
Academy Press, Washington, DC
Faugeron S, Martínez EA, Sánchez PA, Correa JA (2000) Infectious diseases in Mazzaella lami-
narioides (Rhodophyta): estimating the effect of infections on host reproductive potential. Dis
Aquat Org 42:143–148
Gauna M, Parodi E, Cáceres E (2009) Epi-endophytic symbiosis between Laminariocolax aecidi-
oides (Ectocarpales, Phaeophyceae) and Undaria pinnatifida (Laminariales, Phaeophyceae)
growing on Argentinian coasts. J Appl Phycol 21:11–18
Hall-Aspland SA, Rogers TL (2004) Summer diet of leopard seals (Hydrurga leptonyx) in Prydz
Bay, Eastern Antarctica. Polar Biol 27:729–734
Hay ME (1986) Associational plant defenses and the maintenance of species diversity: turning
competitors into accomplices. Am Nat 128:617–641
Hay ME (1997) The ecology and evolution of seaweed-herbivore interactions on coral reefs. Coral
Reefs 16:S67–S76
Hay ME (2009) Marine chemical ecology: chemical signals and cues structure marine populations,
communities, and ecosystems. Annu Rev Mar Sci 1:193–212
Hay ME, Sutherland JP (1988) The ecology of rubble structures in the South Atlantic Bight: a
community profile. US Fish Wildl Serv Biol Rep 85(7.20):1–67
Hay ME, Stachowicz JJ, Cruz-Rivera E, Bullard SB, Deal MS, Lindquist N (1998) Bioassays with
marine and freshwater macroorganisms. In: Haynes KF, Millar JG (eds) Methods in chemical
ecology. Volume 2: Bioassay methods. Kluwer Academic Publishers, Norwell, pp 39–141
Hennebert E, Jangoux M, Flamming P (2013) Functional biology of asteroid tube feet. In:
Lawrence JM (ed) Starfish biology and ecology of the Asteroidea. John Hopkins Press,
Baltimore, pp 24–36
Herms DA, Mattson WJ (1992) The dilemma of plants: to grow or defend. Q Rev Biol 67:285–335
Hommersand MH, Moe RL, Amsler CD, Fredericq S (2009) Notes on the systematics and biogeo-
graphical relationships of Antarctic and Sub-Antarctic Rhodophyta with descriptions of four
new genera and five new species. Bot Mar 52:509–534
360 C. D. Amsler et al.

Huang YM, Amsler MO, McClintock JB, Amsler CD, Baker BJ (2007) Patterns of gammarid
amphipod abundance and species composition associated with dominant subtidal macroalgae
along the western Antarctic Peninsula. Polar Biol 30:1417–1430
Iken K (1999) Feeding ecology of the Antarctic herbivorous gastropod Laevilacunaria antarctica
Martens. J Exp Mar Biol Ecol 236:133–148
Iken K, Barrera-Oro ER, Quartino ML, Casaux RJ, Brey T (1997) Grazing in the Antarctic fish
Notothenia coriiceps: evidence for selective feeding on macroalgae. Antarct Sci 9:386–391
Iken K, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009) Ecological roles of phlorotan-
nins in Antarctic brown algae. Bot Mar 52:547–557
Inderjit, Mallik AU (2002) Chemical ecology of plants: allelopathy in aquatic and terrestrial eco-
systems. Birkhäuser, Basel, p 272
Janosik AM, Halanych KM (2010) Unrecognized Antarctic biodiversity: a case study of the genus
Odontaster (Odontasteridae; Asteroidea). Integr Comp Biol 50:981–992
Jormalainen V, Honkanen T (2008) Macroalgal chemical defenses and their roles in structuring
temperate marine communities. In: Amsler CD (ed) Algal chemical ecology. Springer-Verlag,
Berlin, pp 57–89
Kang S, Kim S, Park H (2015) Transcriptome of the Antarctic amphipod Gondogeneia antarctica
and its response to pollutant exposure. Mar Genomics 24:253–254
Kylin H (1937) Bemerkungen über die Entwicklungsgeschichte eniger Phaeophyceen. Acta Univ
Lund 33:1–34
Lamb IM, Zimmerman MH (1977) Benthic marine algae of the Antarctic Peninsula. Ant Res Ser
5:130–229
Lane AL, Kubanek J (2008) Secondary metabolite defenses against pathogens and biofoulers. In:
Amsler CD (ed) Algal chemical ecology. Springer-Verlag, Berlin, pp 229–243
Littler M, Littler D (1980) The evolution of thallus form and survival strategies in benthic marine
macroalgae: field and laboratory tests of a functional form model. Am Nat 116:25–44
Ma WS, Mutka T, Vesley B, Amsler MO, McClintock JB, Amsler CD, Perman JA, Singh MP,
Maiese WM, Zaworotko MJ, Kyle DE, Baker BJ (2009) Norselic Acids A-E, highly oxidized
anti-infective steroids that deter mesograzer predation, from the Antarctic sponge Crella sp. J
Nat Prod 72:1842–1846
Marcías ML, Deregibus D, Saravia LA, Campana GL, Quartino ML (2017) Life between tides:
spatial and temporal variations of an intertidal macroalgal community at Potter Peninsula,
South Shetland Islands, Antarctica. Estuar Coast Shelf Sci 187:193–203
Maschek JA, Baker BJ (2008) The chemistry of algal secondary metabolism. In: Amsler CD (ed)
Algal chemical ecology. Springer-Verlag, Berlin, pp 1–24
McClintock JB (1994) Trophic biology of Antarctic echinoderms. Mar Ecol Prog Ser 111:191–202
McClintock JB, Baker BJ (eds) (2001) Marine chemical ecology. CRC, Boca Raton
McClintock JB, Amsler CD, Baker BJ (2010) Overview of the chemical ecology of benthic marine
invertebrates along the western Antarctic Peninsula. Integr Comp Biol 50:967–980
McDowell RE, Amsler CD, Dickinson DA, McClintock JB, Baker BJ (2014a) Reactive oxygen
species and the Antarctic macroalgal wound response. J Phycol 50:71–80
McDowell RE, Amsler CD, McClintock JB, Baker BJ (2014b) Reactive oxygen species as a
marine grazing defense: H2O2 and wounded Ascoseira mirabilis both inhibit feeding by an
amphipod grazer. J Exp Mar Biol Ecol 458:34–38
McDowell RE, Amsler MO, Li Q, Lancaster JR, Amsler CD (2015) The immediate wound-­
induced oxidative burst of Saccharina latissima depends on light via photosynthetic electron
transport. J Phycol 51:431–441
McDowell RE, Amsler CD, Amsler MO, Li Q, Lancaster JR Jr (2016) Control of grazing by light
availability via light-dependent, wound-induced metabolites: the role of reactive oxygen spe-
cies. J Exp Mar Biol Ecol 477:86–91
Miller KA, Pearse JS (1991) Ecological studies of seaweeds in McMurdo Sound, Antarctica. Am
Zool 31:35–48
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 361

Moles J, Núñez-Pons L, Taboada S, Figuerola B, Cristobo J, Avila C (2015) Anti-predatory chemi-


cal defences in Antarctic benthic fauna. Mar Biol 162:1813–1821
Müller-Schwarze D (ed) (2006) Chemical ecology of vertebrates. Cambridge University Press,
Cambridge
North AW (1996) Locomotory activity and behaviour of the Antarctic teleost Notothenia coriiceps.
Mar Biol 126:125–132
Núñez-Farfán J, Fornoni J, Valverde PL (2007) The evolution of resistance and tolerance to herbi-
vores. Annu Rev Ecol Evol Syst 38:541–566
Núñez-Pons L, Avila C (2014) Deterrent activities in the crude lipophilic fractions of Antarctic
benthic organisms: chemical defences against keystone predators. Polar Res 33:1
Núñez-Pons L, Rodríguez-Arias M, Gómez-Garreta A, Ribera-Siguán A, Avila C (2012) Feeding
deterrency in Antarctic marine organisms: bioassays with the omnivore amphipod Cheirimedon
femoratus. Mar Ecol Prog Ser 462:163–174
Obermüller B, Puntarulo S, Abele D (2007) UV-tolerance and instantaneous physiological stress
responses of two Antarctic amphipod species Gondogeneia antarctica and Djerboa furcipes
during exposure to UV radiation. Mar Environ Res 64:267–285
Pabis K, Sicinski J (2010) Polychaete fauna associated with holdfasts of the large brown alga
Himantothallus grandifolius in Admiralty Bay, King George Island, Antarctic. Polar Biol
33:1277–1288
Pawson DL (1969) Echinoidea. Antarct Map Folio Ser 11:38–41
Pearse JS, Giese AC (1966) Food, reproduction and organic constitution of the common antarctic
echinoid Sterechinus neumayeri (Meissner). Biol Bull 130:387–401
Peters AF (2003) Molecular identification, taxonomy and distribution of brown algal endophytes,
with emphasis on species from Antarctica. Proc Int Seaweed Symp 17:293–302
Peters KJ, Amsler CD, Amsler MO, McClintock JB, Dunbar RB, Baker BJ (2005) A comparative
analysis of the nutritional and elemental composition of macroalgae from the western Antarctic
Peninsula. Phycologia 44:453–463
Picken GB (1979) Growth, production and biomass of the antarctic gastropod Laevilacunaria
antarctica Martens 1885. J Exp Mar Biol Ecol 40:71–79
Picken GB (1980) The distribution, growth, and reproduction of the Antarctic limpet Nacella
(Patinigera) concinna (Strebel) 1908. J Exp Mar Biol Ecol 42:71–85
Potin P (2008) Oxidative burst and related responses in biotic interactions of algae. In: Amsler CD
(ed) Algal chemical ecology. Springer-Verlag, Berlin, pp 245–271
Puglisi MP, Becerro MA (eds) (2019) Chemical ecology: the ecological impacts of marine natural
products. CRC, Boca Raton, p 400
Purchon RD (1977) Feeding methods and adaptive radiation in the Gastropoda. In: Purchon RD
(ed) The biology of the mollusca, 2nd edn. Pergamon, Amsterdam, pp 41–99
Reichardt W, Dieckmann G (1985) Kinetics and trophic role of bacterial degradation of macro-­
algae in Antarctic coastal waters. In: Siegfried WR, Condy P, Laws RM (eds) Antarctic nutrient
cycles and food webs. Springer-Verlag, Berlin, pp 115–122
Richardson MG (1971) The ecology and physiological aspects of Antarctic weed dwelling amphi-
pods (Preliminary report, II). Brit Antarct Surv Rep N9/1971(-72)/H:1–16
Richardson MG (1977) The ecology including physiological aspects of selected Antarctic marine
invertebrates associated with inshore macrophytes. PhD Dissertation. Department of Zoology,
University of Durham
Rosenthal GA, Berenbaum MR (2012) Herbivores: their interactions with secondary plant metabo-
lites. Ecological and evolutionary processes, 2nd edn. Academic Press, New York, p 493
Santhanam R (2018) Biology and ecology of edible marine gastropod molluscs. Apple Academic
Press, Oakville
Saucède T, Pierrat B, David B (2014) Chapter 5.26. Echinoids. In: De Broyer C, Koubbi P, Griffiths
HJ, Raymond B, d’Udekem d’Acoz C, Van de Putte AP, Danis B, David B, Grant S, Gutt J,
Held C, Hosie G, Huettmann F, Post A, Ropert-Coudert Y (eds) Biogeographic atlas of the
Southern Ocean. Scientific Committee on Antarctic Research, Cambridge, pp 213–220
362 C. D. Amsler et al.

Schoenrock KM, Amsler CD, McClintock JB, Baker BJ (2013) Endophyte presence as a potential
stressor on growth and survival in Antarctic macroalgal hosts. Phycologia 52:595–599
Schoenrock KM, Amsler CD, McClintock JB, Baker BJ (2015a) A comprehensive study of
Antarctic algal symbioses: minimal impacts of endophyte presence in most species of macroal-
gal hosts. Eur J Phycol 50:271–278
Schoenrock KM, Amsler CD, McClintock JB, Baker BJ (2015b) Life history bias in endophyte
infection of the Antarctic rhodophyte, Iridaea cordata. Bot Mar 58:1–8
Schram JB, McClintock JB, Amsler CD, Baker BJ (2015) Impacts of acute elevated seawater
temperature on the feeding preferences of an Antarctic amphipod toward chemically deterrent
macroalgae. Mar Biol 162:425–433
Schram JB, Amsler MO, Amsler CD, Schoenrock KM, McClintock JB, Angus RA (2016)
Antarctic crustacean grazer assemblages exhibit resistance following exposure to decreased
pH. Mar Biol 163:106
Schwarz AM, Hawes I, Andrew N, Norkko A, Cummings V, Thrush S (2003) Macroalgal photo-
synthesis near the southern global limit for growth; Cape Evans, Ross Sea, Antarctica. Polar
Biol 26:789–799
Sotka EE, Whalen KE (2008) Herbivore offense in the sea: the detoxification and transport of
secondary metabolites. In: Amsler CD (ed) Algal chemical ecology. Springer-Verlag, Berlin,
pp 203–228
Sotka EE, Forbey J, Horn M, Poore AGB, Raubenheimer D, Whalen KE (2009) The emerging
role of pharmacology in understanding consumer-prey interactions in marine and freshwater
systems. Integr Comp Biol 49:291–313
Steneck RS, Graham MH, Bourque BJ, Corbett D, Erlandson JM, Estes JA, Tegner MJ (2002) Kelp
forest ecosystems: biodiversity, stability, resilience and future. Environ Conserv 29:436–459
Thomas F, Cosse A, Le Panse S, Kloareg B, Potin P, Leblanc C (2014) Kelps feature systemic
defense responses: insights into the evolution of innate immunity in multicellular eukaryotes.
New Phytol 204:567–576
Valdivia N, Pardo LM, Macaya EC, Huovinen P, Gómez I (2019) Different ecological mechanisms
lead to similar grazer controls on the functioning of periphyton Antarctic and sub-Antarctic
communities. Prog Oceanogr 174:7–16
von Salm JL, Schoenrock KM, McClintock JB, Amsler CD, Baker BJ (2019) The status of marine
chemical ecology in Antarctica: form and function of unique high-latitude chemistry. In:
Puglisi-Weening MP, Becerro MA (eds) Chemical ecology: the ecological impacts of marine
natural products. CRC, Boca Raton, pp 27–69
Waters CM, Bassler BL (2005) Quorum sensing: cell-to-cell communication in bacteria. Annu Rev
Cell Dev Biol 21:319–346
Wiencke C, Amsler CD (2012) Seaweeds and their communities in polar regions. In: Wiencke C,
Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and utilization.
Springer-Verlag, Berlin, pp 265–294
Wiencke C, Clayton MN (2002) Antarctic seaweeds. In: Synopsis of the Antarctic benthos, vol 9.
ARG Gantner Verlag KG, Ruggell
Wiencke C, Amsler CD, Clayton MN (2014) Chapter 5.1. Macroalgae. In: De Broyer C, Koubbi
P, Griffiths HJ, Raymond B, d’Udekem d’Acoz C, Van de Putte AP, Danis B et  al (eds)
Biogeographic atlas of the Southern Ocean. Scientific Committee on Antarctic Research,
Cambridge, 66–73
Wyatt TD (ed) (2014) Pheromones and animal behavior: chemical signals and signatures.
Cambridge Press, Cambridge
Yang EC, Boo SM, Bhattacharya D, Saunders GW, Knoll AH, Fredericq S, Graf L, Yoon HS
(2016) Divergence time estimates and the evolution of major lineages in the florideophyte red
algae. Sci Rep 6:21361
Young RM, von Salm JL, Amsler MO, Lopez-Bautista J, Amsler CD, McClintock JB, Baker BJ
(2013) Site-specific variability in the chemical diversity of the Antarctic red alga Plocamium
cartilagineum. Mar Drugs 11:2126–2139
17  Chemical Mediation of Antarctic Macroalga-Grazer Interactions 363

Zamzow JP, Amsler CD, McClintock JB, Baker BJ (2010) Habitat choice and predator avoidance
by Antarctic amphipods: the roles of algal chemistry and morphology. Mar Ecol Prog Ser
400:155–163
Zamzow JP, Aumack CF, Amsler CD, McClintock JB, Amsler MO, Baker BJ (2011) Gut contents
and stable isotope analyses of the Antarctic fish, Notothenia coriiceps Richardson, from two
macroalgal communities. Antarct Sci 23:107–116
Zenteno L, Cárdenas L, Valdivia N, Gómez I, Höfer J, Garrido I, Pardo LM (2019) Unraveling
the multiple bottom-up supplies of an Antarctic nearshore benthic community. Prog Oceanogr
174:55–63
Chapter 18
Brown Algal Phlorotannins: An Overview
of Their Functional Roles

Iván Gómez and Pirjo Huovinen

Abstract Phlorotannins are polyphenolic compounds, relatively hydrophilic,


formed by polymers of phloroglucinol and found exclusively in brown algae. These
molecules are located in vesicles denominated physodes (the soluble fraction) and
also complexed to polysaccharides in the cell wall (the insoluble fraction). Well
known as potential grazer deterrents, one of the most striking characteristics of
these compounds, due to a number of hydroxyl groups, is their antioxidant poten-
tial, which opens promising perspectives for pharmaceutical and biotechnological
uses. In Antarctic brown algae, especially endemic species of Desmarestiales, con-
stitutively high levels of phlorotannins (up to 12% of dry weight) have been mea-
sured. Although translocation has not been conclusively confirmed, the differential
allocation of phlorotannins in meristematic and reproductive tissues in some species
suggests their involvement in chemical defenses protecting essential metabolic
functions. Due to their UV-absorbing properties and peripheral localization in cells
and tissues, phlorotannins have been related with the increased tolerance to UV
radiation in various Antarctic brown algae. However, no induction of phlorotannins
by UV has been demonstrated, which strongly supports the idea that these mole-
cules are constitutive biochemical components of a suite of mechanisms against
multiple stressors. Due to their structural role as primary compounds, phlorotannins
are essential for various morpho-functional processes that in the case of Antarctic
algae allow them to thrive under extreme conditions. Overall, the significance of
phlorotannins in this group of algae has largely been recognized; however, funda-
mental aspects of their molecular expression, synthesis, and regulation still need to
be addressed, especially considering the climate change-driven environmental
scenarios.

Keywords  Antioxidant activity · Brown algae · Phenolic compounds · Physodes ·


Secondary metabolites · UV-absorbing compounds

I. Gómez (*) · P. Huovinen


Instituto de Ciencias Marinas y Limnológicas, Facultad de Ciencias, Universidad Austral de
Chile, Valdivia, Chile
Research Center Dynamics of High Latitude Marine Ecosystems (IDEAL), Valdivia, Chile
e-mail: igomezo@uach.cl; pirjo.huovinen@uach.cl

© Springer Nature Switzerland AG 2020 365


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6_18
366 I. Gómez and P. Huovinen

Fig. 18.1  Chemical structure of some common phlorotannins extracted from brown algae

18.1  Introduction

Brown algal phenolics belong to a single structural class, the phlorotannins, which
are dehydropolymers of phloroglucinol (1,3,5-trihydroxybenzen). Depending on
the degree of polymerization, phlorotannins present a wide range of molecular
weights (between 140 and 40.000 Da; Ragan and Glombitza 1986). Thus, based
on the number of phloroglucinol units, the profiling of phlorotannins can vary
considerably among species (Steevensz et  al. 2012). Within the phlorotannins
found in brown algae, the most common are fucols, phlorethols, fucophlorethols,
and eckols (Fig.  18.1), which vary depending on the type of chemical linkage,
e.g., ether Aryl-­O-­Aryl linkages in phlorethols or dibenzodioxin linkages in eck-
ols (Ragan and Glombitza 1986; La Barre et al. 2010). As secondary metabolites,
these substances play a series of putative roles in the cell, mainly as anti-herbivory
defense, antifouling activity, UV protectants, and antioxidants. Phlorotannins can
be present as soluble substances sequestered in vesicle-denominated physodes
and as insoluble fraction bound to polysaccharides in the cell wall (denominated
the insoluble fraction) (Fig. 18.2). Due to this, these compounds are regarded also
as primary compounds, essential during cell formation (Schoenwaelder 2002).
This has been corroborated by histological studies indicating that phlorotannins
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 367

Fig. 18.2  Localization of physodes and the different functional forms of phlorotannins in brown
algal cells. The relationship between exuded phlorotannins and extracellular phenolic bodies has
not been conclusively established

participate actively in wall configuration of zygotes (Kevekordes and Clayton


1999), are concentrated at the periphery of multicellular embryos of Fucus
(Schoenwaelder and Clayton 1998), and are actively synthesized during wound
healing and sealing after amputation of thallus regions simulating herbivory
(Lüder and Clayton 2004).
Because of their ubiquity, phlorotannins can be present in high concentrations in
some species. For example, in Antarctic brown algae, particularly endemic species
of Desmarestiales, constitutively high levels of phlorotannins have been measured,
forming up to 12% of dry weight (Table 18.1). However, it has been well demon-
strated that total phlorotannin content can vary significantly depending on a number
of biotic and abiotic factors (Van Alstyne and Pelletreau 2000; Pavia and Toth 2000;
Jormalainen et al. 2003; Gómez and Huovinen 2010).
Apart from their well-known role as herbivore deterrents (reviewed in Chap. 17
in this volume and by Iken et al. 2009), phlorotannins and various of their chemical
derivatives apparently have other important functions, especially in oxidative
metabolism and metal chelation (Stauber and Florence 1987), with far-reaching
implications for ecophysiology of brown algae (La Barre et al. 2010).
368 I. Gómez and P. Huovinen

Table 18.1  Concentration of total phlorotannins (soluble and insoluble) in different Antarctic
brown algae
Phlorotannin
contents
Species Soluble Insoluble Remarks Reference
Percentage (mg g−1 DW)
Desmarestia anceps 117 ± 3 Average of different thallus Fairhead et al.
parts (Anvers Island) (2005a)
60 ± 19 1 and 37 m (Anvers Island) Iken et al. (2007)
60 ± 5.5 28 m (King George Island) Huovinen and
Gómez (2013)
30 10–15 m (Anvers Island) Schoenrock et al.
(2015)
80 ± 18 28 ± 4.9 20 m (King George Island) Gómez and
Huovinen (2015)
52 ± 15 15 m (King George Island) Flores-Molina et al.
(2016)
43–48 33 7–10 m depth (King George Rautenberger et al.
Island) (2015)
Desmarestia 52 ± 2 Average of different thallus Fairhead et al.
menziesii parts (Anvers Island) (2005a)
30 ± 12 1 and 37 m (Anvers Island) Iken et al. (2007)
85 ± 5.6 Apical parts; 17 m (King Huovinen and
George Island) Gómez (2013)
10 10–15 m (Anvers Island) Schoenrock et al.
(2015)
55 ± 15 20 m (King George Island) Gómez and
Huovinen (2015)
55–58 18 7–10 m depth (King George Rautenberger et al.
Island) (2015)
Ascoseira mirabilis 35 ± 25 1 and 37 m (Anvers Island) Iken et al. (2007)
15 ± 12 8 m (King George Island) Huovinen and
Gómez (2013)
12 ± 3 50 ± 10 1 m (King George Island) Huovinen and
Gómez (2015)
13 ± 3 15 ± 2.4 20 m (King George Island) Gómez and
Huovinen (2015)
10–12 20–23 7–10 m depth (King George Rautenberger et al.
Island) (2015)
Himantothallus 63 ± 8 1 and 37 m (Anvers Island) Iken et al. (2007)
grandifolius
120 ± 8.6 30 m (King George Island) Huovinen and
Gómez (2013)
97 ± 20 49 ± 11 20 m (King George Island) Gómez and
Huovinen (2015)
90–99 40 7–10 m depth (King George Rautenberger et al.
Island) (2015)
(continued)
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 369

Table 18.1 (continued)
Phlorotannin
contents
Species Soluble Insoluble Remarks Reference
Cystosphaera 26 ± 16 1 and 37 m (Anvers Island) Iken et al. (2007)
jacquinotii
29 ± 5.2 25 m (King George Island) Huovinen and
Gómez (2013)
75 ± 19 10 ± 3 15–20 m (King George Huovinen and
Island) Gómez (2015)
Desmarestia 42 ± 5.7 20 m (King George Island) Huovinen and
antarctica Gómez (2013)
Adenocystis 52 ± 4.4 Intertidal zone (King George Huovinen and
utricularis Island) Gómez (2013)

18.2  S
 ynthesis and Cellular Localization of Phlorotannins:
Dual Functions as Secondary Metabolites
and Structural Compounds

Phlorotannins are synthesized via the acetate-malonate pathway through a type III
polyketide synthase (Herbert 1989; Meslet-Cladiere et al. 2013), and in terms of
their chemical properties, they differ from the condensed tannins of vascular plants
(Arnold and Targett 2002). It has been suggested that some fatty acids associated
with the Acetyl-CoA, a key intermediate in the polyketide pathway, could be
regarded as precursors of phlorotannins (Steinhoff et al. 2011). However, metabo-
lism of phenolics in algae has been much less studied, and key aspects of biosynthe-
sis and regulation are unknown.
Due to their reactivity, phlorotannins may easily form complexes with macro-
molecules, and they are sequestered in physodes through the formation of covalent,
hydrogen, or ionic bonds (the soluble fraction) (Fig. 18.2). Traditionally physodes
or phenolic precursors are thought to be synthesized in the chloroplast or at the
chloroplast membrane, and various authors have described an osmiophilic material
being released from chloroplasts (Evans and Holligan 1972; Feldmann and
Guglielmi 1972; Pellegrini 1980). An alternative explanation suggests that phenolic
material is produced in the chloroplast endoplasmic reticulum (CER), which may
play a role in the transport of phenolic precursors to cell vacuoles and physodes
(Pellegrini 1980; Clayton and Beakes 1983; Kaur and Vijayaraghavan 1992) or may
directly give rise to physodes (Feldmann and Guglielmi 1972; Oliveira and
Bisalputra 1973). When the physodes make contact with the plasmalemma, phloro-
tannins are released and polymerized in the apoplast (cell wall) forming complexes
with polysaccharides, e.g., alginic-acid-bound phlorotannins (Schoenwaelder and
Clayton 1999) (Fig. 18.2). Some studies have identified peroxidases in the cell wall
of Ascophyllum nodosum suggesting that phlorotannins excreted from the cells may
be modified through the activity of these enzymes (Vilter 1995). In fact, the process
370 I. Gómez and P. Huovinen

of oxidative condensation and the linkage to alginic acids in the apoplast is appar-
ently driven by vanadium-dependent haloperoxidases (Potin and Leblanc 2006;
Salgado et al. 2009). Vreeland and Laetsch (1988) proposed that phenolic cross-­
linking of alginate may occur in early wall formation in Fucus and that peroxidases
may be involved in the catalysis of phenolic condensation into alginate. During
early phases of growth, physode movements to regions of active wall formation
from the cell periphery to the rhizoid tip and to the impending plane of cytokinesis
are dependent on interactions with the cytoskeleton (Schoenwaelder and Clayton
1999). Hence, actin microfilaments may be acting as a general circulatory system
moving physodes around the cell, with microtubules directing physodes (and prob-
ably other wall components) to cell-wall deposition sites, both in the primary wall
and at cross-walls (Kevekordes and Clayton 1999; Schoenwaelder and Clayton
1999). Although the exudation of phlorotannins has been reported (Ragan and
Glombitza 1986; Toth and Pavia 2002; Koivikko et  al. 2005), no clear evidence
exists that they are related with phenolic bodies described in embryos of Durvillaea
antarctica (Kevekordes and Clayton 1999). For example, it was demonstrated that
extracellular excretion of phenolic compounds in Eisenia bicyclis and Ecklonia
kurome corresponded to monomeric bromophenols, while phloroglucinol or poly-
meric phlorotannins were not detected (Shibata et  al. 2006). On the other hand,
phlorotannins are strong chelators of metals, and thus, they are thought to partici-
pate in exudation-based detoxification mechanisms: they may sequester metal ions
in physodes (Smith and Harwood 1986), and through exudation processes, these
metal-complexing compounds may decrease the metal concentration or alter its spe-
ciation in the surrounding water (Gledhill et al. 1999). As physodes are more abun-
dant in peripheral cell layers, their role as a filter stopping metals from entering the
inner cells has been proposed (reviewed by Schoenwaelder 2002).

18.3  Phlorotannins as UV-Screening Substances

In contrast to terrestrial plants, where natural levels of UV radiation do not neces-


sarily result in damage (Paul and Gwynn-Jones 2003; Hideg et al. 2013), aquatic
organisms, especially subtidal seaweeds, can be impaired when they are exposed to
high solar radiation (Bischof et  al. 2006a). Thus, synthesis and accumulation of
UV-screening substances is a common photoprotective strategy observed in several
groups of living organisms (Karentz et  al. 1991; García-Pichel and Castenholz
1993; Cockell and Knowland 1999). In fact, various of these compounds have been
isolated and tested as bioactive substances for use in skin care, cosmetics, and phar-
maceutical products (reviewed in Pangestuti et  al. 2018). Due to their chemical
properties, diverse phenolics are regarded as general anti-stress agents, including
UV protection and antioxidant activity, and apart from brown algae, they have been
reported in green (Menzel et al. 1983; Pérez-Rodriguez et al. 1998, 2001; Gómez
et al. 1998; Ross et al. 2005) and red algae (Athukorala et al. 2003; Yildiz et al.
2011; Heffernan et al. 2014; Cruces et al. 2018).
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 371

Fig. 18.3  Absorption of ethanol-water extract of apical fronds of Lessonia spicata exposed to UV
radiation for 24 h

Phlorotannins have been correlated with an increased tolerance to UV radiation


(Pavia et al. 1997; Swanson and Fox 2007; Gómez and Huovinen 2010; Steinhoff
2010). In general, phlorotannins absorb at wavelengths between 200 and 300 nm,
i.e., well in the UV-C range and at shorter UV-B wavelengths. When phlorotannins
are removed using polyvinylpolypyrrolidone (PVPP), absorption of algal extract
decreases by 70% in the range between 280 and 300 nm (Pavia et al. 1997). Although
the effectiveness of phlorotannins as UV-screening compounds is higher at the
UV-C range, absorption spectra of the different fractions of phlorotannins (phy-
sodes, cell-wall-bound and excreted phlorotannins) may be different. In fact, the
absorption peak of the trihydroxy-coumarin from Dasycladus vermicularis suffers
a shift when excreted to the seawater (Pérez-Rodriguez et  al. 2001). Protection
against UV radiation by phlorotannins in early developmental phases has been dem-
onstrated in Fucus embryos and spores of Laminaria (Schoenwaelder et al. 2003;
Henry and Van Alstyne 2004; Roleda et al. 2010). In the kelp Lessonia spicata, 24-h
exposure to UV radiation increases the synthesis of phlorotannins, compared to
control without UV (Fig. 18.3). In this species, the UV-mediated increase in phloro-
tannins can minimize photodamage of key physiological processes and cellular
components, such as photosynthesis and DNA (Gómez and Huovinen 2010).
Interestingly, differences between species, season, and morpho-functional processes
have raised the question whether the photoprotective role of phlorotannins in brown
algae can be regarded as a constitutive or inducible mechanism. For example, in
Lessonia, the induction of phlorotannins by UV radiation has been shown to occur
only during the period when sporophytes actively grow (Gómez and Huovinen
2010). However, in Fucus vesiculosus, no UV induction of soluble phlorotannins
was found, which was related with a lack of upregulation of pksIII genes (Creis
372 I. Gómez and P. Huovinen

et al. 2015). Overall, the few available studies point to a complex interplay between
the induction of soluble phlorotannins enclosed in physodes and their subsequent
deposition in the cell-wall matrix. Insoluble phlorotannins polymerized in the cell
wall are regarded as primary UV-shielding substances (Gómez and Huovinen 2010),
similar to cell-wall-bound phenolics reported in plants (Clarke and Robinson 2008).
Moreover, photoprotection is conferred not only via intracellular accumulation of
phlorotannins but also as a result of exudation to the surrounding water as has been
suggested for adult thalli and propagules of the giant kelp Macrocystis pyrifera
(Swanson and Druehl 2002).

18.4  Phlorotannins as Active Antioxidant Compounds

The formation of reactive oxygen species (ROS) is one of the primary expressions
of stress in marine algae, but they can also act as signaling molecules in several cel-
lular reactions (revised in Bischof and Rautenberger 2012). Phlorotannins are
known to act, not only as photoprotective substances but also as highly efficient
ROS scavengers (Nakai et al. 2006; Wang et al. 2009; Heffernan et al. 2014). Due
to the presence of various interconnected rings (up to eight) in their chemical struc-
ture, phlorotannins are regarded as potent antioxidants scavenging different types of
ROS, e.g., superoxide anions (O2−), peroxides, singlet oxygen (1O2), and hydroxyl
radicals (•OH) (Ahn et al. 2007; Koivikko et al. 2007). Thus, the hydroxyl groups
present in phlorotannins act as reducing agents, hydrogen donors, and singlet oxy-
gen (reviewed in Michalak 2006). It has been postulated that the relationship
between increased levels of ROS and phlorotannin induction in brown algae can
follow the methyl jasmonate signal transduction pathway, a plant defense-related
pathway reported commonly in plants during high ROS production (Arnold et al.
2001; Küpper et al. 2009). The evidence gained during the last decades appears to
indicate that operation of efficient and rapid ROS scavenging mechanisms based on
phenols can be regarded as an important physiological adaptation in seaweeds when
they are exposed to different environmental stressors, e.g., high solar irradiation,
metal pollution, or high temperature (Aguilera et al. 2002; Contreras et al. 2009;
Cruces et al. 2017).

18.4.1  Phlorotannins and UV-Induced Oxidative Stress

UV radiation is a primary factor inducing ROS in seaweeds, mostly by increasing


the activity of peroxidases and NADPH oxidase (Mackerness et al. 2001). During
exposure to high levels of UV radiation, the xanthophyll cycle is inhibited, which
increases ROS production and impedes an effective dissipation of excess absorbed
excitation energy of photosynthetically active radiation (PAR) (Dring 2005; Bischof
et  al. 2006b; Lesser 2012). Thus, the increased electron fluxes result in a direct
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 373

photoreduction of oxygen in the Mehler reaction in PSI and lower activity of


Rubisco oxygenase, which exacerbate formation of ROS (e.g., superoxides) (Badger
et al. 2000). In cold temperate/Arctic kelp gametophytes, formation of ROS could
be demonstrated to occur predominantly in the peripheral cytoplasm or in plasmatic
vesicles (Müller et al. 2012). Under UV stress, a relationship between the content of
phlorotannins and antioxidant activity in various species of brown algae has been
reported.
Temperature can modify the phlorotannin response to UV radiation. In the sub-­
Antarctic Lessonia spicata, Durvillaea antarctica, and Macrocystis pyrifera from
the coast of Chile, a rapid induction of soluble phlorotannins triggered by UV radia-
tion ameliorated the effects of oxidative stress on photochemical processes after
short-term thermal stress (Cruces et  al. 2012). Interestingly, under elevated tem-
peratures >20 °C, the UV induction of phlorotannins ceases, and lipid peroxidation
increases in D. antarctica and L. spicata, suggesting that the ROS scavenging
potential of these sub-Antarctic species has a geographic component associated
with prevalent UV levels and temperature (Cruces et al. 2013). This can have impor-
tant consequences for stress tolerance in species living at the limits of their geo-
graphic distribution or those exposed to changing conditions of temperature and
solar radiation (e.g., polar, intertidal species). In fact, at a molecular level, oxidative
stress caused by UV radiation in the Arctic kelp Saccharina latissima is higher at
2 °C, which is reflected in upregulation of genes encoding for different ROS defense
mechanisms, especially antioxidant enzymes, but not phlorotannins (Heinrich et al.
2015). On the other hand, UV radiation does not directly regulate the expression of
genes involved in phlorotannin metabolism in the intertidal Fucus vesiculosus, sug-
gesting that UV induction of these substances relies on other processes or their
accumulation represents a constitutive metabolic strategy (Creis et  al. 2015).
Because phlorotannins act also as primary, structural cell components, their accu-
mulation depends on cellular cycles and biomass formation. Hence, the constitutive
nature of phlorotannins confers side-by-side advantages to brown algae under mul-
tiples stress factors, including UV radiation (Arnold and Targett 2003; Gómez and
Huovinen 2010).

18.4.2  Phlorotannins and Their Interaction with Metals

Although the relationship between phlorotannins of brown algae and metals is not
fully understood, increasing evidence indicates that metal tolerance of seaweeds can
be associated with both internal and external metal-complexing ligands (Andrade
et al. 2010; Connan and Stengel 2011). Decreased levels of soluble phenolic com-
pounds in seaweeds (e.g., Scytosiphon lomentaria and Ulva compressa) have been
reported in copper-impacted sites (Ratkevicius et al. 2003; Contreras et al. 2005).
Metals are redox active and also participate in many reactions generating ROS. The
importance of phenolic compounds as key antioxidant agents during metal exposure
has been recognized in plants (reviewed by Sakihama et al. 2002). UV radiation is
374 I. Gómez and P. Huovinen

known to induce or enhance the toxicity of certain organic contaminants (phototox-


icity) (Huovinen et al. 2001), and the presence of various metals has been reported
to have antagonistic effects on seaweeds (Andrade et al. 2006). In copper-impacted
areas, seaweeds (e.g. Ulva compressa) have been shown to develop oxidative stress,
and decreased levels of soluble phenolic compounds have been reported (Ratkevicius
et  al. 2003; Contreras et  al. 2005). Adverse effects of copper in the brown alga
Laminaria digitata were buffered by protective mechanisms regulated by lipid per-
oxide derivatives (Ritter et al. 2008). Proteins potentially involved in the control of
copper-mediated oxidative stress in the brown alga Scytosiphon gracilis were iden-
tified recently (Contreras et al. 2010). Species-specific antioxidant activity of the
soluble phlorotannins and its response to environmental stress (UV radiation, met-
als) has been shown in three Pacific kelps (Huovinen et al. 2010). Here, inorganic
nitrogen was shown to mitigate the adverse effects of copper: the impact of the
interaction of copper, nitrate, and UV radiation was species-specific, Lessonia spi-
cata showing the strongest responses in photosynthetic activity and Durvillaea ant-
arctica the strongest response in phlorotannins and their antioxidant activity.
Macrocystis pyrifera accumulated threefold more copper in its tissues than the other
kelps, but its photosynthetic activity was twofold less inhibited by copper than in
D. antarctica, suggesting higher metal tolerance of M. pyrifera, which was partly
explained by the decreased accumulation of copper in the algal tissues in the pres-
ence of nitrate (Huovinen et al. 2010).
Whether phlorotannins react increasing their ROS scavenging activity after
exposure to metals is not well known. When algae are exposed to metal stress,
increased exudation of organic compounds, including probably phlorotannins, may
retain free metals in form of extracellular complexing ligands. On the other hand,
detoxification of intracellular metals via algal exudates may also increase (Andrade
et al. 2010). Lessonia spicata from uncontaminated sites has been shown to have
capacity to rapidly respond to copper exposure by producing organic ligands that,
due to their complexing capacity in the water, can rapidly attenuate the level of
labile copper (Andrade et al. 2010), thus affecting its bioavailability.

18.5  Phlorotannins in Antarctic Seaweeds

An important feature of various endemic Antarctic brown algae is their high content
of phlorotannins. Different studies have reported total phlorotannin contents in
Antarctic brown algae ranging between 1% and 12% DW. In the case of insoluble
phlorotannins, values are between 1% and 5% DW (Table  18.1). Although some
cold-temperate genera (e.g., Fucus, Ascophyllum) can contain high concentrations
of phlorotannins (>10% DW), normally the maximal values detected in Antarctic
brown algae are higher than most of reported values from temperate, cold-­temperate,
and Arctic species (Connan et al. 2004; Dubois and Iken 2012; Cruces et al. 2013;
Generalíc-Mekiníc et al. 2019). The high concentrations of phlorotannins in some
endemic Antarctic brown algae can be also evidenced by the abundance of physodes
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 375

Fig. 18.4  Ultrastructure and localization of phlorotannin-containing physodes in Antarctic brown


algae. (a) Desmarestia anceps; (b) Phaeurus antarcticus; (c) Halopteris obovata. Left: cross sec-
tions stained with toluidine blue; right: transmission electron microscopy of cortical cells indicat-
ing the presence of physodes (phy). Cell wall (cw) and chloroplasts (chl) are indicated

in the outer cell layers (Fig. 18.4). Herbivory is among the factors that may contrib-
ute to these high levels of phlorotannins (see Chap. 17 in this volume); however,
other intrinsic factors related to biomass formation can also play an important role.
Whether these high phlorotannin levels found in endemic brown algae living
between 5 and 30 m depth can also be associated with photoprotection against high
376 I. Gómez and P. Huovinen

solar radiation has also been evaluated (Gómez and Huovinen 2015). In the follow-
ing sections of the present chapter, the variability in the contents of phlorotannins is
revised in the context of abiotic stress.

18.5.1  Depth Patterns in Phlorotannin Contents

The vertical distribution of endemic Antarctic brown algae can range from 1 to 2 m
down to 40 m or greater depths. This broad distribution is related with a suite of
photobiological adaptations operating in a range of different light fields (see Chap.
11 in this volume). Although Antarctic seaweeds are normally not exposed to high
UV levels, seasonal and oceanographic conditions can increase the eventual inci-
dence of harmful irradiances (the biological impact of this factor on different pro-
cesses related with algal distribution is revised in Chap. 11 by Gómez and Huovinen).
The hypothesis that phlorotannins of Antarctic seaweeds can also be related with the
light acclimation strategies has been tested in algae collected at different depths
(Fairhead et al. 2005a; Huovinen and Gómez 2013; Gómez and Huovinen 2015). In
eight brown algae collected along a depth gradient in King George Island, species
such as Desmarestia anceps, Cystosphaera jacquinotii, and Himantothallus grandi-
folius collected at depths >20 m showed the highest phlorotannin concentrations, in
contrast to shallow water or intertidal species such as Adenocystis utricularis or
Ascoseira mirabilis, which in general had the lowest values (Huovinen and Gómez
2013). However, when intraspecific variability of phlorotannins is examined, this
pattern can be different. In fact, Gómez and Huovinen (2015) analyzed the contents
of phlorotannins in conspecifics of Ascoseira mirabilis, Desmarestia anceps,
D. menziesii and Himantothallus grandifolius collected from 5/10, 20 to 30 m at
King George Island, and found that variation with depth was species-specific. For
example, in A. mirabilis, no changes with depth were detected, while in D. anceps
and D. menziesii, values increased in algae collected at 10 m depth compared to 20
or 30 m. Similar results have been reported in D. anceps from Anvers Island, West
Antarctic Peninsula, where higher phlorotannin contents were measured in shal-
lower locations (3–12  m) compared to samples collected between 18 and 30  m
depth (Fairhead et  al. 2005a). Although many factors can preclude a conclusive
comparison between different studies (e.g., time of collection, study site, and the
characteristics of the depth gradient or differences in depths between samples), the
results appear to indicate that (a) endemic Antarctic brown algae from depth >20 m
in general show constitutively high levels of phlorotannins, (b) phlorotannin con-
tents and their vertical variability mirror differences in life adaptations developed to
cope with multiple abiotic or biotic variables, and (c) phlorotannins form part of a
trade-off between shade adaptation marked by high photosynthetic efficiencies at
low light and tolerance to high solar stress. Thus, phlorotannins act as multifunc-
tional substances that can be “mobilized” in any situation that poses a threat to the
algae (Gómez and Huovinen 2015) (see Sect. 18.5.3).
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 377

18.5.2  Phlorotannin Allocation in Antarctic Seaweeds

Although the Antarctic is devoid of kelps, which resemble plants in being structur-
ally complex, many Antarctic Desmarestiales and Ascoseira and Cystosphaera
show a complex thallus anatomy with morpho-functional processes analogous to
those described in Laminariales and Fucales (see Chap. 11 in this volume). In this
context, it has been commonly observed that concentrations of phlorotannins vary
strongly among different thallus parts (Van Alstyne et al. 1999; Connan et al. 2004;
Iken et al. 2007), stimulating researchers to propose diverse hypothesis explaining
whether this observed variability is related with functional processes at an organis-
mal level. Whether the unequal allocation of phlorotannins to different thallus
regions is related with putative benefits for the alga, e.g., protection of metabolic
performance, reproductive output, or, in general, to guarantee the algal fitness dur-
ing environmental stress, is a relevant question. The optimal defense theory (ODM;
Rhoades 1979) is one of the ecological models used to explain the differential dis-
tribution of phlorotannins in the brown algal thalli, suggesting that chemical
defenses are produced in direct proportion to the risk, i.e., the phenolic compounds
would be produced at a direct expense of other functions (Pavia et al. 2002). High
concentrations of phlorotannins may be expected when an environmental pressure,
e.g., grazing, is high (inducible response) on thallus parts that make an important
contribution to the whole fitness (e.g., meristematic or reproductive regions) or dur-
ing seasonal periods when algae are especially vulnerable. Results in Fucus and
Ecklonia (Steinberg 1985; Yates and Peckol 1993) and the sub-Antarctic kelps
Lessonia spicata and Macrocystis pyrifera (Pansch et al. 2008) indicate that phloro-
tannins could vary as predicted by the ODT. As has been reported for Ascophyllum
nodosum, production of phlorotannins can be highly costly at the expense of growth
(Pavia et al. 1999). Thereby, it has been proposed that due to these costs, synthesis
and accumulation of phenolics could indicate inducible rather than constitutive
defenses (Rhoades 1979), which has been confirmed in some studies of simulated
herbivory (Lüder and Clayton 2004). However, studies carried out in some Antarctic
species indicate that phlorotannin allocation not necessarily confers chemical
defense consistent with the ODM assumptions. For example, regarding the “value”
of different thallus parts in relation with perennial and annual growth strategies,
D. anceps did not show differences in phlorotannin contents between thallus parts
(Fairhead et al. 2005a; Iken et al. 2007); however, there were marked differences in
the toughness, and the chemical defenses in primary stems/stipes were much higher
than the laterals supporting the ODT model (Fairhead et  al. 2005b). In contrast,
D. menziesii and Ascoseira mirabilis had higher phlorotannin concentrations in the
holdfasts compared to the branch or lamina regions. Due to that holdfasts were
regarded here as the most valuable thallus part conferring attachment, the patterns
in these species appear to meet well the ODT (Iken et al. 2007). It must be empha-
sized that the deterrent role of Antarctic phlorotannins against grazers and microbia
or as antifouling agents is species specific and probably depends on the type of
predominant phlorotannins and other not well-known qualitative properties of this
378 I. Gómez and P. Huovinen

compounds (reviewed in Iken et al. 2009). The results agree with longitudinal pro-
files determined in the cold-temperate kelps Laminaria hyperborea and Laminaria
digitata (Connan et al. 2006) and Lessonia spicata (Gómez et al. 2016), where valu-
able regions and basal parts such as haptera or holdfasts and meristematic tissues
allocated the highest phlorotannins compared to the fronds, which can be regarded
as transient structures. These patterns can be associated with various longitudinal
profiles of physiological performance normally described for various kelps (Van
Alstyne et al. 1999; Gómez et al. 2005; Gruber et al. 2011).
In the case of photoprotective responses, it could be reasonable to argue that
valuable thallus regions, e.g., reproductive tissues, should be protected when they
are exposed to UV radiation (Holzinger et al. 2011). This hypothesis has also been
tested in two Antarctic brown algae by exposing reproductive and vegetative thallus
pieces to UV radiation during a short-term period (Huovinen and Gómez 2015). In
the brown alga Cystosphaera jacquinotii, the reproductive structures (receptacles
containing conceptacles) showed higher UV tolerance than its vegetative blades,
whereas in Ascoseira mirabilis, high UV tolerance was demonstrated in both vege-
tative and reproductive tissues. Interestingly, the reproductive structures of both
species of brown algae had higher levels of soluble phlorotannins than the vegeta-
tive tissues, and thus, allocation and proportions of soluble and insoluble, cell-wall-­
bound phlorotannins could be related with the observed patterns of UV tolerance of
the different tissues. Observations of tissue cross sections under violet-blue light
excitation using epifluorescence microscopy confirmed a high allocation of pheno-
lic compounds (as blue autofluorescence) in C. jacquinotii, especially in its repro-
ductive structures (Fig. 18.5a). The study is among the first approaches to address
the defense strategies that Antarctic macroalgae exploit to protect their reproductive
structures. It is likely that the allocation of chemical defenses and UV-absorbing
compounds in reproductive tissues is a widespread strategy to ensure the viability of
spores and gametes during their maturation. For example, blue autofluorescence,
indicating the presence of phenolics compounds in reproductive tissues (carporan-
gia) of the Antarctic red algae Trematocarpus antarcticus (Fig. 18.5b), suggests that
not only phlorotannins but also other phenolics can be allocated providing protec-
tion to reproductive tissues.
It has been suggested that phlorotannins can be remobilized between tissues with
different metabolic demand and age (Arnold and Targett 2000). In Laminariales,
compounds as mannitol, amino acids, and other low-molecular-weight compounds
are transported through specialized cells to power meristematic growth through
translocation processes (Küppers and Kremer 1978; Gómez and Huovinen 2012).
Since phlorotannins are structural components in cells, it may be intuitively sug-
gested that these compounds or some key precursors may be mobilized along the
thallus. The abundance of low-molecular-weight phlorotannins (<1200 Da) in vari-
ous species of brown algae (Steevensz et al. 2012) supports also the idea that these
compounds might be rapidly remobilized. For example, accumulation of phlorotan-
nins in response to artificial wounding in the kelp Ecklonia radiata, including the
presence of physodes in medullary sieve elements (Lüder and Clayton 2004), sug-
gests that these compounds can be “transported” along the thallus. In fact, some
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 379

Fig. 18.5  Blue autofluorescence of phenolic compounds under violet-blue light excitation in (a)
reproductive receptacles of the brown alga Cystosphaera jacquinotii and (b) carposporangium in
the Antarctic red alga Trematocarpus antarcticus; (c) cross section of non-reproductive lamina of
brown alga Ascoseira mirabilis indicating presence of phenolic compounds in medullar “conduct-
ing channels”

Antarctic species such as Ascoseira mirabilis appear to have the anatomical prereq-
uisites as this alga shows “conducting channels” in its medulla (Fig. 18.5c), which
are suggested to have putative functions in remobilization of substances (Clayton
and Ashburner 1990). Overall, phlorotannins are important structural elements in
the algal thallus, and since growth is localized in specific regions, a trade-off
between phlorotannin synthesis, mobilization, and growth might be defined in these
species in a similar way as in terrestrial vascular plants.

18.5.3  Phlorotannins in Response to UV Radiation

Experimental evidence has pointed to a relatively high tolerance of Antarctic sea-


weeds to UV radiation in the short-term, which at least partly can be related to their
chemical defense mechanism based on phenolic substances. Indeed, Antarctic algae
exposed for 2 h to UV radiation at 2 °C showed very low inhibition of photosynthe-
sis measured as maximal quantum yield of fluorescence (Fv/Fm), which can reach up
to 35% in algae collected from depth >20 m. Even in algae growing at 30 m, inhibi-
tion of chlorophyll fluorescence did not exceed 10–15% (Huovinen and Gómez
380 I. Gómez and P. Huovinen

2013). In the case of the brown algae, almost all have high levels of phlorotannins,
which due to their UV-absorbing characteristics are the main candidates conferring
photoprotection in these species (Huovinen and Gómez 2013; Gómez and Huovinen
2015; Núñez-Pons et al. 2018). However, testing these properties experimentally is
not an easy task. In fact, manipulative studies conducted in algae attaining high
concentrations of phenols, e.g., Desmarestia anceps, have not demonstrated induc-
tion in phlorotannins in response to UV (Fairhead et al. 2006; Gómez and Huovinen
2015; Flores-Molina et al. 2016). In contrast, some species with relatively low con-
centrations, such as Ascoseira mirabilis, show a slightly UV-mediated induction of
soluble phlorotannins (Rautenberger et al. 2015). As this species is normally found
at shallower depths (1–10  m), the results suggest that it can become exposed to
harmful solar radiation in summer, thus activating the synthesis of phlorotannins. In
all, photoprotection against excess solar radiation, e.g., via UV shielding, is a col-
lateral function in Antarctic seaweeds as these molecules form part of an integral
defense machinery operating in response to multiple stressors in the polar environ-
ment such as herbivores, antifouling, and changes in temperature or simply they are
synthesized to supply of structural elements during cell growth (as insoluble phlo-
rotannins). These multiple functional roles are explained by their high antioxidant
capacity, the most important chemical property of phlorotannins. In fact, soluble
phlorotannins have been positively correlated with the high antioxidant potential
determined in extracts of various species of Antarctic brown algae. This positive
correlation is observed in algae exposed to different conditions of UV radiation and
temperatures (Gómez and Huovinen 2015; Flores-Molina et al. 2016). Interestingly,
it has been demonstrated that UV effects on photosynthesis in Antarctic macroalgae
are modified by temperature: when algae are incubated at 7 °C, i.e., 5 °C above the
field temperature, inhibition of photosynthesis decreases, and recovery increases,
suggesting that, e.g., the PSII repair cycle is more effective at elevated temperature
resulting in a higher UV tolerance, at least in the short-term (Rautenberger et al.
2015). However, the relationship between the phlorotannin contents and the antioxi-
dant potential of extracts does not change with temperature (Fig. 18.6), reinforcing
the idea that in these species, phlorotannins are not UV-inducible compounds. This
raises questions related to the role of these compounds and their physiological con-
sequences under changing environmental conditions. For example, it is known that
during oxidative stress, Antarctic brown algae can active their enzymatic machinery
(e.g., superoxide dismutase, SOD), whose efficiency varies in response to environ-
mental gradients (Bischof and Rautenberger 2012). Thus, the operation of comple-
mentary mechanisms of ROS detoxifying less affected by, e.g., temperature or UV
radiation could be favored. In this context, it has been reported recently that under
high solar stress conditions, algae display a suite of complementary and consecutive
protective mechanisms based on energy dissipative downregulation of photosynthe-
sis, rapid pigment acclimation and PSII repair mechanisms, synthesis of phenolics
with specific UV absorption characteristics, and complementary ROS scavenging
mediated by antioxidant enzymes and phenols (Cruces et al. 2017).
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 381

Fig. 18.6  Relationship between phlorotannin content and antioxidant activity of extracts from
four Antarctic brown algae after exposure to UV radiation under different temperatures (Adapted
from Rautenberger et al. 2015)

18.6  Concluding Remarks

Antarctic macroalgae are equipped with a suite of anti-stress mechanisms to cope


with the harsh polar environment, probably much more sophisticated and efficient
than previously thought. However, in the case of the putative roles of phlorotannins
in stress tolerance, many gaps still persist, especially those related with their synthe-
sis, the action of different forms of phlorotannins, and their turnover and regulation.
In the case of Antarctic algae, no data on gene expression exist, which precludes an
understanding of crucial aspects related with their multifunctional roles.
In Antarctic brown algae, processes related with synthesis and functional dynam-
ics of phlorotannins take place under constantly low temperatures of 0–2 °C, and
phlorotannin induction can be sensitive to changes in temperate and cold-temperate
species. This raises questions related with the effects of shifts in temperature on the
anti-stress properties of these substances in endemic Antarctic algae, where induc-
tion has not been reported.
Finally, the role of phlorotannins in response to new and emergent stressors, e.g.,
pollutants and acidification (OA), has been very little studied in Antarctic algae.
Although recent experimental essays carried out in Desmarestia anceps and D.
382 I. Gómez and P. Huovinen

menziesii indicated that a combination of low pH and elevated temperature does not
result in marked effects in phlorotannin content, the shifts can have important and
not well-understood ecophysiological consequences (Schoenrock et al. 2015).

Acknowledgments  The authors acknowledge the financial support from Conicyt Chile (FONDAP
15150003, Anillo-PIA ART1101, Fondecyt 1161129) and INACH T-20-09 from the Instituto
Antártico Chileno. The helpful assistance and collaboration of the members of our laboratories at
Universidad Austral de Chile as well as the staff of the Instituto Antártico Chileno during various
Antarctic expeditions are acknowledged.

References

Aguilera J, Bischof K, Karsten U, Hanelt D, Wiencke C (2002) Seasonal variation in ecophysi-


ological patterns in macroalgae from an Arctic fjord. II. Pigment accumulation and biochemi-
cal defence systems against high light stress. Mar Biol 140:1087–1095
Ahn GN, Kim KN, Cha SH, Song CB, Lee J, Heo MS et al (2007) Antioxidant activities of phloro-
tannins purified from Ecklonia cava on free radical scavenging using ESR and H2O2-mediated
DNA damage. Eur Food Res Tech 226:71–79. https://doi.org/10.1007/s00217-006-0510-y
Andrade S, Medina MH, Moffett JW, Correa JA (2006) Cadmium–copper antagonism in seaweeds
inhabiting coastal areas affected by copper mine waste disposals. Env Sci Tech 40:4382–4387
Andrade S, Pulido MJ, Correa JA (2010) The effect of organic ligands exuded by intertidal sea-
weeds on copper complexation. Chemosphere 78:397–401
Arnold TM, Targett NM (2000) Evidence for metabolic turnover of polyphenolics in tropical
brown algae. J Chem Ecol 26:1393–1410
Arnold TM, Targett NM, Tanner CE, Hatch WI, Ferrari KE (2001) Evidence for methyl jasmonate-­
induced phlorotannin production in Fucus vesiculosus (Phaeophyceae). J Phycol 37:1026–1029
Arnold TM, Targett NM (2002) Marine tannins: the importance of a mechanistic framework for
predicting ecological roles. Mini review. J Chem Ecol 28:1919–1934
Arnold TM, Targett NM (2003) To grow and defend: lack of tradeoffs for brown algal phlorotan-
nins. Oikos 100:406–408
Athukorala Y, Lee K-W, Song C, Ahn C-B, Shin T-S, Cha Y-J, Shahidi F, Jeon Y-J (2003) Potential
antioxidant activity of marine red alga Grateloupia filicina extracts. J Food Lipids 10:251–265
Badger MR, von Caemmerer S, Ruuska S, Nakano H (2000) Electron flow to oxygen in higher
plants and algae: rates and control of direct photoreduction (Mehler reaction) and rubisco oxy-
genase. Philos Trans R Soc Lond Ser B Biol Sci 355(1402):1433–1446. https://doi.org/10.1098/
rstb.2000.0704
Bischof K, Gómez I, Molis M, Hanelt D, Karsten U, Lüder U, Roleda MY, Zacher K, Wiencke
C (2006a) Ultraviolet radiation shapes seaweed communities. Rev Environ Sci Biotechnol
5:141–166
Bischof K, Rautenberger R, Brey L, Perez-Llorens JL (2006b) Physiological acclimation to gradi-
ents of solar irradiance within mats of the filamentous green macroalga Chaetomorpha linum
from southern Spain. Mar Ecol Prog Ser 306:165–175
Bischof K, Rautenberger R (2012) Seaweed responses to environmental stress: reactive oxygen
and antioxidative strategies. In: Wiencke C, Bischof K (eds) Seaweed biology: novel insights
into ecophysiology, ecology and utilization, ecological studies, vol 219. Springer-Verlag,
Berlin, pp 109–132
Clarke LJ, Robinson SA (2008) Cell wall-bound ultraviolet-screening compounds explain the high
ultraviolet tolerance of the Antarctic moss, Ceratodon purpureus. New Phytol 179:776–783
Clayton MN, Beakes GW (1983) Effects of fixatives on the ultrastructure of physodes in vegetative
cell of Scytosiphon lomentaria (Scytosiphonaceae, Phaeophyta). J Phycol 19:416
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 383

Clayton MN, Ashburner CM (1990) The anatomy and ultrastructure of “conducting channels” in
Ascoseira mirabilis (Ascoseirales, Phaeophyceae). Bot Mar 33:63–70
Cockell CS, Knowland J (1999) Ultraviolet radiation screening compounds. Biol Rev 74:311–345
Connan S, Goulard F, Stiger V, Deslandes E, Ar Gall E (2004) Interspecific and temporal variation
in phlorotannin levels in an assemblage of brown algae. Bot Mar 47:410–416
Connan S, Delisle F, Deslandes E, Ar Gall E (2006) Intra-thallus phlorotannin content and antioxi-
dant activity in Phaeophyceae of temperate waters. Bot Mar 49:39–46
Connan S, Stengel DB (2011) Impacts of ambient salinity and copper on brown algae: 2. Interactive
effects on phenolic pool and assessment of metal binding capacity of phlorotannin. Aquat
Toxicol 104:1–13
Contreras L, Moenne A, Correa JA (2005) Antioxidant responses in Scytosiphon lomentaria
(Phaeophyta) inhabiting copper-enriched coastal environments. J Phycol 41:1184–1195
Contreras L, Mella D, Moenne A, Correa JA (2009) Differential responses to copper induced
oxidative stress in the marine macroalgae Lessonia nigrescens and Scytosiphon lomentaria
(Phaeophyceae). Aquat Toxicol 94:94–102
Contreras L, Moenne A, Gaillard F, Potin P, Correa JA (2010) Proteomic analysis and identification
of copper stress-regulated proteins in the marine alga Scytosiphon gracilis (Phaeophyceae).
Aquat Toxicol 96:85–89
Creis E, Delage L, Charton S, Goulitquer S, Leblanc C, Potin P, Ar Gall E (2015) Constitutive or
inducible protective mechanisms against UV-B radiation in the brown alga Fucus vesiculosus?
A study of gene expression and phlorotannin content responses. PLoS One 10(6):e0128003.
https://doi.org/10.1371/journal.pone.0128003
Cruces E, Huovinen P, Gómez I (2012) Phlorotannin and antioxidant responses upon short term
exposure to UV radiation and elevated temperature in three South Pacific kelps. Photochem
Photobiol 88:58–66
Cruces E, Huovinen P, Gómez I (2013) Interactive effects of UV radiation and enhanced tempera-
ture on photosynthesis, phlorotannin induction and antioxidant activities of two sub-Antarctic
brown algae. Mar Biol 160:1–13
Cruces E, Rautenberger R, Rojas-Lillo Y, Cubillos VM, Ramirez-Kushel E, Gómez I (2017)
Physiological acclimation of Lessonia spicata to diurnal changing PAR and UV radiation: dif-
ferential regulation among down-regulation of photochemistry, ROS scavenging activity and
phlorotannins as major photoprotective mechanisms. Photosynth Res 131:145–157. https://doi.
org/10.1007/s11120-016-0304-4
Cruces E, Flores MR, Díaz MJ, Huovinen P, Gómez I (2018) Phenolics as photoprotective mecha-
nism against combined action of UV radiation and temperature in the red alga Gracilaria
chilensis? J Appl Phycol 30(2):1247–1257. https://doi.org/10.1007/s10811-017-1304-2
Dring M (2005) Stress resistance and disease resistance in seaweeds: the role of reactive oxygen
metabolism. Adv Bot Res 43:175–207
Dubois A, Iken K (2012) Seasonal variation in kelp phlorotannins in relation to grazer abundance
and environmental variables in the Alaskan sublittoral zone. Algae 27(1):9–19. https://doi.
org/10.4490/algae.2012.27.1.009
Evans LV, Holligan MS (1972) Correlated light and electron microscope studies on brown algae.
II. Physode production in Dictyota. New Phytol 71:1173–1180
Fairhead VA, Amsler CD, McClintock JB, Baker BJ (2005a) Variation in phlorotannin con-
tent within two species of brown macroalgae (Desmarestia anceps and D. menziesii)
from the Western Antarctic Peninsula. Polar Biol 28:680–686. https://doi.org/10.1007/
s00300-005-0735-4
Fairhead VA, Amsler CD, McClintock JB, Baker BJ (2005b) Within-thallus variation in chemical
and physical defenses in two species of ecologically dominant brown macroalgae from the
Antarctic Peninsula. J Exp Mar Biol Ecol 322:1–12
Fairhead VA, Amsler CD, McClintock JB, Baker BJ (2006) Lack of defense or phlorotannins induc-
tion by UV radiation or mesograzers in Desmarestia anceps and D. menziesii (Phaeophyceae).
J Phycol 42:1174–1183
384 I. Gómez and P. Huovinen

Feldmann G, Guglielmi MG (1972) Les physodes et les corps irisants du Dictyota dichotoma
(Hudson) Lamouroux. Comptes Rendus de l’Academie des. Sciences 275:751–754
Flores-Molina MR, Muñoz P, Rautenberger R, Huovinen P, Gómez I (2016) Stress tolerance to
UV radiation and temperature of the endemic Antarctic brown alga Desmarestia anceps is
mediated by high concentrations of phlorotannins. Photochem Photobiol 92:455–466. https://
doi.org/10.1111/php.12580
García-Pichel F, Castenholz RW (1993) Occurrence of UV-absorbing, mycosporine-like com-
pounds among cyanobacterial isolates and an estimate of their screening capacity. Appl
Environ Microbiol 59:170–176
Generalíc-Mekiníc I, Skroza D, Šimat V, Hamed I, Câgalj M, Popovíc-Perkovíc Z (2019) Phenolic
content of brown algae (Phaeophyceae) species: extraction, identification, and quantification.
Biomol Ther 9:244. https://doi.org/10.3390/biom9060244
Gledhill M, Nimmo M, Hill SJ, Brown MT (1999) The release of copper-complexing ligands by
the brown alga Fucus vesiculosus (Phaeophyceae) in response to increasing total copper levels.
J Phycol 35:501–509
Gómez I, Huovinen P (2010) Induction of phlorotannins during UV exposure mitigates inhibition
of photosynthesis and DNA damage in the kelp Lessonia nigrescens. Photochem Photobiol
86:1056–1063
Gómez I, Huovinen P (2012) Morpho-functionality of carbon metabolism in seaweeds. In:
Wiencke C, Bischof K (eds) Seaweed biology: novel insights into ecophysiology, ecology and
utilization, ecological studies, vol 219. Springer-Verlag, Berlin, pp 25–46
Gómez I, Huovinen P (2015) Lack of physiological depth patterns in conspecifics of endemic
Antarctic brown algae: a trade-off between UV stress tolerance and shade adaptation? PLoS
One 10(8):e0134440
Gómez I, Pérez-Rodríguez E, Viñegla B, Figueroa FL, Karsten U (1998) Effects of solar radia-
tion on photosynthesis, UV-absorbing compounds and enzyme activities of the green alga
Dasycladus vermicularis from southern Spain. J Photochem Photobiol B Biol 47:46–57
Gómez I, Ulloa N, Orostegui M (2005) Morpho-functional patterns of photosynthesis ad UV sen-
sitivity in the kelp Lessonia nigrescens (Laminariales, Phaeophyta). Mar Biol 148:231–240
Gómez I, Véliz K, Español S, Huovinen P (2016) Spatial distribution of phlorotannins and its rela-
tionship with photosynthetic UV tolerance and allocation of storage carbohydrates in blades
of the kelp Lessonia spicata. Mar Biol 163:110. https://doi.org/10.1007/s00227-016-2891-1
Gruber A, Roleda MY, Bartsch I, Hanelt D, Wiencke C (2011) Sporogenesis under ultraviolet radi-
ation in Laminaria digitata (Phaeophyceae) reveals protection of photosensitive meiospores
within soral tissue: physiological and anatomical evidence. J Phycol 47:603–614
Heffernan N, Smyth TJ, Soler-Villa A, Fitzgerald RJ, Brunton NP (2014) Phenolic content and
antioxidant activity of fractions obtained from selected Irish macroalgae species (Laminaria
digitata, Fucus serratus, Gracilaria gracilis and Codium fragile). J Appl Phycol 27:519.
https://doi.org/10.1007/s10811-014-0291-9
Heinrich S, Valentin K, Frickenhaus K, Wiencke C (2015) Temperature and light interactively
modulate gene expression in Saccharina latissima (Phaeophyceae). J Phycol 51:93–108
Henry BE, Van Alstyne KL (2004) Effects of UV radiation on growth and phlorotannins in Fucus
gardneri (Phaeophyceae) juveniles and embryos. J Phycol 40:527–533
Herbert RB (ed) (1989) The biosynthesis of secondary metabolites, 2nd edn. Chapman and Hall,
New Yorkk
Hideg É, Jansen MAK, Strid A (2013) UV-B exposure, ROS, and stress: inseparable compan-
ions or loosely linked associates? Trends Plant Sci 18:107–115. https://doi.org/10.1016/j.
tplants.2012.09.003
Holzinger A, Di Piazza L, Lütz C, Roleda MY (2011) Sporogenic and vegetative tissues of
Saccharina latissima (Laminariales, Phaeophyceae) exhibit distinctive sensitivity to experi-
mentally enhanced ultraviolet radiation: photosynthetically active radiation ratio. Phycol Res
59:221–235
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 385

Huovinen P, Gómez I (2013) Photosynthetic characteristics and UV stress tolerance of Antarctic


seaweeds along the depth gradient. Polar Biol 36:1319–1332
Huovinen P, Gómez I (2015) UV Sensitivity of vegetative and reproductive tissues of three
Antarctic macroalgae is related to differential allocation of phenolic substances. Photochem
Photobiol 91:1382–1388. https://doi.org/10.1111/php.12500
Huovinen PS, Soimasuo MR, Oikari AOJ (2001) Photoinduced toxicity of retene to Daphnia
magna under enhanced UV-B radiation. Chemosphere 45:683–691
Huovinen P, Leal P, Gómez I (2010) Impact of interaction of copper, nitrogen and UV radiation on
the physiology of three south Pacific kelps. Mar Freshw Res 61:330–341
Iken K, Amsler CD, Hubbard JM, McClintock JB, Baker BJ (2007) Allocation patterns of phloro-
tannins in Antarctic brown algae. Phycologia 46(4):386–395. https://doi.org/10.2216/06-67.1
Iken K, Amsler CD, Amsler MO, McClintock JB, Baker BJ (2009) Field studies on deterrent roles
of phlorotannins in Antarctic brown algae. Bot Mar 52:547–557
Jormalainen V, Honkanen T, Koivikko R, Eränen J (2003) Induction of phlorotannin production in
a brown alga: defense or resource dynamics? Oikos 103:640–650
Karentz D, McEuen FS, Land MC, Dunlap WC (1991) Survey of mycosporine-like amino acid
compounds in Antarctic marine organisms: potential protection from ultraviolet exposure. Mar
Biol 108:157–166
Kaur I, Vijayaraghavan MR (1992) Physode distribution and genesis in Sargassum vulgare
C. Agardh and Sargassum johnstonii Setchell & Gardner. Aquat Bot 42:375–384
Kevekordes K, Clayton MN (1999) Shedding of the zygote wall by Durvillaea potatorum
(Durvillaeales, Phaeophyta) embryos. Eur J Phycol 34:65–70
Koivikko R, Loponen J, Honkanen T, Jormalainen V (2005) Contents of soluble, cell-wall-bound
and exuded phlorotannins in the brown alga Fucus vesiculosus, with implications on their eco-
logical function. J Chem Ecol 31:195–212
Koivikko R, Loponen J, Pihlaja K, Jormalainen V (2007) High-performance liquid chromato-
graphic analysis of phlorotannins from the brown alga Fucus vesiculosus. Phytochem Anal
18(4):326–332
Küpper FC, Gaquerel E, Cosse A, Adas F, Peters AF et al (2009) Free fatty acids and methyl jas-
monate trigger defense reactions in Laminaria digitata. Plant Cell Physiol 50:789–800
Küppers U, Kremer BP (1978) Longitudinal profiles of carbon dioxide fixation capacities in
marine macroalgae. Plant Physiol 62:49–53
La Barre S, Potin P, Leblanc C, Delage L (2010) The halogenated metabolism of brown algae
(Phaeophyta), its biological importance and its environmental significance. Mar Drugs
8:988–1010. https://doi.org/10.3390/md8040988
Lesser M (2012) Oxidative stress in tropical marine ecosystems. In: Abele D, Vazquez-Medina
JP, Zenteno-Savin T (eds) Oxidative stress in aquatic ecosystems. Blackwell Publishing Ltd,
Oxford, pp 9–19
Lüder UH, Clayton MN (2004) Induction of phlorotannins in the brown macroalga Ecklonia
radiata (Laminariales, Phaeophyta) in response to simulated herbivory – the first microscopic
study. Planta 218:928–937
Mackerness AHS, John CF, Jordan B, Thomas B (2001) Early signaling components in ultraviolet-
­B responses: distinct roles for different reactive oxygen species and nitric oxide. FEBS Lett
489(2–3):237–242
Menzel D, Kazlauskas R, Reichelt J (1983) Coumarins in the siphonalean green algal family
Dasycladaceae Kützing (Chlorophyceae). Bot Mar 29:23–29
Meslet-Cladiere L, Delage L, Leroux CJ, Goulitquer S, Leblanc C, Creis E, Gall EA et al (2013)
Structure/function analysis of a type III polyketide synthase in the brown alga Ectocarpus
siliculosus reveals a biochemical pathway in phlorotannins monomer biosynthesis. Plant Cell
25:3089–3103. https://doi.org/10.1105/tpc.113.111336
Michalak A (2006) Phenolic compounds and their antioxidant activity in plants growing under
heavy metal stress. Polish J Environ Stud 15(4):523–530
386 I. Gómez and P. Huovinen

Müller R, Desel C, Steinhoff FS, Wiencke C, Bischof K (2012) UV-radiation and elevated temper-
atures induce formation of reactive oxygen species in gametophytes of cold temperate/Arctic
kelps (Laminariales, Phaeophyceae). Phycol Res 60:27–36
Nakai M, Kageyama N, Nakahara K, Miki W (2006) Phlorotannins as radical scavengers from the
extract of Sargassum ringgoldianum. Mar Biotechnol 8:409–414
Núñez-Pons L, Avila C, Romano G, Verde C, Giordano D (2018) UV-protective compounds in
marine organisms from the Southern Ocean. Mar Drugs 16(9):336. https://doi.org/10.3390/
md16090336
Oliveira L, Bisalputra T (1973) Studies in the brown alga Ectocarpus in culture. I. General ultra-
structure of the sporophytic vegetative cells. J Submicrosc Cytol 5:107–120
Pangestuti R, Siahaan EA, Kim SK (2018) Photoprotective substances derived from marine algae.
Mar Drugs 16(11):399. https://doi.org/10.3390/md16110399
Pansch C, Gómez I, Rothäusler E, Véliz K, Thiel M (2008) Species-specific defense strategies of
vegetative versus reproductive blades of the Pacific kelps Lessonia nigrescens and Macrocystis
integrifolia. Mar Biol 155:51–62. https://doi.org/10.1007/s00227-008-1006-z
Paul N, Gwynn-Jones D (2003) Ecological roles of solar UV radiation: towards an integrated
approach. Trends Ecol Evol 18:48–55
Pavia H, Toth G (2000) Inducible chemical resistance to herbivory in the brown seaweed
Ascophyllum nodosum. Ecology 81:3212–3225
Pavia H, Cervin G, Lindaren A, Åberg P (1997) Effects of UV-B radiation and simulated herbivory
on phlorotannins in the brown alga Ascophyllum nodosum. Mar Ecol Progr Ser 157:139–146
Pavia H, Toth G, Åberg P (1999) Trade-offs between phlorotannin production and annual growth in
natural populations of the brown seaweed Ascophyllum nodosum. J Ecol 87:761–771
Pavia H, Toth GB, Åberg P (2002) Optimal defense theory: elasticity analysis as a tool to predict
intraplant variation in defenses. Ecology 83:891–897
Pellegrini L (1980) Cytological studies on physodes in the vegetative cells of Cystoseira
stricta Sauvageau (Phaeophyta, Fucales). J Cell Sci 41:209–231
Pérez-Rodríguez E, Gómez I, Karsten U, Figueroa FL (1998) Effects of UV radiation on photosyn-
thesis and excretion of UV-absorbing compounds of Dasycladus vermicularis (Dasycladales,
Chlorophyta) from southern Spain. Phycologia 37:379–387
Pérez-Rodríguez E, Aguilera J, Gómez I, Figueroa FL (2001) Excretion of coumarins by the
Mediterranean green alga Dasycladus vermicularis in response to environmental stress. Mar
Biol 139:633–639
Potin P, Leblanc CL (2006) Phenolic-based adhesives of marine brown algae. In: Smith
AM, Callow JA (eds) Biological adhesive. Springer, Berlin, pp  105–124. https://doi.
org/10.1007/978-3-540-31049-5_6
Ragan MA, Glombitza K-W (1986) Phlorotannins, brown algal polyphenols. Prog Phycol Res
4:130–241
Ratkevicius N, Correa JA, Moenne A (2003) Copper accumulation, synthesis of ascorbate and
activation of ascorbate peroxidase in Enteromorpha compressa (L.) Grev. (Chlorophyta) from
heavy metal-enriched environments in northern Chile. Plant Cell Environ 26:1599–1608
Rautenberger R, Huovinen P, Gómez I (2015) Effects of increased seawater temperature on
UV-tolerance of Antarctic marine macroalgae. Mar Biol 162:1087–1097
Rhoades DF (1979) Evolution of plant chemical defense against herbivores. In: Rosenthal GA,
Janzen DH (eds) Herbivores: their interaction with secondary plant metabolites. Academic
Press, New York, pp 1–55
Ritter A, Goulitquer S, Salaün JP, Tonon T, Correa JA, Potin P (2008) Copper stress induces biosyn-
thesis of octadecanoid and eicosanoid oxygenated derivatives in the brown algal kelp Laminaria
digitata. New Phytol 180(4):809–821. https://doi.org/10.1111/j.1469-8137.2008.02626.x
Roleda MY, Lüder UH, Wiencke C (2010) UV-susceptibility of zoospores of the brown macroalga
Laminaria digitata from Spitsbergen. Polar Biol 33:577–588
Ross C, Küpper FC, Vreeland V, Waite JH, Jacobs RS (2005) Evidence of a latent oxidative burst
in relation to wound repair in the giant unicellular chlorophyte Dasycladus vermicularis. J
Phycol 41:531–541
18  Brown Algal Phlorotannins: An Overview of Their Functional Roles 387

Sakihama Y, Cohen MF, Grace SC, Yamasaki H (2002) Plant phenolic antioxidant and prooxi-
dant activities: phenolics-induced oxidative damage mediated by metals in plants. Toxicology
177:67–80
Salgado LT, Cinelli LP, Viana NB, de Carvalho RT, de Souza Mourão PA et al (2009) A vana-
dium bromoperoxidase catalyzes the formation of high-molecular-weight complexes
between brown algal phenolic substances and alginates. J Phycol 45:193–202. https://doi.
org/10.1111/j.1529-8817.2008.00642.x
Shibata T, Hama Y, Miyasaki T, Ito M, Nakamura T (2006) Extracellular secretion of pheno-
lic substances from living brown algae. J Appl Phycol 18:787–794. https://doi.org/10.1007/
s10811-006-9094-y
Schoenrock KM, Schram JB, Amsler CD, McClintock JB, Angus RA (2015) Climate change
impacts on overstory Desmarestia spp. from the western Antarctic Peninsula. Mar Biol
162:377–389. https://doi.org/10.1007/s00227-014-2582-8
Schoenwaelder MEA (2002) The occurrence and cellular significance of physodes in brown algae.
Phycologia 41:125–139
Schoenwaelder MEA, Clayton MN (1998) Secretion of phenolic substances into zygote wall
and cell plate in embryos of Hormosira and Acrocarpia (Fucales, Phaeophyta). J Phycol
34:969–980
Schoenwaelder MEA, Clayton MN (1999) The presence of phenolic compounds in isolated cell
walls of brown algae. Phycologia 38:161–166
Schoenwaelder MEA, Wiencke C, Clayton MN, Glombitza KW (2003) The effect of elevated UV
radiation on Fucus spp. (Fucales, Phaeophyta) zygote and embryo development. Plant Biol
5:366–377
Smith KL, Harwood JL (1986) The subcellular localization of absorbed copper in Fucus. Physiol
Plant 66:692–698
Stauber JL, Florence TM (1987) Mechanisms of toxicity of ionic copper and copper complexes in
algae. Mar Biol 94:511–519
Steevensz AJ, MacKinnon SL, Hankinson R, Craft C, Connan S, Stengel DB, Melanson JE (2012)
Profiling phlorotannins in brown macroalgae by liquid chromatography–high resolution mass
spectrometry. Phytochem Anal 23(5):547–553
Steinberg PD (1985) Feeding preferences of Tegula funebralis and chemical defences of marine
brown algae. Ecol Monogr 5:333–349
Steinhoff F (2010) Phlorotannins as UV-protective substances in early developmental stages of
brown algae, PhD thesis, University of Bremen
Steinhoff FS, Graeve M, Wiencke C, Wulff A, Bischof K (2011) Lipid content and fatty acid
consumption in zoospores/ developing gametophytes of Saccharina latissima (Laminariales,
Phaeophyceae) as potential precursors for secondary metabolites as phlorotannins. Polar Biol
34:1011. https://doi.org/10.1007/s00300-011-0960-y
Swanson AK, Druehl LD (2002) Induction, exudation and the UV protective role of kelp phloro-
tannins. Aquat Bot 73:241–253
Swanson AK, Fox C (2007) Altered kelp (Laminariales) phlorotannins and growth under elevated
carbon dioxide and ultraviolet-B treatments can influence associated intertidal food webs. Glob
Chang Biol 13:1696–1709
Toth GB, Pavia H (2002) Lack of phlorotannin induction in the kelp Laminaria hyperborea in
response to grazing by two gastropod herbivores. Mar Biol 140:403–409
Van Alstyne KL, Pelletreau KN (2000) Effects of nutrient enrichment on growth and phlorotannin
production in Fucus gardneri embryos. Mar Ecol Prog Ser 206:33–43
Van Alstyne KL, McCarthy JJ, Hustead CL, Kearns LJ (1999) Phlorotannin allocation among
tissues of northeastern Pacific kelps and rockweeds. J Phycol 35:483–492. https://doi.
org/10.1046/j.1529-8817.1999.3530483.x
Vilter H (1995) Vanadium-dependent haloperoxidases. In: Sigel H, Sigel AM (eds) Vanadium and
its role in life. Metal ions in biological systems, vol 31. Dekker, New York, pp 325–352
388 I. Gómez and P. Huovinen

Vreeland V, Laetsch WM (1988) Role of alginate self-associating subunits in the assembly


of Fucus embryo cell walls. In: Varner JE (ed) Self assembling architecture. Alan R.  Liss,
New York, pp 77–96
Wang T, Jónsdóttir R, Ólafsdóttir G (2009) Total phenolic compounds, radical scavenging and
metal chelation of extracts from Icelandic seaweeds. Food Chem 116:240–248
Yates JL, Peckol P (1993) Effects of nutrient availability and herbivory on polyphenolics in the
seaweed Fucus vesiculosus. Ecology 74:1757–1766
Yildiz G, Vatan Ö, Çelikler S, Dere Ş (2011) Determination of the phenolic compounds and anti-
oxidative capacity in red algae Gracilaria bursa-pastoris. Int J Food Prop 14:496
Index

A successional patterns, 243, 244


Abiotic factors, 175 successional processes, 242
Accelerated regional warming, 7 temporal and spatial reduction, 255
Actinobacteria, 286 UV radiation, 255
Adenocystis utricularis, 199 Antarctic benthic system, 14
Algal assemblages, 88 Antarctic biodiversity, 10
Alien competitors, 267, 269 Antarctic biota, 10
Alien macroalgae, 267, 268 Antarctic brown algae, 5
Alien species, 96 Antarctic Circumpolar Current (AAC),
Amphipods 4, 8, 267
assemblages, 269 circumpolar path, 49
chemical defenses, 347, 348 description, 44
description, 342 eddies, 47
feeding bioassays, 342 eastward movement, 84
Lambia antarctica, 349 Gigartina skottsbergii, 95
mortality rates, 270 high-level endemism, 44
populations, 271 mesoscale variability, 47
WAP ocean circulation in SO, 84
coastal food webs, 269 oceanic features, 44
macroalgae, 347 as oceanographic barriers, 46
macroalga-invertebrate primary fronts, 46
interactions, 352–355 Antarctic coastal areas
Antarctic algal succession carbon fluxes, 157, 158
glacier retreat, 252–253 macroalgae, 157, 158
grazing, 249–252 Antarctic coastal ecosystems, 15
UV-B, 248–249 Antarctic coastal systems, 8
Antarctic benthic algae Antarctic coastal zone, 174, 175
biological drivers, 255 Antarctic Convergence, 84
colonization processes, 255 Antarctic costal systems, 6
constraints and difficulties, 254 Antarctic cyanobacteria, 88
in situ succession, 253, 254 Antarctic Desmarestiales, 196
interannual changes, 243 Antarctic environment, 256
macroalgae, 243 algae, 25
sessile faunal assemblages, 242 benthic organisms, 25
structural patterns and changes, 244–248 diversity, 27
subtidal communities, 253 macroalgae, 27

© Springer Nature Switzerland AG 2020 389


I. Gómez, P. Huovinen (eds.), Antarctic Seaweeds,
https://doi.org/10.1007/978-3-030-39448-6
390 Index

Antarctic environment (cont.) B


physical and chemical, 25 Bacterial community diversity
temperature and salinity, 24 Actinobacteria, 286
Antarctic flora, 10 differences, 285
Antarctic high-intertidal tidepools, 272 epiphytic and endophytic bacterial, 286
Antarctic intertidal ecosystems, 52 marine prokaryotic, 284
Antarctic light regime, 4 PacBio SMRT sequencing, 285
Antarctic macroalgae, 89, 158, 252, 267 phylum Ochrophyta, 285
chemical roles, in sensory ecology, 340 predominant bacteria, 285
Antarctic marine ecosystems, 157 proteobacteria and Firmicutes, 286
Antarctic marine flora, 4, 7, 28 Benthic algae, 294
Antarctic marine macroalgae Benthic algal community, 160
ACC, 104 Benthic Antarctic grazers, 274
diversity, 104 Benthic communities
genetic diversity, 109–111 canopy-forming algae physiology, 230
population fragmentation, 109 structure and maintenance, 230
Quaternary glacial events, 106–108 Benthic primary producer communities, 243,
reconstruction, 105 245, 248
Antarctic maritime islands, 37 Biofouling, 340
Antarctic organisms, 16 Biogeographic barriers, 38
Antarctic ozone depletion, 6 Biogeographic processes
Antarctic Polar Front (APF), 9, 62–67, 69, 70, Antarctic Convergence, 84
74, 104–107, 116, 117 barriers and ecological corridors, 84
Antarctic Program, 88 clusters, 87
Antarctic rocky shores, 50, 52 cryptic species, 87
Antarctic seaweeds, 12 diversity and endemism, 86
ACC, 8, 9 eco-regions, isolation and
biogeographical regions, 8 endemism, 93–96
biomass values, 11 faunal provinces, 86
chlorophyceans, 10 GWR model, 86
climate change, 8, 16 meteorological and oceanographic
ecological succession, 12 patterns, 98
ecosystem level, 6 temperature, 86
environmental features, 10 Biological invasions
global climate changes, 7 anthropogenic threats, 266
inorganic and organic pollutants, 16 biological isolation, 266
logistical constraints, 15 isolated ecosystems, 266
long-term assessment, 15 macroalgae, 267
macromorphology, 26 Biomass reduction, 251
molecular ecology, 15 Blue autofluorescence, phenolic compounds, 379
photobiological adaptations, 132 Brown algae
photosynthetic shade adaptation, adverse effects, copper, 374
137, 138 dry weight, 367
phylogenetic relationships, 285 enzymatic machinery, 380
solar radiation, 11 old-temperate genera, 374
Antarctic terrestrial fauna, 45 phlorotannins
Anti-herbivory defences, 269 Cystosphaera jacquinotii, 378
Antioxidant activity, 370, 373, 374 depths, 376 (see also Phlorotannins)
Antioxidant capacity, 14 and metals, 373
Apparent optical properties (AOPs), 133 photoprotective role, 371
Aquatic organisms, 131 UV-absorbing characteristics, 380
Ascendency framework, 311, 317–319 UV protection and antioxidant activity, 370
Ascoseira mirabilis, 196 Brown species, 60, 62
Index 391

C D
Canopy-forming algae Dark respiration, 181
Desmarestiales, 232 Deception Island
functional groups, 231 aerial view, 91
physical disturbance, 230 anthropic activity, 92
taxonomic richness, 230 Chlorophyta, 92
Carbon balance (CB) circular-shaped volcano, 91
daily net CB, 182, 183 Crustose calcareous algae, 92
estimations, 11 definition, 90
glacier influence, 178 diversity, macroalgae, 92
glacier melting, 178 environmental changes, 91
irradiance vs. photosynthesis, 177 filamentous Cyanobacteria, 89
light availability, 179, 180 fumarolic emissions and thermal
mathematical models, 177 springs, 91
newly ice-free areas, 178 fungal species, 93
photosynthesis, 183, 184 geographical position, 85, 90
photosynthetic acclimation, human activity, 91
181, 182 microclimates, 91
positive CB, 177 tourism, 92
Potter Cove, 178 Denaturing-gradient gel electrophoresis
primary production, 176, 177 (DGGE), 283
turbidity, 177 Desmarestia anceps, 196
WAP, 178 Desmarestiales, 245, 247
Carbon flux, 157, 158, 163 Detached seaweeds
Changing light environment, 175 connectivity, 61
Cheirimedon femoratus, 346, 347 D. anceps and D. menziesii, 62
Chemical ecology drift brown algae, 67
defenses, 348, 349 drift D. anceps, 67
in feeding deterrence, macroalga, 341 green algae, 67
mediation, defensive interactions, 340 macroalgae, 62–67
Plocamium cartilagineum, 347 red algae, 62
signals, 340 decomposition rates, 68
Chemical signals, 340 stranded seaweeds, 62, 68, 69
Chilean South Patagonia (CSP), 268 Dispersal of organisms, 46
Chlorophyll fluorescence, 182 DNA damage, 135, 139, 206
Climate changes, 4, 87, 89, 174, 175, 211 Drifting seaweeds
Coastal benthic-pelagic ecosystems, 310, detached macroalgae, 62
319–321, 327, 328, 330 in hollows, 62
Coastal ecosystems, 310 Durvillaea antarctica, 46, 51–52
Coastal marine ecosystems, 156, 157 Dynamic growth models, 162–165
Coastal waters, 50 Dynamic photoinhibition, 138, 139
Colored dissolved organic matter (CDOM),
132, 141, 144
Community structure, 302 E
Constitutive anti-stress mechanisms, 14 Early colonizers, 244
Crude extracts, 341, 343–345, 347, 348, East Antarctic Peninsula (EAP), 37
350, 352 Ecological functions, 12
Cultivation-based methods, 282 Ecological systems, 302
Cyanobacteria, 88, 91, 92 Ecosystem engineering functions, 6
Cyclobutane pyrimidine dimers Ecosystem functioning, 295, 298, 302
(CPDs), 205 Ecosystem functions, 12
Cystosphaera jacquinotii, 64, 74, 75 Eddies, 47, 49
Cystosphaera jacquinotti, 60–62, 72, 74, 75 Ekman transport, 47
392 Index

Electron transport rates (ETR), 204, 205 life strategies and stress tolerance, 227–229
Endemic species, 28, 30, 31, 35 light use characteristics, 225–227
Endemism, 85, 86, 88, 96 morpho-genetic-based program, 218
Energy mass balance, 313 morphological plasticity, 218
Environmental factors phenotypic plasticity, 218
copper, 210 photochemical adjustments, 219
high solar radiation, 205–208 vertical zonation, 222–224
human activities, 210 Functional form groups, 219, 221
ocean acidification, 210 Functional groups, 219
reproduction, 210 Antarctic green algae, 222
salinity, 210 gross morphology, 220
temperature, 208, 209 L. antarctica and M. hariotii, 219
Environmental filtering, 266 Functional traits
Expansion-contraction model, 108 benthic communities, 230
biological interactions, 230
crustose species, 230
F form models, 230
Feeding bioassay, 340–341 low light conditions, 231
Filamentous fungi and yeasts, 282 polar coastal ecosystems, 229
Floating alga temperate ecosystems, 231
abiotic factors, 69–71
biotic factors, 71–72
C. jacquinotti, 60, 61 G
dispersal mechanism, 73 Gametogenesis, 196
kelps, 61 Gastropods, 354–356
M. pyrifera and D. antarctica, 70 Antarctic grazing, 270
physiological responses, 73–75 mesocosm experiments, 269
species, described, 60 GenBank data, 36
See also Detached seaweeds Generalized additive model (GAM), 157
Floating seaweeds, 9 Genetic diversity
biotic factors, 69–71 COI and TufA sequences data sets, 109, 110
cold-temperate, 70 genetic drift, 108
description, 62 Geographically weighted regression (GWR)
drift and stranded seaweeds, 61 model, 86
environmental change, 73 Gigartina skottsbergii, 198
flora and fauna, 61 Glacial refugia
in temperate latitudes, 72 Antarctic macroalgae, 115
nutrient abundance, 70 genetic diversity, 116, 117
Fluorescence in situ hybridization (FISH), 283 geothermal, 107
Fluorescence method, 204 ice coverage, 116
Food webs in situ marine refugia, 107
macroalgae, 294 LGM, 113
macroalgal community, 295 peri-Antarctic islands, 107
modular patterns, 303 population effective size, 116
network, 296, 301 population fragmentation, 108
Potter Cove ecosystem, 296–298 vs. single Antarctic refugium, 108–109
properties, 295, 296 “Glacier influence”, 178
robustness, 300 Glacier melting, 178
trophic levels (TL) of species, 296 Glacier retreat, 162, 164
Food web theory, 295 Global change phenomena, 256
Form and functions Global warming, 46, 166
functional groups, 219–222 Grazing, 13
heteromorphic phase expression, 218 algal communities development, 251
“the holy grail framework”, 219 amphipod, 252
Index 393

Antarctic benthos, 249 K


biotic filtering, 270 “Kelps”, 156
differential susceptibility, 251 Key species, 310
limpet, 252 KeyPlayer Problem (KPP), 324
Greenhouse gases (GHG), 96 Keystoneness, 311
Gross morphology, 219, 220, 223, 232 Keystone species, 310, 311
Gymnogongrus antarcticus, 197 Keystone species complexes (KSCs)
benthic-pelagic ecosystem, 325–327
central set of k nodes, 324
H functional keystone index (KSi), 322, 323
Heavy metals, 143 macroscopic network properties, 327
Herbivory qualitative/semiquantitative loop
macroalga-herbivore interactions, 340 models, 324
macroalgal palatability and resistance structural keystone index (Ki), 323, 324
to amphipods, 342–348
to fish, 349, 350
to sea stars, 350, 351 L
to sea urchins, 351–352 Laboratory culture techniques, 39
tolerating herbivory, 341 Last Glacial Maximum (LGM)
High solar radiation deglaciation, 106
CPDs, 206 genetic diversity, 113, 120
DNA damage, 206 glacial reconstructions, 107
MAA, 207 glacial refugia, 116
propagule physiology, 205 grounded ice, 116
propagules, 205 ice coverage, 116
stress factors, 205 ice sheet expansions, 107
UV-absorbing substances, 207 local refugium, 116
UV radiation, 206 photosynthetic macroalgae, ice impact, 115
UV-stress tolerance, 206 population contraction, macroalgae, 113
UV susceptibility, 207 Latitudinal–temporal scales, 287
UV tolerance, 207 Life history stages
UV wavelengths, 205 Antarctic seaweeds, 194, 195
High-throughput sequencing photosynthetic light requirements, 199–202
(HTS), 282 photosynthetic parameters, 204, 205
Himantothallus grandifolius, 181 season anticipators, 194–198
Hitch-hiking, 46, 50–52, 71, 72 season responders, 198, 199
“The holy grail framework”, 219 Life strategies and stress tolerance
Hymenocladiopsis prolifera, 197, 198 anatomical traits, 228
Hyperbolic tangent function, 181 assemblages, 229
biological indicator, 227
environmental stressors, 229
I intercalary meristem, 228
Ice melting, 162 morpho-functional implications, 228
Ice sheets and icebergs, 7 photoprotection, 228
Inherent optical properties (IOPs), 133 photosynthetic functionality, 228
Internal transcribed spacer region r and k strategies, 227
(ITS), 282 season anticipators, 227
Intertidal filamentous algae, 269 season responders, 227
Invasive and naturally dispersing non-native Light-absorbing impurities (LAIs), 143
species, 46, 48 Light absorption, 132, 133, 225
Invasive macroalgae, 266 Light availability, 179, 180
In vivo absorptance, 225, 226 Light climate, 133–137
Iridaea cordata, 198 See also Underwater light climate
394 Index

Light-independent carbon fixation (LICF), 5 complex, 280


Light penetration, 134, 136, 142, 144 environmental and biological systems, 284
Long-range atmospheric transport eukaryotes, 280
(LRAT), 143 implications, 281
Long-term successional patterns, 253 PCR, 282
Loop Analysis, 313, 315, 324, 329 profiling, 282, 283
Low water transparency, 11 seawater and sediments, 284
16S rRNA, 283
structure and diversity, 282
M Microphytobenthos, 249
Macroalgae, 156, 243 Microplastics, 143
Antarctic coastal food webs, 294 Molecular taxonomy
assemblages, 253 DNA barcodes, 36
common potential consumers, 342, 345 phenotypic plasticity, 35
detritivore pathway, 296 species, 36
direct interactions, 294 UPA, 36
food web, 294 Monostroma hariotii, 110, 111, 114, 117, 246
in Potter Cove, 298 Morpho-functional adaptations, 12, 232
in shallow water, 340 Morpho-functional traits, 222, 223
Macroalga-invertebrate interactions, WAP Multicellular algae, 218
common red macroalgae, 353 Multispecies modelling, 313
community-wide mutualism, 353 Mutualism, 353, 355, 356
endophytism, 354 Mutualistic networks, 302
free-living filamentous algae, 353 Mycosporine-like amino acids (MAAs), 139
gastropods, 354
Notothenia coriiceps, 353
pathogenicity, 354 N
Macroalgal biomass, 156, 159, 160 Nacella concinna, 268
Macroalgal colonization, 162 Net primary production (NPP), 157
Macroalgal community, 295, 304 Newly ice-free areas
Macroalgal rafts, 50–52 algae growth, 176
Macroalgal species, 295 costal shallow areas, 178
Macrocystis pyrifera, 61, 62, 65, 69–73, 75 glacial influence, 185
Macroscopic network properties glacier retreat, 162
A/C ratio, 319 macroalgal community, 162
coastal benthic/pelagic ecological in Potter Cove, 163, 166, 180, 181, 185
system, 319–321 seaweeds, Potter Cove, 185
ecosystem development, 319 WAP, 175, 184
Fildes Bay, 320 water column, 161
KSCs (see Keystone species Non-trophic interactions, 302, 303
complexes (KSCs))
macroscopic descriptors, 317
network analyses, 310 O
Ov/C ratio, 319 Ocean acidification, 210
redundancy values, 319 Ocean circulation, 48
Marine macroalgae, 266, 340 Oceanographic processes
Marine organisms, 45 ACC, 46, 49
Mass balance, 314 eddies, 47
Mesograzers, 268, 269 Ekman transport, 47
Metabolic carbon balance, see Carbon mesoscale variability, 47
balance (CB) ocean circulation, 48
Microbial communities primary fronts, 46
amplicon sequencing, 282 Stokes drift, 47, 48
Antarctic marine environments, 282 Oceanography, 84, 98
Index 395

Ochrophyta, 30–31 Physiognomy, 251


Ozone depletion, 140, 141, 143 Physiological thallus anatomy, 225, 227
Physodes, 366, 367, 369, 370, 372, 374, 378
Phytoplankton, 132, 134, 142
P Pollution, 143
PacBio SMRT sequencing, 285 Polycyclic aromatic hydrocarbons
Palmaria decipiens, 197 (PAHs), 143
Peracarid crustaceans, 50, 51 Polymerase chain reaction (PCR), 282, 284
Persistent organic pollutants (POPs), 143 Population bottleneck, 108, 109, 113, 120
Phaeurus antarcticus, 196 Potter Cove, 160, 162, 166, 296, 298, 299, 301
Phenolic compounds, 370, 373, 374, 377, 378 Primary production, 176
Phlorotannins, 73, 74 Pseudoperennial macroalgae, 244, 245
acetate-malonate pathway, 369
alginic-acid-bound, 369
allocation in Antarctic seaweeds, 377–379 Q
and antioxidant activity of extracts, 381 Quaternary climatic oscillations
in brown algae, 366 (QCO), 10
chemical structure, 366 Quaternary glacial cycles, 106, 107
-containing physodes, 375
in depth patterns, 368–369, 376
exudation, 370 R
herbivory, 375 Rafting, 46, 50
insoluble, 374 and long-distance dispersal, 61
interaction with metals, 373–374 floating seaweeds, 73
physodes, 366, 369 Rafts, 62, 70–73
polymerization, 366 Reactive oxygen species (ROS), 14, 372
response to UV radiation, 379–381 Regional warming, 142
ROS, 372 Rhodophyta, 32–35
scavenging, 372 Roaring Forties, 47
and UV-induced oxidative stress, 372–373 Rocky shores, 158
as UV-screening substances, 370–372 Rocky substrate, 243, 253, 254
Photosynthesis, 177, 183, 184, 208
CDOM, 132
endemic brown seaweeds, 137 S
high photosynthetic tolerance, 139 Salinity, 210
PAR (see Photosynthetically active Sea ice, 96, 97
radiation (PAR)) Seasonal ice cover, 136
shade adaptation, 137 Seasonality, 175
thinner ice cover, 136 Season anticipators
Photosynthesis-irradiance (P-E), 181 A. mirabilis, 196
Photosynthetic acclimation, 181, 182 Antarctic seaweeds, 194, 196
Photosynthetic activity, 209 D. anceps, 196
Photosynthetically active radiation (PAR), fertilization, 196, 197
132–134, 137–139, 176, 177, 179, gametogenesis, 196
180, 183–185 in late winter/spring, 176
Photosynthetic light requirements, 199–203 physiological process, 194
Photosynthetic organisms, 13 reproduction, 194, 197
Photosynthetic oxygen evolution, 181 seasonal growth, 197
Photosynthetic parameters, 204, 205 species, 197
Photosynthetic shade adaptation Season responders, 176, 198, 199
Antarctic seaweeds, 137, 138 Seaweed biomass, 157
Pneumatocysts, 60, 61 Seaweed distribution
Phylum Ochrophyta, 285 benthic communities, 37
Physicochemical changes, 87 macroalgae, 37
396 Index

Seaweed diversity structurally complex habitats, 248


assemblages, 98 subtidal assemblages, 245
biogeographical clusters, 87 succession stages, 246
Deception Island, 92 three-dimensional structure, 247
and endemism, 86 Subtidal macroalgae, 269
macroalgae, 92 Sunlight, 132
macroalgal assemblages, 87
red algae, 87
South Georgia, 94 T
Seaweed microbiomes Temperature, 86, 208, 209
evolutionary trajectories, 287 Terminal restriction fragment length
functional interactions, 281, 282 polymorphism (T-RFLP), 283
holobiont, 280, 287 Terrestrial biodiversity, 44
host-specific microbial taxa, 280 Terrestrial organisms, 44, 45
microbes, 280 Tolerance, 139, 142
structure and diversity, 282–284 Transoceanic dispersal, 46
Seaweed populations, 11 Trematocarpus antarcticus, 197
Seaweed processes, 84 Trophic dynamic, 356
Seaweed production, 166 Trophic levels, 295, 296
Seaweed propagules, 199, 207, 208, 210 Trophic mass balance models, 313
Seaweeds Tropospheric ultraviolet and visible (TUV)
food webs, 295–299 model, 134, 136
interactions, 295 Turbidity, 175, 177, 179, 183, 184, 211
Secondary metabolites, 366, 369–370 Two-species communities, 272, 273
Sedimentation, 270
Short-weighted trophic level, 296
Shotgun metagenomic sequencing, 282 U
Skottsberg, 246 Ultraviolet-B (UV-B)
Solar radiation, 69, 70, 131, 136, 138–140 benthic primary producer community, 248
Southern Ocean (SO) colonization/establishment, 248
ACC, 119 definition, 248
Antarctic Convergence, 84 DNA damage, 249
biogeographic patterns, 85–86 macroalgae, 249
biota distribution, 84 soft-bottom habitats, 248
climate changes, 120 Ultraviolet radiation (UVR), 268
endemism, 104 Ulva intestinalis
macroalgal anatomy, 105 average fitness ratios, 272
non-endemic Antarctic seaweeds, 105 competitive advantages, 270
physical fragmentation, 106 competitive exclusion, 273
waters, cooling and freezing, 120 frequency-dependent consumption, 274
16S ribosomal RNA gene (16S rRNA), gutweed, 270
283, 284 high-intertidal tidepools, 271, 272
Stranded seaweeds, 61–62, 68 invasion process, 272
Stokes drift, 47–49, 54 negative density, 272
Structural patterns and changes negative frequency dependence, 272
Antarctic subtidal sites, 247 two-species communities, 271
benthic diatoms, 245 Underwater light climate
characteristic trait, 247 AOPs, 133
communities, 247 IOPs, 133
diatom assemblages, 244 light attenuation and potential impact, 132
early colonizers, 244 optical properties, 132
filamentous chlorophytes, 246 penetration, UV wavelengths, 133
salinity and light availability, 246 Underwater optics, 132, 141
Skottsberg, 246 UV-absorbing compounds, 380
Index 397

UV radiation, 206, 209 UV tolerance, 72


on aquatic organisms, 135 UV transparency, 134
and CDOM, 141
deleterious effects, 138
harmful levels, 138 W
MAAs, 139 Warming and ozone depletion, 16
ozone hole, 135 Water column optics, 132, 133, 138
TUV model, 134 Western Antarctic Peninsula (WAP),
UV wavebands, 140 174, 175
UV susceptibility, 207 “Wrack”, 158

You might also like