effects of high co levels on growth and calcification of ... · the caribbean coral porites porites...

72
Effects of high CO 2 levels on growth and calcification of the Caribbean coral Porites porites (Pallas, 1766) Rossana Martins de Freitas Masters in Marine Sciences Marine Resources Porto, 2011

Upload: doancong

Post on 10-Dec-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

Effects of high CO2 levels on growth and calcification of

the Caribbean coral Porites porites (Pallas, 1766)

Rossana Martins de Freitas

Masters in Marine Sciences – Marine Resources

Porto, 2011

UNIVERSIDADE DO PORTO

INSTITUTO DE CIÊNCIAS BIOMÉDICAS ABEL SALAZAR

Effects of high CO2 levels on growth and calcification of the

Caribbean coral Porites porites (Pallas, 1766)

Rossana Martins de Freitas

Dissertation application to the

Master‟s Degree in Marine Sciences

– Marine Resources submitted to

Instituto de Ciências Biomédicas de

Abel Salazar of Universidade do

Porto

Master‟s thesis co-ordinate by:

Prof. Dr. Eduardo Rocha (Instituto de

Ciências Biomédicas Abel Salazar –

Universidade do Porto)

Prof. Dr. Chris Langdon (Rosenstiel

School of Marine and Atmospheric

Sciences – University of Miami)

Investigation done in Landgon Lab,

Rosenstiel School of Marine &

Atmospheric Science – University of

Miami

2

Abstract

Ocean acidification is the increase in acidity (decrease in pH) of the ocean‟s

surface waters as result from oceanic uptake of atmospheric carbon dioxide (CO2). The

concentration of this gas is projected to reach twice the preindustrial level by the middle of

the 21st century. In response to this increase, saturation state of surface seawater is

decreasing, causing concern for a global reduction in rates of reef accretion. This also

represents a major threat to coral reefs by reducing the calcification rate of framework

builders.

Here is proposed a study of the effects of high CO2 concentrations on skeletal

growth of the hermatypic coral Porites porites. An experiment was conducted using coral

nubbins at three CO2 levels: control (390 ppm), and two high CO2 scenarios, even if a little

bit higher than the projected ones to this century (1350 and 6000 ppm). In order to see

the effects of ocean acidification on these corals, growth rate was calculated every week

through two different methods: Buoyant Weight and Linear Extension. The Chlorophyll a

concentration and the Zooxanthellar density were also determined.

A decrease in the growth rate was observed for the higher values of CO2 in the

end of the experiment. For the Buoyant Weight, the values recorded ranged from 17.7 ±

0.7 for the low level treatment (390 ppm) to 4.0 ± 1.7 mg d-1 cm-2 for the high level

treatment (6000 ppm). For the Linear Extension, the same answer was observed, i.e., a

decrease in the growth rate for the higher levels of CO2. For the low level treatment, the

growth rate was 98.5 ± 7.2 and for the high level this was 15.6 ± 2.7 µm d-1.

As the Chlorophyll a concentration and Zooxanthellar density, high levels of CO2

showed a decline both in numbers of zooxanthellae cells and chlorophyll per unit surface

area, as expected. Also, after 7 weeks of experiment, the corals that were in high CO2

(6000 ppm) start to bleach, losing their pigments and/or their symbiotic algae.

Considering the results obtained, it can be concluded that Ocean acidification

affects the growth and the health of the hermatypic coral Porites porites. An estimate was

made using values of carbon dioxide concentration of 550 and 950 ppm using the AR4

best and worst CO2 projections for the years 2050 and 2100 respectively, based on

scenarios B1 and A1F1 projected by the Intergovernmental Panel on Climate Change

(IPCC 2007). Growth through buoyant weight is projected to decline by 18% by 2050 and

36% by 2100. The amount of Chla per cm2 of coral tissue is projected to decline 17% by

2050 and 33% by 2100. It is reasonable to expect that these changes will make this

3

species less competitive on the reef and will likely lead to a decline in its abundance on

Caribbean coral reefs.

Little is known about how acidification impacts the physiology of reef builders and

how acidification interacts with warming. So, the next step is to focus on the interaction

between temperature and saturation state, as accurate predictions of global change have

to include the effects of both on coral growth.

Keywords: Porites porites; Ocean acidification; Calcification; Growth rate; Climate

change.

4

Resumo

Acidificação dos oceanos é o aumento da acidez (diminuição do pH) das águas de

superfície dos oceanos como resultado da absorção oceânica do dióxido de carbono

atmosférico (CO2). Projecções feitas dizem que a concentração deste gás alcançará o

dobro do nível pré-industrial até meados do século 21. Em resposta a este aumento, o

estado de saturação da água do mar está a diminuir, causando sérias preocupações em

relação a uma redução nas taxas de acrecção de recifes. Isso representa também uma

grande ameaça para os recifes de coral pois causa uma redução na taxa de calcificação

dos coráis construtores de esqueletos.

O presente estudo foi realizado com o intuito de analisar os efeitos do dióxido de

carbono no crescimento do esqueleto do coral hermatípico Porites porites, quando este

(CO2) se encontra presente na água do mar em elevadas concentrações. Para isso, foi

conduzida uma experiência utilizando “nubbins” de coral e colocando estes em três

diferentes níveis de dióxido de carbono: controlo (390 ppm), e dois cenários de altas

concentrações, mesmo que um pouco superiores aos projectados para este século (1350

e 6000 ppm). Para uma análise dos verdadeiros efeitos da acidificação dos oceanos

nestes coráis, foi calculada a taxa de crescimento todas as semanas através de dois

diferentes métodos: Peso flutuante e Extensão linear. Foram também determinadas a

concentração de Clorofila a e a densidade Zooxantelar.

Em relação ao crescimento, uma diminuição na sua taxa foi observada para os

valores mais elevados de CO2 no final da experiência. Para o Peso flutuante, os valores

registados variaram entre 17,7 ± 0,7 para o tratamento de nível de CO2 baixo (390 ppm) e

4,0 ± 1,7 mg d-1 cm-2 referente ao tratamento de nível mais alto (6000 ppm). Para a

Extensão Linear, a mesma resposta foi observada, ou seja, houve uma diminuição na

taxa de crescimento para os níveis mais elevados de CO2. Para o tratamento de baixo

nível de dióxido de carbono, a taxa de crescimento foi de 98,5 ± 7,2 e para o nível mais

elevado, esta foi de 15,6 ± 2,7 µm d-1.

Quanto à concentração de Clorofila a e à densidade Zooxantelar, os níveis altos

de CO2 mostraram um declínio tanto no número de células de zooxantelas como na

Clorofila, por unidade de área de superfície; como seria esperado. Além destes

resultados, após sete semanas da experiência, os coráis que se encontravam nos

tanques com alta concentração de CO2 (6000 ppm), começaram a perder os seus

pigmentos e/ou algas simbióticas, ou seja, iniciaram o processo de branqueamento.

Considerando os resultados obtidos, pode-se concluir que a acidificação dos

oceanos afecta o crescimento e a saúde do coral hermatípico Porites porites. Uma

5

estimativa foi feita utilizando os valores de concentração de dióxido de carbono de 550 e

950 ppm com base no Quarto Relatório de Avaliação (AR4) - melhores e piores

projecções de CO2 para os anos de 2050 e 2100 respectivamente, de acordo com os

cenários B1 e A1F1 do Painel Intergovernamental de Alterações Climáticas (IPCC 2007).

A taxa de crescimento prevê-se decrescer em 18% até 2050 e 36% em 2100. A

quantidade de clorofila a por cm2 de tecido de coral deverá diminuir cerca de 14% até

2050 e 33% até 2100. Assim, é razoável prever que estas alterações irão tornar esta

espécie menos competitiva no recife, e irão provavelmente levar a um declínio na sua

abundância nos recifes de coral do Caribe.

Apesar de todos os estudos que começam a ser feitos nesta área, pouco se sabe

ainda sobre os impactos da acidificação na fisiologia dos coráis construtores de recifes e

como esta interage com o aquecimento da água do mar. Assim, o próximo passo a

realizar é concentrar-se na interacção entre temperatura e estado de saturação, como

previsões precisas das mudanças globais, que incluem os efeitos de ambas no

crescimento de coráis.

Palavras-Chave: Porites porites; Acidificação dos oceanos; Calcificação; Taxa de

crescimento; Alteração climática.

6

7

Acknowlegments

I would like to start this work with a few words of thanks to everyone who made

this work possible and that made me the person I am today.

Não deixando o português de lado, gostaria de começar este trabalho com

algumas palavras de agradecimento a todos aqueles que tornaram possível este trabalho

e que me tornaram na pessoa que sou hoje.

Thanks to my co-ordinates Professor Dr. Eduardo Rocha from ICBAS – University

of Porto and to Dr. Chris Langdon from RSMAS – University of Miami. Thank you both for

the possibility that you gave me to work with these fantastic animals, that I like and admire

so much. Also thanks for all the wisdom and dedication demonstrated, by your support

and especially for all the patience.

Also a special thanks to Nancy Muehllehner, who have sooo much patience with

me and teach me everything she can. There were many times that instead of being with

her beautiful daughter, she was helping me with my experiment and always with a big

smile in her face! I hope from now on you enjoy more time with Teah and your husband!

You saved me!!

Thanks to everyone that were with me during these six months in Miami. A special

thanks to Carolina to be the fantastic friend she was and still is and to helped me so much

in this journey, and to Teddy that helped me to integrate myself here.

Thanks to everybody that worked with me in Rosenstiel School of Marine &

Atmospheric Science, but especially to Remy, Erica, Rachel, Jay and Evan, for all the

help and friendship.

Last but not least, thank you Gustavo, Rolando, Milan, Andy, Marc, Maria, Matt,

Krystyna, Henry, Bryan, Stephanie, Paul and CS people, that helped me having fun and

enjoyed Miami.

8

Antes de mais, um especial agradecimento à minha família. Aos meus pais e à

minha irmã, que tornaram possível este meu sonho e sempre me apoiaram, mesmo com

todas as adversidades e momentos menos bons. À Palmirinha, ao Jan, à Lina, à Maria

Manuela, à Cristina, à Maria, ao Betxinho por toda a amizade e carinho que me deram e

por acreditarem em mim. Ao resto da família pelos conselhos e toda a força, e

especialmente a força das saudades da minha pequenina, a minha Beatriz.

Ao pessoal todo que me acompanhou durante a licenciatura e fez os meus dias no

Algarve fantásticos e para nunca esquecer, em especial às minhas Marias, claro! Nós as

três éramos a força umas das outras e nada nem ninguém nos vai conseguir separar.

Nem com um oceano entre nós a nossa amizade vai acabar. Adoro-vos para sempre.

Às minhas coisas boas do Puorto, Cláudia, Ângela, Sónia, Sara, Anas, Débora,

Ligia, Lucinha (companheira de quarto mais faladora do mundo) e aos meninos, Hermes,

Francisco, André, Jorge, Ricardo e Ildefonso, que se tornaram tão importantes no espaço

de um ano. Um muito obrigado por todo o tempo que passámos juntos, quer a estudar

quer a sair.

Ao meu pessoal de Aveiro, em especial ao meu Bitor, à Dani, ao Huguinho, ao

Ildefonso. Sem vocês não sou ninguém, são aquelas pessoas que por muito tempo que

estejamos separados, sei que posso sempre contar e vice-versa, são quem me apoia

quando estou em baixo e partem a loiça toda quando estou feliz. E por isso, adoro ter-vos

na minha vida!!

Finalmente quero agradecer a duas pessoas muito importantes na minha vida, a

Sofia e o Garcia. Cada uma à sua maneira, fazem a minha vida muito mais feliz e muito

mais alegre! Duas pessoas com quem a relação não foi nada fácil durante o meu tempo

em Miami, mas tudo se há-de recompor. Não vos quero nunca fora da minha vida e

obrigada por terem entrado nela e a fazerem melhor!

A todos (to all), Muito Obrigada, Thank You.

9

Aos meus pais e irmã,

Olga Maria Soares Martins de Freitas

Sérgio Martins de Freitas e

Isadora Martins de Freitas

10

11

Indice 1. Introduction ...............................................................................................................13

1.1 Biology of Coral Reefs .......................................................................................13

1.2 Ecology and Biodiversity ....................................................................................15

1.3 Threats to reefs of the Caribbean ......................................................................15

1.3.1 Diseases .....................................................................................................16

1.3.2 Sediments ..................................................................................................17

1.3.3 Pollution ......................................................................................................17

1.3.4 Overfishing .................................................................................................18

1.3.5 Coral Bleaching ..........................................................................................19

1.4 Porites porites: Biology ......................................................................................20

1.5 Ocean acidification ............................................................................................21

1.5.1 Ocean carbonate system ............................................................................23

1.5.2 Calcification and Dissolution response from Hermatypic Corals .................24

1.6 Objectives ..........................................................................................................25

2. Material and Methods ...............................................................................................26

2.1 Preparation of corals ..............................................................................................26

2.2 Experimental setup .................................................................................................26

2.3 Measurements........................................................................................................28

2.3.1 Buoyant Weight ...............................................................................................28

2.3.2 Linear Extension ..............................................................................................30

2.3.3 Chlorophyll a concentration ..............................................................................32

2.3.4 Zooxanthellar density .......................................................................................32

2.3.5 Surface area determination ..............................................................................33

2.4 Statistical analyses .................................................................................................34

3. Results .....................................................................................................................35

3.1 Chemical conditions ...............................................................................................35

3.2 Growth rates ...........................................................................................................36

3.2.1 Buoyant Weight (B.W.).....................................................................................36

3.2.2 Linear Extension (L.E.).....................................................................................38

3.2.3 Combined data ................................................................................................39

3.3 Chlorophyll a concentrations and Zooxanthellae density ........................................39

3.4 Light data ...............................................................................................................43

3.5 Bleaching ...............................................................................................................44

4. Discussion ................................................................................................................45

4.1 Growth rates ...........................................................................................................45

12

4.2 Chlorophyll a concentrations and Zooxanthellar densities ......................................49

4.3 Bleaching ...............................................................................................................50

4.4 Projections .............................................................................................................52

4.5 Final Considerations ...............................................................................................53

5. References ...............................................................................................................55

13

1. Introduction

1.1 Biology of Coral Reefs

Corals are invertebrate animals, that is, without spine, belonging to the Order

Scleractinia, where all members have hard limestone skeletons, and one of the wonders

of the underwater world (Veron, 1986). They are divided into two groups: the ones that are

always polyp type and those that pass in their life cycle through a stage called jellyfish,

and may or may not also present a form of a polyp stage (Bacelar, 1997). Sexual

reproduction gives rise to a planula larva and this can or cannot settles onto the

substratum and grow into a benthic polyp phase. This polyp phase multiplies asexually

both to produce other polyps, which may lead an independent life or remain attached to

each to form a colony, as well as the planktonic medusa stage. This, in turn, develops

gonads and reproduces sexually. One of the two phases can be lost from the life history,

and if the medusa stage is absent the polyp stage develops the gonads as well as

multiplying asexually (Barnes, 1998).

Each polyp has a simple, tube-shaped body, with a circular mouth at the top which

also serves as anus. Around the mouth they have numerous tentacles to help capturing

food and inside there is a pharynx that connects the mouth to the coelenterons, and has a

number of mesenteries poking into it (Spalding, 2004). All exchanges of gases and fluids

are happen in coelenterates, because the water enters and leaves the body of the animal

by currents caused by tentacles.

They build exoskeletons, and these may be composed by organic matter or calcium

carbonate (Glynn & Wellington, 1983). They can also be divided in two types: Hermatypic

and Ahermatypic. The first ones can also be called stony corals build reefs and with the

help of the zooxanthellae, they convert surplus food to calcium carbonate, forming a hard

skeleton. The ahermatypic corals have no zooxanthellae and do not build reefs. These are

also known as soft corals and so they are flexible, undulating back and forth in the current,

and their skeletons are proteinaceous instead of calcareous as the stony corals (Wells &

Hanna, 1992). Both types suffer from environmental influences resulting in limiting the

quantity and quality of species.

14

Figure 1 – Basic anatomy of a hard coral polyp. Artwork Credit: NOAA/Gini Kennedy

Corals build colorful and amazing colonies when they find a supportive environment.

These great colonies are the most familiar forms of the animals that we know as corals,

and their skeletons are the ones that help building and defining coral reefs (Spalding,

2004). These clusters in colonies can reach millions of individuals (polyps), forming a

limestone formation, which can have thousands of kilometers. They are home to an

ecosystem with an extraordinary biodiversity and productivity.

The polyps have in their tissues a group of plant cells known as zooxanthellae,

which are the vegetative state of the dinoflagellate Symbiodinium microadriaticum (Wells

et al., 1983).

These zooxanthellae use the sunlight to convert carbon dioxide into carbohydrates,

lipids and aminoacids. These products are passed to the host, which in return, produce a

nitrogen source (ammonia) and phosphate to algae. Thus, the zooxanthellae takes up the

waste of the polyps at the same time that it feeds them, allowing corals to put all their

energy into their growth, and consequently into the formation of reefs (Delbeek & Sprung,

1994).

15

1.2 Ecology and Biodiversity

The diversity of coral reefs is comparable with the most rich and diverse terrestrial

habitats. These species are considered the base of diversity and despite the reefs occupy

a small area of the planet, there are possibly more species per unit area of coral reefs

than in any other ecosystem.

Coral reefs are home to a huge variety of organisms, where can be found more than

4000 species of fish. Among them are tropical or reef fish, such as parrotfish, angelfish

and butterfly fish, sponges, viruses, crustaceans (including prawns, lobsters and crabs),

echinoderms (like starfish, sea urchins and sea cucumbers), sea turtles and sea snakes

(Castro & Huber, 2000; Spalding et al., 2001).

Inside of the reefs, like other diverse communities, there is a huge variety of

interactions between species. Each one occupies a unique role, or niche, on the reef but

because there are so many species, they have to change and adapt. They have to find a

space to live and grow, to eat, to avoid being eaten, to hunt, to reproduce and to hide from

the hunters (Spalding, 2004).

Some organisms have become masters of camouflage to capture or escape. This

co-evolution has also led to a complex two-way interaction between species, including

mutualism, where both organisms benefit. One of the most important examples of this

partnership is the relationship between the corals and their algae, which has led to the

proliferation and success of coral reef builders. Algae use the sunlight and carbon dioxide

to produce organic compounds rich in carbon. These in turn are leaked to the polyp and

may provide almost all of its nutritional requirements (Allen, 2000).

1.3 Threats to reefs of the Caribbean

The coral reef ecosystems are very sensitive to external impacts that violate its

homeostasis, whether natural or human caused. As it is known, almost everywhere there

are far fewer corals today than there were even 10-20 years ago. The main causes for this

decrease are diseases, sedimentation and nutrient impacts, marine-based pollution,

overfishing, direct destruction, coral bleaching and coastal development (Spalding, 2004;

Sulu, 2007).

16

1.3.1 Diseases

There are reports of coral diseases since 1970 but it was only in the 1980s and

1990s that some of them had a dramatic impact. In the Caribbean there are over more

than 150 species infected, and new diseases continue to be added to the ones already

described (Côté & Reynolds, 2006). Infectious diseases can cause population declines or

even species extinctions.

Today most of the reefs in the Caribbean have lost 80-100% of their staghorn and

elkhorn coral through a disease known as white band disease. Some other coral diseases

have been reported and named according to their appearance: white plague, yellow band,

black band, dark spots and rapid wasting disease (Spalding, 2004).

Figure 2 – Some of the major coral diseases: 1) white plague; 2) yellow band; 3) white band; 4) black

band; 5) red band; 6) white pox; 7) dark spots. Image courtesy by Andy Bruckner, NOAA’s National Marine

Fisheries Service

Coral diseases appear everywhere, but they are more common and more

devastating in areas where the corals are already stressed by pollution or sedimentation

(Spalding, 2004). Many pathogens are also sensitive to factors as temperature, humidity,

and rainfall, and these factors can increase the pathogen development and survival rates,

disease transmission and host susceptibility. More research is needed to find if there are

points to potential linkages to other stresses on corals as possible instigators of disease

(Harvell et al., 2002).

17

1.3.2 Sediments

In the Caribbean, there are increasing levels of sediments in coastal waters. The

principal causes of this increase are clearance of forests and movement of agriculture on

the steep slopes, leading to soil erosion and these particles are carried by rivers to the

sea (Spalding, 2004).

Suspended sediments in the water column can affect coral populations and

community structure by suffocating adult corals and enforcing physiological costs by

reducing light availability for photosynthesis or increasing the energy needed for sediment

removal (Riegl & Branch, 1995). They are able to remove these sediments with their

tentacles (sloughing them off with a layer of mucus), but are not able to keep doing it

because it uses up energy and resources.

The real understanding of the effects of sedimentation is not easy because the

interspecific differences in morphology and sediment rejection ability (Stafford-Smith,

1993); not all the corals have the same ability to use heterotrophic modes of nutrition

(Anthony, 2000) and also because sedimentation rates vary with other environmental and

biological factors (Babcock & Smith, 2000).

1.3.3 Pollution

Coral-reef ecosystems are highly vulnerable to environmental perturbations in

general. This sensitivity has three main factors, and they are: corals have small

physiological tolerance for environmental conditions (Johannes, 1975) and any variation

of physical-chemical outside these tolerance ranges could be decisive to coral growth and

survival (Johannes & Betzer, 1975; Endean, 1976; Pearson, 1981). Another factor is the

susceptibility of the interactions of key reef species (like algae-coral competition) to

pollutant stresses. The damage of these corals leads to an eventual end of many reef

species because these depend on living corals for many reasons like shelter, food and

refuge (Johannes, 1975). The third factor that affects corals sensitivity is that the effects of

most toxic substances are enhanced in water with high temperatures, which are easily

found in coral reef environments (Johannes & Betzer, 1975). These toxic wastes come

primarily from detergents, paints, herbicides and pesticides, industrial toxins and oils.

Despite the known toxic wastes, the most widespread form of pollution found on

coral reefs comes from simple nutrients like organic matter, nitrates and phosphates. Two

of the major sources of these are sewage and agriculture (Spalding, 2004). The potential

18

impacts of sewage effluent on coral-reef communities can be classified in three different

categories that represent separate but interacting processes, which are nutrient

enrichment, sedimentation and toxicity (Pastorok & Bilyard, 1985).

1.3.4 Overfishing

Humans have gathered and fished coral reef organisms for thousands of years to

use it as curios, jewelry and food. More recently, this exploitation has also increased as

novel medicinal compounds have been extracted and as reef-related tourism has

expanded. According to published work, in the last 30-40 years, increasing populations of

humans have been damaging coral reefs at an unprecedented rate (Luchavez & Alcala,

1988; Wilkinson et al., 1993; Munro & Munro, 1994; Grigg & Birkeland, 1997).

Overfishing is one of the major threats to the biological integrity of marine habitats

worldwide (Ginsburg, 1994; Dayton et al., 1995; NRC, 1995; Roberts, 1995; Safina,

1995). The removal of grazing species allows algae to overgrow and smother the corals

while the removal of small reef fish for commercial sale means that fewer individuals get

the chance to grow to reach the reproductive age and replenish the population (Karleskint

et al., 2009). Some fishing methods are very destructive, like the careless use of the traps

and nets which can smash corals, the use of explosives or bleach to drive fish out of the

reef and the use of cyanide (Rubec, 1986; McAllister, 1998, Spalding, 2004) to collect

tropical fish for the aquarium trade or fresh fish for local markets. Cyanide is a metabolic

poison, and this not only removes great number of fish from the reef but also poisons

corals and other invertebrates (Karleskint et al., 2009).

Although everybody knows what to do about overfishing, is incredibly difficult to

persuade fishers and politicians to change, because of all the money that is involved.

There are still a lot of fish, but not enough to reproduce and restock the reefs, which

therefore remain quite unproductive. Well managed, in long-term these same reefs could

supply more fish and would also be more attractive to tourists (Spalding, 2004).

19

1.3.5 Coral Bleaching

The phenomenon of coral bleaching affecting extensive reef areas across the

Pacific was first described by Glynn in 1984. This process is portrayed as the whitening of

corals due to loss of symbiotic algae (zooxanthellae) and/or their pigments and is a global

phenomenon that is probably correlated to global climate change and increasing ocean

temperatures (Glynn, 1993; Brown, 1997; Hoegh-Guldberg, 1999). Reef-building corals

are highly dependent on their zooxanthellae because these provide up to 95% of the

corals‟ carbon requirements for growth, reproduction and also maintenance (Muscatine,

1990).

Figure 3 – Partially bleached coral colony (Pocillipora sp.) on a Pacific Reef (A) and close up view of

bleached area (B). Photo courtesy by NOAA.

The most important stressor is the variation of sea surface temperature (Glynn,

1984; Cook et al., 1990; Fitt et al., 1993; Brown et al., 1995; Hoegh-Guldberg & Salvat,

1995; Anthony et al., 2008), however things like freshwater flooding (Goreau, 1964;

Egana & DiSalvo, 1982), pollution (Jones, 1997a; Jones & Steven, 1997), high turbidity

and sedimentation (Meehan & Ostrander, 1997), diseases (Kushmaro et al., 1997; Benin

et al., 2000), high levels of ultraviolet light, low light conditions and abnormal salinity

(Westmacott et al., 2000) can also cause bleaching on corals.

Most reef corals are very sensitive to environmental stressors that can interfere in

the symbiotic relationship with their algae and cause bleaching. Even when corals are

pale or white, that doesn‟t mean they are dead, but without its algae they are severely

weak and more susceptible to diseases. If conditions do not improve, the coral will

inevitably die (Spalding, 2004). Not all coral species are equally affected (Coles & Brown,

2003) and they have different resistance to bleaching (Loya et al., 2001). This difference

20

can be genetically determined or by distribution and environmental variability

(McClanahan et al., 2007).

1.4 Porites porites: Biology

Porites is a genus of sceleterian coral which is characterized by a finger-like

morphology. The members of this genus have widely spaced calices, a well-developed

wall reticulum and are bilaterally symmetrical. Porites porites is a branching species with a

more sprawling appearance than the other branching Porites. Branches may be curved,

hooked or project downwards and sideways, and are thicker with blunt, enlarged tips. The

three branching species are easily confused, and colony shapes overlap. This species

has thick branches (˃ 2cm) may be stubby and little developed with enlarged, blunt tips.

Developed branches upright or sprawling, often grey, occasionally bright blue. The

colonies of this species are common in shallow water near shore where they may form

extensive beds (Voss, 2002).

Figure 4 – Specimen of Porites porites

21

1.5 Ocean acidification

Over the past 250 years, the levels of atmospheric carbon dioxide (CO2) increased

by nearly 40%, from preindustrial levels of approximately 280 ppm (parts per million) to

nearly 384 ppm in 2007 (Solomon et al., 2007). This rate of increase in atmospheric CO2

is approximately 100x faster than has occurred for millions of years (Siegenthaler et al.,

2005) and the current concentration is higher than any level reached in the past 800,000

years (Lüthi et al., 2008). Approximately 25% of the CO2 emitted from all anthropogenic

sources, like human fossil fuel combustion, agriculture, cement production (Houghton,

2001) and deforestation, currently goes to the ocean, where it reacts with water to

produce carbonic acid (Hoegh-Guldberg et al., 2007). The CO2 uptake causes pH

alterations, has this already fallen by 0.1 unit, from approximately 8.21 to 8.10 (Royal

Society, 2005) compared to pre-industrial times, what means a 30% increase in the H+

concentration and is projected to drop another 0.3-0.4 pH units by the end of this century

(Mehrbach et al., 1973; Caldeira & Wickett, 2003; Orr et al., 2005; Loáiciga., 2006; Feely

et al., 2009; Anlauf et al., 2011).

Table 1 – Projected changes in surface ocean carbonate chemistry based on IPCC IS92a

CO2 emission scenario (Guinotte & Fabry, 2008).

Was assumed that PO4 = 0.5 µmol L-1

and Si = 4.8 µmol L-1

, and use the carbonic acid dissociation constants of Mehrbach et al. (1973) as refit by Dickson & Millero (1987). pHT is based on seawater scale. Percent change from preindustrial values are in parenthesis. After Feely et al. (2009).

22

When CO2 is absorbed by seawater, a series of chemical reactions happen. An

increase in atmospheric CO2 concentration lowers the pH and the carbonate ion

concentration (CO32-) available (Kleypas & Langdon, 2002). This reduction also leads to a

reduction in calcium carbonate saturation state (Ω), which has big impacts on marine

organisms (Guinotte & Fabry, 2008). Table 1 lists the projected changes in surface ocean

carbonate chemistry (carbon system parameters and temperature) based on the

Intergovernmental Panel on Climate Change (IPCC) IS92a CO2 emission scenario.

One way to predict climate changes is look backward, for example using ice cores.

A 2083-meter long ice core with 160,000 years was collected by the Soviet Antarctic

Expedition at Vostok, Antarctica (Barnola et al., 1987). There is a close correlation

between carbon dioxide in the air and global temperatures over this time. However, there

are some difficulties in extrapolating these correlations into the future (Krebs, 1994). The

present CO2 levels exceed the levels presented in nature during the last 160,000 years so

is not known if it‟s correct to extrapolate with the same relationship found in the past.

Another problem is the velocity at which CO2 levels are changing. Year to year they are

changing. Lessons learned about the relationship between CO2 and climate in the past will

have to be applied with caution in predicting the future climate because the rate of change

of climate forcing is very different today.

The rate of this change is really concerning biologically as well, since many marine

organisms, particularly the calcifying ones, may not be able to adapt quickly enough to

survive to these changes. Some of the most abundant life forms that will be affected are

plankton. These are very important because they form the base of the marine food chain

and are a major food source for fish and marine mammals (World Bank, 2010).

Acidification may decrease reef growth by reducing calcification rates but also

reproduction and recruitment (NRC, 2010).

Coral reefs are among the marine ecosystems most vulnerable to the changing

climate and atmospheric composition and are threatened by a combination of direct

human impacts and global climate change. Coral reefs are already being pushed to their

thermal limits by recent temperature increases (World Bank, 2009).

Diverse studies identify the ocean acidification as a threat to marine biota for

different factors such as: the calcification rates of shell-forming organisms in response of

supersaturation (Smith & Buddemeier, 1992; Kleypas et al., 1999); aragonite, a mineral

also very important to organisms that calcify, may become undersaturated in the surface

ocean in the early 21st century (Feely & Chen, 1982; Feely et al., 1988; Orr et al., 2005);

and the effects of the decreasing of pH go far beyond limiting calcification.

It is already known that calcification of corals and coral communities is controlled by

saturation of seawater with respect to aragonite (Gattuso et al., 1999; Langdon et al.

23

2000; Leclerq et al., 2000; Leclerq et al., 2002). Several papers have tried to predict the

future response of calcification by coral communities, using the experimental evidence of

previous works, to elevated pCO2 and its biogeochemical significance. Is consensual that

calcification rate will decrease by 14-30% by 2100 (Gattuso et al., 1999; Kleypas et al.,

1999).

One of the most obvious and well documented effects of ocean acidification is, as it

was said before, the decrease of calcification rates, affecting skeletal growth. There are

records that on the Great Barrier Reef, massive corals showed a decrease in calcification

rates by about 14% between 1990 and 2005 (De‟ath et al., 2009). In stony corals, most

studies indicate a 10-60% reduction in calcification rate for a doubling of preindustrial

atmospheric CO2 concentration. The differences between studies may be due to different

species or experimental setups (Doney et al., 2009; Kleypas et al., 2006; Langdon &

Atkinson, 2005).

Growth of the reef structures also rely on the successful recruitment of reef

organisms, which is determined by gamete production, fertilization rates, larval

development and settlement, and post-settlement growth. In general, there are a few data

on any of these aspects for reef-building species, making extrapolation to ecosystem

effects difficult (Jokiel et al., 2008; Albright et al., 2008).

1.5.1 Ocean carbonate system

The carbon system is one of the most important components of oceanography,

primarily because it acts as a governor for the carbon cycle and also because it controls

the acidity of seawater. This system also plays a very important role controlling the

pressure of carbon dioxide in the atmosphere, helping to regulate the temperature of the

planet (Emerson & Edges, 2008).

Seawater carbonate chemistry is a series of chemical equilibria, led mainly by DIC

(Dissolved Inorganic Carbon) and TA (Total Alkalinity) (Kleypas et al., 2006). The surface

water CO2 to atmospheric levels is equilibrated by exchanges between air and sea in a

timescale of approximately one year. When CO2 dissolves in seawater, it reacts with the

water and form carbonic acid (H2CO3), which will lose hydrogen ions to form bicarbonate

(HCO3-) and carbonate (CO3

2-) (Doney et al., 2009).The basic equations describing this

system are:

24

CO2 (g) CO2 (aq)

CO2 (aq) + H2O H2CO3

H2CO3 H+ + HCO3

-

HCO3- H+ + CO3

2-

Processes that increase DIC shift the equilibrium in a way to lower the pH and lower

CO32- concentration; on the other hand, processes that increase TA shift the equilibrium in

a way to increase the pH and CO32- concentration. Calcification and dissolution affect both

DIC and TA, while respiration and photosynthesis primarily affect DIC (Kleypas et al.,

2006).

1.5.2 Calcification and Dissolution response from Hermatypic Corals

The reef-building corals, zooxanthellate, are a very important group with the

capacity to produce massive quantities of calcium carbonate (Stanley & Hardie, 1998). A

lot of experiments were performed to test the response of calcification rate to decreases in

aragonite saturation state, and some evidences indicate that the calcification rate of

tropical reef-building corals will be reduced by 20-60% at double preindustrial CO2

concentrations (Gattuso et al., 1998; Kleypas et al., 1999; Langdon et al., 2000; Kleypas &

Langdon, 2002; Langdon et al., 2003; Reynaud et al., 2003; Langdon & Atkinson, 2005;

Royal Society, 2005; Kleypas et al., 2006).

Calcification rates vary under different environmental conditions and the most

important are light, temperature and nutrients (Doney et al., 2009). Light increases the

primary productivity of coral reefs and provides the precipitation of CaCO3 in

zooxanthellate corals (Falkowski et al., 1990) and coralline algae (Johansen, 1981).

Marubini et al., 2001 found out that calcification increases with increasing light up to a limit

and then saturates. Temperature also affects the calcification rate, as proved in a lot of

literature. Most of species calcify faster with an increase in temperature until an optimal

temperature at or 1º - 2ºC below the normal peak summer temperature, after this point

calcification rates starts to decline (Clausen & Roth, 1975; Jokiel & Coles, 1977). Nutrient

concentration was also found to affect coral calcification in laboratory studies. The

25

principal studied nutrients were nitrate and ammonium. Enrichment of nitrate to levels of

5-20 µM resulted in increased zooxanthellae density and photosynthesis, and so

decreased calcification rate (Marubini & Davies, 1996; Marubini & Thake, 1999). Also an

increase in the ammonium concentration produces similar results (Ferrier-Pages et al.,

2000; Hoegh-Guldberg & Smith, 1989; Stambler et al., 1991).

1.6 Objectives

This study aims to investigate the effects of global environmental changes on the

physiology of scleractinian corals in order to reach better predicting capabilities on their

response to future changes. The effects of increases of CO2 were investigated using

concentrations of 390 ppm (actual concentration found in seawater), 1350 and 6000 ppm.

The experimental coral chosen for this study was the ecological significant species Porites

porites. To achieve that, specific objectives were drawn:

1. To evaluate the effects of high CO2 in corals growth using:

a. Buoyant Weight;

b. Linear Extension;

2. To determine and analyze the zooxanthellar density for corals in high CO2

levels and in control;

3. To calculate the concentration of Chorophyll a for exposed organisms.

26

2. Material and Methods

2.1 Preparation of corals

Branches of Porites porites were brought from the Smithsonian laboratory in Fort

Pierce, FL to RSMAS (Rosenstiel School of Marine & Atmospheric Science) in January

2011. There they were placed in a tank, 380 L, inside the Experimental Corals and

Climate Change Facility and kept at a temperature of 27ºC and a salinity at 36.5. In April

2011, tips were removed from the branches and their cut faces were ground flat. They

were then prepared into coral „nubbins‟ as described in Davies (1995) and glued with All

Game Epoxy II-Part putty on PVC plastic tiles (35 x 37 mm). After the „nubbins‟ were

done, they were placed in a smaller tank with 262 L of seawater, which was called control,

since the CO2 level was 390 ppm. They were kept there for one week so they could

recover and then were switch to the different CO2 levels.

2.2 Experimental setup

An experiment was performed to evaluate the effect of different CO2 levels on

growth and calcification of Porites porites. The experiment was done in the Experimental

Corals and Climate Change Facility of RSMAS and lasted for eight weeks, from April 21st

to June 15th.

This facility is composed by twelve fiberglass tanks measuring 0.6 x 0.5 m filled to a

depth of 0.22 m. The tanks are inside a shadehouse with a polypropylene clear plastic

roof that admits natural sunlight but at the same time, protects the equipment from the

elements. The nominal amount of sunlight reaching the organisms varies seasonally but is

typically 6-8 mol photons m-2 d-1. The tanks are supplied by high quality low nutrient

concentration natural seawater from the nearby Bear Cut entrance to Biscayne Bay. The

salinity also varies seasonally but usually is between 33 and 38. The water from the Bear

Cut is pumped into a 50,000 gallon settling tank and then through a series of filters that

are intended to remove particulates down to a size of 10 µm. This seawater flows to each

of the experimental tanks at a rate of approximately 150-200 ml min-1. These experimental

tanks contain 60 L of seawater and are associated with a 200 L sump tank. The seawater

from the sump tank into the experimental tank is continuously circulated by a magnetically

coupled 500 GPH pump which is in the sump tank. This seawater pumped sprays out

27

through some holes made in a length of ½” PVC pipe that is in the bottom of the

experimental tank, generating lots of turbulence within the tank in proximity to the corals.

The seawater in the experimental tanks is exchanged with the seawater in the sump

approximately every 10 minutes and through a standpipe.

Each sump tank is fitted with a 1500-watt tungsten heater and heat exchanger coil

made of 1” polypropylene irrigation tubing. A solenoid valve is used to block or permit the

circulation of chilled freshwater (20ºC) through the heat exchange coil. There is also an

Omega CN7500 digital temperature controller that is able to hold the average daily

temperature to ±0.1ºC.

The CO level in the experimental tanks is controlled by bubbling the seawater in the

sump tanks with a CO2-enriched air mixture. An air compressor delivers air with a pCO2 of

390 ppm to a centrifugal drier. The dried air at a pressure of 14 psi is connected to several

Sierra Instruments 810C Mass Flow Controllers (MFC). These deliver a steady stream of

dried air at a precisely controlled rate that is adjusted to 6-14 L per minute. A second set

of MFCs are connected to a cylinder of instrument grade CO2 gas. These are adjusted to

a supply stream of CO2 gas at flow rates ranging from 5-100 ml per minute. The outputs of

air and CO2 MFCs are connected in pairs to gas mixing tubes. Adjusting the setting n the

two MFCs can precisely control the pCO2 of the mixture. The flow rate of gas mix going to

each tank is monitored using a rotameter, and by determination a flow rate of 3-5 L per

minute was found to be adequate for controlling the pCO2 in the experimental tanks. From

the rotameter the gas mix goes to a venturi injector. This produces a strong stream of

bubbles that promotes the exchange of CO2 in the gas mix with the seawater in the sump

tanks. To achieve a desired CO2 level, the pCO2 of the gas mix needs to be adjusted

empirically to compensate for the gas exchange and the rate that CO2 is added and

removed from the system by the flow of Bear Cut seawater. An equilibrator/IRGA system

monitors the pCO2 of the seawater actually achieved on the tanks and an Onset Computer

HOBO data logger acquires the temperature of each of the experimental tanks, the tank

pCO2 and the incoming photosynthetically available irradiance every 5 minutes.

In this experiment, three different levels of CO2 were used to test the effects of this

on corals, which were 390 ppm (low), 1350 ppm (medium) and 6000 pm (high).

28

Figure 5 – Tanks inside of the Experimental Corals and Climate Change Facility at RSMAS,

where corals were placed. Photo by the author

The corals were feed every Thursday with a special recipe composed by 134 ml of

phytoplankton, 0.6 g of oystereggs, 0.6 g of larval food supplement 100-150 microns, 0.6

g of larval food supplement 250-450 microns and 50 ml of seawater. Before the feeding

the water flow was turned off and the temperature was take down to 24.8º C.

2.3 Measurements

2.3.1 Buoyant Weight

Coral growth was measured by buoyant weighing (Davies, 1989) once every 2

weeks until May 15th and then weekly. An electronic balance was used, reading to 0.1 mg.

The balance was mounted on a wooden box whose interior formed the weighing chamber.

The front was fitted with a transparent door, which could be closed to exclude air

movements during weighing. After all setup, the coral was handled by the tile to minimize

damage to the tissues and hold inside of a container filled with seawater maintained at the

same temperature and salinity so the density wouldn‟t change.

Final skeletal growth rates were obtained by linear regression of weight against time

over the last 7 weeks of the experiment only, because the first measurement was relative

29

to the week they were under control. At the end of the experiment the surface area of live

tissue of each colony was determined and the growth rate was normalized to the cm2 of

live tissue.

Buoyant weight is the weight of the coral in water, so it was necessary to calculate

the weight of coral in air. The equation used was the one that is below (Eq. 1) and has

been explained in more detail by Jokiel et al., 1978.

Eq. 1 – Equation to calculate the weight of object in air (Davies 1989)

where wt in water = weight (g) of the object in water, Dwater = density (g/cm3) of the

seawater, Dobject = density (g/cm3) of the object being weighed. For the density of the

object was used the value of 2.822 g/cm3, which was direct determined for Porites porites

by Edmund & Davies, 1986. The Dwater was accurately determined from Eq. 2 by weighing

an inert reference object of known air weight and density in the seawater. The object used

was glass and its density was calculated from Eq. 3.

Eq. 2 – Equation to calculate the Density of water (Davies 1989)

30

Eq. 3 – Equation to calculate the Density of object (Davies 1989)

The temperature and the salinity of the seawater were always controlled during the

buoyant weighing and kept at 27ºC and 35 respectively, so the density didn‟t change.

2.3.2 Linear Extension

Another way to measure the coral growth was through linear extension using a

micrometer. The height of the corals was measured almost every week during all

experiment.

The linear extension was measured using an optical micrometer, capable of

measuring linear extension rates to a precision of 0.5 µm. During each measurement

period, the corals were placed into bins filled with seawater and taken to the lab. There

they were briefly removed from the bin, fitted onto the stage of the micrometer, and

advanced through the path of the optical micrometer. Each measurement was conducted

in no longer than 4 minutes, thereby minimizing the time they were out of water.

The average growth rate was obtained from regression equations of maximum point

(maximum height) against time over 5 weeks of the experiment only, because like in the

buoyant weight one of the weeks the corals were still in controls and the other week that

there‟s no data, was because the Micrometer wasn‟t not on the lab.

31

Figure 6 – Micrometer used to the Linear Extension (L.E.). Photo by the author

Figure 7 – Coral being measured by L.E. Photo by the author

Figure 8 – Example of a graph for growth rate by Linear Extension

-25

-20

-15

-10

-5

0

5

10

0 10 20 30 40

He

igh

t (m

m)

Distance (mm)

April-20

May-10

May-20

May-28

June-6

June-13

32

2.3.3 Chlorophyll a concentration

After the end of the experiment, all the corals were brought to the Baker lab in

RSMAS where they were blasted, that is, the tissues were removed from the skeletons

with a jet of filtered seawater using a WaterPikTM (Johannes & Wiebe, 1970). The volume

of the blastate was made up to 50 ml using filtered seawater. Subsamples of the blastate

were taken for analysis of zooxanthellar densities and chl a concentration.

For the chlorophyll a concentration, 1 ml of the blastate was placed in a glass tube,

9 ml of 90% acetone was added with a graduated cylinder and the sample was put in an -

80ºC freezer for about 30 hours. After this time, the sample was brought to room

temperature and vortexed to remove filter fibres from suspension. 0.5 ml of the sample

was removed to a tube and 3 ml of 90% acetone was added (Nakamura et al. 2003). The

sample was put into the Fluorometer TD-700 Turner Designs, and the first read of

absorbance was taken. After this, 1 drop of 10% HCl was added to acidification and the

second read taken.

The follow equation was used to calculate the concentration of Chlorophyll a

Chla = (1x10-3) * D * (K * T) * (Fo-Fa)

where Chla is the amount of the pigment in the coral tissue sample in µg, the factor

10-3 converts from liters to milliliters, D is the overall dilution (50:1)*(10:1)*(7:1) or 3500, K

is the calibration constant (K = 0.0000892904 µg/mL), T is the maximum ratio of

chlorophyll a fluorescence before and after adding acid (T = 1.885908293), Fo is the first

fluorescence reading, Fa is the second fluorescence reading after adding the HCl. Finally

the pigment results were divided by surface area of the coral to obtain the µg Chla per

cm2 of coral tissue.

2.3.4 Zooxanthellar density

1.5 ml of the 50 ml blastate solution was put into Eppendorf tubes and 30 ml of

Lugols added (i.e. 1.5:31.5 dillution). The tubes were stored out of the light. The samples

33

were then vortexed and a portion pulled up into a Pasteur pipette. A haemocytometer was

positioned on edge of microscope platform with coverslip, and filled by capillary action.

A counter was used to count all cells falling within or on the lines defining the 9

squares of the haemocytometer. The counts were recorded in a table under the sample

number and some repeats were done loading the haemocytometer at least 3-5 times. The

volume of liquid counted (9x10-4 ml), the volume of the blastate counted (1.5 ml), the

dilution of the blastate (1.5:31.5) and the surface area of the coral were used to calculate

the density of zooxanthellae cells per cm2.

The formula used to calculate the density of zooxanthellae was the following one:

Z = (N/9x10-4)*D/SA

Where Z is the zooxanthellae density in cell per cm2 of coral tissue, N is the average

number of zooxanthellae counted in the haemocytometer slide, 9x10-4 ml is the volume of

Lugol solution counted, D is the overall dilution factor (50:1)*(31.5:1.5) and SA is the

surface area of the coral.

2.3.5 Surface area determination

Measurements of coral growth rate were made and determinations of zooxanthellae

densities and chlorophyll concentrations. However, the results were taken from different

sizes of coral “nubbins”, which required a method of „normalizing‟ the amount of coral in

the analyses. The most common method is to use the surface area of the colonies to

account for the different amount of biomass in each measurement.

The method used in this specie to measure the surface area was the Aluminum foil

(Marsh, 1970) since it works well for “boulder” corals like Porites porites. To determinate

it, after removal of tissue from the coral, an Aluminium foil was layed over the coral

surface, using scissors to trim the foil to cover the entire blasted area. Then the foil was

stretched and a picture was taken. In the program “Image J” the image was treated and

the surface was calculated.

34

2.4 Statistical analyses

The effect of high levels of CO2 on coral growth rate was tested for each experiment

by separate One-way ANOVAs. Calcification rate by buoyant weight, chlorophyll a

concentration and zooxanthellae density per nubbin were positively correlated to final

surface area, so this was used to normalize the data. Statistical differences were

considered for P ˂ 0.001.

All the statistical analysis were performed using the “data analysis package” from

Microsoft Office, Excel.

35

3. Results

3.1 Chemical conditions

The chemical conditions for the treatment are summarized in Table 2. The

temperature was maintained at 27.1 ± 0.5ºC and the salinity at 36.4 ± 0.2. The Total

Alkalinity (TA) ranged from 2147 ± 115 for the higher level of CO2 to 2282 ± 79 µmol kg-1

for the lower level. The values of pH and Ωaragonite were between 7.1 ± 0.1 and 0.6 ± 0.1 for

the lower level respectively, and 8.1 ± 0.04 and 3.9 ± 0.4 for the higher level.

Table 2 – Physical and chemical conditions during the laboratory experiment 1. Data

given as means ± SD

Salinity

Temp. (ºC)

CO2 (ppm) TA (µmol/kg) pH Ωaragonite

April

Low 36.1 ± 0.64 26.8 ± 0.62 329.4 ± 46.94 2068.8 ± 61.82 8.09 ± 0.08 3.58 ± 0.40

Medium 36.5 ± 0.33 27.0 ± 0.73 1325.5 ± 134.24 2182.4 ± 43.30 7.57 ± 0.03 1.37 ± 0.08

High 36.5 ± 0.26 27.1 ± 0.69 4316.2 ± 671.63 2364.3 ± 96.40 7.72 ± 0.14 0.57 ± 0.06

May

Low 35.7 ± 0.91 27.1 ± 0.49 263.3 ± 7.70 2137.2 ± 59.07 8.15 ± 0.02 3.96 ± 0.18

Medium 36.8 ± 0.35 27.1 ± 0.30 1477.0 ± 156.84 2115.5 ± 45.30 7.52 ± 0.06 1.22 ± 0.15

High 36.6 ±0.45 27.2 ± 0.39 4155.8 ± 224.42 2254.2 ± 25.41 7.12 ± 0.02 0.54 ± 0.02

June

Low 35.6 ± 0.82 27.1 ± 0.06 295.1 ± 17.33 2333.5 ± 148.21 8.15 ± 0.002 4.42 ± 0.31

Medium 36.8 ± 0.33 27.1 ± 0.08 2274.1 ± 1202.61 2242.7 ± 19.90 7.39 ± 0.22 1.06 ± 0.43

High 36.9 ± 0.26 27.3 ± 0.11 3775.3 ± 905.01 2248.9 ± 63.48 7.16 ± 0.11 0.62 ± 0.16

36

3.2 Growth rates

The effects of different levels of CO2 on growth of corals were determined by

buoyant weight and linear extension. The data related to that effect is presented in Tables

3 and 4, for buoyant weight and linear extension respectively.

3.2.1 Buoyant Weight (B.W.)

Table 3 – Mean initial weight, surface area and growth rate for Porites porites calculated

by buoyant weight. Data given as means ± SE

Treatments CO2

level (ppm)

Ωaragonite n Initial

weight (g) Surface area

(cm2)

Growth rate (mg d

-1 cm

-2)

Low 390 3.6 6 14.7 ± 0.9 27.2 ± 3.6 17.7 ± 0.7

Medium 1350 1.3 5 15.9 ± 0.9 27.7 ± 3.5 11.2 ± 0.7

High 6000 0.4 5 15.1 ± 0.8 22.0 ± 2.1 4.0 ±1.7

In the Buoyant weight, the average initial weight of nubbins was 15.2 ± 0.5 g and the

surface area was 25.9 ± 1.9 cm2. A decrease in the growth rate was observed for the

higher values of CO2 in the end of the experiment. The values recorded ranged from 17.7

± 0.7 for the low level treatment (390 ppm) to 4.0 ± 1.7 mg d-1 cm-2 for the high level

treatment (6000 ppm). In a 95% confidence, really small error bars were presented, which

mean that the results are really reliable (Figure 9).

37

Figure 9 – Growth rate by buoyant weight for Porites porites in the end of experiment. Error bars are 95%

confidence

Another very important variable is the aragonite saturation state (Ω). In the 390 ppm

treatment, Ω was kept at 3.6, in the 1350 ppm kept at 1.3 and finally, in the 6000 ppm

maintained at 0.4. The relation between the growth rate and the saturation state it‟s

almost linear, corresponding the lowest growth rate to the lower Ω and the highest growth

rate to the highest Ω (Figure 10).

The growth rate by the technique of buoyant weighing for the control and the two

levels of CO2 was statistically different (P˂0.001), so big differences were found within or

between experiments.

Figure 10 – Scatterplot of average growth rate by Buoyant Weight per tank (g d-1

) to Aragonite saturation

state (Ω)

0

0,05

0,1

0,15

0,2

0,25

0,3

0,35

390 1600 6000

g o

f w

eig

ht

chan

ge in

8 w

ee

ks

CO2 Levels (ppm)

390

1600

6000

0,004

0,0112

0,0177

0

0,002

0,004

0,006

0,008

0,01

0,012

0,014

0,016

0,018

0,02

0 1 2 3 4

Gro

wth

rat

e (

g d

-1)

Aragonite saturation state

6000

1350

390

38

3.2.2 Linear Extension (L.E.)

Table 4 – Mean initial height and growth rate for Porites porites calculated through linear

extension. Data given as means ± SE

Treatments CO2 level

(ppm) n

Initial height (mm)

Growth rate (µm/d)

Low 390 6 16 ± 2 99 ± 7

Medium 1350 5 17 ± 2 45 ± 8

High 6000 5 19 ± 1 16 ± 3

Regarding the linear extension, the results were really similar to the ones obtained

from buoyant weight. A decrease in the growth rate was observed for the higher levels of

CO2. For the Low level treatment the growth rate was 99 ± 7 and for the High level this

was 16 ± 3 µm d-1.

Statistically, the growth rate calculated by linear extension for the control and the

two levels of CO2 was also very different (P˂0.001), which means that big differences

were found within or between experiments.

Figure 11 - Scatterplot of average growth rate by Linear Extension per tank (mm d-1

) to Aragonite saturation

state (Ω)

0,01556

0,04486

0,09852

0

0,02

0,04

0,06

0,08

0,1

0,12

0 1 2 3 4

Gro

wth

rat

e (

mm

d-1

)

Aragonite Saturation State (Ω)

6000

1350

390

39

3.2.3 Combined data

To a better understanding of the effects of CO2 on growth rate of the corals, the two

different methods were analyzed together.

Comparing both methods of growth (Figure 12) was understandable that the growth

rate for the linear extension was higher than the one for the buoyant weight. However, for

all the corals, the growth rate by buoyant weight was much more linear and constant.

Figure 12 – Growth rates through the two methods: Linear extension (blue) and Buoyant weight (red) for the

different levels of CO2

3.3 Chlorophyll a concentrations and Zooxanthellae density

High levels of CO2 showed a decline both in numbers of zooxanthellae and

chlorophyll concentration per unit surface area. The Chlorophyll concentration in 390 ppm

ranged from 4.7 to 9.1 µg cm-2 (Table 5), whereas those exposed to 6000 ppm ranged

from 1.0 to 2.5 µg cm-2. Zooxanthellae densities for the 390 ppm were between 3,500,000

and 3,800,000 cells cm-2 and for the 6000 ppm between 310,000 and 1,100,000 cells cm-

2.

y = 25,35x + 8,13 R² = 0,87

y = 2,52x + 2,54 R² = 0,81

-2

0

2

4

6

8

10

12

14

0

20

40

60

80

100

120

140

0 1 2 3 4

Gro

wth

Rat

e

Aragonite Saturation State (Ω)

Linear Extension (um/d)

Buoyant Weight (mg/d/cm2)

40

Table 5 – Chlorophyll a concentration and number of zooxanthellae per area. Means ± SE

CO2 Level (ppm)

# Coral Chlorophyll a

(µg/cm2)

Average Zooxanthellae

cells/cm2 Average

390

70 4.7

6.7 ± 0.6

3,700,000

3.6 x 106 ± 4.5 x 10

4

73 6.5 3,500,000

74 5.9 3,600,000

79 7.5 3,500,000

85 9.1 3,800,000

88 6.7 3,600,000

1350

72 2.1

3.9 ± 0.5

2,600,000

3.2 x 106 ± 1.5 x 10

5

75 5.0 3,600,000

76 5.0 3,300,000

80 4.1 3,300,000

86 3.6 3,200,000

6000

77 2.5

2.0 ± 0.3

960,000

9.1 x 105 ± 1.5 x 10

5

78 1.0 310,000

83 2.5 1,100,000

84 2.4 1,000,000

87 1.4 1,100,000

For a better understanding of the chlorophyll concentration and number of

zooxanthellae in corals during the experiment, the average per surface area was

calculated for the three different levels of CO2 and represented by the Figure 13 and 14,

respectively.

A One-way ANOVA found that carbon dioxide concentration has a highly significant

effect on chlorophyll a concentration and zooxanthellae density, being P˂0.001.

Figure 13 – Average concentration of Chlorophyll a per unit of area (cm2) for the different levels of CO2. Error

bars are 95% confidence.

0

1

2

3

4

5

6

7

8

9

390 1350 6000

Ch

loro

ph

yll a

co

nce

trat

ion

s (µ

g cm

-2)

CO2 levels (ppm)

390

1350

6000

41

Figure 14 – Concentration of Chlorophyll a per unit of area (cm2) for the different levels of CO2 and

respectively power regression.

Figure 15 – Number of zooxanthellae/cm2 for the different levels of CO

2. Error bars are 95% confidence.

0

500000

1000000

1500000

2000000

2500000

3000000

3500000

4000000

390 1350 6000

Zoo

xan

the

llae

ce

lls c

m-2

CO2 Levels (ppm)

390

1350

6000

y = 101,02x-0,453 R² = 0,9993

0

1

2

3

4

5

6

7

8

0 1000 2000 3000 4000 5000 6000 7000

ug

Ch

la c

m-2

CO2 (ppm)

42

Figure 16 – Number of zooxanthellae/cm2 for the different levels of CO2 and respectively linear regression

The chlorophyll concentrations in zooxanthellae, presented a different result from

the other analysis. A higher value of µg chlorophyll per zooxanthella was observed for the

highest CO2 level (6000 ppm) instead for the control (Figure 17). The values ranged from

1.2 x 10-6 for the medium level (1350 ppm) to 2.3 x 10-6 for the highest level of CO2.

Figure 17 – Chlorophyll a concentrations per number of zooxanthellae for each treatment. Error bars are 95%

confidence.

0

0,0000005

0,000001

0,0000015

0,000002

0,0000025

0,000003

0,0000035

390 1350 6000

µg

Ch

loro

ph

yll a

/Zo

oxa

nth

ella

e

CO2 Levels (ppm)

390

1350

6000

y = -473,64x + 4E+06 R² = 0,9986

0,00E+00

5,00E+05

1,00E+06

1,50E+06

2,00E+06

2,50E+06

3,00E+06

3,50E+06

4,00E+06

0 1000 2000 3000 4000 5000 6000 7000

Zoo

x cm

-2

CO2 (ppm)

43

3.4 Light data

Keeping the amount of light received as close as possible with the natural conditions

on ocean is one of the things to have in attention in an experiment. Light has a central role

in primary productivity of coral reefs and provides the precipitation of CaCO3 skeletons in

hermatypic corals (Falkowski et al., 1990). Figure 18 represents the values obtained for

everyday during the experiment, which ranged from 6.0 to 15.1 mol photons m-2 d-1.

Figure 18 – Total of light (mol photons/m2) emitted per day

0

2

4

6

8

10

12

14

16

04

.21

04

.23

04

.25

04

.27

04

.29

05

.01

05

.03

05

.05

05

.07

05

.09

05

.11

05

.13

05

.15

05

.17

05

.19

05

.21

05

.23

05

.25

05

.27

05

.29

05

.31

06

.02

06

.04

06

.06

06

.08

06

.10

06

.12

06

.14

Tota

l lig

ht

mo

l ph

oto

ns

m-2

pe

r d

ay

Dates

44

3.5 Bleaching

After 7 weeks of exposure to high levels of CO2 (6000 ppm) all the corals that were

in this treatment became white (loss of pigmentation), starting to bleach, as shown in the

picture below (Figure 19). The corals that were in the other treatments didn‟t show any

signal of bleaching.

Figure 19 – The left coral belonged to the control tank (390 ppm), while all the other belonged to the high level

of CO2 (6000). A big difference in color can be seen. Photo by the author.

45

4. Discussion

4.1 Growth rates

In this experiment, the effect of increased CO2 concentration in seawater was tested

for the species Porites porites. The results showed that this increase causes a decrease

in the growth rate of this hermatypic coral skeleton, which means that calcification in this

species depend not only but also on dioxide carbon concentration.

Three different levels of CO2 were used, being one the control (390 ppm) and

another two levels, 1350 and 6000 ppm. These two values are higher than the ones

predicted to the next century, but they were chosen out of curiosity about what may

happen in coming centuries if human society does not bring CO2 levels under control. The

390 ppm corresponded to a saturation state of 3.6, the 1350 ppm to a Ω = 1.3 and finally,

the 6000 ppm to a Ω = 0.4.

Growth rates in this experiment are similar to previous estimates for nubbins of

Porites compressa and Porites porites grown in the laboratory (Marubini & Davies, 1996,

Marubini & Atkinson, 1999, Marubini & Thake, 1999, Marubini et al., 2001, Gattuso et al.,

1999) and in the field (Edmunds & Davies, 1986, Edmunds & Davies, 1989). Since weight

gain is a function of colony size and this is artificially determined when nubbins are cut off

the mother colony, it‟s important to control the size for when measuring growth rate

(Marubini et al., 2001). There are many ways to estimate colony size, being the more

important, surface area (Barnes & Taylor, 1973, Rinkevich & Loya, 1984), skeletal weight,

protein, chlorophyll and tissue content (Marubini et al., 2001). In this study the interest

was the growth response to increased levels of carbon dioxide and decreased saturation

state, as chlorophyll a concentration and zooxanthellae density. For this reason, surface

area was used, since it is linearly correlated to growth rate and controls for colony size.

That is a lot of literature that provides evidence of detrimental effects on corals of

increased pCO2 alone or in combination with other factors (Gattuso et al., 1999, Ohde &

van Woesik, 1999, Leclerq et al., 2000, 2002, Langdon et al., 2003, Marubini et al., 2003,

Reynaud et al., 2003, Renegar & Riegl, 2005). However none of these studies was done

with the species Porites porites neither with such high values of CO2 and low values of

saturation state.

For the buoyant weight, the values recorded ranged from 17.7 ± 0.7 for the control to

4.0 ± 1.7 mg d-1 cm-2 for the high level treatment, which shows a really big decrease in the

growth rate. Results obtained for the species Porites compressa by Marubini et al., 2001,

showed that nubbins grown in conditions of the Last Glacial Maximum (target values: Ω =

46

5.6; pCO2 = 200 µatm) in shallow water had the highest growth rate, 1.8 ± 0.2 mg CaCO3

d-1 cm-2, while the ones grown in conditions of the year 2100 (target values: Ω = 2.4; pCO2

= 700 µatm) at 4 m depth had the lowest growth rate (0.6 ± 0.1 mg CaCO3 d-1 cm-2).

Marubini et al., 2001 studied not also the carbonate ion concentration but also the

influence of the light, using four different depths: 0.5, 2, 4 and 6.5 m. They found that

depth has a strong effect on growth rate, with the nubbins growing at 0.5 m depth having

the fastest calcification rate in all cases.

Marubini & Thake, 1999 tested the hypothesis that the supply of DIC (Dissolved

Inorganic Carbon) in seawater limits the calcification and if limitation of calcification by

nitrogen can be compensated for by DIC enrichment in the hermatypic coral Porites

porites. They found that the nubbins in all treatments in seawater with added DIC have a

higher slope of growth rate. They also reported that in the first 17 days of experiment

there was no significant effect of nitrogen treatment on the total amount of calcium

carbonate deposited. The results were not the same for the last 15 days (which phase

they called established growth rate), when the effect of N on calcification becomes

significant among corals growing in seawater. Corals that were in enriched DIC did not

show a decreased growth rate in response to increased N. The values without DIC were

11.5 ± 1.6 for the control, 5.5 ± 0.8 for NO3- and 8.6 ± 0.9 mg d-1. With the addition of DIC

all these values increased, ranging from 18.6 ± 2.6 for the control + DIC, to 16.0 ± 1.8 for

NO3- + DIC and to 19.0 ± 2.0 mg d-1 for NH4

+ + DIC.

Another study was made to test the effect of nutrient enrichment and elevated CO2

partial pressure for the coral species Acropora cervicornis by Renegar & Riegl, 2005.

Their results showed a growth rate of 1. ± 0.7 for the control period and a growth rate of

0.3 ± 0.1 for the period where the nutrient concentrations were raised to 10 µM NO3- and 4

µM P-PO4-3, and the pCO2 level increased to ~700 to 800 µatm. Even if it is not

appropriate to discount the effects of increased nutrient concentrations on corals, they

found that elevated pCO2 has the most negative effect on growth rate.

We showed that calcification in Porites porites correlates well with saturation state;

as Ω decreased from 3.6 (low concentration of CO2) to 0.4 (high concentration of CO2) the

estimated maximum calcification rate decreased by around 23%. Furthermore, the change

from the medium concentration of CO2 (Ω = 1.3) to those of high treatment (Ω = 0.4)

caused a reduction in calcification.

Looking at the graph in Figure 10, it can be seen the true relationship between the

state of saturation and the rate of coral growth in the different treatments. It can be

concluded that is a linear relationship, being for the lowest saturation state, the lowest

growth rate value and for the highest saturation state, the highest growth rate value.

47

Through the shape of the graph, it can also be seen that the growth rate is directly

proportional to the saturation state and inversely proportional to the concentration of CO2

present in the seawater.

Linear extension is a method of calculation the growth rate that is very new and little

known. However, the results obtained through this process are very reliable and very

clarifying. For this study, it can be seen that, as to the buoyant weight, there was a linear

relationship between the saturation state and the growth rate. The values of growth rate

obtained in this work, ranged from 99 ± 7 for the lowest level of CO2 to 16 ± 3 µg d-1 for the

highest level, with the average of initial height of 17.4 mm. This proves that high

concentration of CO2 not only affect dramatically growth in terms of weight but also in

terms of height of the coral. The observed reductions in linear extension of Porites porites

are consistent with the hypothesis that calcification and, ultimately, growth decline as

pCO2 increases and saturation state decreases (Figure 11). The same results were

obtained by Albright et al. 2010 for the species Acropora palmata. These authors studied

different steps in corals recruitment, concluding that a slowed juvenile growth will translate

into increased juvenile mortality, given that the risk of mortality is inversely proportional to

juvenile growth rate and colony size (Vermeij & Sandin, 2008; Babcock, 1991).

One factor that should also be considered is the temperature increase. Despite not

having been manipulated in this experiment, this factor along with the increase in CO2

concentrations causes further changes in coral growth. Calcification in corals is known to

be enhanced with increasing temperatures (Clausen & Roth, 1975; Reynaud-Vaganay et

al., 1999; Bessat & Buigues, 2001) up to an optimum temperature of ~27ºC (Clausen &

Roth, 1975; Marshall & Clode, 2004; Cooper et al., 2008). Over this temperature,

calcification rates tend to decrease, but dry-weight there is some data that reveal

increased calcification at higher temperatures in the primary polyps of Porites

panamensis. This suggests that this may not be the case in all corals (Anlauf et al., 2011).

Reynaud et al. (2003) showed that the calcification in colonies of Stylophora

pistillata maintained at elevated temperature (28.2ºC), declined by 50% in response to

increased pCO2. However, in colonies maintained at normal temperature (25.2ºC),

calcification was not affected by elevation of pCO2. This last result is not in agreement

with several papers that found a negative relationship between calcification and CO2, or a

positive relationship with the aragonite saturation state (Gattuso et al., 1998; Marubini &

Atkinson, 1999; Langdon et al., 2000, 2003; Leclercq et al., 2000, 2002; Marubini et al.,

2003). These results show that pCO2 and temperature significantly interact to control

calcification in scleractinian corals, which could explain the large range of responses to

elevated pCO2 reported in corals and coral communities.

48

Albright et al., 2008 found that the growth rate of juvenile Porites astreoides

decreased between 45 and 56% when pH was lowered by just 0.07 units. Likewise,

Cohen et al., 2009 reported reduced skeletal growth of the cross-sectional area in newly-

reared primary polyps of Favia fragum under lowered aragonite saturation and calculated

a nearly 80% decrease in the precipitated aragonite.

Marubini & Atkinson (1999) and Langdon et al. (2000) found that the rate of

calcification declined immediately after the rise in pCO2 and did not change afterwards,

demonstrating that no acclimation process occurred. Marubini & Atkinson (1999) reported

for the scleractinian coral Porites compressa that the decrease of calcification in

responses to changes in the carbonate system is immediate and reversible. Also,

Langdon et al. (2000) found that the response of the community calcification rate of the

Biosphere 2 ocean is not significantly different during short-term (days) and long-term

(months) changes in aragonite saturation state.

Reef-building corals can exhibit several responses to reduced calcification, having

all of them deleterious consequences for reef ecosystems. The most direct response is

considered the decreased linear extension rate and skeletal density of coral colonies.

Cooper et al., 2008 found that the massive coral Porites on the Great Barrier Reef has

shown reductions in linear extension rate of 1.02% year-1 and in skeletal density of 0.36%

year-1 during the past 16 years. This is equivalent to a 20.6% drop in growth rate, the

product of linear extension rate and skeletal density, over the 16-year period.

Marubini et al. (2001) made an extrapolation for the effect of seawater Ω on coral

calcification in controlled laboratory setup to the global scale, and concluded that CaCO3

deposition rate in scleractinian corals is expected to decrease around 10% over the next

100 years, as pCO2 will double. Several studies were made to date about the relationship

between Ω and calcification (Gattuso et al., 1999; Schneider & Erez, 2006; Leclercq,

2000) and all of them were consistent with the hypothesis that biogenic calcification is

influenced by the saturation state of seawater. However, the predictions for reduction in

calcification between 2001 and the year 2100 range between 3% (Gattuso et al., 1998)

and 40% (Langdon et al., 2000). This large range in response could be artificially inflated

by the high variability in experimental protocols (combination of many factors like

organismal scale studies with ecosystem scale studies, different taxa of calcifying

organisms, different coral families, different type of chemical manipulation and

experimental duration) or can be due to physiological differences between organisms.

49

4.2 Chlorophyll a concentrations and Zooxanthellar densities

The exposure to high concentrations of CO2 under laboratory conditions severely

affected chlorophyll a concentrations and zooxanthellae densities in Porites porites. The

corals presented in the tanks with high CO2 concentration, showed really low numbers

both in zooxanthellae density and chlorophyll a concentration per unit surface area in the

end of the experiment. The chlorophyll a concentration in control corals ranged from 4.7 to

9.1 µg cm-2 coral surface area, whereas those exposed to high concentrations of carbon

dioxide contained 1.0 – 2.5 µg cm-2. Zooxanthellae densities of the control were 3.4 – 3.8

x 106 cells cm-2. In contrast, densities fell to 3.1 x 105 – 1.1 x 106 cells cm-2 when exposed

to high concentrations. However, the chlorophyll concentration per zooxanthella

decreased as an inverse function, being the highest value to the high CO2 level (2.3 x 10-

6) and the lowest to the medium level (1.2 x 10-6 µg Chl per zooxanthella).

Figure 20 – Relationship between zooxanthellar chl a concentration and zooxanthellar density for all

corals submitted to low, medium and high CO2 concentrations

The results in this study suggest that, for this specie, loss of zooxanthellae can

occur without decreases in zooxanthellar chlorophyll concentrations when exposed to high

levels of CO2. The same results have already been reported in corals exposed to elevated

seawater temperatures during laboratory manipulations (Hoegh-Guldberg & Smith, 1989a;

Glynn et al., 1992; Fitt & Warner, 1995) and in corals sampled after natural bleaching

events (Hoegh-Gulberg & Smith, 1989a; Fitt et al., 1993; Jones, 1997b).

0

0,0000005

0,000001

0,0000015

0,000002

0,0000025

0,000003

0,0000035

0,00E+00 1,00E+06 2,00E+06 3,00E+06 4,00E+06

µg

Ch

loro

ph

yll a

/Zo

oxa

nth

ella

e

Zooxanthellae cm-2

50

Le Tissier & Brown (1996) also found increases in zooxanthellar chla concentrations

in corals (Goniastrea aspera) during stress-related loss of zooxanthellae. In their opinion,

the increase in zooxanthellar chlorophyll concentration occurred because of a greater loss

of zooxanthellae from the apical tissues (leaving “dark-adapted” zooxanthellae located in

the lower tissues) or because a breakdown products of chlorophyll and other pigments

interfering with the absorption peak used to compute chla concentrations.

The inverse relationship between zooxanthellar density and zooxanthellar chl a

concentration (Figure 20) is the opposite of that expected from a photoadaptive response

to zooxanthellar self-shading as zooxanthellar densities increase (Heogh-Guldberg &

Smith, 1989b; Fitt et al., 1993). The Chla concentration of unicellular algae is also a

reliable indicator of nutrient status (Rees, 1991), and the relationship can signify nutrient

limitation at higher zooxanthellar densities (Hoegh-Guldberg & Smith, 1989b; Fitt et al.,

1993; Jones & Steven, 1997). So, the increase in zooxanthellar chla concentration can

represent a response of the remnant zooxanthellar population to increased nutrient ability

through decreased competition.

In general, algal densities in corals are remarkably constant, independently of

genus, depth, growth form or location (Muscatine & Porter, 1977; Porter et al., 1984).

Drew (1972) measured an algal density of 1.45 x 106 cells cm-2 in ten genera of

hermatypic corals and two genera of soft corals. Based on these observations, he

suggested that the area available for the zooxanthellae to occupy directly determines their

densities under steady-state conditions, i.e. a space limitation theory. Dubinsky & Jokiel,

1994 has pointed out that in corals at depths the undersides of the tissues often have no

symbiotic algae. This means that light availability may also limit algal densities and restrict

zooxanthellae from occupying deeper tissues in corals with highly perforate skeletons.

4.3 Bleaching

Bleaching has been described as the dissociation of the symbiotic relationship

between algae and their respective hosts (Iglesias-Prieto et al., 1992). It is important to

understand how each symbiotic partner (both individually and within the intact association)

perceives the environmental stressors that elicit bleaching and how these signals are

transmitted into the observed response. Many authors observed in laboratory experiments

that the response to high temperatures is the paling of the coral through either loss of

zooxanthellae (Hoegh-Guldberg & Smith, 1989a; Glynn & D‟Croz, 1990; Lesser et al.,

51

1990; Fitt & Warner, 1995) and/or the loss of photosynthetic pigment per zooxanthella

(Sharp, 1995; Fitt & Warner, 1995).

It is also known as a generalized stress response resulting from a variety of

environmental conditions outside the normal local range. This can be induced by

unusually high or low temperatures, low light levels or constant light, high levels of UV

radiation, salinity shock or may other chemical factors (Cook et al., 1990; Gates, 1990).

Despite apparent environmental adaptations, different species have different

vulnerabilities to stress-related bleaching and associated mortality at a particular site

(Cook et al., 1990; Hoeksema, 1991). Prevalence of bleaching seems to be higher in

species that have high rates of recruitment and grow rapidly than in other corals in the

same general environment (Gates, 1990; Jokiel & Coles, 1990), and mortality after the

bleaching is higher in some branching species (particularly members of Acropora) than in

many massive corals (Brown & Suharsono, 1990).

The results in this study showed that after 7 weeks of exposure to high

concentrations of CO2, the corals started to bleach (loss of pigmentation). These results

indicate that future predictions of bleaching in response global warming must also take

account of the additional effect of acidification and suggests that any potential adaptation

and acclimatization by coral reef organisms to thermal stress (Baker et al., 2004;

Berkelmans & van Oppen, 2006) may be compensate or overridden by CO2 effects.

Previous studies of CO2 enrichment and also warming in corals and in algae have not

observed a bleaching response (Reynaud et al., 2003; Schneider & Erez, 2006). One

possible explanation is that this experimental period of CO2 dosing in this study was

longer than for example the one in the experiment made by Reynaud et al. 2003, thus

giving time for the buildup of physiological stress. The process by which high CO2 induces

bleaching is still unknown, but according to some authors can involve a number of

possible mechanisms such as photorespiration (Kim et al., 2004), changes to the carbon-

concentrating mechanism (Leggat et al., 1999), and/or direct impacts of acidosis (Gattuso

et al., 1999). According to Anthony et al., 2008, A. Crawley, S.D., & K.R.N.A., unpublished

data, obtained some results indicating that high CO2 and/or lowered pH disrupt the

photoprotective mechanisms of coral symbionts or algal chloroplasts by lowering rates of

photorespiration and the capacity for thermal dissipation.

Anthony et al., 2008 found that, under warm conditions, productivity in Porites

dropped dramatically at the highest CO2 level (1000-1300 ppm), but only 20% (and

nonsignificantly) lower than at control conditions, suggesting a generally weaker response

to acidification than, for example, Acropora.

The productivity responses of the corals to CO2 dosing are likely to be a result of a

series of opposing mechanisms. With a bleaching event, a decline in the symbiont

52

population (or the chlorophyll pool) overrides the increased photosynthetic efficiencies,

leading to a drop in areal productivity. CO2 is the substrate for photosynthesis, and

increasing its supply may increase rates of photosynthesis in organisms that are CO2

limited (Anthony et al., 2008). The potential effects of increasing CO2 and/or impacts of

acidosis on inorganic carbon acquisition are likely to be highly variable in different

organisms, as are the thresholds in temperature that dictate whether this thermal increase

leads to a negative or positive response.

4.4 Projections

Using the data obtained, an estimate was made of the predicted decline in growth in

terms of buoyant weight and linear extension, in chlorophyll a/cm2 and in zooxanthellar

density at CO2 levels 550 and 950 ppm. These values were used because they represent

the levels projected in Fourth Assessment Report (AR4) best and worst CO2 projections

for the years 2050 and 2100 respectively, based on scenarios B1 and A1F1 projected by

the Intergovernmental Panel on Climate Change (IPCC 2007). Table 6 shows the results

of those calculations.

Table 6 – Predicted changes in growth rates, Chlorophyll a and zooxanthellar densities for

the years 2050 and 2100, based on IPCC 2007 B1 and A1F1 CO2 emission scenarios.

Year pCO2 (ppm)

Ωaragonite

Buoyant Weight (mg d-1 cm-2)

Linear Extension

(µm d-1)

Zooxanthellae cm-2

µg Chl a cm-2

Present 390 3.91 12.4 107.2 3.82 x 106 6.8

2050 550 3.04 10.2 85.2 3.74 x 106 5.8

2100 950 2.14 7.9 62.4 3.55 x 106 4.5

% Decrease by 2050 -17.7 -20.6 -2.0 -14.4

% Decrease by 2100 -36.0 -41.8 -7.0 -33.2

Growth through buoyant weight is projected to decline by 18% by 2050 and 36% by

2100, while through linear extension is expected a decline around 21% by 2050 and 42%

by 2100. Marubini et al., 2001 showed that calcification in Porites compressa is correlated

53

with saturation state; as Ω decreased from 5 (Last Glacial Maximum) to 2.3/2.5 (Year

2100), the estimate maximum calcification rate decreased by 30%. These results are

consistent with the ones obtained in this experiment. Anthony et al., 2008 also showed a

significant reduction in calcification for the species Acropora intermedia and Porites

lobata. For the first species, the rate of calcification in the high-CO2 dosing regime (1000-

1300 ppmV), was ~40% lower than at control conditions (380 ppmV). The calcification

response of Porites was almost identical to that of Acropora. In this experiment, these

authors also studied the effects of high temperatures. However, for these two species, the

temperature effect was not very significant. Albright et al., 2010 also showed reductions in

growth rate, but by linear extension, of Acropora palmata juveniles. Comparing with the

control, there was a significant reduction of 39% at the mid-CO2 level (560 µatm) and 50%

at the high-CO2 level (800 µatm). Reynaud et al., 2003 demonstrated that the rate of

calcification of Stylophora pistillata could decrease by 50% between 2000 and 2100.

The amount of Chla per cm2 of coral tissue is projected to decline 14% by 2050 and

33% by 2100. These predictions are quite similar with other studies. So, even if the CO2

levels used here are higher than the ones used in previous studies, the predicted changes

for the years 2050 and 2100 are in line with these. Reynaud et al., 2003 obtained an

increase of photosynthesis with increasing temperature under normal pCO2, which is in

agreement with previous studies performed on scleractinian corals (Coles & Jokiel, 1977,

Coles & Jokiel, 1978, Kajiwara et al., 1995). These last authors shoed that elevated pCO2

did not stimulate photosynthesis, which even decreased slightly. Langdon et al., 2003 also

proved that net community production did not change in response to elevated pCO2.

4.5 Final Considerations

The scientific knowledge of the biological effects of ocean acidification is in its

infancy and the long-term consequences of changing seawater chemistry on marine

ecosystems can only be theorized. However, we know most about the calcification

response for shallow-water scleractinian corals. There are some data sets allowing the

identification of a threshold seawater carbonate chemistry below which ocean acidification

will cause a net calcification rate less than zero (i.e. dissolution) (Hoegh-Guldberg et al.,

2007).

This study did not support the existence of a threshold for Porites porites. Growth

rate decreased with increasing CO2 but remained greater than zero up to the highest level

54

tested, i.e. 6000 ppm. We also observed a decrease in the Chla concentration per unit

area of coral tissue. The zooxanthellae density did not change significantly so the amount

Chla per zooxanthellae cell declined. This could presumably lead to a decrease in the

amount of photosynthate available to the coral host under high CO2 conditions and this

could lead to reduced coral fitness and might explain the observed decrease in

calcification rate. However, photosynthetic rates were not measured in this study so this

hypothesis cannot be confirmed. Anthony et al., 2008 reported a decrease in the pigment

content (based on a semi-quantitative luminance scale) of two Pacific coral species,

Acropora intermedia and Porites lobata at a CO2 level of 1010-1350 ppm. The net

photosynthetic rate of P. lobata declined 30% under normal temperature conditions (27ºC)

relative to the ambient CO2 control while the net photosynthetic rate of the A. intermedia

declined to essentially zero.

Considering the results of this study, the next obvious step is to focus on the

interaction between temperature and saturation state, as accurate predictions of global

change have to include the effects of both saturation state and temperature on coral

growth. Besides the temperature, light and water motion there are other environmental

parameters that should be investigated together with pCO2 to predict the response of

corals calcification and coral communities to global environment changes. Another

important challenge that has received little attention is whether long-term (years to

decades) physiological acclimation and adaptation have the potential to mitigate acute

physiological responses (Gattuso & Buddemeier, 2000).

There is also a need to further investigate the ability of corals to acclimatize and/or

adapt to elevated pCO2 given prolonged exposure, as well as the possibility of taxonomic

differences in sensitivity. Focusing efforts on the protection and cultivation of more

adaptable species may improve the effectiveness of coral preservation and restoration

efforts.

Some priorities must be studies which increase the ecological understanding of the

mechanisms and interactions that structure marine communities; also some experiments

that quantify the capacity of animals to adapt and evolve must be done. To be able to

predict the effects of not only ocean acidification but other, as yet unforeseen,

environmental challenges faced by marine ecosystems, an investment in multidisciplinary

research that brings together community and evolutionary biology, and physiology with

population is necessary.

55

5. References

Albright, R., Mason, B., Langdon, C. (2008). Effect of aragonite saturation state on

settlement and post-settlement growth of Porites astreoides larvae. Coral Reefs 27: 485-

490.

Albright, R., Mason, B., Miller, M., Langdon, C. (2010). Ocean acidification compromises

recruitment success of the threatened Caribbean coral Acropora palmata. PNAS.

Allen, G. (2000). Marine life of the Pacific and Indian Oceans. Periplus Editions (HK) Ltd,

Tuttle Publishing, USA.

Anlauf, H., D‟Croz, L., O‟Dea, A. (2011). A corrosive concoction: The combined effects of

ocean warming and acidification on the early growth of a stony coral are multiplicative. J.

Exp. Mar. Biol. Ecol. 397: 13-20.

Anthony, K.R.N. (2000). Enhanced particle-feeding capacity of corals o turbid reefs (Great

Barrier Reef, Australia). Coral reefs 19: 59-67.

Anthony, K.R.N., Kline, D.I., Diaz-Pulido, G., Dove, S., Hoegh-Guldberg, O. (2008).

Ocean acidification causes bleaching and productivity loss in coral reef builders. PNAS,

Vol. 105, nº 45: 17442-17446.

Babcock, R.C. (1991). Comparative demography of three species of scleractinian corals

using age- and size-dependent classifications. Ecol Monogr 61: 225-244.

Babcock, R.C. & Smith, L. (2000). Effects of sedimentation on coral settlement and

survivorship. Proceedings 9th International Coral Reef Symposium, Bali, Indonesia.

Bacelar, A. (1997). Aquários marinhos de recifes de coráis: montagem e manutenção.

Livraria Nobel, S.A., São Paulo, Brasil, 176 p.

Baker, A., Starger, C.J., McClanahan, T., Glynn, P.W. (2004). Corals‟ adaptive response

to climate change. Nature 430: 741.

56

Barnes, R.S.K. (1998). The diversity of living organisms. Blackwell Science Ltd, University

of Cambridge, U.K., 345 p.

Barnes, D.J. & Taylor, D.L. (1973). In situ studies of calcification and photosynthetic

carbon fixation in the coral Montastrea annularis. Helgol Wiss Meeresunters 24: 284-291.

Barnola, J.M., Raynaud, D., Korotkevich, Y.S., Lorius, C. (1987). Vostok ice core provides

160,000-year record of atmospheric CO2. Nature 329: 408-414.

Banin, E.Y., Ben-Haim, T.I., Israely, T., Loya, Y., Rosenberg, E. (2000). Effect of the

environment on the bacterial bleaching of corals. Water Air Soil Poll. 123: 337–352.

Bessat, F., Buigues, D. (2001). Two centuries of variation in coral growth in a massive

Porites colony from Moorea (French Polynesia): a response of ocean-atmosphere

variability from south central Pacific. Palaeogeogr. Palaeoclimatol. Palaeooecol. 175: 381-

392.

Berkelmans, R. & van Oppen, M.J.H. (2006). The role of zooxanthellae in the thermal

tolerance of corals: A „nugget of hope‟ for coral reefs in an era of climate change. Proc.

Roy. Soc. B 273: 2305-2312.

Brown, B.E. & Suharsono (1990). Damage and recovery of coral reefs affected by El Niño

related seawater warming in the Thousand Islands, Indonesia. Coral reefs 8: 163-170.

Brown, B.E., Le Tissier, M.D.A., Bythell, J.C. (1995). Mechanisms of bleaching deduced

from histological studies of reef corals sampled during a natural bleaching event. Mar.

Biol. 122: 655-663.

Brown, B.E. (1997). Coral bleaching: causes and consequences. Coral Reefs 16, Suppl:

5129-5138.

Caldeira, K. & Wickett, M.E. (2003). Anthropogenic carbon and ocean pH. Nature 425:

365.

Castro, P. & Huber, M.E. (2007). Marine Biology. McGraw-Hill, USA, 459 p.

57

Clausen, C.D. & Roth, A.A. (1975). Effect of temperature and temperature adaptation on

calcification rate in the hermatypic coral Pocillopora damicornis. Mar. Biol. 33: 93-100.

Cohen, A.L., McCorkle, D.C., de Putron, S., Getani, G.A., Rose, K.A. (2009).

Morphological and compositional changes in the skeletons of new coral recruits reard in

acidified seawater: insights into the biomineralization response to ocean acidification.

Geochem. Geophys. Geosyst. 10, Q07005.

Coles, S.L. & Jokiel, P.L. (1977). Effects of temperature on photosynthesis and respiration

in hermatypic corals. Marine Biology 43: 209-216.

Coles, S.L. & Jokiel, P.L. (1978). Synergistic effects of temperature, salinity and light on

the hermatypic coral Montipora verrucosa. Marine Biology 49: 187-195.

Coles, S.L. & Brown, B.E. (2003). Coral bleaching – capacity for acclimatization and

adaptation. Adv. Mar. Biol., Vol. 46: 183-223.

Cook, C.B., Logan, A., Ward, J., Luckhurst, B., Berg, C.J. (1990). Elevated temperatures

and bleaching on a high latitude coral reef: the 1988 Bermuda event. Coral Reefs 9: 45-

49.

Cooper, T.F., De‟ath, G., Fabricius, K.E., Lough, J.M. (2008). Declining coral calcification

in massive Porites in two nearshore regions of the northern Great Barrier Reef. Glob.

Change Biol. 14: 529-538.

Côté, I.M. & Reynolds, J.D. (2006). Conservation Biology 13, Coral reef Conservation.

Cambridge University Press, U.K., 568 p.

Davies, P.S. (1989). Short-term growth measurements of corals using an accurate

buoyant weighing technique. Mar. Biol. 101: 389-395.

Davies, P.S. (1995). Coral nubbins and explants for reef assessment and laboratory

ecotoxicology. Coral Reefs 14: 267-269.

Dayton, P.K., Thrush, S.F., Agardy, M.T., Hofman, R.J. (1995). Environmental effects of

marine fishing. Aquat. Conserve 5: 2-28.

58

De‟ath, G., Lough, J.M., Fabricius, K.E. (2009). Declining coral calcification on the Great

Barrier Reef. Science 323: 116-119.

Delbeek, J.C. & Sprung, J. (1994). The Reef Aquarium: a comprehensive guide to the

identification and care of tropical marine invertebrates (Volume 1). Ricordea Pub.,

Coconut Grove, Florida, 544 p.

Doney, S.C., Fabry, V.J., Feely, R.A., Kleypas, J.A. (2009). Ocean acidification: the other

CO2 problem. Annu. Rev. Mar. Sci. 1: 169-192.

Drew, E.A. (1972). The biology and physiology of algal-invertebrate symbiosis. II. The

density of algal cells in a number of hermatypic hard corals and alcyonarians from various

depths. J. Exp. Mar. Biol. Ecol. 9: 71-75.

Dubinsky, Z. & Jokiel, P.L. (1994). Ratio of energy and nutrient fluxes regulates symbiosis

between zooxanthellae and corals. Pac. Sci. 48: 313-324.

Edmunds, P.J. & Davies, P.S. (1986). An energy budget for Porites porites (Scleractinian).

Mar. Biol. 92: 339-347.

Edmunds, P.J. & Davies, P.S. (1989). An energy budget for Porites porites (Scleractinian),

growing in a stressed environment. Coral Reefs 8: 37-43.

Egana, A.C., & DiSalvo, L.H. (1982). Mass expulsion of zooxanthellae by Easter Island

corals. Pac. Sci. 36: 61–63.

Emerson, S.T. & Hedges, J. (2008). Chemical oceanography and the marine carbon

cycle. Cambridge University Press, Cambridge, U.K., 453 p.

Endean, R. (1976). Destruction and recovery of coral reefs communities. In: Jones, O.A.,

Endean, R. (ed.) Biology and geology of coral reefs. Vol. 3, Biol. 2. Academic Press,

London, pp. 215-254.

Falkowski, P.G., Jokiel, P.L., Kinzie, R.A. (1990). Irradiance and corals. In: Dubinsky Z

(ed) Ecosystems of the world 25: Coral reefs. Elsevier, Amsterdam, pp. 89-107.

59

Feely, R.A., Byrne, R.H., Acker, J.G., Betzer, P.R., Chen C.T.A., Gendron, J.F., Lamb,

M.F. (1988). Winter summer variations of calcite and aragonite saturation in the northeast

Pacific. Mar. Chem. 25: 227-241.

Feely, R.A. & Chen, C.T.A. (1982). The effect of excess CO2 on the calculated calcite and

aragonite saturation horizons in the northeast Pacific. Geophys. Res. Lett. 9: 1294-1297.

Feely, R.A., Orr, J., Fabry, V.J., Kleypas, J.A., Sabine, C.L., Langdon, C. (2009). Present

and future changes in seawater chemistry due to ocean acidification. In AGU Monograph

on the Science and Technology of Carbon Sequestration. ed. B.J. McPherson, E.T.

Sundquist. Am. Geophys. Union. In press.

Ferrier-Pagès, C., Gattuso, J.P., Dallot, S., Jaubert, J. (2000). Effect of nutrient

enrichment on growth and photosynthesis of the zooxanthellae coral Stylophora pistillata.

Coral Reefs 19: 103-114.

Fitt, W.K., Spencer, H.J., Halas, J., White, M.W., Porter, J.W. (1993). Recovery of

Montastrea annularis in the Florida Keys after the 1987 Caribbean „bleaching event‟. Coral

Reefs 12: 57-64.

Fitt, W.K. & Warner, M.E. (1995). Bleaching patterns of 4 species of Caribbean reef

corals. Biol Bull 189: 298-307.

Gates, R. (1990). Seawater temperature and sublethal coral bleaching in Jamaica. Coral

reefs 8: 193-197.

Gattuso, J.P., Frankignoulle, M., Bourge, I., Romaine, S., Buddemeier, R.W. (1998).

Effect of calcium carbonate saturation of seawater on coral calcification. Global Planet

Change 18: 37-46.

Gattuso, J.P., Allemand, D., Frankignoulle, M. (1999). Photosynthesis and calcification at

cellular, organismal and community levels in coral reefs: A review on interactions and

control by carbonate chemistry. Am. Zool. 39: 160-183.

Gattuso, J.P. & Buddemeier, R.W. (2000). Ocean biogeochemistry: calcification and CO2.

Nature 407: 311-313.

60

Ginsburg, R.N., ed. (1994). Proceedings of the Colloquium on Global aspects of coral

reefs: Health, Hazards and History, 1993. Rosenstiel School of Marine and Atmospheric

Sciences, University of Miami, Florida, 420 p.

Glynn, P.W. & Wellington, G.M. (1983). Corals and Coral Reefs of the Galapagos Islands.

University of California Press, USA, 330 p.

Glynn, P.W. (1984). Widespread coral mortality and the 1982/83 El Niño warming event.

Environ. Conserv. 11: 133-156.

Glynn, P.W. & D‟Croz, L. (1990). Experimental evidence for high temperature stress as

the cause of El Niño-coincident coral mortality. Coral reefs 8: 181-192.

Glynn, P.W., Imai, R., Sakai, K., Nakano, Y., Yamazato, K. (1992). Experimental

responses of Okinawan (Ryukyu Islands, Japan) reef corals to high sea temperature and

UV radiation. Proc. 7th Int. Coral Reef Symp, Guam 1: 27-37.

Glynn, P.W. (1993). Coral reef bleaching ecological perspectives. Coral Reefs 12: 1-17

Goreau, T.J. (1964). Mass expulsion of zooxanthellae from Jamaican reef communities

after Hurricane Flora. Science 145: 383–386.

Grigg, R.W. & Birkeland, C. (1997). Status of coral reefs in the Pacific. Sea Grant College

Program, School of Ocean and Earth Science and Technology, University of Hawaii,

Honolulu, HI, USA.

Guinotte, J. M. & Fabry, V. J. (2008). Ocean acidification and its potential effects on

marine ecosystems. Ann. N. Y. Acad. Sci. 1134: 320-342.

Harvell, C.D., Mitchell, C.E., Ward, J.R., Altizer, S., Dobson, A.P., Ostfeld, R.S., Samuel,

M.D. (2002). Climate warning and disease risks for terrestrial and marine biota. Science

296: 2158-2162.

Hoegh-Guldberg, O. & Smith, G.J. (1989a). The effects of sudden changes in light,

temperature and salinity on the population density and export of zooxanthellae from the

reef corals Seriatopora hystrix and Stylophora pistillata. J. Exp. Mar. Biol. Ecol. 129: 279-

303.

61

Hoegh-Gulberg, O. & Smith, G.J. (1989b). Influence of the population density of

zooxanthellae and supply of ammonium on the biomass and metabolic characteristics of

the reef corals Seriatophora hystrix and Stylophora pistillata. Mar. Ecol. Prog. Ser. 57:

173-186.

Hoegh-Guldberg, O. & Salvat, B. (1995). Periodic mass bleaching and elevated seawater

temperatures: bleaching of outer reef slope communities in Moorea, French Polynesia.

Mar. Ecol. Prog. Ser. 121: 181-190.

Hoegh-Guldberg, O. (1999). Climate change, coral bleaching on the future of the world‟s

coral reefs. Mar. Freshwater Res. 50: 839-866.

Hoegh-Guldberg, O., Mumby, P.J., Hooten, A.J., Steneck, R.S., Greenfield, P., Gomez,

E., Harvell, C.D., Sale, P.F., Edwards, A.J., Caldeira, K., Knowlton, N., Eakin, C.M.,

Iglesias-Prieto, R., Muthiga, N., Bradbury, R.H., Dubi, A., Hatziolos, M.E. (2007). Coral

reefs under rapid climate change and ocean acidification. Science 318: 1737-1742.

Hoeksema, B.W. (1991). Control of bleaching in mushroom coral populations

(Scleractinia: Fungiidae) in the Java Sea: stress tolerance and interference by life history

strategy. Mar. Biol. Prog. Ser. 74: 225-237.

Houghton, J.T. (2001). Climate change 2001: The Scientific Basis. Cambridge University

Press, Cambridge, UK., 94 p.

Iglesias-Prieto, R., Matta, J.L., Robins, W.A., Trench, R.K. (1992). Photosynthetic

response to elevated temperature in the symbiotic dinoflagellate Symbiodinium

microadriaticum in culture. Proc. Natl. Acad, Sci. USA 89: 10302-10305.

Johannes, R.E. & Wiebe, W.J. (1970). Method for determination of coral tissue biomass

and composition. Limnol Oceanogr 15: 822-824.

Johannes, R.E. (1975). Pollution and degradation of coral reef communities. In: Ferguson

Wood, E.J., Johannes, R.E. (ed.) Tropical marine pollution. Elsevier, Amsterdam, pp. 13-

51.

62

Johannes, R.E. & Betzer, S.B. (1975). Introduction: marine communities respond

differently to pollution in the tropics than at higher latitudes. In: Ferguson Wood, E.J.,

Johannes, R.E. (ed.) Tropical marine pollution. Elsevier, Amsterdam, pp. 1-12.

Johansen, H.W. (1981). Coralline algae, a first synthesis. CRC Press, Inc., Boca Raton,

FL, 239 p.

Jokiel, P.L. & Coles, S.L. (1977). Effects of temperature on the mortality and growth of

Hawaiian reef corals, Mar. Biol. 43: 201-208.

Jokiel, P.L., Maragos, J.E., Franzisket, L. (1978). Coral growth: buoyant weight technique.

In: Stoddart, D.R., Johannes, R.E. (ed.) Coral reefs: research methods. UNESCO

monographs on oceanographic methodology, Vol. 5, pp. 529-542.

Jokiel, P.L. & Coles, S.L. (1990). Response of Hawaiian and other Indo-Pacific reef corals

to elevated temperature. Coral reefs 8: 155-162.

Jokiel, P.L., Rodgers, K.S., Kuffner, I.B., Andersson, A.J., Cox, E.F., Mackenzie, F.T.

(2008). Ocean acidification and calcifying reef organisms: A mesocosm investigation.

Coral Reefs 27: 473-483.

Jones, R.J. (1997a). Zooxanthellae loss as a bioassay for assessing stress in corals. Mar.

Ecol. Prog. Ser. 149: 163–171.

Jones, R. J. (1997b). Changes in zooxanthellar densities and chlorophyll concentrations in

corals during and after a bleaching event. Mar. Ecol. Prog. Ser. 158: 51-59.

Jones, R.J. & Steven, A.L. (1997). Effects of cyanide on corals in relation to cyanide

fishing on reefs. Mar. Freshwater Res. 48: 517–522.

Kajiwara, K., Nagai, A., Ueno, S., Yokochi, H. (1995). Examination of the effect of

temperature, light intensity, and zooxanthellae concentration on calcification and

photosynthesis of Scleractinian coral Acropora pulchra. Journal of the Schoolof Marine

Science and Technology 40: 95-103.

Karleskint, G., Turner, R., Small, J. (2009). Introduction to Marine Biology, Third Edition.

Cengage Learning, USA, 581 p.

63

Kim, Y., Yakunin, A.F., Kuznetsova, E., Xu, X., Pennycooke, M., Gu, J., Cheung, F.,

Proudfoot, M., Arrowsmith, C.H., Joachimiak, A., Edwards, A.M., Christendat, D. (2004).

Structure- and function-based characterization of a new phosphoglycolate phosphatase

from Termoplasma acidophilum. J. Biol. Chem. 279: 517-526.

Kleypas, J.A., Buddemeier, R.W., Archer, D., Gattuso, J.P., Langdon, C., Opdyke, B.N.

(1999). Geochemical consequences of increased atmospheric carbon dioxide on coral

reefs. Science 284: 118-120.

Kleypas, J.A. & Langdon, C. (2002). Overview of CO2-induced changes in seawater

chemistry. World Coral Reefs in the New Millennium: Bridging Research and Management

for Sustainable Development. In Proceedings of the 9th International Coral Reef

Symposium, 2. M. K. Moosa, S. Soemodihardjo, A. Soegiarto, et al., Eds.: 1085-1089.

Ministry of Environment, Indonesian Institute of Sciences, International Socitey for Reef

Studies, Bali, Indonesia.

Kleypas, J.A., Feely, R.A., Fabry, V.J., Langdon, C., Sabine, C.L., Robbins, L.L. (2006).

Impacts of ocean acidification on coral reefs and other marine calcifiers: a guide for future

research. 88 pp. Report of a workshop sponsored by NSF, NOAA, and the U.S.

Geological Survey. St. Petersburg, Florida.

Krebs, C.J. (1994). Ecology: The experimental analysis of distribution and abundance,

Fourth Edition. Harper Collins College Publishers, New York, 801 p.

Kushmaro, A.E., Rosenberg, M.F., Loya, Y. (1997). Bleaching of the coral Oculina

patagonica by Vibrio AK-1. Mar. Ecol. Prog. Ser. 147: 159–165.

Landgon, C., Takahashi, T., Sweeney, C., Chipman, D., Goddard, J., Marubini, F.,

Aceves, H., Barnett, H., Atkinson, M.J. (2000). Effect of calcium carbonate saturation state

on the calcification rate of an experimental coral reef. Global Biogeochem. Cy., Vol. 14,

NO. 2, 639-654.

Langdon, C., Broecker, W.S., Hammond, D.E., Glenn, E., Fitzsimmons, K., Nelson, S.G.,

Peng, T.H., Hajdas, I., Bonani, G. (2003). Effect of elevated CO2 on the community

metabolism of an experimental coral reef. Global Biogeochem. Cy., 17(1), 1011, doi:

10.1029/ 2002GB001941.

64

Langdon, C. & Atkinson, M. J. (2005). Effect of elevated pCO2 on photosynthesis and

calcification of corals and interactions with seasonal change in temperature/irradiance and

nutrient enrichment. J. Geophys. Res. Vol. 110, C09S07.

Leclercq, N., Gattuso, J., Jaubert, J. (2000). CO2 partial pressure controls the calcification

rate of a coral community. Global Change Biol. 6: 329-334.

Leclercq, N., Gattuso, J.P., Jaubert, J. (2002). Primary production, respiration, and

calcification of a coral reef mesocosm under increased CO2 partial pressure. Limnol.

Oceanogr. 47: 558-564.

Leggat, W., Badger, M.R., Yellowlees, D. (1999). Evidence for an inorganic carbon-

concentrating mechanism in symbiotic dinoflagellate Symbiodinium sp. Plant Physiol. 121:

1247-1255.

Lesser, M.P., Stochaj, W.R., Tapley, D.W., Shick, J.M. (1990). Bleach in coral reef

anthozoans: effects of irradiance, ultraviolet radiation, and temperature on the activities of

protective enzymes against active oxygen. Coral reefs 8: 225-232.

Le Tissier, M.D.A. & Brown, B.E. (1996). Dynamics of solar bleaching in the intertidal reef

coral Goniastrea aspera at Ko Phuket, Thailand. Mar. Ecol. Prog. Ser. 136: 235-244.

Loáiciga, H.A. (2006). Modern-age buildup of CO2 and its effects on seawater acidity and

salinity. Geophys. Res. Lett. 33: L10605. Doi: 10.1029/2006GL026305.

Loya, Y., Sakai, K., Yamazato, K., Nakano, Y., Sambali, H., van Woesik, R. (2001). Coral

bleaching, the winners and the losers. Ecol. Lett. 4: 122-131.

Luchavez, T.F. & Alcala, A.C. (1988). Effect of fishing pressure on coral reef fishes in the

central Philippines. In Proceedings of the Sixth International Coral Reef Symposium, 8-12

August 1988. Townsville, Australia, 1988, (ed). Choat, J.H. et al. vol. 2, pp. 251-254.

Marsh, J.A. (1970). Primary productivity of reef-building calcareous and red algae.

Ecology 55: 225-263.

65

Marshall, A.T. & Clode, P. (2004). Effects of calcium free and low calcium artificial

seawater on polyps of a scleractinian coral Galaxea fascicularis. Coral Reefs 23: 277-280.

Marubini, F. & Davies, P.S. (1996). Nitrate increases zooxanthellae population density

and reduces skeletogenesis in corals. Mar. Biol. 127: 319-328

Marubini, F. & Atkinson, M.J. (1999). Effects of lowered pH and elevated nitrate on coral

calcification. Mar. Ecol. Prog. Ser. 188: 117-121.

Marubini, F. & Thake, B. (1999). Bicarbonate addition promotes coral growth. Limnol.

Oceanogr. 44: 716-720.

Marubini, F., Barnett, H., Langdon, C., Atkinson, M.J. (2001). Dependence of calcification

on light and carbonate ion concentration for the hermatypic coral Porites compressa. Mar.

Ecol. Prog. Ser. 220: 153-162.

Marubini, F., Ferrier-Pages, C., Cuif, J.P. (2003). Suppression of growth in scleractinian

corals by decreasing ambient carbonate ion concentration: a cross-family comparison.

Proc. R. Soc. Lond. Ser. B. Biol. Sci. 270: 179-184.

McAllister, D.E. (1998). Environmental, economic and social costs of coral reef destruction

in the Philippines. Galaxea 7: 161-178.

McClanahan, T.R., Ateweberhan, M., Graham, N.A.J., Wilson, S.K., Ruiz Sebastián, C.,

Guillaume, M.M.M., Bruggemann, J.H. (2007). Western Indian Ocean coral communities:

bleaching responses and susceptibility to extinction. Mar. Ecol. Prog. Ser. 337: 1-13.

Meehan, W.J. & Ostrander, G.K. (1997). Coral bleaching: a potential biomarker of

environmental stress. J. Toxicol. Env. Health 50: 529–552.

Mehrbach, C., Culberson, C.H., Hawley, J.E., Pytkowicz, R.M. (1973). Measurement of

apparent dissociation constants of carbonic acid in seawater at atmospheric pressure.

Limnol. Oceanogr. 18: 897-907.

Munro, J.L. & Munro, P.E. (1994). The Management of Coral Reef Resource Systems.

International Center for Living Aquatic Resources Management Conf. Proc. 44, 124 p.

66

Muscatine, L. & Porter, J.W. (1977). Reef corals: mutualistic symbiosis adapted to nutrient

poor environments. Bio-Science 27: 454-460.

Muscatine, L. (1990). The role of symbiotic algae in carbon and energy flux in reef corals.

Coral Reefs 25: 1–29.

Nakamura, T., Yamasaki, H., van Woesik, R. (2003). Water flow facilitates recovery from

bleaching in the coral Stylophora pistillata. Mar. Ecol. Prog. Ser. 256: 287-291.

NRC (National Research Council) (1995). Understanding Marine Biodiversity: a research

agenda for the nation. Committee on biological diversity in marine systems, National

Academy Press, Washington, D.C., USA, 114 p.

NRC (National Research Council) (2010). Ocean Acidification: A national strategy to meet

the challenges of a changing ocean. National Academy Press, Washington, D.C., USA,

188 p.

Ohde, S., van Woesik, R. (1999). Carbon dioxide flux and metabolic processes of a coral

reef, Okinawa. Bull. Mar. Sci. 65: 559-576.

Orr, J.C., Fabry, V.J., Aumont, O., Bopp, L., Doney, S.C., Feely, R.A., Gnanadesikan, A.,

Gruber, N., Ishida, A., Joos, F., Key, R.M., Lindsay, K., Maier-Reimer, E., Matear, R.J.,

Monfray, P., Mouchet, A., Najjar, R.G., Plattner, G.K., Rodgers, K.B., Sabine, C.L.,

Sarmiento, J.L., Schlitzer, R., Slater, R.D., Totterdell, I.J., Weirig, M.F., Yamanaka, Y.,

Yool, A. (2005). Anthropogenic ocean acidification over the twenty-first century and its

impact on calcifying organisms. Nature 437: 681-686.

Pastorok, R.A. & Bilyard, G.R. (1985). Effects of sewage pollution on coral-reef

communities. Mar. Ecol. Prog. Ser. 21: 175-189.

Pearson, R. (1981). Recovery and recolonization of coral reefs. Mar. Ecol. Prog. Ser. 4:

105-122.

Porter, J.W., Muscatine, L., Dubinsky, Z., Falkowski, P.G. (1984). Primary production and

photoadaptation in light- and shade-adapted colonies of the symbiotic coral Stylophora

pistillata. Proc. R. Soc. Lond. B 222: 161-180.

67

Rees, T.A.V. (1991). Are symbiotic zooxanthellae nutrient deficient? Proc. R. Soc. Lond.

Ser. B. Biol. Sci. 243: 227-233.

Renegar, D.A. & Riegl, B.M. (2005). Effect of nutrient enrichment and elevated CO2 partial

pressure on growth rate of Atlantic scleractinian coral Acropora cervicornis. Mar. Ecol.

Prog. Ser. 293: 69-76.

Reynaud, S., Leclercq, N., Romaine-Lioud, S., Ferrier-Pages, C., Jaubert, J., Gattuso,

J.P. (2003). Interacting effects of CO2 partial pressure and temperature on photosynthesis

and calcification in a scletactinian coral. Glob. Chang. Biol. 9: 1660-1668.

Reynaud-Vaganay, S., Gattuso, J.P., Cuif, J.P., Jaubert, J., Juillet-Leclerc, A. (1999). A

novel culture technique for scleractinian corals: application to investigate changes in

skeletal delta super (18)O as a function of temperature. Mar. Ecol. Prog. Ser. 180: 121-

130.

Riegl, B. & Branch, G.M. (1995). Effects of sedimentation on the energy budget of four

scleractinian (Bourne 1900) and five alcyonacean (Lamouroux 1816) corals. J. Exp. Mar.

Biol. Ecol. 186: 259-275.

Rinkevich, B. & Loya, Y. (1984). Does light enhance calcification in hermatypic corals?

Mar. Biol. 80: 1-6.

Roberts, C.M. (1995). Effects of fishing on the ecosystem structure of coral reefs.

Conserv. Biol. 9: 65-91.

Royal Society (2005). Ocean acidification due to increasing atmospheric carbon dioxide.

London: The Royal Society, 57 p.

Rubec, P.J. (1986). The effects of sodium cyanide on coral reefs and marine fishes in the

Philippines. In MacLean, J.L., Dizon, L.B., Hosillos, L.V. (ed.). The first Asian fisheries

forum. Manilla: Asian Fisheries Society.

Safina, C. (1995). The world‟s imperiled fish. Sci. Am. 273: 46-53.

68

Schneider, K. & Erez, J. (2006). The effect of carbonate chemistry on calcification and

photosynthesis in the hermatypic coral Acropora eurystoma. Limnol. Oceanogr. 51: 1284-

1293.

Sharp, V.A. (1995). Temperature tolerance and heat shock protein production by

temperate and tropical cnidarians. PhD Dissertation, University of Newcastle upon Tyne,

UK.

Siegenthaler, U., Stocker, T.F., Monnin, E., Lüthi, D., Schwander, J., Stauffer, B.,

Raynaud, D., Barnola, J.M., Fischer, H., Masson-Delmotte, V., Jouzel, J. (2005). Stable

carbon cycle-climate relationship during the late Pleistocene. Science 310: 1313-1317.

Smith, S.V. & Buddemeier, R.W. (1992). Global change and coral reef ecosystems. Annu.

Rev. Ecol. Syst. 23: 89-118.

Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M. et al. (2007). Climate change

2007: The Physical Science Basis: Contribution of Working Group I to the Fourth

Assessment Report of the Intergovernmental Panel on Climate Change. New York:

Cambridge Univ. Press., 996 p.

Spalding, M., Ravilious, C., Green, E.P. (2001). World atlas of Coral Reefs. Prepared at

the UNEP World Conservation Monitoring Centre. University of California Press, Berkeley,

USA, 424 p.

Spalding, M.D. (2004). A guide to the Coral Reefs of the Caribbean. University of

California Press, Berkeley, USA, 256 p.

Stafford-Smith, M.G. (1993). Sediment-rejection efficiency of 22 species of Australian

scleractinian corals. Mar. Biol. 115: 229-243.

Stambler, N., Popper, N., Dubinsky, Z., Stimson, J. (1991). Effect of nutrient enrichment

and water motion on the coral Pocillopora damicornis. Pac. Sci. 45: 299-307

Stanley, S.M. & Hardie, L.A. (1998). Secular oscillations in the carbonate mineralogy of

reef-building and sediment-producing organisms driven by tectonically forced shifts in

seawater chemistry. Palaeogeogr. Palaeoclim. Palaeoecol., 144: 3-19.

69

Sulu, R. (2007). Status of coral reefs in the Southwest Pacific, 2004. IPS Publications,

University of South Pacific, Fiji, 266 p.

Vermeij, M.J.A. & Sandin, S.A. (2008). Density-dependent settlement and mortality

structure the earliest life phases of a coral population. Ecology 89: 1994-2000.

Veron, J.E.N. (1986). Corals of Australia and the Indo-Pacific. University of Hawaii Press,

U.S.A., 644 p.

Voss, G.L. (2002). Seashore life of Florida and the Caribbean. Courier Dover Publications,

New York, USA, 207 p.

Wells, J.A., Ferrari, E., Henner, D.J., Estell, D.A., Chen, E.Y. (1983). Cloning, sequencing,

and secretion of Bacillus amyloliquefaciens subtillisin in Bacillus subtilis. Nucleic Acids

Res. 11: 7911-7924.

Wells, S. & Hanna, N. (1992). The Greenpeace book of coral reefs. Sterling Pub. Co.,

New York, 160 p.

Westmacott, S., Teleki, K., Wells, S., West, J.M. (2000). Management of bleached and

severely damaged coral reefs. IUCN, Gland, Switzerland and Cambridge, UK, 36 p.

Wilkinson, C.R., Chou, L.M., Gomez, E., Ridzwan, A.R., Soekarno, S., Sudara, S. (1993).

In: Proceedings of the Colloquium on Global Aspects of Coral Reefs: Health, Hazards and

History, 1993, compiler (ed.) Ginsburg, R.N., pp. 311-317 University of Miami, Florida.

World Bank (2009). World development record 2010: development and climate change.

World Bank Publications, Washington D.C., USA, 417 p.