detailed paper outline on ” Climate change coral reefs algae”

Please submit your detailed outline (2-3 pages + full-text PDF of each of the 10 articles in the references) here follow these instructions:

The outline must have all major sections of the paper and sub topics for each section. See the following link. How to prepare detailed outline Actions
Each statement needs to have a citation. You must include and cite at least 10 peer-reviewed articles, 7 of which must be primary literature. References must be properly formatted: Citing References in Scientific Papers x Actions Recommended: use reference manager software to handle your citations. There are free services such as Mendeley (Links to an external site.).
Please use this Sample outline Actions to help guide you on what your outline should look like.

Don't use plagiarized sources. Get Your Custom Essay on
detailed paper outline on ” Climate change coral reefs algae”
Just from $13/Page
Order Essay
  • Must include 10 full-text PDF of the cited peer-review articles.
  • Topic The topic I’ve chosen is the “climate change coral reefs algae”. I’m going to address the capacity of coral macroalgal difficulties to recover and the importance to marine life. The effect of temperature, carbon dioxide and benthic algae has on the algo community ecological process. And the negative and positive ways humans are contributing to this fact.

    REVIEW

    OCEAN CLIMATE CHANGE, PHYTOPLANKTON COMMUNITY RESPONSES, AND
    HARMFUL ALGAL BLOOMS: A FORMIDABLE PREDICTIVE CHALLENGE1

    Gustaaf M. Hallegraeff 2

    Institute of Marine and Antarctic Studies, and School of Plant Science, University of Tasmania, Private Bag 55, Hobart,

    Tasmania 7001, Australia

    Prediction of the impact of global climate chang

    e

    on marine HABs is fraught with difficulties. How-
    ever, we can learn important lessons from the fossil
    record of dinoflagellate cysts; long-term monitoring
    programs, such as the Continuous Plankton Recor-
    der surveys; and short-term phytoplankton commu-
    nity responses to El Niño Southern Oscillatio

    n

    (ENSO) and North Atlantic Oscillation (NAO) epi-
    sodes. Increasing temperature, enhanced surface
    stratification, alteration of ocean currents, intensifi-
    cation or weakening of local nutrient upwelling,
    stimulation of photosynthesis by elevated CO2,
    reduced calcification through ocean acidification
    (‘‘the other CO2 problem’’), and heavy precipitation
    and storm events causing changes in land runoff
    and micronutrient availability may all produce con-
    tradictory species- or even strain-specific responses.
    Complex factor interactions exist, and simulated
    ecophysiological laboratory experiments rarely allow
    for sufficient acclimation and rarely take into
    account physiological plasticity and genetic strain
    diversity. We can expect: (i) range expansion of
    warm-water species at the expense of cold-water spe-
    cies, which are driven poleward; (ii) species-
    specific changes in the abundance and seasonal
    window of growth of HAB taxa; (iii) earlier timing of
    peak production of some phytoplankton; and (iv)
    secondary effects for marine food webs, notably
    when individual zooplankton and fish grazers are dif-
    ferentially impacted (‘‘match-mismatch’’) by clima

    te

    change. Some species of harmful algae (e.g., toxic
    dinoflagellates benefitting from land runoff and ⁄ or
    water column stratification, tropical benthic dinofla-
    gellates responding to increased water temperature

    s

    and coral reef disturbance) may become more suc-
    cessful, while others may diminish in areas currently
    impacted. Our limited understanding of marine eco-
    system responses to multifactorial physicochemical
    climate drivers as well as our poor knowledge of the
    potential of marine microalgae to adapt genetically
    and phenotypically to the unprecedented pace of
    current climate change are emphasized. The greatest

    problems for human society will be caused by being
    unprepared for significant range expansions or the
    increase of algal biotoxin problems in currently
    poorly monitored areas, thus calling for increased
    vigilance in seafood-biotoxin and HAB monitoring
    programs. Changes in phytoplankton communities
    provide a sensitive early warning for climate-driven
    perturbations to marine ecosystems.

    Key index words: adaptation; algal blooms; climate
    change; continuous plankton recorder; ENSO;
    NAO; ocean acidification; range expansion

    Abbreviations: DMS, dimethylsulfoxide; ENSO, El
    Niño-Southern Oscillation; GOOS, Global Ocean
    Observation Systems; IMOS, Integrated Marine
    Observing System; IPCC, Intergovernmental
    Panel on Climate Change; NAO, North Atlantic
    Oscillation episodes; OOI, Ocean Observatories
    Initiative

    HABs in a strict sense are completely natural phe-
    nomena that have occurred throughout recorded
    history. Even nontoxic algal blooms can have devas-
    tating impacts, for instance, when they lead to kills
    of fish and invertebrates by generating anoxic con-
    ditions in sheltered bays. Other algal species, even
    though nontoxic to humans, can produce exudates
    or reactive oxygen species that can damage the gill
    tissues of fish (raphidophytes Chattonella and Hetero-
    sigma, and dinoflagellates Cochlodinium, Karenia, and
    Karlodinium). Whereas wild fish stocks can swim
    away from problem areas, caged fish in intensive
    aquaculture operations are trapped and thus can
    suffer catastrophic mortalities. Of greatest concern
    to human society are algal species that produce
    potent neurotoxins that can find their way through
    shellfish and fish to human consumers where they
    produce a variety of gastrointestinal and neurologi-
    cal illnesses. One of the first recorded fatal cases of
    food poisoning after eating contaminated shellfish
    happened in 1793, when English surveyor Captain
    George Vancouver and his crew landed in British
    Columbia (Canada) in an area now known as
    Poison Cove. He noted that, for local Indian tribes,

    1Received 29 March 2009. Accepted 10 September 2009.
    2Author for correspondence: e-mail hallegraeff@utas.edu.au.

    J. Phycol. 46,

    220

    –235 (2010)
    � 2010 Phycological Society of America
    DOI: 10.1111/j.1529-8817.2010.00815.x

    220

    it was taboo to eat shellfish when the seawater
    became bioluminescent due to algal blooms by the
    local dinoflagellate Alexandrium catenella, which we
    now know to be a producer of paralytic shellfish
    poisons (PSP) (Dale and Yentsch 1978).

    The increase in shellfish farming worldwide is
    leading to more reports of paralytic, diarrhetic (first
    documented in 1976 in Japan), neurotoxic
    (reported from the Gulf of Mexico as early as
    1840), amnesic (first identified in 1987 in Canada),
    or azaspiracid shellfish poising (first identified in
    1998 in Ireland). The English explorer Captain
    James Cook already suffered from the tropical ill-
    ness of ciguatera fish poisoning when he visited
    New Caledonia in 1774. Worldwide, close to 2,000
    cases of food poisoning from consumption of con-
    taminated fish or shellfish are reported each year.
    Some 15% of these cases will prove fatal. If not con-
    trolled, the economic damage through the slump in
    local consumption and export of seafood products
    can be considerable. Whales and porpoises can also
    become victims when they take up toxins through
    the food chain via contaminated zooplankton or
    fish. In the USA, poisoning of manatees in Florida
    via seagrasses and their faunal epiphytes, and, in
    California, of pelicans and sea lions via contami-
    nated anchovies have also been reported. In the
    past three decades, HABs seem to have become
    more frequent, more intense, and more widespread
    (Hallegraeff 1993, Van Dolah 2000). There is no
    doubt that the growing interest in using coastal
    waters for aquaculture is leading to a greater aware-
    ness of toxic algal species. People responsible for
    deciding quotas for pollutant loadings of coastal
    waters, or for managing agriculture and deforesta-
    tion, should be made aware that one probable out-
    come of allowing polluting chemicals to seep into
    the environment will be an increase in HABs or a
    change in community structure affecting a change
    in food web. In countries that pride themselves on
    having disease- and pollution-free aquaculture, every
    effort should be made to quarantine sensitive aqua-
    culture areas against the unintentional introduction
    of nonindigenous harmful algal species. Nor can
    any aquaculture industry afford not to monitor for
    an increasing number of harmful algal species in
    water samples and for an increasing number of algal
    toxins in seafood products using increasingly sophis-
    ticated analytical techniques. Last but not least, glo-
    bal climate change is now adding a new level of
    uncertainty to many seafood safety and HAB moni-
    toring programs. Whereas in the past two decades
    unexpected new algal bloom phenomena have often
    been attributed to eutrophication (Smayda 1990) or
    ballast water introduction (Hallegraeff 1993, Lilly
    et al. 2002), increasingly novel algal bloom episodes
    are now circumstantially linked to climate change.

    The Intergovernmental Panel on Climate Change
    (IPCC 2008) is planning to include HAB risk fore-
    casts under a range of climate change scenarios. A

    number of scattered publications have started to
    address the topic of HABs and climate change, but
    they usually have focused on single environmental
    factors (e.g., CO2, temperature increase, stratifica-
    tion), single biological properties (photosynthesis,
    Beardall and Stojkovic 2006; calcification, Rost and
    Riebesell 2004; nutrient uptake, Falkowski and Oli-
    ver 2007), or addressed selected species categories
    of regional interest only (Peperzak 2005, Moore
    et al. 2008b). Complex factor interactions are rarely
    considered in climate simulation scenarios, and eco-
    physiological experiments rarely cover the full range
    of genetic diversity and physiological plasticity of
    microalgal taxa. Prediction of the impact of global
    climate change on algal blooms is fraught with
    uncertainties. It is unfortunate that so few long-term
    records exist of algal blooms at any single locality,
    where ideally we need at least 30 consecutive years.
    However, we can learn important lessons from the
    dinoflagellate cyst fossil record (Dale 2001), from
    the few long-term data sets available, such as the
    Continuous Plankton Recorder surveys (Hays et al.
    2005) and short-term phytoplankton community
    responses to ENSO and NAO episodes. Whereas the
    Continuous Plankton Recorder surveys were initially
    designed to primarily sample zooplankton, these
    instruments also collect phytoplankton down to
    even coccolithophorids (Hays et al. 1995). Started
    in 1931 in the North Atlantic, gradually these sur-
    veys have expanded to the North Pacific (since
    1997) and, more recently, also the Western Atlantic,
    Australia, and the Southern Ocean (Richardson
    et al. 2006). The present review seeks to provide a
    broad overview of the complexity of climate variabil-
    ity and factor interactions, examine marine phyto-
    plankton responses with a focus on the HAB species
    niche, and identify major research gaps.

    The global climate system. The term ‘‘climate’’ is
    used here to include both anthropogenic climate
    change as well as the large-scale decadal oceano-
    graphic patterns such as the ENSO, Pacific Decadal
    Oscillation (PDO), and NAO. This use is in contrast
    to ‘‘weather,’’ which occurs over short timescales of
    days to weeks (cf. Moore et al. 2008b). The earth’s
    climate system comprises the atmosphere (air, water
    vapor, constituent gases, clouds, particles), hydro-
    sphere (oceans, lakes, rivers, groundwater), and cry-
    osphere (continental ice sheets, mountain glaciers,
    sea ice, surface snow cover). The oceans are a core
    component of the global climate system because
    they store 93% (=39,100 gigatonnes, Gt) of the
    world’s carbon, but more and more, we are now
    becoming aware of the quantitative contribution to
    climate by marine phytoplankton, accounting for
    50% of global primary productivity (Longhurst et al.
    1995). All the microalgal cells in the world oceans
    could be packed in a plank, 386,000 km long, 7 cm
    thick, and 30 cm wide, that is, stretching from the
    earth to the moon (Andersen 2005). This increased
    recognition of phytoplankton as a climate driver is

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 221

    well demonstrated by the commercial interests in
    ocean fertilization to combat anthropogenic climate
    change (Glibert et al. 2008). Annually, the oceans
    absorb 1.8 Gt of carbon through photosynthesis and
    2 Gt via abiotic absorption. The oceans thus have
    acted as a sink for 30% of all anthropogenic carbon
    emissions since the onset of the Industrial Revolu-
    tion. The oceans are particularly effective in absorb-
    ing heat and have taken up >90% of the increase in
    heat content of the earth since 1961. Climate
    change in the past has occurred naturally due to
    internal fluctuations in the atmosphere, hydro-
    sphere, and cryosphere, but it has also been influ-
    enced by volcanic eruptions, variations in the sun’s
    output, the earth’s orbital variation, and change in
    the solid earth (e.g., continental drift).

    Climate on our planet has been constantly chang-
    ing, over scales of both millions of years (glacial to
    interglacial periods) and short-term oscillations of
    tens of years (ENSO, NAO). The earth’s climate in
    the distant past has at times been subject to much
    higher ultraviolet-B (UVB) levels and CO2 concen-
    trations than we are seeing at present. The first pho-
    tosynthetic cyanobacteria evolved 3.5 billion years
    ago at CO2 levels 1,000· those of the present, fol-
    lowed by green algae 1,000 million years ago (mya;
    500· present) and dinoflagellates 330–400 mya (8·
    present), whereas more recently evolved diatoms
    and haptophytes operated under comparatively low
    CO2 environments (2–3· present) (Beardall and
    Raven 2004; Fig. 1).

    During the past 800,000 years, atmospheric CO2
    has fluctuated between 180 ppm in glacial and
    300 ppm in interglacial periods, but in the past
    200 years, this has increased from 280 ppm to
    >380 ppm at present, with values of 750–1,000 ppm
    predicted by 2100. In the past 1,000 years, our
    planet has gone through episodes warmer than
    present, such as the medieval warm period AD 550–
    1300, and colder than now, such as the little ice age
    AD 1300–1900. Global temperatures in the past 20–
    30 years (Fig. 1, bottom) have increased significantly
    with a further rise of 2�C–4(6)�C predicted over the
    next 100 years. Undoubtedly, climate change of the
    magnitude that we will be experiencing in the next
    100 years has happened before, albeit in the past
    proceeding at a much slower pace and starting from
    a cooler baseline than present (IPCC 2008). Past
    episodes of climate change over long periods of
    geological and evolutionary history allowed organ-
    isms to adapt to their changing environment.
    Because of their short generation times and longev-
    ity, many phytoplankton are expected to respond to
    current climate change with only a very small time
    lag. They are expected to spread quickly with mov-
    ing water masses into climatic conditions that match
    the temperature, salinity, land runoff, and turbu-
    lence requirements of the species. However, our
    knowledge of the potential of marine microalgae to
    adapt is very limited. Collins and Bell (2004) grew

    the freshwater microscopic alga Chlamydomonas over
    1,000 generations at almost 3· present atmospheric
    CO2 concentration. The cells acclimated to the
    change but did not show any genetic mutations that
    could be described as adaptation.

    Defining the niche of HABs. Most HABs are more
    or less monospecific events, and the autecology of
    the causative organisms thus becomes crucial in
    understanding the factors that trigger these phe-
    nomena. Defining the niche of key HAB species is
    crucial when trying to predict winners or losers
    from climate change. An implicit assumption in
    ecological studies is that there exists a critical

    Fig. 1. Climate change is a matter of scale and time and
    can be viewed in terms of thousands of millions of years [evolu-
    tion of life on our planet (top), after Beardall and Raven 2004],
    hundreds of thousands of years [glacial-interglacial periods, from
    Vostok Ice Core data (middle), Lorius et al. 1990], or the past
    hundred years [from Hadley Centre for Climate Prediction and
    Research, (bottom)]. From a geological perspective, there is
    nothing remarkable about the magnitude of climate change we
    are experiencing now, except that it appears to proceed at a fas-
    ter pace and starts from a warmer baseline.

    222 G U S T A A F M . H A L L E G R A E F F

    relationship between form and function in organ-
    isms, and that life-form therefore is a better predic-
    tor of fitness than phylogenetic affinities. Overall,
    morphotaxonomy has worked well with HAB spe-
    cies, but it is increasingly obvious that ecophysiologi-
    cal experiments based on single culture strains can
    be highly misleading (Burkholder and Glibert
    2006). The development in the past three decades
    of the discipline of HAB ecology is evidenced by the
    increased frequency and size of international meet-
    ings since the first HAB meeting in 1974 and the
    creation in 2002 of the dedicated journal Harmful
    Algae. A first major review of harmful algal ecology
    was produced as a result of a Bermuda NATO-ASI
    workshop in 1996 (Anderson et al. 1998), followed
    by an update by Graneli and Turner (2006). The
    brief summary below (see also Table 1) is largely
    based on these two sources.

    The commonality of the PSP-producing dinofla-
    gellates Alexandrium, Pyrodinium, and Gymnodinium
    catenatum lies in the absence of a rapid growth strat-
    egy and reliance on benthic resting cysts in life-
    cycle transitions (Hallegraeff 1998). Alexandrium
    does not usually produce dense biomass blooms
    that persist throughout the year. Instead, seasonal
    bloom events appear to be restricted in time by cyst
    production (Anderson 1997). The persistence of
    these cysts through long-term unfavorable condi-
    tions allows these dinoflagellates to colonize a wide
    spectrum of habitats and hydrographic regimes.
    The tropical dinoflagellate Pyrodinium prefers high
    salinities (30&–35&) and high temperatures
    (25�C–28�C) (Azanza and Taylor 2001). A soil
    extract requirement in culture may explain this spe-
    cies’ association with rainfall events and land runoff
    from mangrove areas. Benthic cyst stages of
    G. catenatum (short dormancy period of 2 weeks) do
    not play a role in seasonal bloom dynamics, and
    their major function is to sustain this species
    through long periods when water column condi-
    tions are unfavorable for bloom formation (Halle-
    graeff et al. 1995). The success of the haptophyte
    Phaeocystis in marine systems has been attributed to
    its ability to form large gelatinous colonies during
    its life cycle. These colonies occupy the same niche
    in turbulent, tidally or seasonally mixed water col-
    umns as colony-forming spring diatom blooms
    (Smayda and Reynolds 2001). The fish killers Hetero-
    sigma, Chattonella, Prymnesium, Chrysochromulina, and
    Karenia mikimotoi have in common the production
    of high biomass blooms together with the produc-
    tion of allelopathic chemicals (including reactive
    oxygen species) that play a role in predator avoid-
    ance (Hallegraeff 1998). Raphidophyte blooms of
    Heterosigma are sensitive to temperature for cyst ger-
    mination, but chemical conditioning of the water
    by land runoff and other growth promoters (e.g.,
    from aquaculture wastes) determines the outcome
    of competition with diatoms. Similarly, the raphido-
    phyte Chattonella includes a benthic cyst stage in its

    life history, but the growth of the germling cells as
    affected by nutrient conditions and the presence of
    diatom competitors holds the key to its bloom
    development. The capacity of Chattonella to undergo
    vertical migration in stratified water columns with a
    shallow nutricline (i.e., nutrients available only at
    depth under dim light) provides a competitive
    advantage (Imai et al. 1998). While harmful marine
    blooms of Chrysochromulina appear to be exceptional
    events (in Scandinavia in 1988 and 1991), fish-kill-
    ing Prymnesium bloom events in inshore (low salin-
    ity), eutrophic waters are recurrent in many parts of
    the world. The expression of toxicity by Chrysochrom-
    ulina and Prymnesium is variable and can be
    enhanced by phosphate limitation (Graneli and
    Turner 2006). The fish-killing dinoflagellate Karenia
    brevis is a K-strategist, adapted to low nutrient, oligo-
    trophic environments. Blooms in the Gulf of Mex-
    ico are initiated offshore before being transported
    into nearshore waters where they cause fish kills,
    discolored water, human respiratory irritation, and
    occasionally neurotoxic shellfish poisoning (NSP)
    in human shellfish consumers. Taxonomically
    related dinoflagellate species of the eurythermal
    and euryhaline K. mikimotoi species complex are
    associated with marine fauna kills but not human
    intoxications. Poorly characterized lipophilic exo-
    toxins and mucus production play an allelopathic
    role against other algae and also act as agents that
    repress zooplankton grazing (Gentien 1998). This
    species is especially successful in frontal regions and
    in stratified water columns where it accumulates in
    the pycnocline (often also a nutricline), thriving on
    regenerated ammonia and benefitting from poly-
    amine growth factors from decaying diatoms.
    Recent success in culturing the dinoflagellate
    Dinophysis has confirmed its mixotrophic feeding
    behavior on cryptomonad and ⁄ or Mesodinium prey
    (Park et al. 2006) and pointed out that the inci-
    dence of occasionally high biomass is the result of
    active growth and not passive cell accumulation.
    The unusual, large phagotrophic dinoflagellate Noc-
    tiluca depends upon high prey biomass (mostly dia-
    toms) and optimal water temperatures during the
    prebloom stage, with starved cells coming to the
    surface and aggregating at fronts during calm
    weather conditions and wind mixing terminating
    the blooms. Diatom blooms of the cosmopolitan
    genus Pseudo-nitzschia are common in coastal waters
    all over the world. Blooms generally occur during
    colder seasons, and seed populations can derive
    from both inshore or offshore waters (Bates et al.
    1998). The community dynamics of epiphytic ⁄
    benthic tropical Gambierdiscus ciguatera dinoflagel-
    lates and their associated macroalgal canopy are dic-
    tated to a large extent by the degree of water
    movement, with other physical and chemical factors
    such as temperature, salinity, gases, and inorganic
    and organic nutrients only playing a role with
    diminishing hydrodynamics (Bagnis et al. 1985).

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 223

    T
    a

    b
    l
    e

    1
    .

    S
    u

    m
    m

    a
    ry

    o
    f

    p
    h

    ys
    ic

    o
    ch

    e
    m

    ic

    a
    l

    a
    n

    d
    b

    io
    lo

    g
    ic

    a
    l

    n
    ic

    h
    e
    -d

    e
    fi

    n
    in

    g
    fa

    ct
    o

    rs
    fo

    r
    k

    e
    y

    H
    A

    B
    sp

    e
    ci

    e
    s.

    K
    e
    y

    H
    A
    B
    sp
    e
    ci

    e
    s

    a
    n

    d
    th

    e
    ir

    i

    m
    p

    a
    ct

    T
    e
    m

    p

    e
    ra

    tu
    re

    S
    a
    li

    n
    it

    y
    In

    o
    rg

    a
    n

    ic
    n

    u
    tr

    i

    e
    n

    ts
    M

    ic
    ro

    n
    u

    tr
    ie

    n
    ts

    (l
    a
    n

    d
    ru

    n
    o

    ff
    )

    O
    rg

    a
    n
    ic
    n
    u
    tr

    it
    io

    n
    (m

    ix
    o

    t

    r
    o

    p
    h

    y)
    A

    ll
    e
    lo

    p
    a
    th

    y;
    g

    ra
    ze

    r
    a
    vo

    id
    a
    n

    ce
    L

    if
    e

    h
    is

    to
    ry

    S
    e

    n

    si
    ti

    ve
    to

    tu
    rb

    u
    le

    n
    ce

    A
    le

    xa
    n

    d
    ri

    u
    m

    (

    P
    S

    P
    )

    C
    ys

    t
    g

    e
    rm

    i

    n
    a

    ti

    o
    n

    ;
    b

    lo
    o

    m
    w

    i

    n
    d

    o
    w

    F

    e
    ,

    C
    u

    +

    N

    o
    ta

    b
    ly

    ch
    a
    in

    fo
    rm

    e
    rs

    G
    ym

    n
    od

    in
    iu

    m
    ca

    te
    n

    a
    tu

    m
    (P

    S
    P

    )
    B

    lo
    o
    m
    w
    in
    d
    o
    w
    S
    e

    +
    +

    P
    yr

    od
    in

    iu
    m

    ba
    h

    a
    m

    en
    se

    (P
    S
    P
    )

    T
    ro

    p
    ic

    a
    l

    S
    e
    ;

    m
    a
    n

    g
    ro

    ve
    ru

    n
    o

    ff
    +

    +

    C
    h

    a
    tt

    on
    el

    l

    a
    (fi

    sh
    k

    il
    ls

    )
    C

    ys
    t

    g
    e

    rm

    in
    a
    ti
    o
    n

    M
    n

    ,
    F

    e
    ,
    C
    u

    R

    e
    sp

    o
    n

    d
    s

    to
    a
    q

    u
    a
    cu

    lt
    u

    re
    w

    a
    st

    e
    s
    +

    H
    et

    er
    os

    ig
    m

    a
    (fi
    sh
    k
    il
    ls
    )
    C
    ys
    t
    g
    e
    rm
    in
    a
    ti
    o
    n

    F
    re

    s

    h
    w

    a
    te

    r
    o

    n
    to

    p
    o

    f
    w

    a
    te

    r
    co

    lu
    m

    n
    M
    n
    ,
    F
    e
    ,
    C
    u
    R
    e
    sp
    o
    n
    d
    s
    to
    a
    q
    u
    a
    cu
    lt
    u
    re
    w
    a
    st
    e
    s
    +
    +

    C
    ya

    n
    o

    b
    a
    ct

    e
    ri

    a
    B

    lo
    o
    m
    w
    in
    d
    o
    w

    P
    h

    o
    sp

    h
    a
    te

    +
    M

    ic
    ro

    cy
    st

    is
    (m

    ic
    ro
    cy
    st

    in
    s)

    F
    re
    sh
    w
    a
    te

    r
    su

    m
    m

    e
    r

    b
    lo

    o
    m

    s
    P
    h
    o
    sp
    h
    a
    te
    +

    N
    od

    u
    la

    ri
    a

    sp
    u

    m
    ig

    en
    a

    (n
    o

    d
    u

    la
    ri

    n
    )

    B
    ra

    ck
    is

    h
    w
    a
    te

    r
    P

    h
    o

    sp
    h

    a
    te

    T
    ri

    ch
    od

    es
    m

    iu
    m

    (‘
    ‘r

    e
    d

    -t
    id

    e
    ’’

    )
    T

    r

    o
    p

    ic
    a
    l

    F
    e

    fo
    r

    N
    fi

    x
    a
    ti

    o
    n
    +

    K
    a
    re

    n
    ia

    br
    ev

    is
    (N

    S
    P

    ,
    fi

    sh
    k
    il
    ls

    )
    D

    e
    m

    is
    e

    o
    f
    b
    lo
    o
    m

    R
    iv

    e
    r

    fl
    o

    w
    F

    e
    ,
    M
    n

    V
    it

    a
    m

    in
    s,

    a
    m

    in
    o

    a
    ci

    d
    s
    +
    +
    K
    a
    re
    n
    ia

    m
    ik

    im
    ot

    oi
    (fi

    sh
    k
    il
    ls

    )
    P

    o
    ly

    a
    m

    in
    e
    s

    +
    P

    yc
    n

    o
    cl

    in
    e
    s

    fr
    o

    n
    ts

    N
    oc

    ti
    lu

    ca
    (‘

    ‘r
    e
    d

    -t
    id
    e
    ’’
    )
    B
    lo
    o
    m
    w
    in
    d
    o
    w

    In
    d

    ir
    e
    ct

    ;
    h

    ig
    h

    p
    re

    y
    b

    io
    m

    a
    ss

    P
    h

    a
    g

    o
    tr

    o
    p

    h
    y

    P
    se

    u
    d
    o-

    n
    it

    zs
    ch

    ia
    (A

    S
    P

    )
    ?R

    e
    sp
    o
    n
    d
    s

    to
    ra

    in
    fa

    ll
    N

    it
    ra

    te

    P
    fi

    es
    te

    ri
    a

    a
    C

    ry

    p
    to

    p
    er

    id
    in

    io
    p
    si

    sa
    C

    h
    a
    n

    g
    e

    in
    n

    u
    tr

    ie
    n

    t
    re

    g
    im

    e

    C
    ry

    p
    to

    m
    o

    n
    a
    d

    p
    re

    y
    +

    K
    a
    rl

    od
    in
    iu
    m

    (

    fi
    sh

    k
    il

    ls
    )

    C
    ry
    p
    to
    m
    o
    n
    a
    d
    p
    re

    y
    P

    h
    a
    eo

    cy
    st

    is
    (g

    e
    la

    ti
    n

    o
    u

    s
    b

    lo
    o

    m
    s)

    C
    h
    a
    n
    g
    e
    in
    n
    u
    tr
    ie
    n
    t
    re
    g
    im
    e

    F
    e
    ?

    +

    C
    h
    ry

    so
    ch

    ro
    m

    u
    li

    n
    a

    (r
    a
    re

    ly
    fi

    sh
    k
    il
    ls

    )
    S

    e
    ,

    F
    e
    ,

    C
    u

    ,
    V

    it
    B

    1
    2

    +
    +

    P
    ry

    m
    n

    es
    iu

    m
    (fi

    sh
    k
    il
    ls

    )
    +

    D
    in

    op

    h
    ys

    is
    (D

    S
    P
    )
    +
    +
    +

    G
    a
    m

    bi
    er

    d
    is

    cu
    s

    (

    C
    F

    P
    )
    T
    ro
    p
    ic
    a
    l

    M
    a
    cr

    o
    a
    lg

    a
    l

    su
    b

    st
    ra

    te

    T
    a
    x

    a
    id

    e
    n

    ti
    fi

    e
    d

    a
    s

    re
    sp

    o
    n

    si
    ve

    to
    te

    m
    p
    e
    ra
    tu
    re

    ,
    la

    n
    d

    ru
    n

    o
    ff

    ,
    n

    u
    tr
    ie
    n

    ts
    a
    n

    d
    m

    ix
    e
    d

    -l
    a
    ye

    r
    d

    e
    p

    th
    ,

    a
    n

    d
    p

    h
    ys
    ic
    a
    l
    tu
    rb
    u
    le
    n
    ce

    a
    p

    p
    e
    a
    r

    m
    o

    st
    vu

    ln
    e
    ra

    b
    le

    to
    cl

    im
    a
    te

    ch
    a
    n

    g
    e
    .

    ‘‘
    ?’


    in

    d
    i-

    ca
    te

    s
    li

    m
    it

    e
    d

    e
    vi

    d
    e
    n

    ce
    .

    P
    S

    P
    ,

    p
    a
    ra

    ly
    ti

    c
    sh

    e
    ll

    fi
    sh
    p
    o

    is
    o

    n
    in

    g
    ;

    N
    S

    P
    ,

    n
    e
    u

    ro
    to

    x
    ic

    sh
    e
    ll

    fi
    sh
    p
    o
    is
    o
    n
    in
    g
    ;

    A
    S

    P
    ,
    a
    m

    n
    e
    si

    c
    sh
    e
    ll
    fi
    sh
    p
    o
    is
    o
    n
    in
    g
    ;

    D
    S

    P
    ,

    d
    ia

    rr
    h

    e
    ti

    c
    sh
    e
    ll
    fi
    sh
    p
    o
    is
    o
    n
    in
    g
    ;
    C
    F
    P
    ,

    ci
    g

    u
    a
    te

    ra
    fo

    o
    d

    p
    o
    is
    o
    n
    in

    g
    .

    a
    D

    is
    p

    u
    te

    d
    H

    A
    B

    sp
    e
    ci

    e
    s.

    224 G U S T A A F M . H A L L E G R A E F F

    Most HAB species have been demonstrated to
    have either some capability of mixotrophy ⁄ organic
    nutrient uptake or a requirement for micronutrients
    (Graneli and Turner 2006). Temperature plays a
    crucial role in the bloom dynamics of the cyst-form-
    ing PSP dinoflagellates and raphidophytes, as well
    as for species such as G. catenatum, Noctiluca, and
    many cyanobacteria, which have well-defined sea-
    sonal temperature windows. However, once cells of
    these species enter the water column, other factors
    such as nutrients, turbulence, and grazing deter-
    mine the outcome of competition. HAB species
    show a perplexing diversity of biomass and toxicity
    patterns (Cembella 2003), ranging from species
    such as Dinophysis and Chrysochromulina, which can
    cause toxicity problems even at very low cell concen-
    trations, to species such as Phaeocystis and Noctiluca,
    which are basically nontoxic but whose nuisance
    value derives from their high biomass production.
    Persistent near-monospecific algal blooms of, for
    example, Aureococcus, Chrysochromulina, Prymnesium,
    and Nodularia have recently been referred to as eco-
    system disruptive algal blooms (EDABs), in which
    toxic or unpalatable algal species disrupt grazing
    and thus diminish nutrient supply via recycling
    (Sunda et al. 2006). Could climate perturbations
    perhaps create a niche for such HAB species?

    From progress in the past three decades, it has
    become abundantly clear that the niche of HAB spe-
    cies is much wider than originally envisaged. HAB
    species are not restricted to dinoflagellates but also
    include diatoms, haptophytes, raphidophytes, and
    cyanobacteria. Furthermore, they cover the com-
    plete range from r-strategists (e.g., Pseudo-nitzschia,
    Chattonella), whose success is due to their high
    growth rates (r) and efficient use of nutrients, to
    K-strategists (e.g., G. catenatum), which can achieve
    high biomass levels by being energy (light) efficient,
    for example, by vertical migration (Margalef 1978,
    Smayda and Reynolds 2001). Taxa identified in
    Table 1 as responsive to temperature, land runoff,
    nutrients and mixed-layer depth, and physical turbu-
    lence appear most vulnerable to climate change.
    When and where climate-driven perturbations open
    a new ‘‘niche,’’ any number of ecologically similar
    organisms have the opportunity to emerge from the
    background to become a HAB phenomenon.

    Algal bloom range expansions and climate change. For
    many HAB species, significant bloom episodes can
    serve as stepping stones toward range expansions via
    natural current systems, sometimes facilitated by
    local climate events or ship ballast water dispersal.
    The dinoflagellate Pyrodinium bahamense is presently
    confined to tropical, mangrove-fringed coastal
    waters of the Atlantic and Indo-West Pacific. A sur-
    vey of cyst fossils (named Polysphaeridium zoharyii)
    going back to the warmer Eocene 50 mya indicates
    a much wider range of distribution in the past. For
    example, in the Australasian region at present, the
    alga is not found farther south than Papua New

    Guinea, but some 120,000 years ago, the alga ran-
    ged as far south as Bulahdelah (32�S) just north of
    Sydney (McMinn 1988, 1989). There is concern
    that, with increased warming of the oceans, this spe-
    cies may return to Australian waters (Fig. 2). In the
    tropical Atlantic, in areas such as Bahia Fosforescen-
    te in Puerto Rico and Oyster Bay in Jamaica, the
    bioluminescent blooms of Pyrodinium are a major
    tourist attraction, but Pyrodinium blooms gained a
    more sinister reputation in 1972 in Papua New Gui-
    nea after red-brown water discolorations coincided
    with the fatal food poisoning of three children in a
    seaside village, diagnosed as PSP. Since then, the
    incidence of toxic blooms has spread to Brunei and
    Sabah (1976), the central (1983) and northern Phil-
    ippines (1987), and Indonesia (North Mollucas).
    Pyrodinium is a serious public health and economic
    problem for tropical countries, all of which depend
    heavily on seafood for protein. In the Philippines
    alone, Pyrodinium has now been responsible for
    >2,000 human illnesses and 100 deaths resulting
    from the consumption of contaminated shellfish as
    well as sardines and anchovies (Hallegraeff and
    Maclean 1989). There exists circumstantial but
    debated evidence of a coincidence between Pyrodi-
    nium blooms and the ENSO (Maclean 1989, Azanza
    and Taylor 2001). In the Pacific Basin, trade winds
    and strong equatorial currents normally flow
    westward, and cold upwelling occurs off Peru. In
    contrast, during an ENSO event, trade winds are
    weak, and anomalously warm equatorial water flows
    eastward, and stratification is enhanced. Erickson

    Fig. 2. Global distribution of Pyrodinium bahamense in recent
    plankton (A) and much wider distribution in the fossil cyst
    record (B) (after Hallegraeff 1993).

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 225

    and Nishitani (1985) similarly reported exceptional
    PSP episodes by Alexandrium tamarense ⁄ catenella in
    the Pacific Northwest during seven out of nine
    ENSO events between 1941 and 1984 (but see
    Moore et al. 2008a for an alternative interpreta-
    tion). An exceptional Karenia digitata red-tide event
    in Hong Kong in 1998 (HK$250 M loss to aquacul-
    ture) was associated with El Niño and altered ocean-
    ographic conditions (Yin et al. 1999). In the North
    Atlantic, the NAO reflects a north–south oscillation
    in atmosphere mass between the Iceland-low and
    the Azores high-pressure center. A high NAO means
    increased westerly winds and milder temperatures
    over northern Europe, and a low NOA causes cooler
    temperatures due to decreased westerly winds. Bel-
    grano et al. (1999) found significant correlations
    between NAO, phytoplankton biomass, primary pro-
    duction, and Dinophysis concentrations off Sweden.

    Until recently, NSP by the dinoflagellate K. brevis
    was considered to be endemic to the Gulf of Mexico
    and the east coast of Florida, where red tides had
    been reported as early as 1844. An unusual feature
    of NSP is the formation of toxic aerosols by wave
    action, which can lead to respiratory asthma-like
    symptoms in humans. In 1987, a major Florida
    bloom was dispersed by the Gulf Stream northward
    into North Carolina waters, even though it has not
    persisted there (Tester et al. 1991, 1993). Unexpect-
    edly, in early 1993, >180 human NSPs were reported
    from New Zealand. Most likely, this mixed bloom of
    K. mikimotoi and related species was again triggered
    by the unusual weather conditions at the time,
    including higher than usual rainfall and lower than
    usual temperature, which coincided with El Niño
    (Rhodes et al. 1993, Chang et al. 1998).

    Ciguatera caused by the benthic dinoflagellate
    Gambierdiscus toxicus is a food-poisoning syndrome
    caused by ingesting tropical fish and is well known
    in coral reef areas in the Caribbean, Australia, and
    especially French Polynesia (Fig. 3). Whereas, in a
    strict sense, this is a completely natural phenome-
    non, from being a rare disease two centuries ago,

    ciguatera has now reached epidemic proportions in
    French Polynesia. From 1960 to 1984, >24,000
    patients were reported from this area, which is more
    than six times the average for the Pacific as a whole
    (Bagnis et al. 1985). Evidence is accumulating that
    reef disturbance by hurricanes, military and tourist
    developments, as well as coral bleaching (linked to
    global warming), increased water temperatures
    (>29�C preferred in culture), and perhaps in future
    increasing coral damage due to ocean acidification
    (Hoegh-Guldberg 1999) are increasing the risk of
    ciguatera by freeing up space for macroalgae for
    Gambierdiscus to colonize upon. During El Niño
    events, ciguatera increased on Pacific islands where
    sea surface temperatures increased (Hales et al.
    2001). In the Australian region, G. toxicus is well
    known from the tropical Great Barrier Reef and
    southward to just north of Brisbane (25�S), but in
    the past 5 years, this species has undergone an
    apparent range expansion into southeast Australian
    seagrass beds as far south as Merimbula (37�S),
    aided by a strengthening of the East Australian
    Current (S. Brett, M. de Salas, and G. Hallegraeff,
    unpublished data). A similar expansion of Gambier-
    discus into the Mediterranean and eastern Atlantic
    has been reported (Aligizaki et al. 2008), and
    blooms of the associated benthic dinoflagellate
    genus Ostreopsis are also an increasingly common
    phenomenon in temperate regions worldwide
    (Shears and Ross 2009).

    In the same Australian region, the red-tide dino-
    flagellate Noctiluca scintillans (known from Sydney
    as early as 1860) has expanded its range from Syd-
    ney into Southern Tasmanian waters since 1994
    where it has caused problems for the salmonid fish
    farm industry (Fig. 4). In the North Sea, an analo-
    gous northward shift of warm-water phytoplankton

    Fig. 3. Current global distribution of ciguatera food poison-
    ing from fish (after Hallegraeff 1993).

    Fig. 4. Apparent range expansion of Noctiluca scintillans in the
    Australian region, comparing distribution records in 1860–1950,
    1980–1993 (expansion of blooms in the Sydney region),
    1994–2005 (range extension into Tasmania), and 2008 (first
    reports in Queensland, West Australia, and South Australia).
    After Hallegraeff et al. (2008).

    226 G U S T A A F M . H A L L E G R A E F F

    has occurred due to regional climate warming
    (Edwards and Richardson 2004, Richardson and
    Schoeman 2004). For example, Ceratium trichoceros,
    previously found only south of the British Isles, has
    expanded its geographic range to the west coast of
    Scotland and the North Sea, and the subtropical
    Ceratium hexacanthum moved 1,000 km northward in
    40 years (Hays et al. 2005). At the same time, Proro-
    centrum, Ceratium furca, and Dinophysis increased
    along the Norwegian coast, and Noctiluca increased
    in the southern North Sea (Fig. 5). It is difficult to
    untangle the role of climate change and eutrophi-
    cation in some of these species patterns. Dale
    (2009) used the dinoflagellate cyst record from the
    last 100 years to discriminate between the role of
    local eutrophication events within the Skagerrak
    (indicated by a shift to heterotrophic dinocysts,
    reflecting increased diatom prey) and the role of
    regional variation in the NAO, thought to have
    increased transport of relatively nutrient-rich North
    Sea water into the system (indicated by increased
    Lingulodinium benefitting from added nutrients dur-
    ing warm summers and increased stratification). In
    the same Skagerrak area, Thorsen et al. (1995),
    working with a much longer sediment core, docu-
    mented the immigration into the region �6000
    B.P. (before present) of Gymnodinium nolleri (ini-
    tially reported as G. catenatum), which achieved
    bloom proportions from 2000 to 500 B.P. during a
    warming period, followed by a near extinction dur-
    ing the cooling period that commenced in 300 B.P.
    (Fig. 6). Reduction between 1997 and 2002 of Arc-
    tic ice by 33% allowed greater flow through the
    trans-Arctic pathway from the Pacific to the Atlantic
    and was associated with the first appearance in
    800,000 years in the Labrador Sea of the diatom
    Neodenticula seminae (Reid et al. 2007). While warm-
    water species can be expected to expand their dis-
    tribution, cold-water species will contract their
    range (compare Beaugrand et al. 2002 for zoo-
    plankton). For example, the cold-water dinoflagel-
    late cyst Bitectatodinium tepikiense currently is
    confined to Tasmania (43�S), but in the last inter-
    glacial period (120,000 years ago), it was found as
    far north as Sydney (34�S) (McMinn and Sun

    1994). On top of range extensions driven by grad-
    ual climate change, ship ballast water translocations
    continue to alter species distributions. The two
    mechanisms interact since ecosystems disturbed by
    pollution or climate change are more prone to
    ballast water invasions (Stachowicz et al. 2002).
    Similarly, melting of Arctic sea ice and opening of
    new Arctic shipping lanes will encourage range
    expansions via both natural current dispersal as
    well as ballast water invasions.

    Fig. 5. Decadal anomaly maps (difference between long-term 1960–1989 mean and the 1990–2002 period) for four common HAB spe-
    cies (from left to right): Prorocentrum, Ceratium furca, Dinophysis, and Noctiluca in the North Atlantic. Note the increase in Prorocentrum, C.
    furca, and Dinophysis along the Norwegian coast, and increase in Noctiluca in the southern North Sea, reportedly associated with a contraction
    of the Subpolar Gyre to the west allowing subtropical water to penetrate farther north (adapted from Edwards et al. 2008, with permission).

    Fig. 6. Quantitative distribution of Gymnodinium nolleri cysts
    and total dinoflagellate cysts (cysts Æ g)1 dry sediment) in an
    860 cm long sediment core from the southern Kattegat (after
    Thorsen et al. 1995).

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 227

    Impact of global warming and sea surface temperature
    change. Phytoplankton grow over a range of tem-
    peratures characteristic of their habitat, and growth
    rates are usually higher at higher temperature, but
    considerably lower beyond an optimal temperature
    (Eppley 1972). Natural populations of phytoplank-
    ton often occur at temperatures suboptimal for pho-
    tosynthesis, and it is believed that this distribution is
    designed to avoid risking abrupt declines in growth
    associated with the abrupt incidence of warmer tem-
    peratures (Li 1980). Temperature effects on phyto-
    plankton growth and composition are more
    important in shallow coastal waters, which experi-
    ence larger temperature fluctuations than oceanic
    waters. Predicted increasing sea surface tempera-
    tures of 2�C–4�C may shift the community composi-
    tion toward species adapted to warmer temperatures
    as observed in the temperate North Atlantic
    (Edwards and Richardson 2004). Several well-stud-
    ied PSP dinoflagellates, such as A. catenella in Puget
    Sound (Moore et al. 2008b) and G. catenatum in
    Tasmania, Australia (Hallegraeff et al. 1995), bloom
    in well-defined seasonal temperature windows
    (>13�C and >10�C, respectively). Climate change
    scenarios are predicted to generate longer-lasting
    bloom windows (Fig. 7).

    In the North Sea, the NAO has been shown to
    affect the length of the phytoplankton growing sea-
    son, which has increased in parallel with the warm-
    ing of sea surface temperatures (Barton et al. 2003).
    Seasonal timing of phytoplankton blooms is now
    occurring there up to 4–6 weeks earlier (Fig. 8).
    However, where individual zooplankton or fish graz-
    ers are differentially impacted by ocean warming,
    this may have cascading impacts on the structure of
    marine food webs (‘‘match-mismatch’’ sensu Cush-

    ing 1974). Replacement in the North Sea of the
    cold-water copepod Calanus finmarchicus by the
    warm-water Calanus helgolandicus has been circum-
    stantially associated with the decline of cod
    (Edwards et al. 2008). Similarly, Attrill et al. (2007)
    predict an increasing occurrence of jellyfish in the
    central North Sea over the next 100 years related to
    the increased Atlantic inflow to the northern North
    Sea.

    Sea-level rise, wind, and mixed-layer depth. Increasing
    sea surface temperature and water column stratifica-
    tion (shallowing of the mixed layer) can be
    expected to have a strong impact on phytoplankton
    because of the resource requirements and tempera-
    ture ranges that species are adapted to. Wind deter-
    mines the incidence of upwelling and downwelling,
    which in turn strongly affect the supply of macronu-
    trients to the surface (recognized as drivers of
    G. catenatum blooms off Spain; Fraga and Bakun
    1990). Climate change may thus affect the timing
    and strength of coastal upwellings. Broad changes
    in ocean circulation such as those comprising the
    deep-ocean conveyor belt (Rahmstorf 2002) can also
    cause displacements to current systems and associ-
    ated algal bloom phenomena. Wind-driven currents
    can transport phytoplankton away from a region
    and affect the size and frequency of formation of
    mesoscale features such as fronts and eddies.
    Locally, wind intensity strongly influences depth
    and intensity of vertical mixing in the surface layer,
    thereby affecting phytoplankton access to nutrients,
    light availability for algal photosynthesis, and phyto-
    plankton exposure to potentially harmful UVB radi-
    ation. Winds can also influence the supply of iron
    to the surface ocean through aeolian transport of
    dust from land to sea, contributing micronutrients
    such as iron, which has been shown to stimulate
    K. brevis blooms off Florida (Walsh and Steidinger
    2001). Extreme climate events such as hurricanes

    Fig. 8. Long-term monthly values of ‘‘phytoplankton color’’
    in the central North Sea from 1948 to 2001. Circles denote >2 SD
    above the long-term monthly mean and a major regime shift
    �1998. Note an apparent shift toward earlier spring and autumn
    phytoplankton blooms (after Edwards 2004, with permission).

    Fig. 7. Scenarios for warmer sea surface temperature condi-
    tions in Puget Sound by 2, 4, and 6�C would widen the >13�C
    window (in gray) of accelerated growth for the PSP dinoflagellate
    Alexandrium catenella. After Moore et al. (2008b). PSP, paralytic
    shellfish poisoning.

    228 G U S T A A F M . H A L L E G R A E F F

    are also known to expand the existing distribution
    of cyst-producing toxic dinoflagellates (e.g., A. tama-
    rense in New England after a 1972 hurricane; Ander-
    son 1997). Sea-level rises of 18 to 59 cm (up to
    1 m) predicted by 2100 (IPCC 2008) have the
    potential to increase the extent of continental shelf
    areas, providing shallow, stable water columns favor-
    ing phytoplankton growth. The proliferation of coc-
    colithophorids in the Cretaceous geological period
    has been attributed to expanding continental shelf
    areas (Bown et al. 2004).

    Finally, increasing temperature driven by climate
    change is predicted to lead to enhanced surface
    stratification, more rapid depletion of surface nutri-
    ents, and a decrease in replenishment from deep
    nutrient-rich waters (Fig. 9A). This in turn will lead
    to changes in phytoplankton species, with smaller
    nano- and picoplankton cells with higher surface
    area:volume ratios (better able to cope with low
    nutrient levels) favored over larger cells. A decline
    in silica concentrations is widely expected in a
    warming world (reported, e.g., from the Mediterra-
    nean; Goffart et al. 2002), and this is anticipated to
    restrict the abundance of diatoms. Mixing depth
    affects sea surface temperature, the supply of light
    (from above) and nutrients (from below), and phy-
    toplankton sinking losses within the surface layer.
    Climate models predict changes in mixed-layer
    depth in response to global warming for large
    regions of the global ocean. In the North Pacific,
    decadal-scale climate and mixed-layer variability,
    (Hayward 1997) and, in the North Atlantic, longer-
    term changes in wind intensity and stratification
    since the 1950s have been associated with consider-
    able changes in phytoplankton community structure
    (Richardson and Schoeman 2004). In regions with
    intermediate mixing depth, increased stratification

    is expected to result in decreased phytoplankton
    biomass due to reductions in nutrient supply. The
    observed reductions in open-ocean productivity
    (‘‘desertification’’) during ENSO warming events
    provide insight on how future climate change can
    alter marine food webs (Behrenfeld et al. 2006).
    Conversely, in high-latitude regions with relatively
    deep mixing and nonlimiting nutrients, decreasing
    mixing depth is expected to result in higher phyto-
    plankton biomass because of increased light avail-
    ability (Fig. 9B; Doney 2006).

    Impact of heavy precipitation and storm events and flash
    floods. Episodic storm events affect the timing of
    freshwater flow, residence time, and magnitude and
    time of nutrient pulses. Changes in the amount or
    timing of rainfall and river runoff affect the salinity
    of estuaries and coastal waters. Salinity is relatively
    constant throughout the year in most oceanic waters
    and in coastal areas that receive little freshwater
    input. Coastal phytoplankton is subject to more vari-
    ation in salinity than phytoplankton in oceanic
    waters. While some species grow well over a wide
    range of salinities, other species grow best only at
    salinities that are low (estuarine), intermediate
    (coastal), or high (oceanic species). Freshwater also
    modifies the stratification of the water column,
    thereby affecting nutrient resupply from below.
    While diatoms seem to be negatively affected by the
    inhibition of mixing associated with river discharge,
    dinoflagellates often benefit as this usually increases
    stratification and the availability of humic substances
    for growth (Graneli and Moreira 1990, Doblin et al.
    2005). PSP dinoflagellate blooms of G. catenatum (in
    Tasmania; Hallegraeff et al. 1995) and A. tamarense
    (off Massachusetts; Anderson 1997) tend to be clo-
    sely associated with land runoff events. In Hiro-
    shima Bay, blooms of the fish-killing raphidophyte

    Fig. 9. Predicted phytoplank-
    ton response to increased tem-
    perature in ocean surface waters:
    (A) reduced productivity in the
    thermally stratified water of tropi-
    cal and midlatitudes caused by
    reduced nutrient supply; (B)
    increased productivity at polar
    and subpolar salt-stratified oceans
    where reduced mixing keeps
    plankton closer to the well-lit
    nutrient-sufficient surface layers.
    Adapted from Doney (2006).

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 229

    Chattonella marina followed typhoon-induced accre-
    tion of nutrient-rich land runoff (Kimura et al.
    1973). Climate change is predicted to cause rainfall
    to occur in more concentrated bursts followed by
    long dry periods, thus favoring dinoflagellates.

    Increased CO2 and ocean acidification. Increasing
    atmospheric CO2 is leading to ocean acidification,
    which could potentially have an adverse impact on
    calcifying organisms, the most important of which
    in terms of biomass and carbon sequestration is the
    coccolithophorid Emiliania huxleyi (Riebesell et al.
    2000). Calculations based on CO2 measurements of
    the surface oceans indicate that uptake by the
    oceans of approximately half the CO2 produced by
    burning fossil fuels has already led to a reduction of
    surface pH by 0.1 unit. Under the current scenario
    of continuing global CO2 emissions from human
    activities, average ocean pH is predicted to fall by
    0.4 units by the year 2100 (Orr et al. 2005). Such
    pH is lower than has been experienced for millen-
    nia, and, critically, this rate of change is 100 times
    faster than ever experienced in the known history
    of our planet (Raven et al. 2005b). Experimental
    manipulations of pH in E. huxleyi cultures have both
    produced reduced (Riebesell et al. 2000) and
    enhanced calcification and growth (Iglesias-
    Rodriguez et al. 2008). Strikingly, depression of
    calcification under high CO2 is most expressed
    under high-light conditions (Feng et al. 2008;
    Fig. 10), emphasizing the need to carefully consider
    factor interactions in ecophysiological experiments.
    Decreasing pH < 8.0 has been observed to nega- tively affect nitrification in marine bacteria and therefore could potentially reduce nitrate availability for plankton algae. The nitrogen-fixing tropical cya- nobacterium Trichodesmium may be a beneficiary of ocean acidification, however (Hutchins et al. 2007). Decreasing pH has also been found to increase the availability of toxic trace elements such as copper. As the relative consumption of HCO3

    ) and CO2
    differs between phytoplankton species, changes in
    their availability may affect phytoplankton at the cel-
    lular, population, and community levels. Most HAB
    species tested thus far lack carbon concentrating
    mechanisms (CCMs), and hence their photosyn-
    thetic performance may benefit from increased
    atmospheric CO2 (e.g., E. huxleyi in Fig. 11). How-
    ever, photosynthetic performance in diatom species
    such as Skeletonema, for which photosynthesis is
    already CO2 saturated, will remain constant (Beardall
    and Raven 2004). In bioassays in the equatorial
    Pacific, high CO2 (750 ppm) favored the hapto-
    phyte Phaeocystis at the expense of diatoms, whereas
    at low CO2 (150 ppm), diatom growth was stimu-
    lated (Tortell et al. 2002). Riebesell et al. (2007),
    working with mesocosms dominated by diatoms and
    coccolithophorids, observed increases in productiv-
    ity of 27% and 39% when CO2 levels were elevated
    to 700 and 1,050 ppm, respectively. Similarly, Schip-
    pers et al. (2004) predict in nutrient-replete systems

    a 10%–40% increase of marine productivity for spe-
    cies with low bicarbonate affinity, thus potentially
    aggravating some coastal algal blooms.

    UV radiation. Although the implementation of
    the Montreal Protocol has done much to slow the
    build-up of chlorofluorocarbons in the stratosphere,
    elevated UVB levels from the Arctic and Antarctic

    Fig. 10. Calcification of the coccolithophorid Emiliania huxleyi
    (expressed as the particulate inorganic carbon [PIC] to particu-
    late organic carbon ratio [POC]) as a function pCO2 (ambient
    375 ppm or high 750 ppm), temperature (ambient 20�C or high
    24�C), and light (low = black bars; high = open bars). Neither
    pCO2, light, nor temperature influences PIC:POC under low
    light, but calcification is most reduced under a combination of
    high light · high pCO2 (after Feng et al. 2008).

    Fig. 11. Photosynthesis of three different phytoplankton spe-
    cies (diatom Skeletonema costatum, haptophytes Phaeocystis globosa
    and Emiliania huxleyi) with respect to CO2 sensitivity (adapted
    from Rost and Riebesell 2004). Microalgal species differ in their
    responses to CO2, which implies that a high-CO2 ocean will
    induce shifts in phytoplankton species composition.

    230 G U S T A A F M . H A L L E G R A E F F

    ozone holes are expected to persist until at least
    2050. UVB can negatively affect several physiological
    processes and cellular structures of phytoplankton,
    including photosynthesis, nutrient uptake, cell
    motility and orientation, algal life span, and DNA
    (Häder et al. 1991). Whereas shorter wavelengths
    generally cause greater damage per dose, inhibition
    of photosynthesis by ambient UVB increases linearly
    with increasing total dose. In clear oceanic waters,
    UVB radiation can reach depths of at least 30 m.
    Although some phytoplankton may acclimate to,
    compensate for, or repair damage by UVB, this
    involves metabolic costs, thereby reducing energy
    available for cell growth and division. Raven et al.
    (2005a) suggest that UVB intensity affects the size
    ratio in phytoplankton communities because small
    cells are more prone to UVB and have compara-
    tively high metabolic costs to screen out damaging
    UVB. Many surface-dwelling red-tide species of raph-
    idophytes and dinoflagellates possess UVB-screening
    pigments, which give them a competitive advantage
    over species lacking such UV protection (Jeffrey
    et al. 1999). In some species, nutrient limitation of
    either N or P (from increased water column stratifi-
    cation) can enhance the sensitivity of cells to UVB
    damage (Shelly et al. 2002).

    Feedback mechanisms. One cannot talk about the
    impact of climate on phytoplankton without also
    considering the impact of phytoplankton on
    climate (Fig. 12). Phytoplankton play a key role in
    several global biogeochemical cycles and thereby
    exert important feedback effects on climate by
    influencing the partitioning of climate-relevant
    gases between the ocean and the atmosphere.
    Some species (e.g., Emiliania, Phaeocystis) are pro-
    ducers of dimethylsulfonium propionate, a precur-
    sor of dimethylsulfoxide (DMS), which in the
    atmosphere is oxidized into sulfate, which forms
    condensation nuclei for clouds (Charleson et al.
    1987). Subsequent work on DMS has clarified that
    it is not just phytoplankton, however, but also zoo-
    plankton and bacterial food-web structure and
    dynamics that drive oceanic production of atmo-
    spheric sulfur. Phytoplankton, therefore, indirectly
    affect albedo and precipitation and hence coastal
    runoff, salinity, water column stratification, and
    nutrient supply.

    Through the process of photosynthesis, phyto-
    plankton constitute a major consumer of CO2.
    The ability of the oceans to act as a sink for
    anthropogenic CO2 largely relies on the conver-
    sion of this gas by phytoplankton into particulate

    Fig. 12. Summary diagram of known feedback mechanisms between physicochemical climate variables and biological properties of
    marine phytoplankton systems. Red: Greenhouse warming raises surface temperatures and causes a shoaling of mixed-layer depths but can
    also have broader impacts on global currents, upwelling, and even the deep-ocean conveyor belt. Blue: Increased atmospheric CO2 drives
    the biological pump, can alter phytoplankton species composition, and can alter ocean pH, influencing calcification of coccolithophorids
    but also nutrient availability. Green: Nutrient impacts from water column stratification, as well as linked to shifts in marine food-web struc-
    ture, influenced by fishing, eutrophication, and even ship ballast water invasions. Yellow: Marine food-web structure, including top-down
    as well as bottom-up influences on phytoplankton species composition. Orange: Selected phytoplankton such as coccolithophorids pro-
    duce dimethylsulfoxide (DMS), acting as cloud condensation nuclei, thereby reducing solar irradiation. Other anthropogenic influences
    in terms of eutrophication, shipping (ballast water introductions), and fishing are also indicated. Without exception, all perturbations will
    drive changes in phytoplankton species composition.

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 231

    organic matter and subsequent partial loss to the
    deep ocean (so-called biological pump). Any
    reduction in net ocean CO2 uptake caused by
    shifts in ocean circulation or reduced phytoplank-
    ton growth in surface waters reducing the export
    of organic matter to the deep sea via the biologi-
    cal pump could lead to an acceleration in the
    rate of atmospheric CO2 increase and global
    warming. Models have estimated that a 50%
    decrease in oceanic calcification from ocean acidi-
    fication thus would reduce atmospheric CO2 by
    10–40 ppm (Heinze 2004, Munhoven 2007), equiv-
    alent to 5–20 years of industrial emissions. Con-
    versely, an increase in calcification would increase
    CO2 levels by a similar amount (carbonate count-
    erpump). Coupled climate-carbon models are
    increasingly revealing feedback mechanisms, which
    were completely unpredicted from first principles.
    Stratosphere ozone depletion increases the
    strength of Southern Ocean winds and thereby
    the ventilation of carbon-rich deep water, with
    consequences of reduced ocean carbon uptake
    and enhanced ocean acidification (Lenton et al.
    2009).

    Woods and Barkmann’s (1993) ‘‘plankton
    multiplier’’ is an example of a positive feedback
    mechanism linking greenhouse warming to the
    biological pump. Enhanced greenhouse CO2
    induces ocean surface warming, diminishing winter
    convection and nutrient availability and thereby
    primary production, thus weakening the biological
    pump and further enhancing atmospheric CO2.
    These authors suggested that a similar mechanism
    may have underpinned global warming at the end
    of the ice ages when the Milankovich effect
    enhanced greenhouse warming. Surface phyto-
    plankton blooms influence the oceanic heat bud-
    get, and this is dependent not only on the chl
    biomass but also on its precise vertical distribution
    (Frouin and Lacobellis 2002).

    Finally, another powerful mechanism for algal
    bloom formation occurs through ‘‘top-down con-
    trol’’ of the marine food web (Turner and Grane-
    li 2006). Overfishing removes top fish predators,
    stimulating small fish stocks, which graze away
    zooplankton, thus relieving phytoplankton grazing
    pressure. Differential impacts of climate change
    on individual zooplankton or fish grazers (uncou-
    pling between trophic levels) thus can result in
    stimulation of HABs. Figure 12 summarizes known
    feedback mechanisms between physicochemical cli-
    mate variables and biological properties of marine
    phytoplankton systems, altogether confronting us
    with a formidable predictive challenge. Without
    exception, all physicochemical climate stressors
    drive changes in phytoplankton species composi-
    tion, but the precise direction of such changes
    (i.e., whether they may lead to HABs) remains lar-
    gely unpredictable in view of our current incom-
    plete knowledge of phytoplankton ecophysiology.

    CONCLUSIONS

    Climate change confronts marine ecosystems with
    multifactorial stressors, such as increased tempera-
    ture, enhanced surface stratification, alteration of
    ocean currents, intensification or weakening of
    nutrient upwelling, stimulation of photosynthesis by
    elevated CO2, reduced calcification from ocean acid-
    ification, and changes in land runoff and micronu-
    trient availability. Complex factor interactions are
    rarely covered by simulated ecophysiological experi-
    ments, and the full genetic diversity and physiologi-
    cal plasticity of phytoplankton taxa are rarely
    considered. Traditional experimental challenges last
    days to weeks and impose new growth conditions
    rather quickly, thus only allowing for limited accli-
    mation (testing short-term physiological plasticity
    but without genetic changes). Predicted global
    change will occur gradually over decades, allowing
    for adaptation of species to perhaps become geneti-
    cally and phenotypically different from the present
    population. Laboratory studies should aim to mimic
    environmental conditions as closely as possible
    (Rost et al. 2008). A typical example is the problem
    of the potential impact of increased CO2 on the
    coccolithophorid E. huxleyi. Initial concerns focused
    on reduced calcification (Riebesell et al. 2000), but
    we now recognize that increased CO2 at the same
    time stimulates photosynthesis (Iglesias-Rodriguez
    et al. 2008). Complex factor interactions between
    increased CO2, light, and temperature on the calci-
    fication versus photosynthesis dynamics of E. huxleyi
    have been demonstrated by Feng et al. (2008), while
    geographic strain variability of this ‘‘cosmopolitan’’
    taxon has confounded the extensive literature on
    this taxon (Langer et al. 2009). At the same time,
    field observations of E. huxleyi are suggesting an
    apparent range expansion in the past two decades
    toward both the Arctic (Bering Sea; Merico et al.
    2003) and Antarctic (Cubillos et al. 2007), but the
    environmental drivers underpinning this are by no
    means clear. Undoubtedly, there will be winners
    and losers from climate change, and one thing we
    can be certain about is local changes in species
    composition, abundance, and timing of algal
    blooms.

    The greatest problems for human society will be
    caused by being unprepared for significant range
    extensions of HAB species or the increase of algal
    biotoxin problems in currently poorly monitored
    areas. While, for example, ciguatera contamination
    would be expected and monitored for in tropical
    coral reef fish, with the apparent range extension of
    the causative benthic dinoflagellate into warm-tem-
    perate seagrass beds of Southern Australia, other
    coastal fisheries unexpectedly could be at risk.
    Range expansion of Noctiluca from Sydney to Tasma-
    nia exposed the salmonid aquaculture industry to a
    novel HAB problem. Polar expansion of domoic-
    acid-producing Pseudo-nitzschia australis could pose a

    232 G U S T A A F M . H A L L E G R A E F F

    novel threat to krill-feeding whales (Lefebvre et al.
    2002). Similarly, incidences of increased surface
    stratification in estuaries or heavy precipitation or
    extreme storm events are all warning signs that call
    for increased vigilance of monitoring seafood prod-
    ucts for algal biotoxins even in areas not currently
    considered to be at risk. Changes in phytoplankton
    communities provide a sensitive early warning for
    climate-driven perturbations to marine ecosystems.

    Only with improved global ocean observation sys-
    tems (GOOS) can we hope to quantitatively monitor
    the key variables identified in this review. New,
    improved, and expanded ocean sensor capabilities
    (e.g., argo floats, ocean gliders, coastal moorings
    and coastal radar, multiwavelength and variable flu-
    orometers, optical sensors) are necessary to realize
    the full potential of in situ ocean-observing net-
    works in support of integrated satellite-derived
    ‘‘ocean color’’ maps and expanded biological and
    biogeochemical observations (continuous plankton
    recorder, ecogenomics). Observations from the indi-
    vidual components of such systems must be inte-
    grated through data management and
    communication capabilities that provide open
    searchable access and routine delivery to all users.
    Sustained observations, process research, and mod-
    eling should determine fluxes and cycling of bio-
    geochemical variables, identify impacts on
    ecosystems, and resolve feedback from ecosystems
    on climate. Achieving this will require extensive
    infrastructure investment and poses a major chal-
    lenge for the marine science community. It is pleas-
    ing to see that a number of national (e.g., the U.S.
    NSTC Joint Subcommittee on Ocean Science and
    Technology 2007 Ocean Observatories Initiative
    [OOI], the Australian Integrated Marine Observing
    System [IMOS 2009]) and international programs
    (e.g., the Intergovernmental Oceanographic Com-
    mission of UNESCO’s GEOHAB) are actively pursu-
    ing these ambitious goals, but necessary if we wish
    to define management options, forecast ocean-
    related risks to human health and safety, and shed
    light on the impact of climate variability on marine
    life and humans in general.

    An earlier version appeared as a section of FAO Assessment
    and Management of Fish Safety and Quality Technical Paper, and I
    am grateful to Dr. Iddya Karunasagar for inviting me to con-
    tribute to that effort. I thank my Tasmanian colleagues Prof.
    Andrew McMinn, Prof. Harvey Marchant, and Prof. Tom
    Trull for valuable discussions on the ocean carbon pump;
    Prof. Chris Reid and Dr. Anthony Richardson (Sir Alister
    Hardy Foundation for Ocean Science, Plymouth) for insights
    into the Continuous Plankton Recorder program; Prof. Barrie
    Dale for discussions on the fossil dinoflagellate cyst record;
    and Dr. Stephanie Moore (University of Washington) for clar-
    ifying the intricacies of ENSO, PDO, and NAO.

    Aligizaki, K., Nikolaidis, G. & Fraga, S. 2008. Is Gambierdiscus
    expanding to new areas? Harmful Algae 36:6–7.

    Andersen, R. A. [Ed.] 2005. Algal Culturing Techniques. Elsevier, New
    York, 596 pp.

    Anderson, D. M. 1997. Bloom dynamics of toxic Alexandrium species
    in the northeastern U.S. Limnol. Oceanogr. 42:1009–22.

    Anderson, D. M., Cembella, A. D. & Hallegraeff, G. M. [Eds.] 1998.
    Physiological Ecology of Harmful Algal Blooms. Proc. NATO-ASI
    Workshop, Bermuda. Springer Verlag, Heidelberg, Germany,
    662 pp.

    Attrill, M., Wright, J. & Edwards, M. 2007. Climate-related increases
    in jellyfish frequency suggest a more gelatinous future for the
    North Sea. Limnol. Oceanogr. 52:480–5.

    Azanza, R. V. & Taylor, F. J. R. 2001. Are Pyrodinium blooms in
    Southeast Asia recurring and spreading? A view at the end of
    the millennium. Ambio 30:356–64.

    Bagnis, R., Bennett, J. & Barsinas, M. 1985. Epidemiology of
    ciguatera in French Polynesia from 1960 to 1984. In Gabrie, C.
    & Salvat, B. [Eds.] Proc. 5th Int. Coral Reef Congress. Antenne
    Museum, Ephe Moorea, Tahiti, pp. 475–82.

    Barton, A. D., Greene, C. H., Monger, B. C. & Pershing, A. J. 2003.
    Continuous plankton recorder survey of phytoplankton mea-
    surements and the North Atlantic Oscillation: interannual to
    multidecadal variability. Prog. Oceanogr. 58:337–58.

    Bates, S. S., Garrison, D. L. & Horner, R. A. 1998. Bloom dynamics
    and physiology of domoic-acid producing Pseudo-nitzschia spe-
    cies. In Anderson, D. M., Cembella, A. D. & Hallegraeff, G. M.
    [Eds.] Physiological Ecology of Harmful Algal Blooms. NATO ASI
    Series, Vol. G41. Springer Verlag, Heidelberg, Germany, pp.
    267–92.

    Beardall, J. & Raven, J. A. 2004. The potential effects of global
    climate change on microalgal photosynthesis, growth and
    ecology. Phycologia 43:26–41.

    Beardall, J. & Stojkovic, S. 2006. Microalgae under global environ-
    mental change: implications for growth and productivity,
    populations and trophic flow. Science Asia 32(Suppl. 1):1–10.

    Beaugrand, G., Reid, P. C., Ibanez, F., Lindley, J. A. & Edwards, M.
    2002. Reorganization of North Atlantic marine copepod bio-
    diversity and climate. Science 296:1692–4.

    Behrenfeld, M. J., O’Malley, R. T., Siegel, D. A., McClain, C. R.,
    Sarmiento, J. L., Feldman, G. C., Milligan, A. J., Falkowski, P. G.,
    Letelier, R. M. & Boss, E. S. 2006. Climate-driven trends in
    contemporary ocean productivity. Nature 444:752–5.

    Belgrano, A., Lindahl, O. & Hernroth, B. 1999. North Atlantic
    Oscillation, primary productivity and toxic phytoplankton in
    the Gullmar Fjord, Sweden (1985–1996). Proc. R. Soc. Lond. B
    Biol. Sci. 266:425–30.

    Bown, P. R., Lees, J. A. & Young, J. R. 2004. Calcareous nano-
    plankton evolution and diversity through time. In Thierstein,
    H. R. & Young, J. R. [Eds.] Coccolithophores: From Molecular
    Processes to Global Impact. Springer Verlag, Heidelberg, Ger-
    many, pp. 481–508.

    Burkholder, J. M. & Glibert, P. M. 2006. Intraspecific variability: an
    important consideration in forming generalizations about
    toxigenic algal species. Afr. J. Mar Sci. 28:177–80.

    Cembella, A. D. 2003. Chemical ecology of eukaryotic microalgae
    in marine ecosystems. Phycologia 42:420–47.

    Chang, F. H., Sharples, J., Grieve, J. M., Miles, M. & Till, D. G. 1998.
    Distribution of Gymnodinium cf. breve and shellfish toxicity
    from 1993 to 1995 in Hauraki Gulf, New Zealand. In Reguera,
    B., Blanco, J., Fernandez, M. L. & Wyatt, T. [Eds.] Harmful
    Algae. Xunta de Galicia and IOC of UNESCO, Grafisant, San-
    tiago de Compostela, pp. 139–42.

    Charleson, R. J., Lovelock, J. E., Andreae, M. O. & Warren, S. G.
    1987. Oceanic phytoplankton, atmospheric sulphur, cloud al-
    bedo and climate. Nature 326:655–61.

    Collins, S. & Bell, G. 2004. Phenotypic consequences of 1000 gen-
    erations of selection at elevated CO2 in a green alga. Nature
    431:566–9.

    Cubillos, J. C., Wright, S. W., Nash, G., de Salas, M. F., Griffiths, B.,
    Tilbrook, B., Poisson, A. & Hallegraeff, G. M. 2007. Calcifica-
    tion morphotypes of the coccolithophorid Emiliania huxleyi in
    the Southern Ocean: change in 2001–2006 compared to his-
    toric data. Mar. Ecol. Prog. Ser. 384:47–54.

    Cushing, D. H. 1974. The natural regulation of fish populations. In
    Harden-Jones, F. R. [Ed.] Sea Fisheries Research. Elek Science,
    London, pp. 399–412.

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 233

    Dale, B. 2001. The sedimentary record of dinoflagellate cysts:
    looking back into the future of phytoplankton blooms. Sci.
    Mar. 65:257–72.

    Dale, B. 2009. Eutrophication signals in the sedimentary record of
    dinoflagellate cysts in coastal waters. J. Sea Res. 61:103–13.

    Dale, B. & Yentsch, C. M. 1978. Red tide and paralytic shellfish
    poisoning. Oceanus 2:41–9.

    Doblin, M., Thompson, P. A., Revill, A. T., Butler, E. C. V., Black-
    burn, S. I. & Hallegraeff, G. M. 2005. Vertical migration of the
    toxic dinoflagellate Gymnodinium catenatum under different
    concentrations of nutrients and humic substances in culture.
    Harmful Algae 5:665–77.

    Doney, S. C. 2006. Plankton in a warmer world. Nature 444:695–6.
    Edwards, M. 2004. Phytoplankton blooms in the North Atlantic:

    results from the Continuous Plankton Recorder survey
    2001 ⁄ 2002. Harmful Algae News 25:1–3.

    Edwards, M., Johns, D. G., Beaugrand, G., Licandro, P., John, A. W.
    G. & Stevens, D. P. 2008. Ecological status report: results from
    the CPR survey 2006 ⁄ 2007. SAHFOS Tech. Rep. 5:1–8.

    Edwards, M. & Richardson, A. J. 2004. The impact of climate
    change on the phenology of the plankton community and
    trophic mismatch. Nature 430:881–4.

    Eppley, R. W. 1972. Temperature and phytoplankton growth in the
    sea. Fish. Bull. 70:1063–85.

    Erickson, G. & Nishitani, L. 1985. The possible relationship of El
    Nino ⁄ Southern Oscillation events to interannual variation in
    Gonyaulax populations as shown by records of shellfish toxicity.
    In Wooster, W. S. & Flaherty, D. L. [Eds.] El Niño North: Niño
    Effects in the Eastern Subarctic Pacific. Washington Seagrant
    Program, University of Washington, Seattle, pp. 283–90.

    Falkowski, P. G. & Oliver, M. J. 2007. Mix and match: how climate
    selects phytoplankton. Nat. Rev. 5:813–9.

    Feng, Y., Warner, M. E., Zhang, Y., Sun, J., Fu, F.-X., Rose, J. M. &
    Hutchins, D. A. 2008. Interactive effects of increased pCO2,
    temperature and irradiance on the marine coccolithophore
    Emiliania huxleyi (Prymnesiophyceae). Eur. J. Phycol. 43:87–98.

    Fraga, S. & Bakun, A. 1990. Global climate change and harmful
    algal blooms: the example of Gymnodinium catenatum on the
    Galician coast. In Smayda, T. J. & Shimizu, Y. [Eds.] Toxic
    phytoplankton blooms in the sea. Dev. Mar. Biol. 3:59–65.

    Frouin, R. & Lacobellis, S. F. 2002. Influence of phytoplankton on
    the global radiation budget. J. Geographic Res. 107:5-1–-10.

    Gentien, P. 1998. Bloom dynamics and ecophysiology of the
    Gymnodinium mikimotoi species complex. In Anderson, D. M.,
    Cembella, A. D. & Hallegraeff, G. M. [Eds.] Physiological Ecology
    of Harmful Algal Blooms. NATO ASI Series, Vol. G41. Springer
    Verlag, Heidelberg, Germany, pp. 155–73.

    Glibert, P. M., Azanza, R., Burford, M., Furuya, K., Abal, E., Al-Azri, A.,
    Al-Yamani, F., et al. 2008. Ocean urea fertilization for carbon
    credits poses high ecological risks. Mar. Pollut. Bull. 56:1049–56.

    Goffart, A., Hecq, J. H. & Legendre, L. 2002. Changes in the
    development of the winter-spring phytoplankton bloom in the
    Bay of Calvi (northwestern Mediterranean) over the last two
    decades: a response to changing climate. Mar. Ecol. Prog. Ser.
    236:45–60.

    Graneli, E. & Moreira, M. O. 1990. Effects of river water of different
    origin on the growth of marine dinoflagellates and diatoms in
    laboratory cultures. J. Exp. Mar. Biol. Ecol. 136:89–106.

    Graneli, E. & Turner, J. T. [Eds.] 2006. Ecology of Harmful Algae.
    Ecological Studies Series 189. Springer Verlag, Heidelberg,
    Germany, 403 pp.

    Häder, D. P., Worrest, R. C. & Kumar, H. D. 1991. Aquatic eco-
    systems. In van der Leun, J. & Tevini, M. [Eds.] Environmental
    Effects of Ozone Depletion. United Nations Environment Pro-
    gramme, Nairobi, Kenya, pp. 33–40.

    Hales, S., Weinstein, P. & Woodward, A. 2001. Ciguatera (fish
    poisoning), El Niño and Pacific sea surface temperatures.
    Ecosyst. Health 5:20–5.

    Hallegraeff, G. M. 1993. A review of harmful algal blooms and their
    apparent global increase. Phycologia 32:79–99.

    Hallegraeff, G. M. 1998. Concluding remarks on the autecology of
    harmful algal blooms. In Anderson, D. M., Cembella, A. D. &
    Hallegraeff, G. M. [Eds.] Physiological Ecology of Harmful Algal

    Blooms. NATO ASI Series, Vol. G41. Springer Verlag, Heidel-
    berg, Germany, pp. 371–8.

    Hallegraeff, G., Hosja, W., Knuckey, R. & Wilkinson, C. 2008. Re-
    cent range expansion of the red-tide dinoflagellate Noctiluca
    scintillans in Australian coastal waters. IOC-UNESCO Harmful
    Algae Newsletter 38:10–11.

    Hallegraeff, G. M. & Maclean, J. L. 1989. Biology, Epidemiology and
    Management of Pyrodinium Red Tides. International Centre for
    Living Aquatic Resources Management, Manila, Conf. Proc.
    21, 286 pp.

    Hallegraeff, G. M., McCausland, M. J. & Brown, R. K. 1995. Early
    warning of toxic dinoflagellate blooms of Gymnodinium catena-
    tum in southern Tasmanian waters. J. Plankton Res. 17:1163–76.

    Hays, G. C., Richardson, A. J. & Robinson, C. 2005. Climate change
    and marine plankton. Trends Ecol. Evol. 20:337–44.

    Hays, G. C., Warner, A. J., John, A. W. G., Harbour, D. S. &
    Holligan, P. M. 1995. Coccolithophores and the continuous
    plankton recorder survey. J. Mar. Biol. Assoc. U. K. 75:503–6.

    Hayward, T. 1997. Pacific Ocean climate change: atmospheric
    forcing, ocean circulation and ecosystem response. Trends Ecol.
    Evol. 12:150–4.

    Heinze, C. 2004. Simulating oceanic CaCO3 export production in
    the greenhouse. Geophys. Res. Lett. 31: L16308, doi: 10.1029/
    2004GL020613.

    Hoegh-Guldberg, Ø. 1999. Climate change, coral bleaching and the
    future of the world’s coral reefs. J. Mar. Freshw. Res. 50:839–66.

    Hutchins, D. A., Fu, F. X., Zhang, Y., Warner, M. E., Feng, Y.,
    Portune, K., Bernhardt, P. W. & Mulholland, M. R. 2007. CO2
    control of Trichodesmium N2 fixation, photosynthesis, growth
    rates, and elemental ratios: implications for past, present, and
    future ocean biogeochemistry. Limnol. Oceanogr. 52:1293–304.

    Iglesias-Rodriguez, M. D., Halloran, P. R., Rickaby, R. E. M., Hall, I.
    R., Colmenero-Hidalgo, E., Gittins, J. R., Green, D. R. H., et al.
    2008. Phytoplankton calcification in a high-CO2 world. Science
    320:336–40.

    Imai, I., Yamaguchi, M. & Watanabe, M. 1998. Ecophysiology, life
    cycle, and bloom dynamics of Chattonella in the Seto Inland
    Sea, Japan. In Anderson, D. M., Cembella, A. D. & Hallegraeff,
    G. M. [Eds.] Physiological Ecology of Harmful Algal Blooms. NATO
    ASI Series, Vol. G41. Springer Verlag, Heidelberg, Germany,
    pp. 95–112.

    Integrated Marine Observing System (IMOS). 2009. Available at:
    http://imos.org.au/ (last accessed 14 July 2009).

    IPCC. 2008. Climate Change 2007 – Impacts, Adaptation and Vulnera-
    bility. Working Group II contribution to the Fourth Assessment
    Report of the IPCC Intergovernmental Panel on Climate
    Change. Available at: http://www.ipcc.ch/ (last accessed 16
    January 2010).

    Jeffrey, S. W., MacTavish, H. S., Dunlap, W. C., Vesk, M. & Gro-
    enewoud, K. 1999. Occurrence of UV A- and UV B-absorbing
    compounds in 152 species (206 strains) of marine microalgae.
    Mar. Ecol. Prog. Ser. 189:35–51.

    Kimura, T., Mizokami, A. & Hashimoto, T. 1973. The red tide that
    caused severe damage to the fishery resources in Hiroshima
    Bay: outline of its occurrence and environmental conditions.
    Bull. Plankton Soc. Jpn. 19:82–96.

    Langer, G., Nehrke, G., Probert, I., Ly, J. & Ziveri, P. 2009. Strain-
    specific responses of Emiliania huxleyi to changing seawater
    carbonic chemistry. Biogeosci. Disc. 6:4361–83.

    Lefebvre, K. A., Bargu, S., Kieckhefer, T. & Silver, M. W. 2002. From
    sanddabs to blue whales: the pervasiveness of domoic acid.
    Toxicon 40:971–7.

    Lenton, A., Codron, F., Bopp, L., Metzl, N., Cadule, P., Tagliabue,
    A. & Le Sommer, J. 2009. Stratospheric ozone depletion re-
    duces ocean carbon uptake and enhances ocean acidification.
    Geophys. Res. Lett. 36:L12606. doi: 10.1029/2009GL038227.

    Li, W. K. W. 1980. Temperature adaptation in phytoplankton: cel-
    lular and photosynthetic characteristics. In Falkowski, P. [Ed.]
    Primary Productivity in the Sea. Plenum Press, New York, pp.
    259–79.

    Lilly, E. L., Kulis, D. M., Gentien, P. & Anderson, D. M. 2002.
    Paralytic shellfish poisoning toxins in France linked to a
    human-introduced strain of Alexandrium catenella from the

    234 G U S T A A F M . H A L L E G R A E F F

    western Pacific: evidence from DNA and toxin analysis.
    J. Plankton Res. 24:443–52.

    Longhurst, A., Sathyendranath, S., Platt, T. & Caverhill, C. 1995. An
    estimate of global primary production in the ocean from sa-
    tellite radiometer data. J. Plankton Res. 17:1245–71.

    Lorius, C., Jouzel, J., Raynaud, D., Hansen, J. & Le Treut, H. 1990.
    The ice-core record: climate sensitivity and future greenhouse
    warming. Nature 347:139–45.

    Maclean, J. L. 1989. Indo-Pacific red tides, 1985–1988. Mar. Pollut.
    Bull. 20:304–10.

    Margalef, R. 1978. Life-forms of phytoplankton as survival alterna-
    tives in an unstable environment. Oceanologia Acta 1:493–509.

    McMinn, A. 1988. A Late Pleistocene dinoflagellate assemblage
    from Bulahdelah, N.S.W. Proc. Linn. Soc. N. S. W. 109:175–81.

    McMinn, A. 1989. Late Pleistocene dinoflagellate cysts from Botany
    Bay, New South Wales, Australia. Micropaleontology 35:1–9.

    McMinn, A. & Sun, X. 1994. Recent dinoflagellates from the
    Chatham Rise, east of New Zealand; a taxonomic study. Paly-
    nology 18:41–53.

    Merico, A., Tyrrell, T., Brown, C. W., Groom, S. B. & Miller, P. I.
    2003. Analysis of satellite imagery for Emiliania huxleyi blooms
    in the Bering Sea before 1997. Geophys. Res. Lett. 30:1337. doi:
    10.1029/2002GL016648.

    Moore, S. K., Mantua, J. M., Hickey, B. & Trainer, V. L. 2008a.
    Recent trends in paralytic shellfish toxins in Puget Sound,
    relationship to climate, and capacity for prediction of toxic
    events. Harmful Algae 8:463–77.

    Moore, S. K., Trainer, V. L., Mantua, N. J., Parker, M. S., Laws, E. A.,
    Baker, L. C. & Fleming, L. E. 2008b. Impacts of climate vari-
    ability and future climate change on harmful algal blooms and
    human health. Environ. Health 7(Suppl. 2):S4. doi: 10.1186/
    1476-069X-7-S2-S4.

    Munhoven, G. 2007. Glacial-interglacial rain ratio changes: impli-
    cations for atmospheric CO2 and ocean-sediment interaction.
    Deep-Sea Res. Part I Oceanogr. Res. Pap. 54:722–46.

    NSTC Joint Subcommittee on Ocean Science and Technology.
    2007. Charting the Course for Ocean Science in the United States for
    the Next Decade: An Ocean Research Priorities Plan and Implemen-
    tation Strategy. Available at: http://ocean.ceq.gov/about/docs/
    orppfinal (last accessed 14 July 2009).

    Orr, J. C., Fabry, V. J., Aumont, O., Bopp, L., Doney, S. C., Feely, R.
    A., Gnanadesikan, A., et al. 2005. Anthropogenic ocean acid-
    ification over the twenty-first century and its impact on calci-
    fying organisms. Nature 437:681–6.

    Park, M. G., Kim, S., Kang, Y. G. & Yih, W. 2006. First successful
    culture of the marine dinoflagellate Dinophysis acuminata.
    Aquat. Microb. Ecol. 45:101–6.

    Peperzak, L. 2005. Future increase in harmful algal blooms in the
    North Sea due to climate change. Water Sci. Technol. 51:31–6.

    Rahmstorf, S. 2002. Ocean circulation and climate during the past
    120,000 years. Nature 419:207–14.

    Raven, J., Caldeira, K., Elderfield, H., Hoegh-Guldberg, O., Liss, P.,
    Riebesell, U., Shepherd, J., Turley, C. & Watson, A. 2005b.
    Ocean Acidification Due to Increasing Atmospheric Carbon Dioxide.
    The Royal Society, London, 68 pp.

    Raven, J. A., Finkel, Z. V. & Irwin, A. J. 2005a. Picophytoplankton:
    bottom-up and top-down controls on ecology and evolution.
    Vie Milieu 55:209–15.

    Reid, P. C., Johns, D. G., Edwards, M., Starr, M., Poulin, M. &
    Snoeijs, P. 2007. A biological consequence of reducing Arctic
    ice cover: arrival of the Pacific diatom Neodenticula seminae in
    the North Atlantic for the first time in 800,000 years. Glob.
    Change Biol. 13:1910–21.

    Rhodes, L. L., Haywood, A. J., Ballantine, W. J. & MacKenzie, A. L.
    1993. Algal blooms and climate anomalies in North-east New
    Zealand, August–December 1992. N. Z. J. Mar. Freshw. Res.
    27:419–30.

    Richardson, A. J. & Schoeman, D. S. 2004. Climate impact on
    plankton ecosystems in the Northeast Atlantic. Science
    305:1609–12.

    Richardson, A. J., Walne, A. W., John, A. W. G., Jonas, T. D.,
    Lindley, J. A., Sims, D. W., Stevens, D. & Witt, M. 2006. Using
    continuous plankton recorder data. Prog. Oceanogr. 68:27–74.

    Riebesell, U., Schulz, K. G., Bellerby, R. G. J., Botros, M.,
    Fritsche, P., Meyerhöfer, M., Neill, C., et al. 2007. Enhanced
    biological carbon consumption in a high CO2 ocean. Nature
    450:545–8.

    Riebesell, U., Zondervan, I., Rost, B., Tortell, P. D., Zeebe, R. E. &
    Morel, F. M. M. 2000. Reduced calcification of marine plank-
    ton in response to increased atmospheric CO2. Nature
    407:364–7.

    Rost, B. & Riebesell, U. 2004. Coccolithophores and the biological
    pump: responses to environmental changes. In Thierstein, H.
    R. & Young, J. R. [Eds.] Coccolithophores: From Molecular Processes
    to Global Impact. Springer Verlag, Heidelberg, Germany, pp.
    99–125.

    Rost, B., Zondervan, I. & Wolf-Gladrow, D. 2008. Sensitivity of
    phytoplankton to future changes in ocean carbonate chemis-
    try: current knowledge, contradictions and research direc-
    tions. Mar. Ecol. Prog. Ser. 373:227–37.

    Schippers, P., L}urling, M. & Scheffer, M. 2004. Increase of atmo-
    spheric CO2 promotes phytoplankton productivity. Ecol. Lett.
    7:446–51.

    Shears, N. T. & Ross, P. M. 2009. Blooms of benthic dinoflagellates
    of the genus Ostreopsis: an increasing and ecologically impor-
    tant phenomenon on temperate reefs in New Zealand and
    worldwide. Harmful Algae 8:916–25.

    Shelly, K., Heraud, P. & Beardall, J. 2002. Nitrogen limitation in
    Dunaliella tertiolecta Butcher (Chlorophyceae) leads to in-
    creased susceptibility to damage by ultraviolet-B radiation but
    also increased repair capacity. J. Phycol. 38:1–8.

    Smayda, T. J. 1990. Novel and nuisance phytoplankton blooms in
    the sea: evidence for a global epidemic. In Granéli, E., Sun-
    dström, B., Edler, L. & Anderson, D. M. [Eds.] Toxic Marine
    Phytoplankton. Elsevier, New York, pp. 29–40.

    Smayda, T. J. & Reynolds, C. S. 2001. Community assembly in
    marine phytoplankton; application of recent models to
    harmful dinoflagellate blooms. J. Plankton Res. 23:447–61.

    Stachowicz, J. J., Terwin, J. R., Whitlatch, R. B. & Osman, R. W.
    2002. Linking climate change and biological invasions: ocean
    warming facilitates non-indigenous species invasion. Proc. Natl.
    Acad. Sci. U. S. A. 99:15497–500.

    Sunda, W. G., Graneli, E. & Gobler, C. J. 2006. Positive feedback
    and the development and persistence of ecosystem disruptive
    algal blooms. J. Phycol. 42:963–74.

    Tester, P. A., Geesey, M. E. & Vukovich, F. M. 1993. Gymnodinium
    breve and global warming: what are the possibilities? In Smayda,
    T. J. & Shimizu, Y. [Eds.] Toxic phytoplankton blooms in the
    sea. Dev. Mar. Biol. 3:67–72.

    Tester, P. A., Stumpf, R. P., Vukovich, F. M., Folwer, P. K. & Turner,
    J. T. 1991. An expatriate red tide bloom: transport, distribu-
    tion, and persistence. Limnol. Oceanogr. 36:1053–61.

    Thorsen, T., Dale, B. & Nordberg, K. 1995. Blooms of the toxic
    dinoflagellate Gymnodinium catenatum as evidence of climatic
    fluctuations in the Late Holocene of southwestern Scandina-
    via. The Holocene 5:435–46.

    Tortell, P. D., Giocoma, R. D., Sigman, D. M. & Morel, F. M. M.
    2002. CO2 effects on taxonomic composition and nutrient
    utilization in an equatorial Pacific phytoplankton assemblage.
    Mar. Ecol. Prog. Ser. 236:37–43.

    Turner, J. T. & Graneli, E. 2006. ‘‘Top-down’’ predation control on
    marine harmful algae. In Graneli, E. & Turner, J. T. [Eds.]
    Ecology of Harmful Algae. Ecological Studies Series 189.
    Springer, Verlag, Heidelberg, Germany, pp. 355–66.

    Van Dolah, F. V. 2000. Marine algal toxins: origins, health effects,
    and their increased occurrence. Environ. Health Perspect.
    108(Suppl. 1):133–41.

    Walsh, J. J. & Steidinger, K. A. 2001. Saharan dust and Florida red
    tides: the cyanophyte connection. J. Geophys. Res. 106:11597–
    612.

    Woods, J. & Barkmann, W. 1993. The plankton multiplier –
    positive feedback in the greenhouse. J. Plankton Res. 15:
    1053–74.

    Yin, K. D., Harrison, P. J., Chen, J., Huang, W. & Qian, P. Y. 1999.
    Red tides during spring in 1998 in Hong Kong: is El Niño
    responsible? Mar. Ecol. Prog. Ser. 187:289–94.

    C L I M A T E C H A N G E A N D A L G A L B L O O M S 235

    REVIEW

    TESTING THE EFFECTS OF OCEAN ACIDIFICATION ON ALGAL METABOLISM:
    CONSIDERATIONS FOR EXPERIMENTAL DESIGNS

    1

    Catriona L. Hurd,2 Christopher D. Hepburn

    Department of Botany, University of Otago, PO Box 56, Dunedin 9054, New Zealand

    Kim I. Currie

    National Institute for Water and Atmospheric Research Ltd., Centre of Excellence for Chemical and Physical Oceanography,

    Department of Chemistry, University of Otago, PO Box 56, Dunedin 9054, New Zealand

    John A. Raven

    Division of Plant Sciences, Scottish Crop Research Institute, University of Dundee at SCRI, Invergowrie, Dundee DD2 5DA, UK

    and Keith A. Hunter

    Department of Chemistry, University of Otago, PO Box 56, Dunedin 9054, New Zealand

    Ocean acidification describes changes in the car-
    bonate chemistry of the ocean due to the increa-
    sed absorption of anthropogenically released CO2

    .

    Experiments to elucidate the biological effects of
    ocean acidification on algae are not straightforward
    because when pH is altered, the carbon speciation
    in seawater is altered, which has implications for
    photosynthesis and, for calcifying algae, calcifica-
    tion. Furthermore, photosynthesis, respiration, and
    calcification will themselves alter the pH of the sea-
    water medium. In this review, algal physiologist

    s

    and seawater carbonate chemists combine their
    knowledge to provide the fundamental information
    on carbon physiology and seawater carbonate chem-
    istry required to comprehend the complexities of
    how ocean acidification might affect algae metabo-
    lism. A wide range in responses of algae to ocean
    acidification has been observed, which may be
    explained by differences in algal physiology, time-
    scales of the responses measured, study duration,
    and the method employed to alter pH. Two meth-
    ods have been widely used in a range of experimen-
    tal systems: CO2 bubbling and HCl ⁄ NaOH
    additions. These methods affect the speciation of
    carbonate ions in the culture medium differently; we
    discuss how this could influence the biological
    responses of algae and suggest a third method based
    on HCl ⁄ NaHCO3 additions. We then discuss eight
    key points that should be considered prior to setting
    up experiments, including which method of manipu-
    lating pH to choose, monitoring during experiments,
    techniques for adding acidified seawater, biological

    side effects, and other environmental factors. Finally,
    we consider incubation timescales and prior condi-
    tioning of algae in terms of regulation, acclimation,
    and adaptation to ocean acidification.

    Key index words: algae; bicarbonate; calcium car-
    bonate; carbon; carbon dioxide; climate change;
    ocean acidification; phytoplankton; seawater car-
    bonate system; seaweed

    Abbreviations: AT, total alkalinity; CA, carbonic
    anhydrase; CCM, carbon-concentrating mecha-
    nism; CT, total inorganic carbon; pCO2, partial
    pressure of CO2(g

    )

    The term ‘‘ocean acidification’’ describes changes
    in the carbonate chemistry of the ocean due to
    increased CO2 absorption since the Industrial Revo-
    lution (The Royal Society 2005, Doney et al. 2009

    ).

    Phytoplankton and macroalgae have key ecological
    roles as primary producers in coastal and open
    oceans, supplying fixed carbon to the entire marine
    food web, recycling nutrients, and modifying global
    climate (Smith 1981, Duggins et al. 1989, Field et al.
    1998, Gattuso et al. 1998a, Zondervan 2007). Addi-
    tionally, calcareous algae (e.g., planktonic cocco-
    lithophores and benthic calcifying macroalgae) are
    a major source of marine carbonates and sediments
    (Gattuso et al. 1998b, Feely et al. 2004, Balch et al.
    2007), and the coralline macroalgae fulfill impor-
    tant ecological processes, including reef building
    (Adey 1998, Chisholm 2003), and are the preferred
    sites for settlement of invertebrate larvae (Roberts
    2001).

    A major focus of research into ocean acidification
    has been the effects on calcifying organisms

    1Received 28 November 2008. Accepted 6 May 2009.
    2Author for correspondence: e-mail catriona.hurd@botany.otago.

    ac.nz.

    J. Phycol. 45,

    1236

    –1251 (2009)
    � 2009 Phycological Society of Americ

    a

    DOI: 10.1111/j.1529-8817.2009.00768.x

    1236

    (animals and algae). Elevated seawater CO2 concen-
    trations will lower carbonate saturation states, which
    in turn may reduce the ability of calcifiers to main-
    tain existing, and build new, carbonate skeletons
    (Bijma et al. 1999, Riebesell et al. 2000, Orr et al.
    2005, Shirayama and Thornton 2005, De’ath et al.
    2009). These effects will occur in high latitudes first;
    for example, in Southern Ocean surface waters,
    undersaturation of aragonite is predicted between
    2030 and 2038 (McNeil and Matear 2008). Such
    reduced ability to calcify may decrease their compet-
    itive fitness (Kuffner et al. 2007, McNeil and Matear
    2008). In addition to influencing calcification, the
    changed speciation of dissolved inorganic carbon in
    seawater and decreased pH via the carbonate buffer
    system, and the differing abilities of algae to utilize
    CO2 and HCO3

    ), ocean acidification has the poten-
    tial to affect the metabolism and growth rates of all
    algae, both noncalcareous and calcareou

    s.

    Ocean acidification is an emerging field of
    research. Experiments to elucidate the biological
    effects of increased CO2 on marine organisms are
    not straightforward because when pH is altered, the
    carbon speciation in seawater is modified, which has
    strong implications for photosynthesis, respiration,
    and calcification. Furthermore, these same three
    metabolic processes themselves alter the pH of sea-
    water medium surrounding the algae. Therefore, an
    understanding of both seawater carbonate chemistry
    and physiological processes related to carbon metab-
    olism and calcification is required to design experi-
    ments without artifacts, which can be carefully
    replicated and test the impacts of increased CO

    2

    (i.e., lowered pH) on algae. This review is in two
    sections. In the first section, we highlight key
    aspects of seawater carbonate chemistry, algal car-
    bon acquisition, and calcification and consider the
    wide range of biological responses by different algae
    to ocean acidification. On the basis of this appraisal,
    we consider reasons for the observed broad range
    of physiological responses, which include physiology
    of different species, timescales of studies, and tech-
    niques used to modify seawater pH (CO2 bubbling
    vs. HCl ⁄ NaOH additions). In section two, we focus
    on how bubbling with CO2 and adding HCl ⁄ NaOH
    each modifies the carbonate chemistry of seawater
    during incubation studies, discuss the probable bio-
    logical responses of algae to each technique, and
    recommend a series of steps that should be consid-
    ered when designing experiments to test the effects
    of ocean acidification on algae.

    Carbon chemistry and biological responses to its manip-
    ulation. Seawater carbonate chemistry: Since the Indus-
    trial Revolution, levels of CO2 in the atmosphere
    have increased at rates 100-fold greater than prein-
    dustrial times and have caused a rise in atmospheric
    CO2 levels from 280 to 384 ppmv (Solomon et al.
    2007). CO2 in the atmosphere is increasing at rates
    faster than that predicted as a ‘‘worst-case scenario’’

    by the International Panel for Climate Change
    (IPCC) in 2000 (Raupach et al. 2007). The world’s
    oceans have absorbed up to 50% of anthropogeni-
    cally derived CO2, and modeling studies suggest a
    0.1 unit decline in surface seawater pH since 1750
    (Caldiera and Wickett 2003).

    In seawater, free CO2(aq) is in equilibrium with a
    small concentration of the carbonic acid species
    H2CO3, but it is conventional to regard both species
    as stoichiometrically equivalent with respect to sub-
    sequent acid-base reactions and denote this combi-
    nation by a hypothetical species H2CO3

    * where
    [H2CO3

    *] = [H2CO3] + [CO2(aq)]. The effects of
    ocean acidification cannot be described in a simple
    way using just the parameter pH, and it is necessary
    to consider the effect of CO2 uptake on the entire
    CO2 equilibrium system. CO2 in the gas phase equil-
    ibrates with H2CO3

    * in seawater through the well-
    known Henry’s law equilibrium:

    CO2ðgÞ$ H2CO3� KH ¼
    ½H2CO3��

    pCO2
    ð1Þ

    where pCO2 is the partial pressure of CO2(g) and
    KH, the Henry’s law equilibrium constant, is a func-
    tion of temperature (T) and salinity (S). This rela-
    tionship means that at a given T and S, pCO2 and
    [H2CO3

    *] are linearly related to each other. It is
    most common to use pCO2 as a parameter because
    this allows a simple comparison with the actual
    atmospheric CO2 partial pressure when air and
    water are not in equilibrium.

    The acid dissociation reactions of H2CO3
    * are as

    follows:

    H2CO3
    �$HCO3�þHþ K1¼

    ½HCO3��½Hþ�
    ½H2CO3��

    ð2Þ

    HCO3
    �$CO32�þHþ K2¼

    ½CO32��½Hþ�
    ½HCO3��

    ð3Þ

    where HCO3
    ) and CO3

    2) are the bicarbonate and car-
    bonate ions, respectively, and K1 and K2 are the first
    and second acid dissociation constants of H2CO3

    *.
    Equations (2) and (3) show that the concentrations
    of the three CO2 species and that of H

    + are inextrica-
    bly linked, meaning that it is physically impossible to
    vary systematically any one of these while at the same
    time holding all of the others constant. Equation (1)
    shows that this relationship also extends to pCO2.
    This fact complicates understanding the underlying
    chemistry affecting ocean acidification.

    Concentrations of the individual CO2 species in
    seawater cannot be directly measured. Instead,
    changes in the speciation of the CO2 system in sea-
    water are normally described, and measured, using
    the following two parameters (Mackenzie and
    Lerman 2006): (i) Total dissolved CO2 (usually
    symbolized as CT for the total inorganic carbon in
    solution or DIC for dissolved inorganic carbon),

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1237

    which is the stoichiometric sum of all dissolved inor-
    ganic carbon species

    C T ¼ ½H2CO3��þ ½HCO3��þ ½CO32�� ð4Þ

    (ii) Total alkalinity, AT, which is the total concentra-
    tion of titratable weak bases in seawater relative to
    the reference proton condition comprising pure
    CO2 in seawater

    AT¼½HCO3��þ2½CO32��þ½OH���½Hþ�þð…Þ ð5Þ
    where (…) represents various minor acid-base spe-
    cies, such as borate ion. Both these parameters have
    the advantage of being independent of changes in
    temperature and pressure and are conserved during
    the mixing of different seawater masses. Useful soft-
    ware programs for calculation of CO2 speciation in
    seawater have been presented by Lewis and Wallace
    (1998) and Hunter (2007).

    For calcification, the removal of CO3
    2) ions by

    precipitation of calcium carbonate (CaCO3) causes
    HCO3

    ) ions to dissociate to restore the CO3
    2) ions

    lost. The H+ ion released by this dissociation gener-
    ates additional H2CO3

    * by combining with a second
    HCO3

    ) ion (Frankignoulle and Canon 1994). The
    overall stoichiometric change is therefore:

    Ca2þ þ 2HCO3� ! CaCO3ðsÞþ H2CO3� ð6Þ

    At today’s pH (�8.07), 91% of CT is as bicarbon-
    ate ions (2,200 lM), 1% as H2CO3

    * (14 lM), and
    8% as CO3

    2). The predicted decrease to pH 7.65 by
    2100 will result in a 300% increase in H2CO3

    * con-
    centration, a 9% increase in that of HCO3

    ), and a
    56% decrease in that of CO3

    2) (from table 1 of The
    Royal Society 2005). These changes will affect the
    ability of algae to acquire carbon and ⁄ or produce
    and maintain calcium carbonate structures.

    Physiological basis for algal carbon acquisition and
    calcification: Most marine algae can acquire the
    CO2 required as a substrate for RUBISCO via active
    uptake from seawater of CO2 and ⁄ or bicarbonate;
    the active transport of either of these species, or in
    some cases of protons, constitutes a carbon-concen-
    trating mechanism (CCM; Giordano et al. 2005).
    The photosynthetic rates of algae that have CCMs
    are not generally carbon limited under most envi-
    ronmental conditions (Giordano et al. 2005). Some
    bicarbonate-using algae convert HCO3

    ) to CO2
    using extracellular carbonic anhydrase (CA); the
    CO2 then enters the cell by active transport or by
    diffusion (if there are zones of surface acidification
    where the steady-state CO2 concentration exceeds
    that in the medium). Other bicarbonate-using algae
    with CCMs actively take up the HCO3

    ) ion across
    the cell membrane, and CA acts intracellularl

    y.

    Some algae (e.g., some dinoflagellates) have little or
    no capacity to use bicarbonate, and their CCM relies
    on active CO2 uptake (Dason et al. 2004). CA syn-
    thesis and CCM activities in eukaryotes are con-

    trolled by the concentration of H2CO3
    *, in the few

    cases examined (Giordano et al. 2005).
    Energy and nutrients are required to operate

    active transport and to make the CCMs (generally
    including CAs), whereas the alternative of diffusive
    H2CO3

    * has energy and nutrient costs of operating
    photorespiration and making the relevant enzymes
    and additional RUBISCO (Raven et al. 2000). Some
    algae adapted to low light levels lack CCMs and rely
    on diffusive H2CO3

    * entry (e.g., the red seaweed
    Lomentaria, Kübler et al. 1999). Under low irradianc-
    es, energy limitation outweighs limitation by CO2,
    and the use of diffusive H2CO3

    * entry has energetic
    advantages (see Raven et al. 2000, 2005). Growth at
    low irradiances cannot explain all cases of the
    absence of detectable CCMs; for example, some
    strains of coccolithophores rely on diffusive uptake
    of H2CO3

    * and cannot utilize HCO3
    ) or carry out

    active CO2 transport (Nimer and Merrett 1993).
    Algae that rely on H2CO3

    * diffusion alone are gen-
    erally carbon-limited under today’s seawater concen-
    trations (Kübler et al. 1999).

    Calcification is the biogenic formation of calciu

    m

    carbonate (Borowitzka 1987). The most common
    forms of calcium carbonate (CaCO3) synthesized by
    algae are calcite, aragonite, or high-magnesium cal-
    cite (Adey 1998). High magnesium calcite is the
    most soluble form of these three (Chave et al.
    1962), and therefore algae with high-magnesium
    calcite are most susceptible to predicted decreases
    in pH due to ocean acidification (The Royal Society
    2005). Algae have a range of mechanisms for calcify-
    ing. For example, coccolithophores produce calcite
    coccoliths intracellularly and extrude them to the
    cell’s surface, whereas in the tropical green seaweed
    Halimeda, aragonite mineralizes in the intercellular
    space between tightly appressed utricles, and red
    coralline seaweeds deposit high-magnesium calcite
    into their cell walls (Borowitzka 1987).

    Algae themselves modify the pH of seawate

    r.

    When algae photosynthesize, the removal of CO2 in
    assimilation by RUBISCO usually occurs faster than
    CO2 can be resupplied from the atmosphere or dee-
    per water, so that there is a reequilibration among
    the inorganic species that yields a decrease in
    HCO3

    ) and an increase in CO3
    2) and pH. In nat-

    ure, this results in significant pH increases most
    especially in isolated habitats like high-intertidal
    rock pools (Midelboe and Hansen 2007) but also in
    coastal waters (Hinga 2002). Indeed, such effects
    are the principle underpinning pH-drift experi-
    ments where a rapid increase in pH as a result of
    photosynthesis is used to elucidate mechanisms of
    carbon acquisition by seaweeds and microalgae
    (Maberly 1990, Chen et al. 2006). Calcification and
    respiration alter the seawater carbonate system in
    ways that decrease the pH of the culture medium.
    Respiration results in CO2 being released into the
    surrounding medium, and when algae are in the
    dark, the pH of the culture medium will decline.

    1238 C A T R I O N A L . H U R D E T A L .

    T
    a

    b
    l
    e

    1
    .

    N
    o

    n
    ca

    lc

    a
    re

    o
    u

    s

    a
    lg

    a
    e

    .

    E
    x

    a
    m

    p
    le

    s
    o

    f

    e
    x

    p
    e
    ri

    m

    e
    n

    t

    s
    co

    n
    d

    u
    ct

    e
    d

    to
    te

    st

    th

    e

    e
    ff

    e
    ct

    s
    o

    f
    C

    O
    2

    c

    o
    n

    ce
    n

    tr

    a
    ti

    o
    n

    in
    se

    a
    w

    a
    te

    r
    o

    n
    a

    lg

    a
    l

    g
    ro

    w
    th

    a
    n

    d
    ⁄o

    r
    m

    e
    ta

    b
    o

    li
    sm

    .
    U

    n
    it

    s
    fr

    o
    m

    th
    e

    o
    ri

    g
    in

    a
    l

    p
    a
    p

    e

    rs

    a
    re

    r

    e
    p

    o
    rt

    e
    d

    .
    L

    😀
    ,

    li
    g

    h
    t:

    d

    a
    rk

    .

    S
    p

    e
    ci

    e
    s

    ⁄s
    ys

    t

    e
    m

    M
    e
    th

    o
    d

    s
    T

    im
    e

    fr
    a
    m

    e
    p

    H
    ⁄p

    C
    O

    2

    tr
    e

    a
    tm

    e
    n

    ts
    R

    e
    p

    li
    ca

    ti
    o

    n
    R

    e
    su

    lt
    s

    A
    u

    t

    h
    o

    rs

    P
    h

    y

    t
    o

    p
    la

    n
    k

    to
    n

    A
    st

    er
    io

    n
    el

    la
    gl

    a
    ci

    a
    li

    s,
    T

    h
    a
    la

    ss
    io

    si
    ra

    p
    u

    n
    ct

    ig
    er

    a
    ,

    C
    os

    ci
    n

    od
    is

    c

    u
    s

    w
    a
    il

    es
    ii

    ,
    P

    h
    a
    eo

    d
    a
    ct

    yl
    u

    m
    tr

    ic
    or

    n
    u

    t

    u
    m

    ,
    S
    cr

    i

    p
    p

    si

    el
    la

    tr
    oc

    h
    oi

    d
    ea

    H
    C

    l⁄
    N

    a
    O

    H
    ,

    d
    il

    u
    te

    b

    a
    tc

    h
    cu

    lt
    u

    re
    s,

    1
    5

    0
    l

    m
    o

    l
    p

    h
    o
    to
    n

    s
    Æ

    m
    )

    2
    Æ

    s

    )
    1

    ,

    1
    6

    :8
    o

    r
    2

    4
    :0

    L
    😀

    ,
    f
    ⁄2

    n
    u

    tr
    ie

    n
    ts

    .

    C
    o

    m
    p

    le
    te

    a
    cc

    li
    m

    a
    ti
    o
    n

    o
    f

    cu
    lt

    u
    re

    s
    to

    g
    ro
    w
    th

    co
    n

    d
    it

    io
    n

    s
    li

    k
    e
    ly

    .

    S
    ix

    C
    O
    2
    co
    n
    ce
    n

    tr
    a
    ti

    o
    n

    s
    ra

    n
    g

    in
    g

    fr
    o

    m
    1

    to
    3

    8
    l

    m
    o

    l
    Æ

    k
    g

    )
    1
    N
    o

    t
    cl

    e
    a
    r

    N
    o
    e
    ff
    e
    ct
    o
    f

    tr

    e
    a
    tm

    e
    n
    t
    o
    n
    g

    r

    o
    w

    th
    ra

    te
    o

    n
    a

    c

    e
    ll

    in
    cr

    e
    a
    se

    o
    r

    o
    rg

    a
    n

    ic
    ca

    rb
    o

    n
    b

    a
    si

    s.

    B
    u

    rk
    h

    a
    rd

    t

    e
    t

    a
    l.

    (1
    9

    9
    9

    )
    P
    h

    yt
    o

    p
    la
    n
    k
    to
    n

    a
    ss

    e
    m

    b
    la

    g
    e

    s,

    e
    q

    u
    a
    to

    ri
    a

    l

    P
    a
    ci

    fi
    c

    o
    ff

    P
    e
    ru

    C
    O

    2
    b

    u
    b

    b
    li

    n
    g

    ,
    se

    m
    ic

    o
    n

    ti
    n

    u
    o

    u
    s

    b
    a
    tc

    h
    cu
    lt
    u

    re
    ,

    3
    0

    %
    su

    rf

    a
    ce

    ir
    ra

    d
    ia

    n
    ce

    ,
    su

    rf
    a
    ce

    se
    a
    w

    a
    te

    r
    te

    m
    p

    e
    ra

    tu
    re

    ,
    n

    a
    tu

    ra
    l

    n
    u
    tr
    ie
    n
    ts
    .
    E
    x
    p
    e
    ri

    m
    e
    n

    ts
    la

    st
    e
    d

    1
    1

    d
    .

    p
    C

    O
    2
    1
    5

    0
    (j

    u
    st

    le
    ss

    th
    a
    n

    l

    a
    st

    g
    la

    ci
    a
    l

    m
    a
    x

    i

    m
    u

    m
    )
    a
    n

    d
    7

    5
    0

    p
    p
    m

    n
    =

    3
    N

    o
    si

    g
    n

    ifi
    ca

    n
    t

    e
    ff
    e
    ct

    ch
    a
    n

    g
    e
    d

    C
    O

    2
    o

    n
    p

    ri
    m

    a
    ry

    p
    ro

    d
    u

    ct
    io

    n
    (1

    4
    C

    )
    o

    f
    th

    e
    a
    ss

    e
    m
    b
    la

    g
    e
    .

    N
    a
    n

    o
    fl

    a
    g

    e
    ll
    a
    te

    s
    (u

    n
    ic

    e
    ll

    u
    la

    r
    P

    h
    a
    eo

    cy
    st

    is
    )

    le
    ss

    si
    g

    n
    ifi

    ca
    n

    t
    re

    la
    ti

    ve
    to

    d
    ia

    to
    m

    s
    in

    h
    ig

    h
    C

    O

    2
    .

    T
    o

    rt
    e
    ll

    e
    t
    a
    l.

    (

    2
    0

    0
    2

    )

    E
    m

    il
    ia

    n
    ia

    h
    u

    xl
    ey

    i
    (n

    o
    n

    c

    a
    lc

    if
    yi

    n
    g

    s

    t
    ra

    in
    )

    C
    O
    2
    b
    u
    b
    b
    li
    n
    g

    ,
    ch

    e
    m

    o
    st

    a
    t

    a
    n

    d
    b

    a
    tc
    h
    cu
    lt
    u
    re
    s,

    8
    0

    o
    r
    5
    0
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ

    s)
    1

    ir
    ra
    d
    ia
    n
    ce

    tr
    e
    a
    tm

    e
    n

    ts
    ,

    1
    4

    :

    1
    0

    L
    😀

    ,
    1

    8
    �C

    ,
    a
    ls

    o
    m
    a
    n

    ip
    u

    la
    te

    d
    n

    u
    tr

    ie
    n

    ts
    .

    1
    p

    lu
    s

    3
    w

    e
    e
    k

    s
    a
    cc

    li
    m
    a
    ti
    o
    n

    to
    e
    x

    p
    e
    ri
    m
    e
    n

    ta
    l

    tr
    e
    a
    tm
    e
    n
    ts
    .
    p
    C
    O
    2

    =
    3

    6
    0

    a
    n

    d
    2

    ,

    0
    0

    0
    p

    p
    m

    n
    =

    3
    In

    cr
    e
    a
    se

    d
    g

    ro
    w

    th
    a
    t

    h
    ig

    h

    e
    r

    p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    t
    ra

    te
    p

    e
    r
    o
    rg
    a
    n

    ic
    m

    a
    tt

    e
    r,

    b
    u

    t
    n

    o
    in

    cr
    e
    a
    se
    o
    n
    a
    ce

    ll
    n

    u
    m

    b
    e
    r

    b
    a
    si

    s.

    L
    e
    o

    n
    a
    rd

    o
    s

    a
    n

    d
    G

    e
    id

    e
    r

    (2
    0

    0
    5

    )

    P
    ro

    ch
    lo

    ro
    co

    cc
    u

    s
    C

    C
    M

    P
    1

    9
    8

    6
    ,

    S
    yn

    ec
    h
    oc

    cu
    s

    sp
    .

    C
    C

    M
    P

    1
    3

    3
    4

    C
    O
    2
    b
    u
    b
    b
    li
    n
    g
    ,
    se
    m
    ic
    o
    n
    ti
    n
    u
    o
    u
    s
    cu
    lt
    u
    re

    ,
    4

    5
    l

    m
    o
    l
    Æ
    m
    )
    2
    Æ

    s)

    1
    ,

    1
    2

    :1
    2

    L
    😀

    ,
    2

    0
    �C

    a
    n
    d
    2

    4
    �C

    ,
    o

    li
    g

    o
    tr

    o
    p

    h
    ic

    n
    a
    tu

    ra
    l
    se
    a
    w
    a
    te

    r
    w

    it
    h

    A
    Q

    U
    IL

    n
    u
    tr
    ie
    n
    ts
    .

    A
    t

    le
    a
    st

    2
    w

    e
    e
    k

    s
    (a

    t
    le

    a
    st

    7
    g

    e
    n
    e
    ra
    ti
    o

    n
    s)

    a
    cc
    li
    m
    a
    ti
    o
    n
    .
    p
    C
    O
    2

    3
    8

    0
    a
    n

    d
    7
    5
    0
    p
    p

    m
    n

    =
    3
    N
    o
    si
    g
    n
    ifi
    ca
    n

    t
    e
    ff

    e
    ct
    o
    n
    g
    ro
    w
    th

    ra
    te

    o
    f

    e
    it

    h
    e
    r

    o
    rg
    a
    n

    is
    m

    o
    f
    in
    cr
    e
    a
    se

    d
    C

    O
    2
    a
    t
    a
    g

    i

    v

    e
    n

    te
    m

    p
    e
    ra

    tu
    re
    ,
    o
    r
    o

    n
    o

    rg
    a
    n

    ic
    C

    p
    e
    r

    ce
    ll

    in
    P

    ro
    ch

    lo
    ro

    co
    cc

    u
    s.

    In
    cr

    e
    a
    se

    d
    o

    rg
    a
    n
    ic
    C
    p
    e
    r
    ce
    ll

    w
    it

    h
    in

    cr
    e
    a
    se
    d
    C
    O
    2

    in
    S
    yn

    ec
    h
    oc

    cu
    s.

    F
    u

    e
    t
    a
    l.
    (2
    0

    0
    7

    )

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1239

    T
    a
    b
    l
    e
    1
    .
    C
    o

    n
    ti

    n
    u
    e
    d
    .
    S
    p
    e
    ci
    e
    s
    ⁄s
    ys
    te
    m
    M
    e
    th
    o
    d
    s
    T
    im
    e
    fr
    a
    m
    e
    p
    H
    ⁄p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    R
    e
    p
    li
    ca
    ti
    o
    n
    R
    e
    su
    lt
    s
    A
    u

    th
    o

    rs
    P
    h
    yt
    o
    p
    la
    n
    k
    to
    n
    a
    ss
    e
    m
    b
    la
    g
    e
    s,

    R
    o

    ss
    S

    e
    a
    ,

    A
    n

    ta
    rc

    ti
    ca

    C
    O
    2
    b
    u
    b
    b
    li
    n
    g

    ,
    sh

    ip
    b

    o
    a
    rd

    ,
    se
    m
    ic
    o
    n
    ti
    n
    u
    o
    u
    s
    b
    a
    tc
    h
    cu
    lt
    u
    re
    ,
    3
    0

    %
    a
    m

    b
    ie

    n
    t
    ir
    ra
    d
    ia
    n
    ce

    ,
    0

    �C
    ,

    n
    a
    tu
    ra
    l

    le
    ve

    ls
    o

    f
    m

    a
    cr

    o
    n
    u
    tr
    ie
    n
    ts
    ,

    1
    n

    M
    F

    e
    C

    l 3
    .

    E
    x
    p
    e
    ri
    m
    e
    n
    ts
    la
    st
    e
    d
    1
    0
    o
    r

    1
    8

    d
    .
    p
    C
    O
    2

    =
    1

    0
    0

    ,
    3

    8
    0

    ,
    a
    n

    d
    8

    0
    0
    p
    p
    m
    n
    =
    3
    N
    o

    v.

    D

    e
    c.

    e
    x
    p
    e
    ri
    m
    e
    n

    t:
    1
    4
    C

    u
    p

    ta
    k

    e
    (2

    4
    h

    )
    a
    n

    d
    g
    ro
    w
    th
    ra

    te
    s

    h
    ig
    h
    e
    r

    in
    3

    8
    0
    a
    n
    d
    8
    0
    0
    tr
    e
    a
    tm
    e
    n

    ts
    co

    m
    p
    a
    re

    d
    to

    1
    0
    0
    p
    p
    m

    .
    Ja

    n
    .

    e
    x
    p
    e
    ri
    m
    e
    n
    t:
    1
    4
    C
    u
    p
    ta
    k

    e
    in

    cr
    e
    a
    se
    s
    li

    n
    e

    a
    rl

    y
    w

    it
    h
    in
    cr

    e
    a
    si

    n
    g
    p
    C

    O
    2
    ,

    a
    n
    d
    g
    ro
    w
    th
    ra

    te
    h

    i

    g
    h

    e
    r

    in
    8

    0
    0
    p
    p

    m
    co

    m
    p
    a
    re
    d
    to
    1
    0
    0
    a
    n

    d
    3

    8
    0
    p
    p

    m
    .

    T
    o
    rt
    e
    ll
    e
    t
    a
    l.
    (2
    0

    0
    8

    )

    M
    a
    c
    ro

    a
    lg
    a
    e

    L
    om

    e

    n
    ta

    ri
    a

    a
    rt

    ic
    u

    la
    ta

    B
    u

    b
    b

    li
    n

    g
    w

    it
    h
    C
    O

    2
    ⁄O

    2
    ⁄N

    2
    m

    ix
    ,

    0
    .0

    6
    L

    cu
    lt
    u
    re

    ve
    ss

    e
    l,

    4
    0


    5

    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ

    s)
    1
    ,

    1
    2
    :1
    2
    L
    😀
    ,
    1
    0
    �C
    ,
    P

    ro
    va

    s

    o
    li

    e
    n

    ri
    ch

    e
    d
    se
    a
    w
    a
    te
    r.

    1
    -w

    e
    e
    k
    a
    cc
    li
    m
    a
    ti
    o
    n
    ,
    3

    -w
    e
    e
    k

    e
    x
    p
    e
    ri
    m
    e
    n

    t.

    p

    C

    O
    2

    =
    2

    3
    5

    ,
    3
    5
    0

    ,
    7

    0
    0
    a
    n

    d
    1

    ,

    7
    5

    0
    p
    p
    m
    n
    =

    4
    5

    2
    %

    in
    cr
    e
    a
    se
    in
    g
    ro
    w
    th
    ra

    te
    .

    K

    b
    le

    r
    e
    t

    a
    l.
    (1
    9
    9
    9
    )

    P
    or

    p
    h

    yr

    a
    li

    n
    ea

    ri
    s

    In
    je

    ct
    io
    n
    o
    f
    C
    O
    2

    g

    a
    s

    u
    si

    n
    g

    p
    H

    co
    n

    tr
    o

    ll
    e
    r

    a
    n

    d
    T

    A
    P

    S
    b

    u
    ff

    e
    r

    fo
    r

    h
    ig

    h
    p

    H
    ,

    0
    .5

    L
    cu

    lt
    u

    re
    ve

    ss
    e
    l,

    7
    0

    l
    m

    o
    l

    p
    h

    o
    to

    n
    Æ

    m
    )
    2
    Æ
    s)
    1
    ,
    1
    0

    :1
    4

    L
    😀
    ,
    1

    5
    �C

    ,
    P
    ro
    va
    so
    li

    E
    n

    ri
    ch
    m
    e
    n

    t
    M

    e
    d

    ia
    p

    lu
    s

    M
    e
    ta

    l
    M

    ix
    II

    .
    1
    -w
    e
    e
    k
    a
    cc
    li
    m
    a
    ti
    o
    n
    ,
    2
    -w
    e
    e
    k
    e
    x
    p
    e
    ri
    m
    e
    n
    t.
    p

    H
    =

    6
    .0

    ,
    6

    .5
    ,

    7
    .0

    ,
    8

    .7
    N

    o
    t

    cl
    e
    a
    r

    N
    o
    cl
    e
    a
    r

    d
    if

    fe
    re

    n
    ce

    s
    w

    it
    h

    in
    re

    a
    li

    st
    ic

    ra
    n

    g
    e

    (7
    .5


    8

    ).

    Is
    ra

    e
    l

    e
    t
    a
    l.
    (1
    9
    9
    9
    )
    C
    o

    n
    st

    a
    n

    t
    a
    e
    ra

    ti
    o

    n
    w

    it
    h
    C
    O

    2
    ⁄a

    ir
    m

    ix
    ,
    4
    0
    L
    cu
    lt
    u

    re
    ta

    n
    k

    s,
    o

    th
    e
    r

    co
    n
    d
    it
    io
    n

    s
    n

    o
    t

    st
    a
    te

    d
    .
    3
    w
    e
    e
    k
    s
    C
    o
    n
    tr
    o

    l
    =

    3
    6

    0
    m

    g
    Æ

    L
    )

    1

    T
    re

    a
    tm
    e
    n

    t
    =

    7
    5
    0
    m
    g
    Æ
    L
    )
    1
    n
    =
    3
    5

    0
    %

    re
    d

    u
    ct
    io
    n
    in
    g
    ro
    w
    th
    a
    t
    h
    ig
    h
    C
    O
    2
    .

    1240 C A T R I O N A L . H U R D E T A L .

    T
    a
    b
    l
    e
    1
    .
    C
    o
    n
    ti
    n
    u
    e
    d
    .
    S
    p
    e
    ci
    e
    s
    ⁄s
    ys
    te
    m
    M
    e
    th
    o
    d
    s
    T
    im
    e
    fr
    a
    m
    e
    p
    H
    ⁄p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    R
    e
    p
    li
    ca
    ti
    o
    n
    R
    e
    su
    lt
    s
    A
    u
    th
    o
    rs
    1
    4

    g
    re

    e
    n

    ,
    re

    d
    ,

    a
    n
    d
    b
    ro
    w

    n
    m

    a
    cr

    o
    a
    lg

    a
    e
    C
    o
    n
    st
    a
    n
    t
    a
    e
    ra
    ti
    o
    n
    w
    it
    h
    C
    O
    2
    ⁄a
    ir
    m
    ix
    ,
    4
    0
    L
    cu
    lt
    u
    re
    ta
    n
    k

    s,
    0

    .2
    m

    M
    p

    h
    o

    sp
    h

    a
    te
    ,
    2

    m
    M

    a
    m
    m
    o

    n
    iu

    m
    .

    4


    1

    5
    w

    e
    e
    k
    s
    C
    o
    n
    tr
    o
    l
    =
    3
    6
    0
    l
    m
    o
    l
    Æ
    m
    o

    l)
    1

    T
    re
    a
    tm
    e
    n
    t
    =
    7
    5
    0
    l
    m
    o
    l
    Æ
    m
    o
    l)
    1
    N
    o
    t
    cl
    e
    a
    r
    N
    o
    cl
    e
    a
    r
    e
    ff
    e
    ct
    o
    f

    e
    le

    va
    te

    d
    C
    O
    2
    o
    n
    g
    ro
    w
    th
    o
    r
    p
    h
    o
    to

    sy
    n

    th

    e
    si

    s.
    L

    o
    w
    p
    H

    m
    a
    y

    h
    a
    ve

    d
    a
    m

    a
    g
    e
    d

    so
    m

    e
    sp

    e
    ci

    e
    s.

    Is
    ra
    e
    l
    a
    n

    d
    H

    o
    p

    h
    y

    (2
    0
    0
    2
    )

    P
    u

    re
    C

    O
    2

    in
    je

    ct
    e
    d

    a
    t
    th
    e

    b
    a
    se

    m

    e
    sh

    b
    a
    g

    s
    co
    n
    ta

    i

    n
    in

    g
    se

    a
    w

    e
    e
    d

    s
    in

    si
    tu

    ,
    a
    ve

    ra
    g

    e
    1

    ,

    4
    9

    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,

    2
    8

    �C
    ,
    a
    m
    b
    ie
    n
    t
    n
    u
    tr
    ie
    n
    ts
    .

    5
    d

    C
    o

    n
    tr

    o
    l
    p
    H

    =
    8

    .1
    8


    8

    .1
    4

    T
    re
    a
    tm
    e
    n

    t
    p

    H
    =

    7
    .4


    7

    .6

    3

    N
    o
    t
    cl
    e
    a
    r
    N
    o
    cl
    e
    a
    r
    e
    ff
    e
    ct
    s
    o

    f
    e
    le

    va
    te
    d
    C
    O
    2
    .

    S
    a
    cc

    h

    a
    ri

    n
    a
    la
    ti

    ss
    im

    a
    a
    n

    d
    N

    er
    eo

    cy
    st

    is
    lu

    et
    ke

    a
    n
    a
    C
    o
    n
    st
    a
    n
    t
    a
    e
    ra
    ti
    o
    n
    w
    it
    h
    C
    O
    2
    ⁄a
    ir
    m

    ix
    (3

    ,0
    0

    0
    p
    p
    m

    )
    d

    ir
    e
    ct

    ly
    in

    o
    u

    td
    o

    o
    r
    7
    5

    0
    L

    ta
    n

    k
    s,

    m
    a
    x

    .
    2

    ,2
    0

    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,
    1
    1

    �C

    1

    5
    �C

    ,
    a
    m

    b
    ie
    n
    t
    n
    u
    tr
    ie
    n
    ts
    .

    5
    5

    d
    C
    o
    n
    tr
    o
    l
    p

    H

    8
    .0

    2

    8

    .1
    6

    ,
    tr

    e
    a
    tm
    e
    n

    t

    7
    .1

    7

    7

    .9
    9

    a
    t

    m
    id

    d
    a
    y.

    p
    H

    co
    u

    ld
    b

    e
    a
    s

    m
    u

    ch
    a
    s

    0
    .8

    u
    n

    it
    s

    lo
    w

    e
    r
    a
    t

    d
    a
    w

    n
    .

    P
    se

    u
    d

    o
    -r

    e
    p
    li
    ca

    te
    d

    3
    ·

    in
    cr
    e
    a
    se
    d
    g
    ro
    w

    th
    in

    S
    .

    la
    tt

    is
    m
    a
    ,

    4
    ·

    d

    e
    cr

    e
    a
    se

    in
    N

    .
    lu

    et
    ke
    a
    n

    a
    .

    S
    w

    a
    n

    so
    n

    a
    n

    d
    F

    o
    x

    (2
    0
    0
    7
    )

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1241

    Calcification also results in CO2 production (see
    eq. 6).

    Biological responses of algae to pH manipulation: The
    goal of most experiments investigating ocean acidifi-
    cation is to increase the concentration of H2CO3

    *

    and decrease the pH in seawater to mimic the
    increase in H2CO3

    * predicted to occur as the ocean
    takes up anthropogenic carbon and examine the
    effects (physiological, ecological, biogeochemical)
    of such manipulations on algae. Two techniques
    have been used to manipulate seawater pH in the
    majority of biological perturbation experiments:
    CO2 bubbling and HCl ⁄ NaOH additions. These
    have been used in a variety of experimental setups
    over various incubation timescales (Tables 1 and 2).

    Across the range of calcareous and noncalcareous
    algae tested, there are no clear patterns regarding
    the responses of primary production, growth, or cal-
    cification rates to ocean acidification (Tables 1 and
    2). The growth rate of some species is unchanged
    by altered CO2 treatments—for example, Thalassios-
    ira pseudonana (Pruder and Bolton 1980), four dia-
    toms and one dinoflagellate (Burkhardt et al.
    1999), the coccolithophores Calcidiscus leptoporus
    and Coccolithus pelagica (Langer et al. 2006), and
    Emiliania huxleyi (Feng et al. 2008). For T. pseudo-
    nana, the lack of change in growth rate following
    CO2 treatment is consistent with inorganic carbon
    concentrations being saturating for growth (Clark
    and Flynn 2000). Species that responded to
    CO2 ⁄ pH treatments include E. huxleyi, in which
    increased H2CO3

    * concentrations resulted in
    increased organic carbon content per cell, but no
    increase in the number of cells (Leonardos and
    Geider 2005, Iglesias-Rodriguez et al. 2008a).
    Growth rates of Antarctic phytoplankton assem-
    blages were also affected by pCO2, but the response
    to treatments varied depending on the time of year
    when the experiment was conducted (Tortell et al.
    2008). The marine diazotrophic cyanobacterium
    Trichodesmium also demonstrated significant
    increases in the rate of carbon assimilation (and of
    diazotrophic nitrogen assimilation) with substantial
    increases in CO2 concentration, in each of three
    studies (Barcelos e Ramos et al. 2007, Hutchins
    et al. 2007, Levitan et al. 2007). Similar results were
    reported for the unicellular marine diazotrophic
    cyanobacterium Crocosphaera under iron-sufficient,
    but not iron-limiting conditions (Fu et al. 2008).

    For calcareous E. huxleyi and Ca. leptoporus, there
    was a decrease in rates of calcification and ⁄ or mal-
    formed coccoliths at pCO2 values greater (or lower)
    than the present day; however, for Co. pelagica, there
    was no effect of pCO2 treatment on lith formation
    (Riebesell et al. 2000, Langer et al. 2006). Interest-
    ingly, when nanofossil records from cores from the
    last glacial maximum (�18,000 years ago, atmo-
    spheric CO2 180–200 lmol Æ mol

    )1 total gas) were
    examined, there was no evidence of malformed or
    incomplete liths for Ca. leptoporus or Co. pelagica;

    Langer et al. (2006) suggest adaptation (see below
    for definition) by these species to the pCO2 environ-
    ment they inhabit. Iglesias-Rodriguez et al. (2008a)
    found no decrease in calcification or lith malforma-
    tion in E. huxleyi grown at pCO2 higher than the
    present day.

    Relatively few studies have determined the likely
    impact of elevated CO2 on calcareous and noncal-
    careous macroalgae. Earlier works (unrelated to
    ocean acidification) used manipulations of seawater
    pH and carbon chemistry to unravel mechanisms of
    carbon acquisition and calcification (Smith and
    Roth 1979, Borowitzka 1981, Gao et al. 1993). As for
    phytoplankton, there are a wide range of responses
    by macroalgae to elevated CO2 concentration. A
    52% increase in growth in response to a doubling
    of pCO2 was observed for Lomentaria, a species that
    uses only CO2 (Kübler et al. 1999). This increase in
    growth rate is consistent with the idea that algae
    without CCMs are likely to respond to increased
    pCO2. Some macroalgal studies have suggested a
    negative effect of acidification on particular species,
    while other species show positive or no response to
    elevated CO2 (Israel et al. 1999, Israel and Hophy
    2002, Swanson and Fox 2007). Tropical macroalgal
    assemblages have shown positive influences of ele-
    vated CO2 on recruitment of noncalcifying macroal-
    gae, while inhibiting recruitment of corallines
    (Kuffner et al. 2008). It is not clear if these differ-
    ences in response were due to reduced survivorship
    or competitive ability of calcifying recruits, and ⁄ or
    increased competitive ability of noncalcifying algae,
    or other factor(s) (Kuffner et al. 2008). Hall-Spen-
    cer et al. (2008) demonstrated that within 120 m of
    cold CO2 vents (average pH 7.83), macroalgal com-
    munities are dominated by fleshy seaweeds, whereas
    calcareous seaweeds dominated farther from the
    vent (average pH 8.14).

    Methods: an appraisal. There are clearly a range
    of biological responses to pH manipulation treat-
    ments. This observed spectrum of responses may be
    due to inherent differences in algal physiology, the
    environment in which the algae have grown prior to
    experiments (e.g., light climate), the timescale of
    the physiological response measured (e.g., short-
    term estimates of photosynthesis vs. integrated
    growth), duration of the study (days vs. months), or
    time of year (Tortell et al. 2008). Another key influ-
    ential factor could be the method of pH manipula-
    tion because CO2 bubbling and HCl ⁄ NaOH affect
    carbonate chemistry differently (see below); there
    have been heated discussions on which method is
    most suitable (Iglesias-Rodriguez et al. 2008a,b,
    Riebesell et al. 2008). Critically, it is difficult to tease
    apart the relative importance of each of these
    potentially influential factors on the outcome of
    experiments. Here, we focus on the different ways
    in which carbonate chemistry is altered during pH
    manipulation experiments and how this might

    1242 C A T R I O N A L . H U R D E T A L .

    T
    a
    b
    l
    e
    2
    .

    C
    a
    lc

    a
    re
    o
    u

    s
    a
    lg

    a
    e
    .

    E
    x
    a
    m
    p
    le
    s
    o

    f
    e
    x

    p
    e
    ri
    m
    e
    n
    ts
    co
    n
    d
    u
    ct
    e
    d
    o
    n
    e
    ff
    e
    ct
    s
    o
    f
    C
    O
    2
    co
    n
    ce
    n
    tr
    a
    ti
    o
    n
    o
    n
    a
    lg
    a
    l
    g
    ro
    w
    th

    ,
    ca

    lc
    ifi

    ca
    ti

    o
    n
    ,
    o
    r
    m
    e
    ta
    b
    o

    li
    c

    ra
    te

    s.
    U

    n
    it
    s
    fr
    o
    m
    th
    e
    o
    ri
    g
    in
    a
    l
    p
    a
    p

    e
    rs

    a
    re

    re
    p

    o
    rt
    e
    d
    .
    S
    p
    e
    ci
    e
    s
    ⁄s
    ys
    te
    m
    M
    e
    th
    o
    d
    s
    T
    im
    e
    fr
    a
    m
    e
    p
    H
    ⁄p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    R
    e
    p
    li
    ca
    ti
    o
    n
    R
    e
    su
    lt
    s
    A
    u
    th
    o
    rs
    P
    h
    yt
    o
    p
    la
    n
    k
    to
    n
    C
    o

    cc
    o

    li
    th

    o
    p
    h
    o

    re
    s

    E
    m
    il
    ia
    n
    ia
    h
    u
    xl
    ey

    i
    a
    n

    d
    G

    ep
    h
    yr

    oc
    a
    p
    sa

    oc
    ea

    n
    ic
    a
    H
    C
    l⁄
    N
    a
    O
    H
    ,
    d
    il
    u
    te
    b
    a
    tc
    h
    cu
    lt
    u
    re
    ,
    3
    0

    1
    5
    0
    l
    m
    o
    l
    p
    h
    o
    to

    n
    s

    Æ
    m

    )
    2

    Æ
    s)

    1
    ,
    1
    6
    :8
    o
    r
    2
    4
    :0
    L
    😀
    ,
    1
    5
    �C
    ,
    1
    0
    0
    l
    M
    n
    it
    ra
    te
    ,
    6

    .5
    l

    M
    p
    h
    o
    sp
    h
    a
    te
    ,
    f
    ⁄2

    tr
    a
    ce

    e
    le
    m
    e
    n
    ts
    .

    A
    cc

    li
    m
    a
    ti
    o
    n
    fo
    r

    7


    9

    g
    e
    n

    e
    ra
    ti
    o
    n
    s
    th
    e

    8
    ce

    ll
    d

    iv
    is

    io
    n

    s
    d

    u
    ri

    n
    g
    e
    x
    p
    e
    ri
    m
    e
    n
    t.
    p
    C
    O
    2
    =
    2
    8
    0

    7
    5
    0

    .
    n

    =
    3

    R
    e
    d

    u
    ce

    d
    ca

    lc
    ifi
    ca
    ti
    o
    n
    o
    f
    1
    5

    .7
    %

    in
    E

    .
    h
    u

    xl
    ey
    i
    a
    n

    d
    4

    4
    .7

    %
    in

    G
    .

    oc
    ea
    n
    ic
    a
    .

    M
    a
    lf

    o
    rm

    e
    d
    co
    cc
    o
    li

    th
    s,

    sl
    ig

    h
    t

    in
    cr
    e
    a
    se

    in
    p

    h
    o

    to
    sy

    n
    th

    e
    si
    s
    in
    b
    o

    th
    .

    R
    ie

    b
    e
    se

    ll
    e
    t

    a
    l.
    (2
    0
    0
    0
    )
    N
    o

    rt
    h

    P
    a
    ci
    fi
    c
    p
    h
    yt
    o
    p
    la
    n
    k
    to
    n
    a
    ss
    e
    m
    b
    la

    g
    e
    s

    B
    u
    b
    b
    li
    n
    g
    w
    it
    h
    C
    O

    2
    a
    ir

    m
    ix

    tu
    re
    s
    o

    r
    H

    C
    l⁄

    N
    a
    O

    H
    ,

    sh
    ip

    b
    o
    a
    rd
    ,
    3
    0
    %

    in
    ci

    d
    e
    n

    t
    ir

    ra
    d

    ia
    n

    ce
    ,

    1
    3
    �C
    ,
    n
    u
    tr
    ie
    n
    t
    co
    n
    ce
    n
    tr
    a
    ti
    o
    n
    s
    n
    o
    t
    re
    p
    o
    rt
    e
    d
    .

    1
    .5


    9

    d
    p

    C

    O
    2
    =
    2
    5
    0

    ,
    p

    H
    =

    8
    .2

    a
    n

    d
    p
    C

    O
    2
    =
    8
    0
    0
    ,
    p

    H
    7

    .7
    5

    n
    =

    3
    S

    ig
    n

    ifi
    ca
    n
    t
    re
    d
    u
    ct
    io
    n

    s
    (3

    6
    %


    8

    3
    %

    )
    in

    ca
    lc

    ifi
    ca
    ti
    o
    n
    in

    4
    in

    d
    e
    p

    e
    n
    d
    e
    n

    t
    e
    x

    p
    e
    ri
    m
    e
    n
    ts
    .
    N
    o
    rt
    h

    A
    tl

    a
    n

    ti
    c

    p
    h
    yt
    o
    p
    la
    n
    k
    to
    n
    a
    ss
    e
    m
    b
    la
    g
    e
    s
    fr
    o

    m
    su

    rf
    a
    ce
    a
    n

    d
    d

    e
    p
    th
    H
    C
    l⁄
    N
    a
    O
    H
    ,
    sh
    ip
    b
    o
    a
    rd
    ,
    3
    5
    0
    a
    n

    d
    5

    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,
    a
    m
    b
    ie
    n
    t
    te
    m
    p
    e
    ra
    tu
    re
    ,
    n
    u
    tr
    ie
    n

    t
    co

    n
    ce
    n
    tr
    a
    ti
    o
    n
    s
    n
    o
    t
    re
    p
    o
    rt
    e
    d
    .

    2
    h

    in
    cu

    b
    a
    ti

    o
    n

    3
    ,

    1
    0
    ,
    3
    6
    ,
    a
    n

    d
    9

    1
    l

    M
    C

    O
    2
    n
    =
    3
    S
    ig
    n
    ifi
    ca
    n
    t
    in
    cr
    e
    a
    se

    in
    1
    4
    C

    in
    co

    rp
    o

    ra
    ti

    o
    n
    w
    it

    h
    3

    6
    a
    n

    d
    9
    1
    l

    M
    tr

    e
    a
    tm
    e
    n
    t
    co
    m
    p
    a
    re

    d
    w

    it
    h
    1
    0
    l
    M

    .
    D

    e
    cr
    e
    a
    se
    fo
    r

    3
    l

    M
    tr
    e
    a
    tm
    e
    n
    t.

    H
    e
    in

    a
    n

    d
    S

    a
    n

    d
    -J

    e
    n

    se
    n

    (1
    9

    9
    7

    )
    E
    m
    il
    ia
    n
    ia
    h
    u
    xl
    ey

    i
    (c

    a
    lc
    if
    yi
    n
    g

    st
    ra

    in
    )
    C
    O
    2
    b
    u
    b
    b
    li
    n
    g
    ,
    se
    m
    ic
    o
    n
    ti
    n
    u
    o
    u
    s
    cu
    lt
    u
    re

    ,
    5

    0
    a
    n
    d
    4
    0
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,
    2
    0

    �C
    a
    n

    d
    2
    4
    �C
    ,
    f
    ⁄2
    n
    u
    tr
    ie
    n
    ts
    .
    A
    cc
    li
    m
    a
    ti
    o
    n

    in
    tr

    e
    a
    tm
    e
    n
    t
    co
    n
    d

    it
    io

    n
    s
    fo
    r
    7
    g
    e
    n
    e
    ra
    ti
    o

    n
    s.

    p
    C
    O
    2
    =
    3

    5
    7

    a
    n
    d
    7
    5
    0
    p
    p
    m
    n
    =
    3
    N
    o
    e
    ff
    e
    ct
    o
    f
    p
    C
    O
    2
    o
    n
    g
    ro
    w
    th
    ra
    te
    o
    r
    o
    rg
    a
    n
    ic
    C
    p
    e
    r
    ce
    ll
    a
    t

    g
    iv

    e
    n
    ir
    ra
    d
    ia
    n
    ce
    a
    n

    d
    te

    m
    p
    e
    ra
    tu
    re
    tr
    e
    a
    tm
    e
    n

    t.
    P

    IC
    w

    a
    s
    re
    d
    u
    ce

    d
    in

    th
    e
    h
    ig

    h
    li

    g
    h

    t
    ⁄7

    5
    0
    p
    p
    m
    tr
    e
    a
    tm
    e
    n
    t.

    F
    e
    n

    g
    e
    t

    a
    l.
    (2
    0
    0
    8
    )
    C
    a
    lc

    id
    is

    cu
    s

    le

    p
    to

    p
    or

    u
    s
    a
    n
    d
    C

    oc
    cl

    it
    h
    u

    s

    p
    el

    a
    gi

    cu
    s
    H
    C
    l⁄
    N
    a
    O
    H
    ,
    d
    il
    u
    te
    b
    a
    tc
    h
    cu
    lt
    u
    re
    s,
    3
    5
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,
    1
    6

    :8
    L

    😀
    ,
    2
    0

    �C
    fo

    r
    C

    .
    le

    p
    to
    p
    or

    u
    s,

    1
    7

    �C
    fo
    r
    C

    .
    p
    el

    a
    gi

    ca
    ,

    1
    0
    0
    l

    M
    n

    it
    ra

    te
    ,

    6
    .5

    l
    M
    p
    h

    o
    sp

    h
    a
    te

    ,
    f
    ⁄2
    tr
    a
    ce
    e
    le
    m
    e
    n
    ts
    .
    A
    cc
    li
    m
    a
    ti
    o
    n
    in
    tr
    e
    a
    tm
    e
    n
    t
    co
    n
    d
    it
    io
    n
    s
    fo
    r
    1
    0
    g
    e
    n
    e
    ra
    ti
    o
    n
    s.

    C
    .

    le
    p
    to

    p
    or

    u
    s:

    6
    p
    C

    O
    2
    tr
    e
    a
    tm
    e
    n

    ts
    fr

    o
    m
    9
    8

    9
    2
    0

    l
    a
    tm

    .
    C

    .
    p
    el
    a
    gi

    ca
    :

    3
    tr

    e
    a
    tm
    e
    n

    ts
    1

    4
    9
    ,
    3
    4
    5
    ,
    a
    n
    d
    9
    1
    5
    l
    a
    tm
    .
    N
    o
    t
    cl
    e
    a
    r
    C
    .
    le
    p
    to
    p
    or
    u
    s:

    m
    a
    lf

    o
    rm
    e
    d
    co
    cc
    o
    li

    th
    s

    a
    t
    lo
    w
    a
    n

    d
    h

    ig
    h

    p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    co
    m
    p
    a
    re
    d
    to
    co
    n
    tr
    o

    l
    (3

    4
    5
    p
    p

    m
    ).

    C
    .
    p
    el
    a
    gi

    cu
    s:

    n
    o
    ch
    a
    n
    g
    e

    i

    n
    li

    th
    m

    o
    rp

    h
    o

    lo
    g

    y.
    B

    o
    th

    sp
    e
    ci

    e
    s:

    n
    o
    e
    ff
    e
    ct
    o
    f
    C
    O
    2
    o
    n
    p
    o
    p
    u
    la
    ti
    o
    n
    g
    ro
    w
    th
    ra

    te
    b

    u
    t

    in
    cr
    e
    a
    se
    d
    o
    rg
    a
    n

    ic
    co

    n
    te

    n
    t
    o
    f
    ce
    ll

    s
    u

    n
    d
    e
    r
    h
    ig
    h
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    .

    L
    a
    n

    g
    e
    r

    e
    t
    a
    l.
    (2
    0

    0
    6

    )
    E
    m
    il
    ia
    n
    ia
    h
    u
    xl
    ey

    i
    C

    O
    2
    b
    u
    b
    b
    li
    n

    g
    ,

    d
    il
    u
    te
    b
    a
    tc
    h
    cu
    lt
    u
    re
    s,
    1
    5
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,
    1
    2
    :1
    2
    L
    😀
    ,
    1

    9
    �C

    ,
    f
    ⁄2

    m
    e
    d

    iu
    m

    .

    A
    lg

    a
    e
    a
    cc
    li
    m
    a
    te

    d
    a
    t

    e
    x
    p
    e
    ri
    m
    e
    n
    ta
    l
    co
    n
    d
    it
    io
    n

    s
    fo

    r
    9

    g
    e
    n
    e
    ra
    ti
    o
    n
    s.
    C
    O

    2
    =

    2
    8

    0
    ,

    3
    0
    0
    ,
    4
    9
    0
    ,
    6
    0
    0
    ,
    7
    5
    0
    p
    p
    m
    v
    n
    =

    6
    D

    o
    u
    b
    li
    n
    g
    in
    p
    a
    rt
    ic
    u
    la
    te

    in
    o

    rg
    a
    n
    ic
    ca
    rb
    o

    n
    a
    n

    d
    p
    a
    rt
    ic
    u
    la
    te
    o
    rg
    a
    n
    ic
    ca
    rb
    o

    n
    ,

    a
    n
    d
    d
    e
    cr
    e
    a
    se
    in
    g
    ro
    w
    th
    ra
    te
    s
    a
    t
    7
    5

    0
    co

    m
    p
    a
    re
    d
    to
    2
    8
    0
    p
    p
    m

    v.
    N

    o
    e
    ff

    e
    ct
    o
    f
    d
    o
    u
    b

    le
    d

    C
    O
    2
    o
    n
    li
    th
    m
    o
    rp
    h
    o
    lo
    g
    y.

    Ig
    le

    si
    a
    s-

    R
    o

    d
    ri

    g
    u

    e
    z

    e
    t
    a
    l.
    (2
    0
    0
    8

    a
    )

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1243

    T
    a
    b
    l
    e
    2
    .
    C
    o
    n
    ti
    n
    u
    e
    d
    .
    S
    p
    e
    ci
    e
    s
    ⁄s
    ys
    te
    m
    M
    e
    th
    o
    d
    s
    T
    im
    e
    fr
    a
    m
    e
    p
    H
    ⁄p
    C
    O
    2
    tr
    e
    a
    tm
    e
    n
    ts
    R
    e
    p
    li
    ca
    ti
    o
    n
    R
    e
    su
    lt
    s
    A
    u
    th
    o
    rs
    M
    a
    c
    ro
    a
    lg
    a
    e

    C
    or

    a
    ll

    in
    a

    p
    il

    u
    li

    fe
    ra

    C
    o
    n
    st
    a
    n
    t
    a
    e
    ra
    ti
    o
    n
    w
    it
    h
    C
    O
    2
    ⁄a
    ir
    m
    ix
    ,
    1
    5

    cm
    h

    ig
    h
    w
    it

    h
    1

    4
    cm

    d
    ia

    m
    e
    te

    r
    2
    0
    �C
    ,
    3
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ

    s)
    1
    .

    1
    2

    h
    p
    C

    O
    2
    =
    3
    5
    0
    ,
    p

    H
    8

    .2
    a
    n

    d
    p
    C
    O
    2

    1
    ,5

    0
    0
    a
    n
    d
    p
    H
    7
    .6
    P
    se
    u
    d

    o

    re
    p
    li
    ca
    te
    d

    4
    4

    %
    re

    d
    u
    ct
    io
    n
    in
    ca
    lc
    ifi
    ca
    ti
    o
    n
    .

    G
    a
    o

    e
    t
    a
    l.
    (1
    9

    9
    3

    )

    C
    ru

    st
    o

    se
    C

    or
    a
    ll

    in
    e

    re
    cr

    u
    it

    s
    In
    je
    ct
    io
    n
    o

    f
    1

    0
    %
    H
    C

    l⁄
    fr

    e
    sh

    w
    a
    te

    r
    m

    ix
    a
    t

    in
    fl

    o
    w

    to
    5

    0
    0

    L
    o

    u
    td

    o
    o

    r
    m

    e
    so

    co
    sm

    s,
    a
    m

    b
    ie
    n
    t
    ir
    ra
    d
    ia
    n
    ce
    ,
    se
    a
    w
    a
    te
    r
    te
    m
    p
    e
    ra
    tu
    re
    ,
    a
    n
    d
    n
    u
    tr
    ie
    n
    ts
    .

    7
    w

    e
    e
    k
    s
    C
    o
    n
    tr
    o

    l
    p
    C

    O
    2

    =
    4

    0
    0
    l
    a
    tm
    ,
    p
    H
    8

    .1
    7

    a
    n

    d
    tr

    e
    a
    tm
    e
    n

    t
    p
    C

    O
    2

    =
    7

    6
    5

    l
    a
    tm
    ,
    p
    H
    7

    .9
    1

    a
    t
    m
    id
    d
    a
    y.

    D
    iu

    rn
    a
    l

    fl
    u

    ct
    u

    a
    ti
    o
    n
    s
    o

    f
    p

    H
    o

    f

    0
    .5
    u
    n

    it
    s.

    n
    =

    3
    M

    a
    rk
    e
    d
    re
    d
    u
    ct
    io
    n
    in
    re

    cr
    u

    it
    m

    e
    n

    t
    a
    n

    d
    %

    co
    ve

    r
    o

    f
    cr

    u
    st

    o
    se

    c

    o
    ra

    ll
    in

    e
    s

    in
    h

    ig
    h


    C

    O
    2
    tr
    e
    a
    tm
    e
    n

    t.
    In

    cr
    e
    a
    se

    in
    n

    o
    n
    ca
    lc
    if
    yi
    n
    g

    m
    a
    cr

    o
    a
    lg
    a
    e
    .

    K
    u

    ff
    n

    e
    r
    e
    t
    a
    l.
    (2
    0
    0
    8
    )
    P
    or

    ol
    it

    h
    on

    on
    ko

    d
    es

    C
    O
    2
    b
    u
    b
    b
    le
    d
    in
    to
    m

    ix
    in

    g
    ta

    n
    k

    s,
    fe

    d
    in

    to
    2

    0
    L

    a
    q

    u
    a
    ri

    a
    ,
    a
    m
    b
    ie
    n
    t
    ir
    ra
    d
    ia
    n
    ce
    7
    0

    0

    1

    ,2
    0
    0
    l
    m
    o
    l
    p
    h
    o
    to
    n
    s
    Æ
    m
    )
    2
    Æ
    s)
    1
    ,

    tw
    o

    te
    m
    p
    e
    ra
    tu
    re
    tr
    e
    a
    tm
    e
    n

    ts
    :

    2
    5

    �C


    2

    6
    �C

    a
    n
    d
    2
    8
    �C

    2
    9
    �C
    ,
    a
    m
    b
    ie
    n
    t
    n
    u
    tr
    ie
    n
    ts
    .

    8
    w

    e
    e
    k
    s
    C
    o
    n
    tr
    o
    l
    p
    H
    =
    8
    .0

    8

    .4
    (3

    8
    0
    p
    p

    m
    v)

    ;
    tr

    e
    a
    tm
    e
    n

    ts
    =

    p
    H

    7
    .8

    5

    7

    .9
    5

    (C
    O

    2
    5
    2
    0

    7
    0
    5
    p
    p
    m
    v)
    a
    n
    d
    p
    H
    7

    .6
    0


    7

    .7
    0

    (1
    ,0

    0
    0

    1

    ,3
    0

    0
    p
    p
    m

    v)
    .

    n
    =

    5
    B

    le
    a
    ch

    in
    g

    ,
    m

    a
    rk
    e
    d

    d
    e
    cr

    e
    a
    se
    in
    n
    e
    t
    p
    h
    o
    to
    sy
    n

    th
    e
    si

    s
    a
    n

    d
    ca
    lc
    ifi
    ca
    ti
    o
    n
    w
    it

    h
    d

    e
    cr
    e
    a
    si
    n
    g
    p
    H
    tr
    e
    a
    tm
    e
    n
    t.
    A
    n
    th
    o

    n
    y

    e
    t
    a
    l.
    (2
    0
    0
    8
    )

    S
    u

    b
    ti

    d
    a
    l

    a
    n
    d
    in

    te
    rt

    id
    a
    l

    co
    m

    m
    u
    n
    it

    ie
    s

    ‘‘
    N

    a
    tu
    ra
    l

    la
    b

    o
    ra

    to
    ry

    ’’
    (i

    .e
    .,

    a
    co

    ld
    -w

    a
    te

    r,
    su

    b
    m

    a
    ri
    n
    e
    C
    O
    2

    ve
    n

    t)
    .

    C
    o

    m
    m

    u
    n

    it
    y

    su
    rv

    e
    ys

    o
    f
    sp
    e
    ci
    e
    s

    p
    re

    se
    n

    ce
    a
    n

    d
    %
    co
    ve

    r
    a
    t

    va
    ri

    o
    u
    s
    d

    is
    ta

    n
    ce
    s
    fr
    o
    m
    C
    O
    2
    ve
    n
    ts
    .
    p
    H

    g
    ra

    d
    ie

    n
    t
    o
    f

    8
    .1

    4

    6

    .5
    7

    R
    e
    p

    li
    ca
    te
    d

    %
    co

    ve
    r

    o
    f
    ca
    lc
    a
    re
    o
    u

    s
    se

    a
    w
    e
    e
    d
    s
    d
    e
    cr
    e
    a
    se

    d
    cl

    o
    se

    r
    to

    ve
    n
    ts
    ,
    %
    co
    ve
    r
    n
    o
    n
    ca

    lc
    a
    re

    o
    u
    s
    se
    a
    w
    e
    e
    d
    s
    in
    cr
    e
    a
    se
    d
    .

    H
    a
    ll

    -S
    p

    e
    n

    ce
    r

    e
    t
    a
    l.
    (2
    0
    0
    8
    )

    1244 C A T R I O N A L . H U R D E T A L .

    impact the physiological response of algae. We then
    consider the steps required to design and imple-
    ment improved experiments on ocean acidification,
    and finally, we discuss the implications of a lack of
    standardization in the other environmental factors
    on making a direct comparison between the results
    from different published studies.

    Comparison of HCl additions versus CO2 bubbling on
    seawater carbonate chemistry: The most commonly
    used means to simulate the effects of acidification
    involves an initial adjustment of the CO2 speciation
    of a seawater culture medium to achieve the desired
    degree of acidification and then maintaining those
    conditions as growth of the alga(e) proceeds. There
    are two methods of decreasing seawater pH, by bub-
    bling the seawater with CO2 gas or by the addition
    of acid (commonly HCl).

    Each method produces a specific target CO2 spe-
    ciation, but they achieve this in different ways
    (Langdon 2003). The equilibration of CO2 gas with
    seawater fixes the equilibrium pCO2 to that of the
    control gas, without any change in total alkalinity,
    AT. As a result, CT will increase from its initial value.
    In contrast, addition of HCl to seawater involves a
    known decrease in the original AT, but no change
    in CT.

    Figure 1 shows simulated values of pCO2,
    [HCO3

    )], [CO3
    2)], and calcite saturation (X)

    achieved by acidifying a seawater sample of typical
    surface-water composition using each method.
    There is very little difference in the two methods
    for a given pH with respect to [CO3

    2)] and X, but
    at pH 7.5, pCO2 and HCO3

    ) are 23% and 22%
    lower, respectively, using HCl additions compared
    to CO2 bubbling. This is because when seawater is
    acidified with HCl, CT remains constant (provided

    CO2 does not escape into the gas phase), and so
    [HCO3

    )] increases slightly as a small amount of
    CO3

    2) is converted to HCO3
    ) by reaction with H+

    ions. However, when CO2(g) is added, CT is
    increased, mainly in the form of HCO3

    ).
    CO2 bubbling is arguably much closer to the

    actual ocean acidification that is currently affecting
    the surface ocean. Here, we suggest a third method
    that will achieve the same effect without bubbling
    with CO2, which is by adding, separately, equivalent
    concentrations of HCl and sodium bicarbonate solu-
    tions. In this case, the NaHCO3 exactly neutralizes
    the alkalinity decrease caused by the HCl and sup-
    plies the increased CT (mostly as HCO3

    )) that
    would be the result of CO2 gas bubbling. While this
    method has not been used in published biological
    experiments to date, the use of separate HCl and
    NaHCO3 additions has the advantage that it avoids
    any possible physiological effects that bubbling itself
    might cause (see later). The changes in CO2 specia-
    tion induced by this method are identical to those
    achieved by bubbling with CO2 gas, as shown in
    Figure 1.

    How might the method of pH manipulation affect algal
    physiology? The question is whether the different
    methods that have been used to manipulate seawa-
    ter pH could affect the outcome of experiments.
    For algae without CCMs that rely on diffusive
    uptake of CO2, the 23% difference in H2CO3

    *

    between the CO2-bubbling and the HCl-addition
    methods at pH 7.5 (and indeed a 17% difference at
    pH 7.75; Fig. 1) could cause significantly lower rates
    of photosynthesis and growth with the HCl method.
    For noncalcifying algae with CCMs, the concentra-
    tion of CO2 at the alga’s surface controls the activity
    of the CCM and carbonic anhydrase (Giordano
    et al. 2005); the 23% higher H2CO3

    * concentration
    (with CO2 bubbling) may be sufficient to cause a
    down-regulation of the CCM and CA relative to the
    HCl method. Such a down-regulation in the CCM
    and the CA could result in the reallocation of cellu-
    lar energy to, for example, more rapid growth.

    At the current oceanic pH of �8.1, HCO3)
    concentrations in seawater are sufficient to saturate
    photosynthesis in most organisms with CCMs
    (Giordano et al. 2005). Therefore, for algae with
    CCMs that utilize bicarbonate, the 22% difference
    that results from the two methods of pH manipula-
    tion should not affect photosynthesis per se because
    HCO3

    ) is not limiting. The saturation state of cal-
    cite (X) is key in determining if organisms can lay
    down calcium carbonate, and this is similar for both
    methods of pH manipulation (Fig. 1). Whether the
    22% difference in HCO3

    ) concentration between
    the two methods might influence calcification is not
    clear.

    A direct comparison of the two methods (CO2
    bubbling or HCl ⁄ NaOH additions) of CO2 manipu-
    lation on the responses of algae has not been made.
    However, the responses of calcifying strains of coc-

    Fig. 1. Calculated equilibrium CO2 parameters as a function
    of seawater pH for a seawater sample having initial
    CT = 1,900 lmol Æ kg

    )1 and AT = 2,300 lmol Æ kg
    )1, salinity 35

    and temperature 10�C, that has been acidified either with strong
    acid HCl or CO2 gas: (a) equilibrium CO2 partial pressure pCO2,
    (b) bicarbonate ion concentration [HCO3

    )], (c) carbonate ion
    concentration [CO3

    2)], and (d) saturation ratio X for calcite. It is
    assumed that the HCl addition involves insignificant dilution of
    the seawater.

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1245

    colithophores to ocean acidification have been
    tested (by different research groups and often,
    when the same species was employed, using differ-
    ent strains) using the two methods. Feng et al.
    (2008) used the CO2-bubbling method to show that
    the particulate inorganic carbon (PIC) production
    rate of E. huxleyi remains unaffected by high CO2
    concentration at low irradiances for growth, but it is
    decreased by high CO2 concentration at high
    growth irradiances. The particulate organic carbon
    (POC) per cell increased in high CO2(aq) at the
    higher growth irradiances. Iglesias-Rodriguez et al.
    (2008a) also used the CO2-bubbling method on
    E. huxleyi, working at light saturation, and deter-
    mined that while increased CO2 concentrations
    increased the POC and PIC per cell, there was no
    change in the PIC:POC ratio with increasing CO2,
    and a significant decrease in the specific growth rate
    of the cells at the highest CO2 concentration used.

    Using the HCl ⁄ NaOH method of CO2 manipula-
    tion, Riebesell et al. (2000) showed that high CO2
    concentration decreased the rate of calcification
    (on a per cell basis) in E. huxleyi and Gephyrocapsa
    oceanica, with increased calcification at lower than
    present CO2 levels. Again, using the HCl ⁄ NaOH
    technique, Langer et al. (2006) observed no
    increase in inorganic carbon precipitation by Co.
    pelagicus by increasing CO2 concentrations from that
    occurring during the last glacial maximum level to
    more than twice present-day CO2 levels, while Ca.
    leptoporus had the highest rate of calcification at the
    present CO2 level, with lower rates at more than
    twice the present-day level and an even greater
    decrease at the last glacial maximum level. Clearly,
    there are intergeneric variations in the response of
    calcification to CO2 concentrations in all experi-
    ments using the HCl ⁄ NaOH method. Whether the
    difference in response of E. huxleyi to CO2 variations
    reported by Riebesell et al. (2000), Feng et al.
    (2008), and Iglesias-Rodriguez et al. (2008a) is a
    function of the different methods of changing CO2
    or of other experimental differences or the use of dif-
    ferent algal strains awaits further experimentation.

    Methodological considerations: Here we discuss
    eight points that we consider essential in designing
    experiments that examine the impacts of ocean
    acidification on algae. Most considerations are rele-
    vant to studies involving either macroalgae or micro-
    algae.

    1. Which method of manipulating pH to choose? This
    will depend on the research question being asked.
    If the question is related to carbon acquisition and
    rates of photosynthesis and growth, then adding
    CO2 to shift seawater pH most closely mimics the
    changes in seawater carbon speciation predicted in
    the future due to climate change (changing CT,
    constant alkalinity). For phytoplankton, however,
    there may be side effects of directly bubbling with
    CO2, such as damage to fragile phytoplankton
    through the effects of small-scale turbulence with

    increasing sensitivity in the order green algae >
    cyanobacteria > diatoms > dinoflagellates (Thomas
    and Gibson 1990). The HCl method is relatively
    simple to implement and can be used for investigat-
    ing the effects of changes in pH, H2CO3

    *, or CO3
    2)

    as long as the resulting changes in AT (and lack of
    change in CT) are of no concern. In addition, the
    scale of experiment might affect the pH manipula-
    tion used. For small-scale laboratory cultures and
    shipboard experiments, pH can be manipulated
    using CO2 bubbling, but the HCl-addition tech-
    nique may be more practicable for large-scale meso-
    cosm experiments (Kuffner et al. 2008). Whichever
    method is chosen, clear interpretation of physiologi-
    cal responses of algae to pH manipulation will be
    facilitated by a thorough understanding of the
    entire carbonate system in the culture medium. To
    this end, it is essential to conduct regular monitor-
    ing of at least two of the carbonate analytical param-
    eters (see below).

    2. Methods of pH control. For CO2 bubbling, pertur-
    bations can take place using pure CO2 gas (Leclercq
    et al. 2000, Israel and Hophy 2002), CO2 ⁄ air mix-
    tures (Gao et al. 1993, Riebesell et al. 2000, Engel
    et al. 2005), or CO2 mixed with other gases (O2,
    N2) (the latter to permit an analysis of CO2 ⁄ O2
    interaction effects on growth, Kübler et al. 1999).
    CO2 ⁄ air and gas mixtures at certified concentrations
    are expensive, and using gas mixers to produce
    appropriate CO2 ⁄ gas mixtures is an option (Kübler
    et al. 1999, Engel et al. 2005). Parsons et al. (1992)
    describe how to build an inexpensive gas-mixing sys-
    tem that was used effectively to grow a macroalga at
    a range of pCO2 (Kübler et al. 1999). Careful bub-
    bling of small amounts of inexpensive food-grade
    100% CO2 while monitoring pH is also effective in
    providing required pH ⁄ CO2 concentration (C. D.
    Hepburn, K. Currie, and C. L. Hurd, unpublished
    data).

    When using HCl, it is important not to introduce
    any other perturbation to the system. The HCl
    should be made up in NaCl solution such that the
    total ionic strength is similar to that of the seawater,
    that is, 0.7 M (e.g., 0.1 M HCl and 0.6 M NaCl).
    The HCl solution should not be contaminated with
    trace metals if this is important (e.g., in iron-limited
    waters).

    Higher pH and reduced pCO2 (i.e., mimicking
    the preindustrial era) can also be achieved through
    bubbling seawater with another gas (e.g., N2), CO2-
    depleted air (produced by pumping ambient air
    through Na2CO3 traps) (Leclercq et al. 2000, Engel
    et al. 2005), or CO2-depleted gas mixtures (Kübler
    et al. 1999). Due to seawater’s strong buffering
    capacity, bubbling with CO2-free or CO2-depleted
    gases can take some time and a significant amount
    of depleted gas before the required pCO2 is
    reached. Shifting pH using concentrated NaOH is a
    faster and straightforward method, but, as for HCl,
    the resulting carbonate speciation is different from

    1246 C A T R I O N A L . H U R D E T A L .

    that which occurs in situ in response to changing
    pCO2.

    3. Monitoring pH and carbonate analytical parameters
    during experiments. Algal metabolism will alter the
    pH (and thus carbon speciation) of the incubation
    medium. It is essential to measure at least daily pH
    and one other carbonate analytical parameter of the
    incubation medium. Measurement of two of the car-
    bonate analytical parameters (pH, CT, alkalinity, or
    pCO2) allows determination of the other two param-
    eters and the concentrations of the species when
    combined with knowledge of the carbonate equilib-
    rium constants (Lewis and Wallace 1998, Hunter
    2007). pH measurements can be made either poten-
    tiometrically, using high precision and carefully
    maintained electrode ⁄ meter combinations, or opti-
    cally, using spectrophotometric measurement of a
    dye ⁄ seawater mixture (Tapp et al. 2000, Ohline
    et al. 2007). Care must be taken not to lose (or
    gain) CO2 by exchange with the atmosphere during
    the subsampling and measurement process. Careful
    calibration of the pH measurement system is
    required using appropriate buffer solutions made
    up in synthetic seawater (Dickson 1993a,b, Dickson
    et al. 2007). The defined pH value of the buffer is
    temperature-dependent, so careful temperature con-
    trol is required. The measurement of other carbon-
    ate analytical parameters is described by Dickson
    et al. (2007), and pCO2 can be measured directly
    using membrane inlet mass spectrometry (MIMS;
    Gueguen and Tortell 2008).

    4. Modification of seawater pH ⁄ pCO2 within an experi-
    mental system. Modification of pH ⁄ pCO2 should
    occur in separate (mixing) containers, not directly
    in seawater in contact with the study organisms.
    This can be achieved in flow-through systems by
    adding the amendment solution or gas to indepen-
    dent header tanks (e.g., collapsible bags to prevent
    gas exchange with headspace), to tubing (Kuffner
    et al. 2007), or in mixing chambers (Leclercq et al.
    2000) upstream of culture containers in flow-
    through systems. The bubbling of CO2 ⁄ air mixtures
    directly into culture containers is also acceptable for
    macroalgae (as long as it is at the correct pCO2 for
    the treatment) but could damage delicate phyto-
    plankton (Thomas and Gibson 1990, Berdalet et al.
    2007). Direct addition of acid or concentrated CO2
    into the culture medium surrounding experimental
    subjects makes it difficult to separate damage due to
    direct shock to the organisms from localized spikes
    of pH from the cumulative effect of altered
    pH ⁄ pCO2 due to the treatment.

    5. Biological side effects and pH range during experi-
    ments. Photosynthesis (increasing pH), respiration,
    and calcification (lowering pH) by algae can
    strongly modify the pH and carbonate chemistry of
    seawater in culture containers (Israel and Hophy
    2002). Such effects are particularly important when
    using macroalgae that are often large and have
    rapid metabolic rates, and in mesocosms where bio-

    mass levels, and hence biological activity, can be
    high. For experiments that require a constant pH,
    variation due to photosynthesis or respiration can
    be reduced by high seawater to macroalgal tissue
    ratios and ⁄ or systems that have seawater flow from
    reservoirs with fixed pH levels, and for microalgae,
    continuous and semicontinuous culture systems can
    be advantageous over batch cultures because of the
    constant replenishment with fresh media. Critically,
    care must be taken not to mistake the effects of
    short periods of unrealistic pH resulting from meta-
    bolic processes to the effect of different pCO2 and
    pH treatments that simulate acidification.

    Seawater pH naturally varies on timescales from
    diurnal to seasonal. An understanding of this natu-
    ral variability is important when designing an
    experiment and interpreting results. Culture cham-
    bers with inflow from surrounding coastal waters
    can exhibit pH fluctuations of up to 0.6 or 1 pH
    unit due to natural diurnal variations in the seawa-
    ter source (Swanson and Fox 2007, Anthony et al.
    2008). Prior to experiments that use seawater
    pumped from coastal waters, seawater pH should
    be monitored at least over a daily cycle so that nat-
    ural fluctuations experienced by the algae are
    known. In experiments to examine community-level
    effects of pH, the goal may be to achieve a variable
    pH that reflects that of the natural environment.
    For example, a diurnal cycle of �0.6 pH units was
    evident for mesocosm cultures of coralline algae
    (Kuffner et al. 2007), while seawater pH increased
    by 0.2–0.3 units gradually over the first 11 d of the
    21 d long PeECE III mesocosm experiments (Bell-
    erby et al. 2008).

    6. Overcoming chemical artifacts. In both methods
    for simulating acidification, secondary changes in
    CO2 speciation are possible through the following
    processes. For calcifying algae, the most important
    will be dissolution of biogenic CaCO3 in the experi-
    mental chamber as a result of acidification. This
    phenomenon will increase both CT and AT. How-
    ever, in a realistic experimental setup, one would
    want to know about dissolution of biogenic CaCO3
    as an outcome, so it is likely that this would be mon-
    itored, either by measuring weight loss of the
    CaCO3 or by parallel measurements of any two of
    the CO2 system parameters. For example, one could
    monitor pH or pCO2 continuously during the cul-
    ture experiment and also take samples for CT
    and ⁄ or AT measurements. This approach would
    enable any changes arising from CaCO3 dissolution
    to be corrected for.

    Another secondary effect is loss of CO2 gas
    because the equilibrium of pCO2 in the chamber is
    greater than that of the ambient atmosphere. For
    culture experiments controlled by addition of CO2
    gas, this can be minimized by partly enclosing the
    ambient air so that both air and water phases
    remain in equilibrium. The disadvantage of this
    approach is that each chamber requiring a different

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1247

    pCO2 condition must be supplied with its own stan-
    dard air-CO2 mixture.

    For a system maintained by periodic additions of
    HCl, the extent of CO2 loss can be monitored by
    measurement of CT before and after. In practice,
    the loss of CO2 under realistic pH conditions
    (pH > 7.5) does not appear to be very large if the
    culture chamber is fitted with an inflated plastic bag
    that has a volume similar to, or smaller than, the
    volume of water in the chamber. This is because the
    quantity of CO2 in a head space is extremely small
    compared to that contained in an equivalent vol-
    ume of seawater.

    7. Replication. A problem with some studies on
    ocean acidification is that they exhibit low levels of
    replication and ⁄ or that replicates are not truly inde-
    pendent of each other; this is especially the case for
    macroalgae (Tables 1 and 2). Care must be taken to
    provide independent replicates required for the cor-
    rect application of statistical tests (i.e., each repli-
    cate culture tank should have its seawater modified
    to the appropriate pH independently, not in one
    header tank per treatment). Psuedoreplication (i.e.,
    growing ‘‘replicate’’ algae in the same treatment
    container) must be avoided (see Hurlbert 1984).
    Obtaining appropriate levels of independent repli-
    cation is especially difficult for the larger macroal-
    gae (e.g., Fucales and Laminariales), which can be
    problematic to maintain in culture long-term, and
    enclosing individuals or populations for field manip-
    ulations is extremely difficult. Experiments using
    macroalgae naturally suffer from high levels of stan-
    dard deviation between replicates because each rep-
    licate comprises one individual (compared to a
    phytoplankton culture of millions of cells); a popu-
    lation response of macroalgae to a treatment is
    therefore difficult to achieve.

    8. Other environmental factors. A preoccupation
    with mimicry of seawater carbonate chemistry may
    result in overlooking other factors important for
    algal growth (e.g., temperature, UV radiation, light
    climate [photon flux density, light:dark cycle], nutri-
    ent concentrations, water motion). It is evident
    from Tables 1 and 2 that these factors vary widely
    across the range of studies considered. Rates of
    calcification and carbon acquisition are energy (i.e.,
    light) dependent. Light limitation may result in
    algae taking up CO2 in preference to HCO3

    ); light
    levels that are too high might induce photoinhibi-
    tion and cause redirection of energy to cellular
    repair mechanisms and away from growth. Suitable
    irradiances can be determined from the results of
    photosynthesis versus irradiance curves. Tempera-
    ture influences all aspects of algal growth and physi-
    ological rates. Inadequate water motion will cause
    diffusion-limited growth, particularly for macroalgae
    (Hurd 2000). Ideally, standardization between
    studies would be valuable and would permit direct
    comparison of the results from different studies.
    There is also a need for experiments that test the

    interactive effects of ocean acidification and other
    predicted changes in climate (Feng et al. 2008). We
    recommend providing experimental conditions that
    are as similar to those in the natural environment
    as possible. Nevertheless, it is essential to report all
    the experimental growth conditions listed above to
    permit critical evaluation of the results.

    Incubation timescales and prior conditioning of organ-
    isms: Several important issues are apparent from
    the CO2-manipulation experiments conducted to
    date. A major issue is the degree of physiological
    response to altered environmental conditions, and
    this is strongly influenced by the timescale of exper-
    iments. The outcomes of these studies have mainly
    been at the level of acclimation, as defined by Raven
    and Geider (2003), that is, studies of organisms that
    have had time to show qualitative or quantitative
    changes in gene expression during growth in
    response to the experimental treatments. An impli-
    cit or explicit assumption in interpreting the results
    is that the experiments did not last long enough to
    permit adaptation (genetic change) of the a strain
    in unialgal cultures of the kind investigated over
    1,000 generations of the freshwater alga Chlamydo-
    monas reinhardtii by Collins and Bell (2004) using
    the CO2-enrichment method. ‘‘Natural laborato-
    ries’’ such as underwater CO2 vents also provide
    opportunities to study adaptation (Hall-Spencer
    et al. 2008). Very short-term experiments (e.g., mea-
    surements of short-term inorganic 14C assimilation,
    net oxygen exchange, or chlorophyll-fluorescence-
    derived electron transport rates), involving exposure
    of organisms grown at the present day, or some
    other CO2 concentration to step changes in CO2,
    only permit the operation of cellular mechanisms
    termed regulation (preexisting metabolic machin-
    ery; Raven and Geider 2003) but are essential in
    defining the kinetic properties of the inorganic car-
    bon acquisition mechanisms under a given set of
    growth conditions. Short-term (2 h) photosynthesis
    experiments on natural marine phytoplankton
    assemblages do not give time for complete acclima-
    tion to new experimental conditions and may (as
    acknowledged by Hein and Sand-Jensen 1997, see
    also Schippers et al. 2004) overestimate the longer-
    term (days) effect of the increased CO2 on meta-
    bolic rates.

    Research using laboratory cultures suffers from
    selection of genotypes favored by the maintenance
    conditions for the isolate. These conditions include
    the absence of UV radiation, low PAR fluxes, and
    unnatural nutrient solute concentrations (including
    inorganic carbon) in the medium if the isolates have
    been in culture for a long time (months to years).
    Against this, there is the possibility of using data from
    other experiments involving the same algal strain in
    planning and interpreting experiments for the organ-
    isms that were isolated and cultured a long time ago.
    In some cases, laboratory experiments used recently
    isolated strains when these were available: Burkhardt

    1248 C A T R I O N A L . H U R D E T A L .

    et al. (1999) used Asterionella glacialis, Coscinodiscus
    wailesii, Thalassiosira punctigera, and Scrippsiella
    trochoidea strains that had recently been isolated from
    the North Sea but obtained Phaeodactylum tricornutom
    from a culture collection. For mesocosms, there is
    the advantage that the algae examined have not spent
    a long period in culture, but there is the problem of
    separating the main species contributing to the algal
    biomass from other species if more than cell counts
    are needed, and the requirement to check by molecu-
    lar phylogenetic means that the same genotype is
    involved not only in the various CO2 treatments but
    also within replicates of a given treatment. This prob-
    lem of intraspecific genotypic variability also applies
    to the use of macroalgae and seagrasses from natural
    populations for CO2 manipulations in the field or in
    the laboratory. For natural vents, the presence of
    chemicals in addition to CO2 in the vent fluids, plus
    the problem of advection of parcels of water and
    their contained biota to and away from the vent site,
    are complicating factors for planktonic organisms
    (Dando et al. 2000), and of genotypic selection of
    adjacent macroalgae and seagrasses.

    In conclusion, there is a need to run experi-
    ments using both approaches to altering CO2
    chemistry, so that any differences in approaches as
    a contributing factor to the wide range of
    responses reported for both micro- and macroalgae
    upon alteration of CO2 chemistry can be
    accounted for. It is essential that the carbon specia-
    tion within culture vessels is carefully monitored at
    least daily during incubation experiments by mea-
    suring pH and one other analytical parameter (CT,
    alkalinity or pCO2) of the seawater carbonate sys-
    tem. There is also a need to better standardize
    across the scientific community the timescales of
    preconditioning of samples (e.g., natural communi-
    ties vs. laboratory cultures) and of incubations dur-
    ing experiments (from 2 h, Schippers et al. 2004,
    to 2 years, Collins and Bell 2004, 2006). Finally,
    independent replication of experimental pCO2
    treatments is an essential prerequisite for statisti-
    cally meaningful results, and other incubation con-
    ditions should mimic natural environmental
    conditions (e.g., light climate, inorganic nutrient
    concentrations) wherever possible.

    This work was funded by University of Otago Research Grants
    to C. L. H., C. D. H., and K. A. H., and a Royal Society of
    New Zealand ISAT-linkages grant to C. L. H. J. A. R.’s work
    on calcified algae is supported by the Natural Environment
    Council (UK). The University of Dundee is a registered Scot-
    tish charity, No. SC015096. We thank Philip Boyd for his
    insightful comments, and Daniel Pritchard and Christopher
    Cornwall for helpful discussions. We are grateful to three
    anonymous reviewers for their perceptive and generous
    reviews. This manuscript is dedicated to our colleague and
    mentor Prof. Peter Bannister.

    Adey, W. H. 1998. Coral reefs: algal structured and mediated eco-
    systems in shallow, turbulent, alkaline waters. J. Phycol. 34:393–
    406.

    Anthony, K. R. N., Kline, D. I., Diaz-Pulido, G., Dove, S. & Hoegh-
    Gludberg, O. 2008. Ocean acidification causes bleaching and
    productivity loss in coral reef builders. Proc. Natl. Acad. Sci.
    U. S. A. 105:17442–6.

    Balch, W., Drapeau, D., Bowler, B. & Booth, E. B. 2007. Prediction
    of pelagic calcification rates using satellite measurements.
    Deep-Sea Res. Part II Top. Stud. Oceanogr. 54:478–95.

    Barcelos e Ramos, J., Biswas, H., Schulz, K. G., La Roche, J. &
    Riebesell, U. 2007. Effect of rising carbon dioxide on the
    marine nitrogen fixer Trichodesmium. Glob. Biogeochem. Cycles
    21:GB2028.

    Bellerby, R. G. J., Schulz, K. G., Riebesell, U., Neill, C., Nondal, G.,
    Heegaard, E., Johannssen, T. & Brown, K. R. 2008. Marine
    ecosystem community carbon and nutrient uptake stoichiom-
    etry under varying ocean acidification during the PeACE III
    experiment. Biogeosciences 5:1517–27.

    Berdalet, E., Peters, F., Koumandou, V. L., Roldan, C., Guadayol,
    O. & Estrada, M. 2007. Species-specific responses of dinofla-
    gellates to quantified small-scale turbulence. J. Phycol.
    439:965–77.

    Bijma, J., Spero, H. J. & Lea, D. W. 1999. Reassessing foraminiferal
    stable isotope geochemistry: impact of the oceanic carbonate
    system (experimental results). In Fisher, G. & Wefer, G. [Eds.]
    Use of Proxies in Paleoceanography: Example from the South Atlantic.
    Springer-Verlag, Berlin, pp. 498–512.

    Borowitzka, M. A. 1981. Photosynthesis and calcification in the
    articulated coralline red algae Amphiroa anceps and A. foliacea.
    Mar. Biol. 62:17–23.

    Borowitzka, M. A. 1987. Calcification in algae: mechanisms and the
    role of metabolism. CRC Crit. Rev. Plant Sci. 6:1–45.

    Burkhardt, S., Riebesell, U. & Zondervan, I. 1999. Effects of growth
    rate, CO2 concentration, and cell size on the stable carbon
    isotope fractionation in marine phytoplankton. Geochim. Cos-
    mochim. Acta 63:3729–41.

    Caldiera, K. & Wickett, M. E. 2003. Anthropogenic carbon and
    ocean pH. Nature 425:365.

    Chave, K. E., Deffeyes, K. S., Weyl, P. K., Garrels, R. M. & Thomp-
    son, M. E. 1962. Observations on the solubility of skeletal
    carbonates in aqueous solutions. Science 137:33–4.

    Chen, X., Qiu, C. E. & Shao, J. Z. 2006. Evidence for K+-dependent
    HCO3

    ) utilization in the marine diatom Phaeodactylum tricor-
    nutum. Plant Physiol. 141:731–6.

    Chisholm, J. R. M. 2003. Primary productivity of reef-building
    crustose coralline algae Limnol. Oceanogr. 48:1376–87.

    Clark, D. R. & Flynn, K. J. 2000. The relationship between dissolved
    inorganic carbon concentration and growth rate in marine
    phytoplankton. Phil. Trans. R. Soc. Lond. B 267:953–9.

    Collins, S. & Bell, G. 2004. Phenotypic consequences of 1000 gen-
    erations of selection at elevated CO2 in a green alga. Nature
    431:137–45.

    Collins, S. & Bell, G. 2006. Evolution of natural algal populations at
    elevated CO2. Ecol. Lett. 9:129–35.

    Dando, P. R., Aliani, S., Arab, H., Bianchi, C. N., Brehmer, M.,
    Cociti, S., Fowler, S. W., et al. 2000. Hydrothermal studies in
    the Aegean Sea. Phys. Chem. Earth B 25:1–8.

    Dason, J. S., Huertas, I. E. & Colman, B. 2004. Source of inorganic
    carbon for photosynthesis in two marine dinoflagellates.
    J. Phycol. 40:285–92.

    De’ath, G., Lough, J. M. & Fabricius, K. E. 2009. Declining
    coral calcification on the Great Barrier Reef. Science
    323:116–9.

    Dickson, A. G. 1993a. The measurement of pH in seawater. Mar.
    Chem. 44:131–42.

    Dickson, A. G. 1993b. pH buffers for sea water media based on the
    total hydrogen ion concentration scale. Deep-Sea Res. 40:107–
    18.

    Dickson, A. G., Sabine, C. L. & Christian, J. R. [Eds.] 2007. Guide to
    Best Practices for Ocean CO2 Measurements. PICES Special Publi-
    cation 3. North Pacific Marine Science Organisation (PICES)
    Sidney, British Columbia, Canada, 191 pp.

    Doney, S. C., Fabry, V. J., Feely, R. A. & Kleypas, J. A. 2009. Ocean
    acidification: the other CO2 problem. Annu. Rev. Mar. Sci.
    1:169–92.

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1249

    Duggins, D. O., Simenstad, C. A. & Estes, J. A. 1989. Magnification
    of secondary production by kelp detritus in coastal marine
    ecosystems. Science 245:170–3.

    Engel, A., Zondervan, I., Aerts, K., Beaufort, L., Benthien, A., Chou,
    L., Delille, B., et al. 2005. Testing the effect of CO2 concen-
    trations on a bloom of the coccolithophorid Emiliania huxleyi
    in mesocosm experiments. Limnol. Oceanogr. 50:493–507.

    Feely, R. A., Sabine, C. L., Lee, K., Berelson, W., Kleypas, J., Fabry,
    V. J. & Millero, F. J. 2004. Impact of anthropogenic CO2 on the
    CaCO3 system of the oceans. Science 305:362–6.

    Feng, Y., Warner, M. E., Zhang, Y., Sun, J., Fu, F.-X., Rose, J. M. &
    Hutchins, D. A. 2008. Interactive effects of increased pCO2,
    temperature and irradiance on the marine coccolithophore
    Emiliania huxleyi (Prymnesiophyceae). Eur. J. Phycol. 43:87–98.

    Field, C. B., Behrenfeld, M. J., Randerson, J. T. & Falkowski, P.
    1998. Primary production of the biosphere: integrating ter-
    restrial and oceanic components. Science 281:237–40.

    Frankignoulle, M. & Canon, C. 1994. Marine calcification as a
    source of carbon dioxide: positive feedback of increasing
    atmospheric CO2. Limnol. Oceanogr. 39:458–62.

    Fu, F.-X., Mulholland, M. R., Garcia, N. S., Beck, A., Bernhardt, P.
    W., Warner, M. E., Sañudo-Wilhelmy, S. A. & Hutchins, D. A.
    2008. Interactions between changing pCO2, N2 fixation, and
    Fe limitation in the marine unicellular cyanobacterium Cro-
    cosphaera. Limnol. Oceanogr. 53:2472–82.

    Fu, F.-X., Warner, M. E., Zhang, Y., Feng, Y. & Hutchins, D. A. 2007.
    Effects of increased temperature and CO2 on photosynthesis,
    growth, and elemental rations in marine Synechococcus and
    Prochlorococcus (cyanobacteria). J. Phycol. 43:485–96.

    Gao, K. Y., Aruga, Y., Asada, K. & Kiyohara, M. 1993. Calcification in
    the articulated coralline algae Corallina pilulifera, with special
    reference to the effect of elevated CO2 concentration. Mar.
    Biol. 117:129–32.

    Gattuso, J.-P., Allemand, D. & Frankignoulle, M. 1998a. Effect of
    calcium carbonate saturation of seawater on coral calcifica-
    tion. Glob. Planetary Change 18:37–46.

    Gattuso, J.-P., Frankignoulle, M. & Wollast, R. 1998b. Carbon and
    carbonate metabolism in coastal aquatic systems. Annu. Rev.
    Ecol. Syst. 29:405–34.

    Giordano, M., Beardall, J. & Raven, J. A. 2005. CO2 concentrating
    mechanisms in algae: mechanisms, environmental modulation
    and evolution. Annu. Rev. Plant Biol. 56:99–131.

    Gueguen, C. & Tortell, P. D. 2008. High-resolution measurement of
    Southern Ocean CO2 and O2 ⁄ Ar by membrane inlet mass
    spectrometry. Mar. Chem. 108:184–94.

    Hall-Spencer, J. M., Rodolfo-Metalpa, R., Martin, S., Ransome, E.,
    Fine, M. E., Turner, S. M., Rowley, S. J., Tedesco, D. & Buia,
    M.-C. 2008. Volcanic carbon dioxide vents show ecosystem
    effects of ocean acidification. Nature 454:96–9.

    Hein, M. & Sand-Jensen, K. 1997. CO2 increases oceanic primary
    productivity. Nature 388:526–7.

    Hinga, K. R. 2002. Effects of high pH on coastal marine phyto-
    plankton. Mar. Ecol. Prog. Ser. 238:281–300.

    Hunter, K. A. 2007. SWCO2 software for computation of equilib-
    rium composition of carbon dioxide in seawater. Available at:
    http://neon.otago.ac.nz/research/mfc/people/keith_hunter/
    software/software.htm.

    Hurd, C. L. 2000. Water motion, marine macroalgal physiology,
    and production. J. Phycol. 36:453–72.

    Hurlbert, S. H. 1984. Pseudoreplication and the design of ecolog-
    ical field experiments. Ecol. Monogr. 54:187–211.

    Hutchins, D. A., Fu, F.-X., Zhang, Y., Warner, M. E., Feng, Y.,
    Portune, K., Bernhardt, P. W. & Mulholland, M. R. 2007. CO2
    control of Trichodesmium N2 fixation, photosynthesis, growth
    rates, and elemental ratios: implications for past, present and
    future ocean biogeochemistry. Limnol. Oceanogr. 52:1293–304.

    Iglesias-Rodriguez, M. D., Buitenhuis, E. T., Raven, J. A., Schofield,
    O., Poulton, A. J., Gibbs, S., Halloran, P. R. & de Baar, H. J. W.
    2008b. Response to comment on ‘‘Phytoplankton calcification
    in a high-CO2 world.’’ Science 322:1466c.

    Iglesias-Rodriguez, M. D., Halloran, P. R., Rickaby, R. E. M., Hall, I.
    R., Colmenero-Hidalgo, E., Gittins, J. R., Green, D. R. H., et al.

    2008a. Phytoplankton photosynthesis in a high-CO2 world.
    Science 320:336–40.

    Israel, A. & Hophy, M. 2002. Growth, photosynthetic properties and
    Rubisco activities and amounts of marine macroalgae grown
    under current and elevated seawater CO2 concentrations. Glob.
    Change Biol. 8:831–40.

    Israel, A., Katz, S., Dubinsky, Z., Merrill, J. E. & Friedlander, M.
    1999. Photosynthetic inorganic carbon utilization and
    growth of Porphyra linearis (Rhodophyta). J. Appl. Phycol.
    11:447–53.

    Kübler, J. E., Johnston, A. M. & Raven, J. A. 1999. The effects of
    reduced and elevated CO2 and O2 on the seaweed Lomentaria
    articulata. Plant. Cell Environ. 22:1303–10.

    Kuffner, I. B., Andersson, A. J., Jokiel, P. L., Rodgers, K. S. &
    Mackenzie, F. 2007. Decreased abundance of crustose coral-
    line algae due to ocean acidification. Nat. Geosci. 1:114–7.

    Langdon, C. 2003. Review of experimental evidence for effects of
    CO2 on calcification of reef builders. Glob. Biogeochem. Cycles
    17:1011.

    Langer, G., Geisen, M., Baumann, K.-H., Kläs, J., Riebesell, U.,
    Thoms, S. & Young, J. R. 2006. Species-specific responses of
    calcifying algae to changing seawater carbonate chemistry.
    Geochem. Geophys. Geosyst. 7:1–12.

    Leclercq, N., Gattuso, J.-P. & Jaubert, J. 2000. CO2 partial pressure
    controls the calcification rate of a coral community. Glob.
    Change Biol. 6:329–34.

    Leonardos, N. & Geider, R. J. 2005. Elevated atmospheric carbon
    dioxide increases organic carbon fixation by Emiliania huxleyi
    (Haptophyta), under nutrient-limited high-light conditions.
    J. Phycol. 41:1196–203.

    Levitan, O., Rosenberg, G., Setlik, I., Setilkova, E., Grigel, J.,
    Klepetar, J., Prasilm, O. & Berman-Frank, I. 2007. Ele-
    vated CO2 enhances nitrogen fixation and growth in the
    marine cyanobacterium Trichodesmium. Glob. Change Biol.
    13:531–8.

    Lewis, E. & Wallace, D. W. R. 1998. Program Developed for CO2 System
    Calculations. ORNL ⁄ CDIAC-105. Carbon Dioxide Information
    Analysis Center, Oak Ridge National Laboratory, U.S.
    Department of Energy, Oak Ridge, Tennessee.

    Maberly, S. C. 1990. Exogenous Sources on inorganic carbon for
    photosynthesis by marine macroalgae. J. Phycol. 26:439–49.

    Mackenzie, F. T. & Lerman, A. 2006. Carbon in the Geobiosphere –
    Earth’s Outer Shell. Springer, Dordrecht, the Netherlands, 402
    pp.

    McNeil, B. I. & Matear, R. J. 2008. Southern Ocean acidification: a
    tipping point at 450-ppm atmospheric CO2. Proc. Natl. Acad.
    Sci. U. S. A. 105:18860–4.

    Midelboe, A. L. & Hansen, P. J. 2007. High pH in shallow-water
    macroalgal habitats. Mar. Ecol. Prog. Ser. 338:107–17.

    Nimer, N. A. & Merrett, M. J. 1993. Calcification rate in Emiliania
    huxleyi Lohmann in response to light, nitrate and availability of
    inorganic carbon. New Phytol. 123:673–7.

    Ohline, S. M., Reid, M. R., Husheer, S. L. G., Currie, K. & Hunter,
    K. A. 2007. Spectrophotometric determination of pH in sea-
    water off Taiaroa Head, Otago, New Zealand: full-spectrum
    modeling and accurate prediction of pCO2 levels. Mar. Chem.
    107:143–55.

    Orr, J. C., Fabry, V. J., Aumont, O., Bopp, L., Doney, S. C., Feely, R.
    A., Gnanadesikan, A., et al. 2005. Anthropogenic ocean acid-
    ification over the twenty-first century and its impact on calci-
    fying organisms. Nature 437:681–6.

    Parsons, R., Raven, J. A. & Sprent, J. I. 1992. A simple flow system
    used to measure acetylene reduction activity of Sesbania rostrata
    stem and root nodules. J. Exp. Bot. 43:595–604.

    Pruder, G. D. & Bolton, E. T. 1980. Differences between cell divi-
    sion and carbon fixation rates associated with light intensity
    and oxygen concentration: implications in the cultivation of
    an estuarine diatom. Mar. Biol. 59:1–6.

    Raupach, M. R., Marland, G., Ciais, P., Le Quere, C., Canadell, J. G.,
    Klepper, G. & Field, C. B. 2007. Global and regional drivers
    of accelerating CO2 emissions. Proc. Natl. Acad. Sci. U. S. A.
    104:10288–93.

    1250 C A T R I O N A L . H U R D E T A L .

    Raven, J. A., Ball, L. A., Beardall, J., Giordano, M. & Maberly, S. C.
    2005. Algae lacking carbon concentrating mechanisms. Can. J.
    Bot. 83:879–90.

    Raven, J. A. & Geider, R. D. 2003. Adaptation, acclimation and
    regulation of photosynthesis in algae. In Larkum, A. W. D.,
    Douglas, S. E. & Raven, J. A. [Eds.] Photosynthesis in Algae.
    Kluwer Academic Publishers, Dordrecht, the Netherlands,
    pp. 385–412.

    Raven, J. A., Kübler, J. & Beardall, J. 2000. Put out the light, and
    then put out the light. J. Mar. Biol. Assoc. U. K. 80:1–25.

    Riebesell, U., Bellerby, R. G. J., Engel, A., Fabry, V. J., Hutchins, D.
    A., Reusch, T. B. H., Shulz, K. G. & Morel, F. M. M. 2008.
    Comment on ‘‘Phytoplankton calcification in a high-CO2
    world.’’ Science 322:1466b.

    Riebesell, U., Zondervan, I., Rost, B., Tortell, P. D., Zeebe, R. E. &
    Morel, F. M. M. 2000. Reduced calcification on marine
    plankton in response to increased atmospheric CO2. Nature
    407:364–7.

    Roberts, R. A. 2001. Review of settlement cues for larval abalone
    (Haliotis spp.). J. Shellfish Res. 20:571–86.

    Schippers, P., Lürling, M. & Scheffer, M. 2004. Increased atmo-
    spheric CO2 promotes phytoplankton productivity. Ecol. Lett.
    7:446–51.

    Shirayama, Y. & Thornton, H. 2005. Effect of increased atmo-
    spheric CO2 on shallow water marine benthos. J. Geophys. Res.
    Oceans 110:C09S08.

    Smith, S. V. 1981. Marine macrophytes as a global carbon sink.
    Science 211:838–40.

    Smith, A. D. & Roth, A. A. 1979. Effect of carbon dioxide con-
    centration on calcification in the red coralline alga Bossiella
    orbigniana. Mar. Biol. 52:217–25.

    Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt,
    K. B., Tignor, M. & Miller, H. L. [Eds.] 2007. Climate Change
    2007: The Physical Science Basis. Contribution of Working Group I to
    the Fourth Assessment Report of the Intergovernmental Panel on Cli-
    mate Change. Cambridge University Press, Cambridge, UK, 996
    pp.

    Swanson, A. K. & Fox, C. H. 2007. Altered kelp (Laminariales)
    phlorotannins and growth under elevated carbon dioxide and
    ultraviolet-B treatments can influence associated intertidal
    food webs. Glob. Change Biol. 13:1696–709.

    Tapp, M., Hunter, K. A., Currie, K. & Macaskill, J. 2000. Apparatus
    for continuous-flow underway spectrophotometric measure-
    ment of surface water pH. Mar. Chem. 72:193–202.

    The Royal Society. 2005. Ocean Acidification Due to Increasing Atmo-
    spheric Carbon Dioxide. Policy Document 12 ⁄ 05. The Royal
    Society, London, 57 pp.

    Thomas, W. H. & Gibson, C. H. 1990. The effects of small-scale
    turbulence on microalgae. J. Appl. Phycol. 2:71–7.

    Tortell, P. D., DiTullio, G. R., Sigman, D. M. & Morel, F. M. M.
    2002. CO2 effects on taxonomic composition and nutrient
    utilization in an equatorial Pacific phytoplankton assemblage.
    Mar. Ecol. Prog. Ser. 236:37–43.

    Tortell, P. D., Payne, C. D., Li, Y., Trimborn, S., Rost, B., Smith, W. O.,
    Riesselman, C., Dunbar, R. B., Sedwick, P. & DiTullio, G. R.
    2008. CO2 sensitivity of Southern Ocean phytoplankton.
    Geophys. Res. Lett. 35:L04605.

    Zondervan, I. 2007. The effect of light, macronutrients, trace
    metals and CO2 on the production of calcium carbonate and
    organic carbon in coccolithophores – a review. Deep-Sea Res.
    Part II Top. Stud. Oceanogr. 54:521–37.

    O C E A N A C I D I F I C A T I O N A N D A L G A L M E T A B O L I S M 1251

    Ocean Acidification and Its Potential Effects
    on Marine Ecosystems

    John M. Guinottea and Victoria J. Fabryb

    aMarine Conservation Biology Institute, Bellevue, Washington, USA
    bCalifornia State University San Marcos, San Marcos, California, USA

    Ocean acidification is rapidly changing the carbonate system of the world oceans.
    Past mass extinction events have been linked to ocean acidification, and the current
    rate of change in seawater chemistry is unprecedented. Evidence suggests that these
    changes will have significant consequences for marine taxa, particularly those that
    build skeletons, shells, and tests of biogenic calcium carbonate. Potential changes in
    species distributions and abundances could propagate through multiple trophic levels
    of marine food webs, though research into the long-term ecosystem impacts of ocean
    acidification is in its infancy. This review attempts to provide a general synthesis of
    known and/or hypothesized biological and ecosystem responses to increasing ocean
    acidification. Marine taxa covered in this review include tropical reef-building corals,
    cold-water corals, crustose coralline algae, Halimeda, benthic mollusks, echinoderms,
    coccolithophores, foraminifera, pteropods, seagrasses, jellyfishes, and fishes. The risk
    of irreversible ecosystem changes due to ocean acidification should enlighten the ongo-
    ing CO2 emissions debate and make it clear that the human dependence on fossil fuels
    must end quickly. Political will and significant large-scale investment in clean-energy
    technologies are essential if we are to avoid the most damaging effects of human-induced
    climate change, including ocean acidification.

    Key words: ocean acidification; climate change; carbonate saturation state; seawater
    chemistry; marine ecosystems; anthropogenic CO2

    Introduction

    The carbonate system (pCO2, pH, alkalin-
    ity, and calcium carbonate saturation state) of
    the world oceans is changing rapidly due to
    an influx of anthropogenic CO2 (Skirrow &
    Whitfield 1975; Whitfield 1975; Broecker &
    Takahashi 1977; Broecker et al. 1979; Feely
    & Chen 1982; Feely et al. 1984; Kleypas et al.
    1999a; Caldeira & Wickett 2003; Feely et al.
    2004; Orr et al. 2005). Ocean acidification
    may be defined as the change in ocean chem-
    istry driven by the oceanic uptake of chemi-
    cal inputs to the atmosphere, including carbon,
    nitrogen, and sulfur compounds. Today, the

    Address for correspondence: John M. Guinotte, Marine Conserva-
    tion Biology Institute, 2122 112th Avenue NE, Suite B-300, Belle-
    vue, WA 98004-2947. Voice: +1-425-274-1180; fax: +1-425-274-1183.
    john@mcbi.org

    overwhelming cause of ocean acidification is
    anthropogenic atmospheric CO2, although in
    some coastal regions, nitrogen and sulfur are
    also important (Doney et al. 2007). For the past
    200 years, the rapid increase in anthropogenic
    atmospheric CO2, which directly leads to de-
    creasing ocean pH through air–sea gas ex-
    change, has been and continues to be caused
    by the burning of fossil fuels, deforestation, in-
    dustrialization, cement production, and other
    land-use changes. The current rate at which
    ocean acidification is occurring will likely have
    profound biological consequences for ocean
    ecosystems within the coming decades and
    centuries.

    Presently, atmospheric CO2 concentration is
    approximately 383 parts per million by volume
    (ppmv), a level not seen in at least 650,000 years,
    and it is projected to increase by 0.5% per year

    Ann. N.Y. Acad. Sci. 1134: 320–342 (2008). C© 2008 New York Academy of Sciences.
    doi: 10.1196/annals.1439.013 320

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 321

    TABLE 1. Projected changes in surface ocean carbonate chemistry based on IPCC IS92a CO2 emission
    scenario (Houghton et al. 2001)a

    Parameter Symbol Unit Glacial Preindustrial Present 2 × CO2 3 × CO2
    Temperature T ◦C 15.7 19 19.7 20.7 22.7
    Salinity S 35.5 34.5 34.5 34.5 34.5
    Total alkalinity AT µmol kg−1 2356 2287 2287 2287 2287
    pCO2 in seawater pCO2 µatm 180 280 380 560 840

    (−56) (0) (35.7) (100) (200)
    Carbonic acid H2CO3 µmol kg−1 7 9 13 18 25

    (−29) (0) (44) (100) (178)
    Bicarbonate ion HCO3 − µmol kg−1 1666 1739 1827 1925 2004

    (−4) (0) (5) (11) (15)
    Carbonate ion CO3 2− µmol kg−1 279 222 186 146 115

    (20) (0) −(16) (−34) (−48)
    Hydrogen ion H+ µmol kg−1 4.79 × 10−3 6.92 × 10−3 8.92 × 10−3 1.23 × 10−2 1.74 × 10−2

    (−45) (0) (29) (78) (151)
    Calcite saturation �calc 6.63 5.32 4.46 3.52 2.77

    (20) (0) (−16) (−34) (−48)
    Aragonite saturation �arag 4.26 3.44 2.9 2.29 1.81

    (19) (0) (−16) (−33) (−47)
    Dissolved inorganic DIC µmol kg−1 1952 1970 2026 2090 2144

    carbon
    (−1) (0) (2.8) (6.1) (8.8)

    Total pH pHT 8.32 8.16 8.05 7.91 7.76

    aWe assume 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 and Millero (1987). pHT is based on seawater scale. Percent change from
    preindustrial values are in parentheses. After Feely et al. (2008).

    throughout the 21st century (Petit et al. 1999;
    Houghton et al. 2001; Augustin et al. 2004;
    Siegenthaler et al. 2005; Meehl et al. 2007). The
    rate of current and projected increases in atmo-
    spheric CO2 is approximately 100× faster than
    has occurred in at least 650,000 years (Siegen-
    thaler et al. 2005). In recent decades, only half of
    anthropogenic CO2 has remained in the atmo-
    sphere; the other half has been taken up by the
    terrestrial biosphere (ca. 20%) and the oceans
    (ca. 30%) (Feely et al. 2004; Sabine et al. 2004).
    Since the Industrial Revolution, a time span of
    less than 250 years, the pH of surface oceans has
    dropped by 0.1 pH units (representing an ap-
    proximately 30% increase in hydrogen ion con-
    centration relative to the preindustrial value)
    and is projected to drop another 0.3–0.4 pH
    units by the end of this century (Mehrbach et al.
    1973; Lueker et al. 2000; Caldeira & Wickett

    2003; Caldeira et al. 2007; Feely et al. 2008).
    [Note: The pH scale is logarithmic, and as a
    result, each whole unit decrease in pH is equal
    to a 10-fold increase in acidity.] A pH change
    of the magnitude projected by the end of this
    century probably has not occurred for more
    than 20 million years of Earth’s history (Feely
    et al. 2004). The rate of this change is cause
    for serious concern, as many marine organ-
    isms, particularly those that calcify, may not be
    able to adapt quickly enough to survive these
    changes.

    A series of chemical reactions is initiated
    when CO2 is absorbed by seawater. � is the
    calcium carbonate saturation state:

    � = [Ca2+][CO2−3 ]/K∗sp
    where K∗sp is the stoichiometric solubility

    product for CaCO3 and [Ca2+] and [CO
    2−
    3 ]

    322 Annals of the New York Academy of Sciences

    TA
    B
    LE

    2
    .

    C
    h

    a
    n

    g
    e
    s

    in
    ca

    lc
    ifi

    ca
    ti

    o
    n

    in
    re

    sp
    o
    n
    se

    to
    in

    cr
    e
    a
    se

    d

    p
    C

    O
    2

    co
    n

    ce

    n
    tr

    a
    ti
    o
    n
    s

    re
    la

    ti
    ve

    to
    ca

    lc
    ifi
    ca
    ti
    o
    n

    a
    t

    p
    re

    se
    n
    t-

    d
    a
    y

    p
    C
    O
    2

    (∼
    3

    8
    0

    p
    p

    m

    v)
    a

    A
    p

    p
    ro

    x.
    %

    ch
    an

    g

    e
    in

    ca
    lc

    ifi
    ca

    ti
    on

    w
    h

    en
    p

    C
    O

    2
    is

    O
    rg

    a

    n
    is

    m
    /S

    ys
    te

    m
    M

    in
    er

    al
    og

    y

    p
    re

    in
    d

    u
    st

    ri
    al


    p

    re
    in

    d
    u

    st
    ri

    al
    R

    ef
    er

    en
    ce

    s
    N

    ot
    es

    C
    o

    cc
    o

    li
    th

    o
    p

    h
    o

    re
    s

    E
    m

    il
    ia

    n
    ia

    h
    u
    xl

    ey
    i

    C
    al

    ci
    te

    −2
    5

    S
    ci

    an
    d

    r

    a
    et

    a
    l.

    20
    03

    E
    .
    h
    u
    xl

    ey
    i

    ′′
    −9

    −1
    8

    R
    ie

    b
    es

    el
    le

    t
    a
    l.

    20
    00

    ;
    D

    ec
    re

    as
    e

    in
    C

    aC
    O

    3
    /c

    el
    l

    Z
    on

    d
    er

    va
    n

    et
    a
    l.

    20
    01

    E
    .
    h
    u
    xl
    ey
    i

    ′′
    −4

    0
    D

    el
    ill

    e
    et

    a
    l.

    20
    05

    I

    n
    it

    ia
    lp

    C
    O

    2
    =

    71
    3

    p
    p

    m
    v.

    D
    ec

    re
    as

    e
    in

    n
    et

    co
    m

    m
    u

    n
    it

    y
    ca

    lc
    ifi
    ca
    ti

    on
    G

    ep
    h
    yr

    oc
    a
    p
    sa

    oc
    ea

    n
    ic

    a
    ′′

    −2
    9

    −6
    6

    R
    ie
    b
    es
    el
    le
    t
    a
    l.
    20
    00
    ;
    D
    ec
    re
    as
    e
    in
    C
    aC
    O
    3
    /c
    el
    l
    Z
    on
    d
    er
    va
    n
    et
    a
    l.
    20
    01

    C
    oc

    co
    li
    th

    u
    s

    p
    el

    a
    gi

    cu
    s

    ′′
    0

    0
    L

    an
    ge

    r
    et

    a
    l.

    20
    06

    N
    o

    ch
    an

    ge
    in

    C
    aC

    O
    3
    /c

    el
    l

    C
    a
    lc

    id
    is

    cu
    s

    le
    pt

    op
    or

    u
    s
    −2
    5

    L
    an

    ge
    r

    et
    a
    l.
    20
    06
    D
    ec
    re
    as
    e
    in
    C
    aC

    O
    3
    ;d

    ec
    re
    as
    e

    al
    so

    ob
    se

    rv
    ed

    at
    18

    0
    p

    p
    m

    v
    F

    o
    ra

    m
    in

    if
    er

    a
    O

    rb
    u
    li
    n
    a

    u
    n
    iv

    er
    sa

    C
    al
    ci
    te

    −8
    −1

    4
    S

    p
    er

    o
    et

    a
    l.

    19
    97

    ;
    D
    ec
    re
    as
    e

    in
    sh

    el
    lw

    ei
    gh

    t
    B

    ijm
    a

    et
    a
    l.

    19
    99

    ;
    B

    ijm
    a
    et
    a
    l.

    20
    02

    G
    lo

    b
    ig

    er
    in

    oi
    d
    es

    sa
    cc

    u
    li
    fe

    r
    ′′

    −4
    to

    −6
    −6

    to
    −8

    B
    ijm

    a
    et
    a
    l.
    19
    99
    ;
    D
    ec
    re
    as
    e
    in
    sh
    el
    lw
    ei
    gh
    t
    B
    ijm
    a
    et
    a
    l.
    20
    02

    S
    cl

    er
    a

    ct
    in

    ia
    n

    co
    ra

    ls
    S
    ty

    lo
    p
    h
    or

    a
    p
    is

    ti
    ll
    a
    ta

    ′′
    −1

    4
    −2

    0
    G

    at
    tu

    so
    et

    a
    l.

    19
    98

    C
    u

    rv
    e

    w
    as

    n
    on

    lin
    ea

    r,
    lin

    ea
    r

    fit
    w

    as
    m

    a

    d
    e

    fo
    r

    ar
    ag

    fr
    om

    2.
    0−

    3.
    9

    S
    .
    p
    is

    ti
    ll
    a
    ta
    ′′
    0

    to
    −5

    0
    R

    ey
    n

    au
    d

    et
    a
    l.
    20
    03

    L
    ev

    el
    of

    re
    sp

    on
    se

    is
    te

    m
    p

    er
    at

    u
    re


    d

    ep
    en

    d
    en

    t
    A

    cr
    op

    or
    a

    ce
    rv

    ic
    or

    n
    is

    A
    ra

    go
    n

    it
    e

    −4
    0

    −5
    9

    R
    en

    eg
    ar

    &
    R

    ie
    gl

    20
    05

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 323

    TA
    B
    LE
    2
    .
    A
    p
    p
    ro
    x.
    %
    ch
    an
    ge
    in
    ca
    lc
    ifi
    ca
    ti
    on
    w
    h
    en
    p
    C
    O
    2
    is
    O
    rg

    an
    is

    m
    /S
    ys
    te
    m
    M
    in
    er
    al
    og
    y

    p
    re
    in
    d
    u
    st
    ri
    al

    p
    re
    in
    d
    u
    st
    ri
    al
    R
    ef
    er
    en
    ce
    s
    N
    ot
    es

    A
    cr

    op
    or

    a
    eu

    ry
    st

    om
    a

    ′′
    −5

    5
    S

    ch
    n

    ei
    d

    er
    &

    E
    re

    z
    20

    06
    A

    cr
    op
    or
    a

    ve
    rw

    ey
    i
    ′′
    −1

    2
    −1

    8
    M

    ar
    u

    b
    in

    ie
    t
    a
    l.

    20
    03

    P.
    co

    m
    p
    re

    ss
    a

    +

    M
    on

    ti
    p
    or

    a
    ca

    pi
    ta

    ta
    ′′

    −4
    0
    −5
    9
    L
    an

    gd
    on

    &
    A

    tk
    in

    so
    n

    20
    05

    P
    or

    it
    es

    co
    m
    p
    re
    ss
    a
    ′′
    −1

    7
    −2

    5
    M

    ar
    u
    b
    in
    ie
    t
    a
    l.
    20
    01

    P.
    lu

    te
    a

    ′′
    −3

    8
    −5

    6
    O

    h
    d

    e
    &

    H
    os

    sa
    in

    20
    04

    P.
    lu
    te
    a
    ′′
    −3

    3
    −4

    9
    H

    os
    sa

    in
    &

    O
    h

    d
    e
    20
    06

    P.
    p
    or

    it
    es
    ′′
    −1

    6
    M

    ar
    u
    b
    in

    i

    &
    T

    h
    ak

    e
    19

    99
    P

    av
    on

    a
    ca

    ct
    u
    s

    ′′
    −1
    4
    −2

    0
    M

    ar
    u
    b
    in
    ie
    t
    a
    l.
    20
    03

    F
    u
    n
    gi

    a
    sp

    .
    ′′

    −4
    7

    −6
    9

    H
    os
    sa
    in
    an
    d
    O
    h
    d
    e
    20
    06

    G
    a
    la

    xe
    a

    fa
    sc

    ic

    u
    la

    ri
    s

    ′′
    −1
    2
    −1
    8
    M
    ar
    u
    b
    in
    ie
    t
    a
    l.
    20
    03

    G
    .
    fa

    sc
    ic

    u
    la
    ri
    s
    ′′
    −5

    6
    −8

    3
    M

    ar
    sh

    al
    l&

    C
    lo

    d
    e
    20
    02

    T
    u
    rb

    in
    a
    ri

    a
    re

    n
    if
    or

    m
    is

    ′′
    −9

    −1
    3

    M
    ar

    u
    b

    in
    ie

    t
    a
    l.
    20
    03
    C
    o

    ra
    ll

    in
    e

    re
    d

    a
    lg

    a
    e

    P
    or

    ol
    it
    h
    on

    ga
    rd

    in
    er

    i
    H

    ig
    h

    -M
    g

    ca
    lc
    it
    e
    −2
    5

    A
    ge

    gi
    an

    19
    85

    M
    es

    o
    co

    s

    m
    s

    a
    n

    d
    fi

    el
    d

    st
    u

    d
    ie

    s
    B

    io
    sp

    h
    er

    e
    2

    M
    ix

    ed
    −5

    6
    −8

    3
    L

    an
    gd

    on
    et

    a
    l.
    20
    00

    D
    om

    in
    at

    ed
    by

    co
    ra

    lli
    n

    e
    re

    d
    al

    ga
    e

    M
    on

    a
    co

    m
    es

    oc
    os

    m
    ′′

    −2
    1

    L
    ec

    le
    rc

    q
    et

    a
    l.
    20
    00
    M
    on
    a
    co
    m
    es
    oc
    os
    m
    ′′

    −1
    5

    L
    ec
    le
    rc
    q
    et
    a
    l.
    20
    02

    B
    a
    h
    a
    m

    a
    s

    B
    a
    n
    k

    ′′
    −5

    7
    −8

    5
    B

    ro
    ec

    ke
    r

    &
    T

    ak
    ah

    as
    h

    i1
    96

    6
    R

    u
    ka

    n
    -s

    h
    o,

    O
    ki

    n
    a
    w

    a
    ′′

    −4
    5

    −6
    7

    O
    h
    d
    e

    &
    va

    n
    W

    oe
    si

    k
    19

    99
    N

    a
    tu

    re
    R

    es
    er

    ve
    R

    ee
    f,

    R
    ed

    S
    ea

    ′′
    −5
    5
    S

    ilv
    er

    m
    an

    et
    a
    l.

    20
    07

    a
    N

    ea
    rl

    y
    al

    lo
    rg

    an
    is
    m
    s

    te
    st

    ed
    to

    d
    at

    e
    sh

    ow
    re

    d
    u

    ce
    d

    ca
    lc
    ifi
    ca
    ti
    on
    in
    re

    sp
    on

    se
    to

    el
    ev

    at
    ed

    p
    C
    O
    2
    an
    d

    d
    ec

    re
    as

    ed
    ca

    rb
    on

    at
    e

    io
    n

    co
    n

    ce
    n

    tr
    at

    io
    n
    an
    d

    ca
    rb

    on
    at

    e
    sa

    tu
    ra

    ti
    on

    st
    at

    e.
    M

    od
    ifi

    ed
    fr

    om
    K

    le
    yp

    as
    et

    a
    l.

    (2
    00

    6)
    .

    324 Annals of the New York Academy of Sciences

    are the in situ calcium and carbonate concen-
    trations, respectively. The end products of these
    reactions are an increase in hydrogen ion con-
    centration (H+), which lowers pH (making wa-
    ters more acidic), and a reduction in the num-
    ber of carbonate ions (CO2−3 ) available. This
    reduction in carbonate ion concentration also
    leads to a reduction in calcium carbonate sat-
    uration state (�), which has significant impacts
    on marine calcifiers. Table 1 lists carbon sys-
    tem parameters and temperature changes for
    surface waters based on the Intergovernmental
    Panel on Climate Change (IPCC) IS92a CO2
    emission scenario.

    A reduction in the number of carbon-
    ate ions available will make it more diffi-
    cult and/or require marine calcifying organ-
    isms to use more energy to form biogenic
    calcium carbonate (CaCO3). Many marine
    organisms form biogenic calcium carbonate
    including: crustose coralline algae (the pri-
    mary cementer that makes coral reef formation
    possible), Halimeda (macroalgae), foraminifera,
    coccolithophores, tropical reef-building corals,
    cold-water corals, bryozoans, mollusks, and
    echinoderms. The majority of marine calci-
    fiers tested to date are sensitive to changes
    in carbonate saturation state and have shown
    declines in calcification rates in laboratory
    and mesocosm studies (Table 2). These or-
    ganisms are affected and will continue to be
    affected by ocean acidification, but less well
    known are the ecosystem impacts on higher
    trophic-level organisms that rely on these cal-
    cifiers for shelter, nutrition, and other core
    functions.

    Decreasing pH is not the only effect on the
    inorganic carbon system in seawater that re-
    sults from the ocean’s uptake of anthropogenic
    CO2. Calcite and aragonite are the major
    biogenically formed carbonate minerals pro-
    duced by marine calcifiers, and the stability
    of both minerals is affected by the amount
    of CO2 in seawater, which is partially deter-
    mined by temperature. Colder waters natu-
    rally hold more CO2 and are more acidic than
    warmer waters. The depths of the aragonite

    and calcite saturation horizons are important
    to marine calcifiers because the depth of these
    horizons determines the limit at which pre-
    cipitation of biogenic calcium carbonate by
    marine organisms is favored (shallower than
    the saturation horizon) and at which they will
    experience dissolution (deeper than the satu-
    ration horizons) in the absence of protective
    mechanisms.

    The aragonite and calcite saturation hori-
    zons of the world’s oceans are moving to
    shallower depths due to the rapid influx of an-
    thropogenic CO2 to the oceans (Fig. 1). This
    process has been well documented and mod-
    eled at the global scale (Skirrow & Whitfield
    1975; Broecker & Takahashi 1977; Feely &
    Chen 1982; Feely et al. 1984, 1988; Kleypas
    et al. 1999a; Broecker 2003; Caldeira & Wickett
    2003; Feely et al. 2004; Caldeira & Wickett
    2005; Orr et al. 2005). Future estimates of arag-
    onite saturation horizon depth indicate that
    shoaling will occur in the North Pacific, North
    Atlantic, and Southern Ocean within the cen-
    tury (Orr et al. 2005). The aragonite and cal-
    cite saturation horizons in the North Pacific
    are currently very shallow (Feely et al. 2004)
    and are moving toward the surface at a rate of
    1–2 m per year (R.A. Feely pers. comm. 2007).
    Many of the areas where shoaling is predicted
    to occur within the century are highly produc-
    tive and home to many of the world’s most
    important and economically lucrative commer-
    cial fisheries.

    It is clear that human-induced changes in
    atmospheric CO2 concentrations are funda-
    mentally altering ocean chemistry from the
    shallowest waters to the darkest depths of the
    deep sea. The chemistry of the oceans is ap-
    proaching conditions not seen in many mil-
    lions of years, and the rate at which this is
    occurring is unprecedented (Caldeira & Wick-
    ett 2003). Caldeira and Wickett (2003, p. 365)
    state “Unabated CO2 emissions over the com-
    ing centuries may produce changes in ocean
    pH that are greater than any experienced
    in the past 300 million years, with the pos-
    sible exception of those resulting from rare,

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 325

    Figure 1. (A) Depth of the aragonite saturation horizon (ASH), locations of deep-sea
    bioherm-forming corals, and diversity contours for 706 species of azooxanthellate corals.
    Projected ASH depth for year 1765 (preindustrial); pCO2 = 278 ppmv. Green triangles
    are locations of six deep-sea, scleractinian, bioherm-forming coral species (Lophelia pertusa,
    Madrepora oculata, Goniocorella dumosa, Oculina varicosa, Enallopsammia profunda, and
    Solenosmilia variabilis). Numerals not falling on diversity contours indicate number of azoox-
    anthellate coral species. Reprinted with permission from Guinotte et al. (2006). (B) Projected
    ASH depth for year 2040; pCO2 = 513 ppmv. (C) Projected ASH depth for year 2099; pCO2
    = 788 ppmv. Black areas appearing in the Southern Ocean and North Pacific indicate areas
    where the ASH depth has reached the surface. (In color in Annals online.)

    326 Annals of the New York Academy of Sciences

    catastrophic events in Earth history” (Caldeira
    and Rampino 1993; Beerling and Berner
    2002). Recent evidence suggests ocean acid-
    ification was a primary driver of past mass
    extinctions and reef gaps, which are time
    periods on the order of millions of years
    that reefs have taken to recover from mass
    extinctions (Stanley 2006; Veron 2008). Za-
    chos and colleagues (2005) calculated that if
    the entire fossil fuel reservoir (ca. 4500 GtC)
    were combusted, the impacts on deep-sea pH
    and biota would probably be similar to those
    in the Paleocene–Eocene Thermal Maximum
    (PETM), 55 million years ago. The PETM
    likely caused a mass extinction of benthic
    foraminifera (Zachos et al. 2005). Projected an-
    thropogenic carbon inputs will occur within just
    300 years, which is thought to be much faster
    than the CO2 release during the PETM and too
    rapid for dissolution of calcareous sediments to
    neutralize anthropogenic CO2. Consequently,
    the ocean acidification-induced impacts on sur-
    face ocean pH and biota will probably be more
    severe than during the PETM (Zachos et al.
    2005).

    While it is apparent changing seawater
    chemistry will have serious consequences for
    many marine calcifiers, the effects of ocean
    acidification on noncalcifiers and the ecosys-
    tem responses to these changes will be com-
    plex and difficult to quantify. Assessing whether
    ocean acidification is the primary driver of
    a species’ population decline will be diffi-
    cult due to the multitude of ongoing phys-
    ical and chemical changes currently occur-
    ring in the ocean. Ocean acidification is oc-
    curring in synergy with significant ongoing
    environmental changes (e.g., ocean temper-
    ature increases), and these cumulative im-
    pacts or interactive effects of multiple stressors
    may have more significant consequences for
    biota than any single stressor. Thus, research
    into the synergistic effects of these changes
    on marine organisms and the consequent
    ecosystem responses is critical but still in its
    infancy.

    Calcification and Dissolution
    Response

    Hermatypic Corals (Zooxanthellate)

    The calcification response of reef-building
    corals to decreases in aragonite saturation state
    has been well documented for a handful of se-
    lect species. These experiments have been con-
    ducted in laboratory tanks and mesocosms, but
    to date have not been conducted in in situ field
    experiments under “natural” conditions. Evi-
    dence from species tested to date indicate that
    the calcification rates of tropical reef-building
    corals will be reduced by 20–60% at double
    preindustrial CO2 concentrations (pCO2 ca.
    560 ppmv) (Gattuso et al. 1998; Kleypas et al.
    1999a; Langdon et al. 2000; Kleypas & Lang-
    don 2002; Langdon et al. 2003; Reynaud et al.
    2003; Langdon & Atkinson 2005; c.f. Royal
    Society 2005; c.f. Kleypas et al. 2006) (see
    Table 2). Figure 2 illustrates the projected re-
    duction in surface-water aragonite saturation
    state through the year 2069. A reduction in cal-
    cification of this magnitude could fundamen-
    tally alter the current structure and function of
    coral-reef ecosystems, as their growth is depen-
    dent on their ability to accrete at faster rates
    than erosional processes can break them down.
    Reef accretion will become increasingly more
    critical in the coming decades, as global sea
    levels rise and available light for photosynthe-
    sis becomes a limiting factor for corals at the
    deepest reaches of the photic zone.

    A substantial decrease in the number of car-
    bonate ions available in seawater will have se-
    rious implications for coral calcification rates
    and skeletal formation. Weaker coral skele-
    tons will probably result from a reduction in
    carbonate ions, enabling erosional processes
    to occur at much faster rates than have oc-
    curred in the past, and slower growth rates
    may also reduce corals’ ability to compete for
    space and light, though no studies have been
    conducted to test this hypothesis (reviewed in
    Kleypas et al. 2006). Biosphere II mesocosm ex-
    periments suggest that net reef dissolution will

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 327

    Figure 2. (A) Surface aragonite saturation state. Calculated preindustrial (1870) �arag values; pCO2
    = 280 ppmv. Green dots represent present-day distribution of zooxanthellate coral reefs. Figure modified
    from Guinotte et al. (2003). Figure legend classification from Kleypas et al. (1999b). (B) Surface aragonite
    saturation state. Projected �arag values, 2060–2069; pCO2 = 517 ppmv. (In color in Annals online.)

    outpace net reef calcification when carbonate
    ion concentration decreases to about 150 to
    110 µmole kg−1, a range that corresponds to
    atmospheric CO2 concentrations of 560–840
    ppmv (C. Langdon pers. comm. 2007). Hoegh-
    Guldberg and colleagues (2007) stated that
    aragonite saturation values will favor erosion
    when the carbonate ion concentration ap-
    proaches 200 µmole kg−1 (atmospheric CO2
    concentration = 480 ppmv). The effects of a
    reduction in calcification rates on recruitment,
    settlement, and juvenile life stages of most ma-
    rine calcifiers, including the majority of scler-
    actinian corals, are not well known. However,

    Edmunds (2007) documented a decline in the
    growth rates of juvenile scleractinian corals in
    the U.S. Virgin Islands and raised the possi-
    bility that the effects of global climate change
    (increased seawater temperatures and decreas-
    ing aragonite saturation state) have already re-
    duced the growth rate of juvenile corals.

    Fine and Tchernov (2007a) reported two
    species of scleractinian corals were able to sur-
    vive corrosive water conditions (pH values of
    7.3–7.6), which caused their skeletons to dis-
    solve completely, leaving the coral polyps ex-
    posed. When water chemistry returned to nor-
    mal/ambient conditions, the coral polyps were

    328 Annals of the New York Academy of Sciences

    able to recalcify their skeletons without any ob-
    vious detrimental effects. These findings shed
    new light on the hypothesis that corals have a
    means of alternating between soft bodies and
    skeletal forms, which are absent from the fos-
    sil record during reef gaps (Stanley & Fautin
    2001; Medina et al. 2006; Stanley 2006; Fine
    & Tchernov 2007a). Fine and Tchernov’s re-
    sults offer some hope for the future of corals
    in a high CO2 world, but caution should be
    exercised as these manipulative experiments
    did not include the effects of predation on the
    “naked” coral polyps. Hard skeletons also pro-
    vide another core function for coral polyps by
    protecting them from periodic natural events
    such as tsunamis and cyclones, which can cause
    significant damage to coral colonies and reef
    systems.

    There is some discrepancy regarding the rep-
    resentativeness of the coral species used in the
    Fine and Tchernov calcification experiments.
    Stanley (2007) stated the experiments may
    not be representative of all coral species, par-
    ticularly zooxanthellate reef-building species,
    which might have responded quite differently
    to the experiments because of the complex
    nature of their photosymbiosis. This assertion
    was challenged by Fine & Tchernov (2007b)
    in the statement that the evolution and physi-
    ology of the studied species are indistinguish-
    able from tropical reef-building species. Rep-
    resentativeness aside, both parties agree that
    ocean acidification poses a significant threat
    to coral-reef ecosystems and the services they
    provide.

    Calcifying Macroalgae

    Coralline Algae

    Scleractinian corals are not the only reef-
    calcifying organisms that are sensitive to de-
    creasing saturation states. Crustose coralline al-
    gae (CCA) are a critical player in the ecology of
    coral-reef systems as they provide the “cement”
    that helps stabilize reefs, make significant sed-
    iment contributions to these systems, and are
    important food sources for sea urchins, par-

    rot fish, and several species of mollusks (Littler
    & Littler 1984; Chisholm 2000; Diaz-Pulido
    et al . 2007). CCA also provide important hard
    settlement substrate for coral larvae (Heyward
    & Negri 1999; Harrington et al. 2005; Diaz-
    Pulido et al. 2007). Coralline algae produce cal-
    cium carbonate in the form of high-magnesium
    calcite, a more soluble form of calcium carbon-
    ate than either calcite or aragonite, which make
    these species particularly sensitive to decreasing
    carbonate saturation states.

    Mesocosm experiments exposing CCA to el-
    evated pCO2 (2 × present day) indicate up to a
    40% reduction in growth rates, 78% decrease
    in recruitment, 92% reduction in total area cov-
    ered by CCA, and a 52% increase in noncal-
    cifying algae (Buddemeier 2007; Kuffner et al.
    2008). Agegian (1985) also reported a reduc-
    tion in recruitment when CCA were exposed to
    elevated pCO2 in aquarium experiments. Bud-
    demeier (2007) states, “The combined effects of
    reduced carbonate production and diminished
    stabilization (cementation) of coasts and shal-
    low seafloors by encrusting calcifiers are likely
    to lead to more rapid erosion and ecosystem
    transitions (macroalgal takeover) than would
    be expected on the basis of decreases in coral
    growth alone.” The ecological importance of
    coralline algae to reef systems and the effects
    decreasing carbonate saturation state will have
    on these organisms have been overlooked to a
    significant degree, and more research is needed
    to document CCA response to reduced car-
    bonate saturation states and in turn how these
    responses will impact reef ecosystems.

    Halimeda

    Halimeda is a genus of green, calcifying
    macroalgae that forms extensive beds in cer-
    tain regions of the world’s oceans. Some of
    the most well-developed Halimeda beds occur
    off the northeast coast of Australia, and es-
    timates of total area covered by Halimeda in
    the Great Barrier Reef region are upwards
    of 2000 km2 (reviewed by Diaz-Pulido et al.
    2007). Halimeda, along with other calcareous
    algae (Udotea, Amphiroa, and Galaxaura), are

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 329

    important producers of marine sediments
    and contribute to reef accretion by filling
    voids in the reef matrix with their sediments
    (Hillis-Colinvaux 1980; Davies & Marshall
    1985; Drew & Abel 1988; Diaz-Pulido et al.
    2007). Reefs and Halimeda bioherms have high
    calcification rates and are responsible for the
    majority of CaCO3 production and accumu-
    lation on the continental shelf (Milliman &
    Droxler 1996; Kleypas et al. 2006).

    The three-dimensional structures Halimeda
    form, which can be 20 m in height, provide im-
    portant habitat for adult fishes and may serve as
    nursery grounds for juvenile fishes and inverte-
    brates (Beck et al. 2003). Calcifying macroalgae
    produce biogenic calcium carbonate in three
    forms: high-magnesium calcite, aragonite, and
    calcite; all of these forms are susceptible to the
    negative effects of decreasing carbonate satura-
    tion states (Littler & Littler 1984). Few species
    of Halimeda have been exposed to high pCO2
    in lab experiments, but one species from the
    Great Barrier Reef, Halimeda tuna, displayed a
    negative calcification response when exposed
    to a pH drop of 0.5 units (8 to 7.5) (Borowitzka
    & Larkum 1986).

    Cold Water Corals (Azooxanthellate)

    Cold water corals and the ecologically
    rich bioherms they form are widely dis-
    tributed throughout the world oceans (see
    Fig. 1). A great number of these highly pro-
    ductive ecosystems have been discovered only
    in the last decade, and it is thought that the
    area covered by these organisms may surpass
    the total area of tropical zooxanthellate reef
    systems (Mortensen et al. 2001; Freiwald et al.
    2004; Freiwald & Roberts 2005; Guinotte et al.
    2006; Turley et al. 2007). Cold water corals are
    azooxanthellate, which means they do not con-
    tain photosynthetic algae, and thus are not lim-
    ited to the photic zone. The majority of cold
    water corals are found in depths of 200–1000
    m or more, and some solitary colonies have
    been found at depths of several thousand me-
    ters (Freiwald 2002; Freiwald et al. 2004). There

    are six species of azooxanthellate, bioherm-
    forming, scleractinian corals (Lophelia pertusa,
    Madrepora oculata, Goniocorella dumosa, Oculina

    varicosa, Enallopsammia profunda, and Solenosmilia
    variabilis), all of which produce calcium carbon-
    ate skeletons of aragonite. Cold water corals
    bioherms have extremely high biodiversity and
    provide habitat and nursery areas for many
    deep-sea organisms, including several commer-
    cially important fish species (Rogers 1999; Fossa
    et al. 2002; Husebo et al. 2002). Scleractinian
    cold water corals are not the only azooxanthel-
    late habitat formers.

    The “coral gardens” of the North Pacific are
    biodiversity hotspots dominated by octocorals
    (soft corals, stoloniferans, sea fans, gorgonians,
    and sea pens) and stylasterids, the majority of
    which produce calcite spicules and holdfasts
    (Cairns & Macintyre 1992; Guinotte et al. 2006;
    Stone 2006). Stone (2006) reported that 85%
    of the economically important fish species ob-
    served on submersible transects in waters off
    the Aleutian Islands were associated with corals
    and other emergent epifauna. The waters off
    the Aleutian Islands have the highest abun-
    dance and diversity of cold-water corals found
    to date in high-latitude ecosystems (Heifetz et al.
    2005; Stone 2006), but well-developed scler-
    actinian bioherms are curiously absent from
    this region even though scleractinian bioherm-
    forming species are found in North Pacific wa-
    ters (Guinotte et al. 2006).

    The reason scleractinian bioherms are not
    present in North Pacific waters could be a
    function of the shallow depth of the arago-
    nite saturation horizon and high dissolution
    rates throughout the region (Guinotte et al.
    2006). If this hypothesis is true, then decreas-
    ing carbonate saturation state will probably
    impact scleractinian cold-water corals earlier
    than shallow-water reef builders. Cold-water
    corals are bathed in cold, deep waters that
    have naturally high levels of CO2 (global av-
    erage �arag = 2). The low carbonate saturation
    state environment in which they live probably
    contributes to their slow growth/calcification
    rates, which are an order of magnitude slower

    330 Annals of the New York Academy of Sciences

    than tropical zooxanthellate corals (global aver-
    age �arag = 4). Indeed, some deeper cold-water
    coral bioherms could already be experiencing
    corrosive conditions with respect to aragonite
    saturation state (�arag < 1), though no evidence of this has been documented.

    Greater than 95% of the present day
    distribution of bioherm-forming scleractinian
    species occur in waters that are supersaturated
    with aragonite (Guinotte et al. 2006). Future
    aragonite saturation state projections from Orr
    and co-authors (2005) indicate that 70% of scle-
    ractinian cold-water coral bioherms could be
    in undersaturated water with respect to arago-
    nite by the end of the century (Guinotte et al.
    2006; Turley et al. 2007) (see Fig. 1). Labora-
    tory experiments are currently being conducted
    to test whether cold water corals scleractinians
    (Lophelia pertusa) are sensitive to decreasing arag-
    onite saturation state (Riebesell pers. comm.),
    but no lab experiments have been conducted
    to test the sensitivity of cold-water octocorals
    and stylasterids to decreasing carbonate satu-
    ration states. Manipulative CO2 experiments
    to determine cold-water coral sensitivity and
    calcification response to decreasing carbonate
    saturation states are a top priority for future
    research (Guinotte et al. 2006; Kleypas et al.
    2006; Roberts et al. 2006; Turley et al. 2007).

    Benthic Mollusks, Bryozoans,
    and Echinoderms

    The physiological and ecological impacts of
    increasing pCO2 on benthic mollusks, bry-
    ozoans, and echinoderms are not well known,
    and few manipulative experiments have been
    carried out to determine sensitivity to ele-
    vated pCO2 (Kleypas et al. 2006). The nega-
    tive effects of acidic waters on bivalves have
    been investigated in a small number of stud-
    ies (Kuwatani & Nishii 1969; Bamber 1987,
    1990; Michaelidis et al. 2005; Berge et al. 2006),
    and only one investigated the negative calci-
    fication response to pCO2 levels within the
    range predicted by the IPCC (Gazeau et al.
    2007). Gazeau and colleagues (2007) found

    that calcification rates of the mussel (Mytilus
    edulis) and Pacific oyster (Crassostrea gigas) can
    be expected to decline linearly with increas-
    ing pCO2, 25% and 10% respectively, by the
    end of the century (ca. 740 ppmv, IPCC IS92a
    scenario). Both species are important coastal
    ecosystem engineers and represent a signif-
    icant portion of global aquaculture produc-
    tion (Gazeau et al. 2007). Bivalves that settle
    in coastal estuarine areas may be particularly
    vulnerable to anthropogenic ocean acidifica-
    tion. These organisms naturally experience ex-
    tremely high mortality rates (>98%) in their
    transition from larvae to benthic juveniles (re-
    viewed by Green et al. 2004), and any increase
    in juvenile mortality due to ocean acidification
    could have serious effects on estuarine bivalve
    populations.

    Kurihara and colleagues (2007) demon-
    strated that increased pCO2 of seawater pro-
    jected to occur by the year 2300 (pH 7.4) will
    severely impact the early development of the
    oyster Crassostrea gigas and highlighted the im-
    portance of acidification effects on larval de-
    velopment stages of marine calcifiers. Because
    early life stages appear to be more sensitive
    to environmental disturbance than adults and
    most benthic calcifiers possess planktonic lar-
    val stages, fluctuations in larval stages due to
    high mortality rates may exert a strong influ-
    ence on the population size of adults (Green
    et al. 2004). Kurihara and co-authors (2004)
    investigated the effects of increased pCO2 on
    the fertilization rate and larval morphology of
    two species of sea urchin embryos (Hemicentrotus
    pulcherrimus and Echinometra mathaei) and found
    the fertilization rate of both species declined
    with increasing CO2 concentration. In addi-
    tion, the size of pluteus larvae decreased with
    increasing CO2 concentration and malformed
    skeletogenesis was observed in larval stages of
    both species. Kurihara and Shirayama (2004)
    concluded that both decreasing pH and altered
    carbonate chemistry affect early development
    and life history of many marine organisms,
    which will result in serious consequences for
    marine ecosystems.

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 331

    Experiments focusing on the direct effects
    of increasing ocean acidification on marine
    calcifiers have been the dominant activity to
    date, but numerous and ecologically significant
    indirect effects are probable. Bibby and col-
    leagues (2007) documented interesting behav-
    ioral, metabolic, and morphological responses
    of the intertidal gastropod Littorina littorea to
    acidified seawater (pH = 6.6). This marine snail
    produced thicker shells when exposed to preda-
    tion (crab) cues in control experiments, but this
    defensive response was disrupted when pH was
    decreased. The snails also displayed reduced
    metabolic rates and an increase in avoidance
    behavior, both of which could have significant
    ecosystem implications via organism interac-
    tions, energy requirements, and predator–prey
    relationships. This study investigated only one
    species of mollusk, but other marine organisms
    will probably have indirect responses to ocean
    acidification (Bibby et al. 2007).

    Coccolithophores, Foraminifera,
    and Pteropods

    The major planktonic producers of CaCO3
    are coccolithophores (single-celled algae),
    foraminifera (protists), and euthecosomatous
    pteropods (planktonic snails). Coccol-
    ithophores and foraminifera secrete CaCO3 in
    the form of calcite, whereas pteropods secrete
    shells made of aragonite, which is about 50%
    more soluble in seawater than calcite (Mucci
    1983). These planktonic groups differ with
    respect to their size, trophic level, generation
    time, and other ecological attributes. High
    quality, quantitative data on the latitudinal and
    vertical distributions and abundances of these
    calcareous taxa are lacking, and estimates of
    their contributions to global calcification rates
    are poorly constrained.

    The calcification response of coccol-
    ithophores, foraminifera, and pteropods to
    ocean acidification has been investigated to
    date in very few species. Most studies have in-
    volved bloom-forming coccolithophores, and
    these species (Emiliania huxleyi and Geophyro-

    capsa oceanica) show decreased calcification
    rates ranging from 25 to 66% when pCO2
    is increased to 560–840 ppmv, respectively
    (TABLE 2) in lab and mesocosm experiments. In
    lab experiments with the coccolithophore Coc-
    colithus pelagicus, however, Langer et al. (2006)
    found that calcification did not change with in-
    creased CO2. Moreover, there is evidence sug-
    gesting that at least one coccolithophore species
    may have the capacity to adapt to changing
    pCO2 over long periods. Experimental manip-
    ulations show that Calcidiscus leptoporus exhibits
    highest calcification rates at present-day CO2
    levels, with malformed coccoliths and cocco-
    spheres at both lower and higher pCO2 (Langer
    et al. 2006). Because no malformed coccol-
    iths were observed in sediments from the Last
    Glacial Maximum (when pCO2 levels were
    about 200 ppmv), the authors concluded that
    C. leptoporus has adapted to present-day CO2
    levels.

    In lab experiments with two species of plank-
    tonic foraminifera, shell mass decreased as the
    carbonate ion concentration of seawater de-
    creased (Spero et al. 1997; Bijma et al. 1999,
    2002). When grown in lab experiments in
    seawater chemistry equivalent to pCO2 val-
    ues of 560 and 740 ppmv, shell mass of the
    foraminifera Orbulina universa and Globigerinoides
    sacculifer declined by 4–8% and 6–14%, respec-
    tively, compared to the shell mass secreted at
    the preindustrial pCO2 value.

    Data for a single species of shelled pteropods
    suggest that net shell dissolution occurs in live
    pteropods when the aragonite saturation is
    forced to <1.0 (Orr et al. 2005; Fabry et al. 2008). When live pteropods (Clio pyramidata) were collected in the subarctic Pacific and ex- posed to a level of aragonite undersaturation similar to that projected for Southern Ocean surface waters by the year 2100 under the IS92a emissions scenario, shell dissolution occurred within 48 hours, even though animals were ac- tively swimming.

    The response of planktonic calcifying or-
    ganisms to elevated pCO2 may not be uni-
    form among species or over time. To date,

    332 Annals of the New York Academy of Sciences

    published research indicates that most cal-
    careous plankton show reduced calcification
    in response to decreased carbonate ion con-
    centrations; however, the limited number of
    species investigated precludes identification of
    widespread or general trends. All studies thus
    far on the impacts of ocean acidification
    on calcareous plankton have been short-term
    experiments, ranging from hours to weeks.
    Nothing is known about the long-term impacts
    of elevated pCO2 on the reproduction, growth,
    and survivorship of planktonic calcifying or-
    ganisms or their ability to adapt to changing
    seawater chemistry. Chronic exposure to in-
    creased pCO2 may have complex effects on the
    growth and reproductive success of calcareous
    plankton or may induce adaptations that are
    absent in short-term experiments. No studies
    have investigated the possibility of differential
    impacts with life stage or age of the organism.
    Additional experimental evidence from plank-
    tonic calcifiers is urgently needed if we are
    to develop a predictive understanding of the
    impacts of ocean acidification on planktonic
    communities.

    Physiological Reponses

    Fishes

    Elevated CO2 partial pressures (hyper-
    capnia) will affect the physiology of water-
    breathing animals by inducing acidosis in the
    tissues and body fluids of marine organisms,
    including fishes (Roos & Boron 1981; Portner
    et al. 2004). pH, bicarbonate, and CO2 lev-
    els within the organism are altered with long-
    term effects on metabolic functions, growth,
    and reproduction, all of which could be harm-
    ful at population and species levels (Portner
    et al. 2004). Short-term effects of elevated CO2
    on fishes include alteration of the acid–base
    status, respiration, blood circulation, and ner-
    vous system functions, while long-term effects
    include reduced growth rate and reproduc-
    tion (Ishimatsu & Kita 1999; Ishimatsu et al.

    2004, 2005). Most experiments undertaken to
    date involved altering pH to levels consistent
    with conditions that would be present if CO2
    were to be directly injected to the seafloor
    (pH ca. 5.8–6.2). These experiments have
    shown adverse negative effects of acidified sea-
    water on fish throughout their entire life cycle
    (eggs, larvae, juveniles, and adults) (Kikkawa
    et al. 2003, 2004; Ishimatsu et al. 2004; Portner
    et al. 2004).

    Fish in early developmental stages are more
    sensitive to environmental change than adults
    and a limited number of studies have shown
    this to be true when fish eggs, larvae, and juve-
    niles were exposed to elevated CO2 (McKim
    1977; Kikkawa et al. 2003, 2004; Ishimatsu
    et al. 2004). Ishimatsu and co-authors (2004)
    state, “Even if the severity of environmental
    hypercapnia due to CO2 sequestration is made
    tolerable to adults, a gradual reduction in pop-
    ulation size and changes in marine ecosystem
    structures are unavoidable consequences when
    young individuals cannot survive” (p. 732). The
    long-term effects and adaptation potential of
    fishes experiencing future pCO2 levels consis-
    tent with IPCC scenarios are not known.

    Photosynthetic Organisms

    Phytoplankton and Cyanobacteria

    Most species of marine phytoplankton have
    carbon-concentrating mechanisms that accu-
    mulate inorganic carbon either as CO2 or
    HCO−3 or both (Giordano et al. 2005). Owing
    in large part to their carbon-acquisition mech-
    anisms and efficiencies, most marine phyto-
    plankton tested to date in single-species lab ex-
    periments or natural community-perturbation
    experiments show either no change or small
    increases (generally ≤ 10%) in photosynthetic
    rates when grown under high pCO2 condi-
    tions equivalent to ca. 760 micro atmosphere
    (µatm) (Tortell et al. 1997; Hein & Sand-Jensen
    1997; Burkhardt et al. 2001; Tortell and Morel
    2002; Rost et al. 2003; Beardall & Raven 2004;

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 333

    Schippers et al. 2004; Giordano et al. 2005;
    Martin & Tortell 2006). Unlike other major
    phytoplankton groups investigated thus far, the
    coccolithophorid Emiliania huxleyi has low affin-
    ity for inorganic carbon and could be carbon-
    limited in today’s ocean (Rost & Riebesell
    2004). Whether E. huxleyi will show increased
    rates of photosynthesis with progressive oceanic
    uptake of atmospheric CO2, however, may de-
    pend on nutrient availability and light condi-
    tions (Zondervan 2007). In a recent mesocosm
    CO2 manipulation, study, Riebesell and col-
    leagues (2007) reported that CO2 uptake by a
    phytoplankton community (primarily diatoms
    and coccolithophores) in experimental pCO2
    treatments of 700 and 1050 µatm was 27% and
    39% higher, respectively, relative to the pCO2
    treatment of 350 µatm.

    Ocean acidification will be accompanied
    by climate warming in large expanses of the
    oceans. Higher sea-surface temperatures in-
    crease thermal stratification of the upper ocean,
    thereby reducing the vertical mixing of nutri-
    ents to surface waters, and have been linked to
    observed decreases in phytoplankton biomass
    and productivity, particularly at low and mid-
    latitudes (Behrenfeld et al. 2006). In warm,
    nutrient-poor tropical and subtropical regions,
    however, continued ocean absorption of an-
    thropogenic CO2 may enhance fixation of
    atmospheric nitrogen and could lead to in-
    creased total primary productivity. Nitrogen-
    fixing cyanobacteria in the genus Trichodesmium,
    which support a large portion of primary pro-
    ductivity in such low-nutrient areas of the
    world’s oceans, show increased rates of nitro-
    gen and carbon fixation under elevated pCO2
    (Hutchins et al. 2007; Barcelos e Ramos et al.
    2007). At CO2 levels of 750 ppmv, Trichodesmium
    increased N2 fixation rates by 35–100% and
    CO2 fixation rates by 15–128%, relative to
    present-day CO2 conditions (Hutchins et al.
    2007).

    In a review of coastal marine phytoplankton,
    Hinga (2002) found that while some species
    grow well at a wide range of pH, others have
    growth rates that vary greatly over a 0.5 to

    1.0 pH unit change. He concluded that small
    changes in ambient seawater pH could affect
    species growth rates, abundances, and succes-
    sion in coastal phytoplankton communities. Eu-
    trophication and ocean acidification may act
    in concert to amplify the pH range found in
    coastal habitats, which in turn could lead to
    increased frequency of blooms of those species
    with tolerance to extreme pH (cf. Hinga 2002).
    In both coastal and open ocean environments,
    ocean acidification could also affect primary
    productivity through pH-dependent speciation
    of nutrients and metals (Zeebe & Wolf-Gladrow
    2001; Huesemann et al. 2002).

    Seagrasses

    Seagrasses represent one of the most bio-
    logically rich and productive marine ecosys-
    tems in the ocean. They create critical nursery
    grounds for juvenile fishes and important habi-
    tat for adult fishes, invertebrates, and mollusks.
    Several higher order and endangered species
    rely on seagrasses for a significant portion of
    their diet (e.g., dugongs, manatees, and green
    sea turtles). Seagrass ecosystems are a critical
    component to maintaining the biological diver-
    sity of the oceans and could be one of the few
    ecosystems that stand to benefit from increas-
    ing levels of CO2 in seawater. Seagrasses are
    capable of dehydrating HCO−3 , but many ap-
    pear to use CO2 (aq) for at least 50% of their
    carbon requirements used for photosynthesis
    (Palacios & Zimmerman 2007). Zimmerman
    and colleagues (1997) found that short-term
    (ca. 45 days) CO2 (aq) enrichment increased
    photosynthetic rates and reduced light require-
    ments for eelgrass (Zostera marina L) shoots in
    laboratory experiments.

    Longer-term (1 year) experiments expos-
    ing Zostera marina L to CO2 (aq) concentra-
    tions of 36–1123 µM (pH 7.75–6.2) conducted
    by Palacios and Zimmerman (2007) resulted
    in higher reproductive output, an increase in
    below-ground biomass, and vegetative prolifer-
    ation of new shoots when light was in abundant
    supply. These findings suggest that as the CO2

    334 Annals of the New York Academy of Sciences

    content of the surface ocean rises, so too will
    the productivity of seagrass meadows, which
    in turn may positively influence invertebrate
    and fish populations. This increase in produc-
    tivity will probably be true for other seagrass
    species as most appear to be photosynthetically
    limited by the present-day availability of CO2
    (Durako 1993; Invers et al. 2001; Palacios and
    Zimmerman 2007). Palacios and Zimmerman
    (2007) noted that a significant indirect effect
    of increased eelgrass density could be an in-
    crease in sediment retention, which could lead
    to increased water clarity and an expansion in
    the depth distribution of eelgrasses to deeper
    waters.

    Community Impacts

    Seagrasses, Coral Reefs, and Fishes

    Seagrass meadows and mangroves provide
    important nursery areas for juvenile fishes,
    many of which migrate to coral reefs as adults,
    and enhance fish diversity and abundance on
    coral reefs adjacent to these ecosystems (Pollard
    1984; Parrish 1989; Beck et al. 2001; Sheridan
    & Hays 2003; Mumby et al. 2004; Dorenbosch
    et al. 2005). The net effect of increasing CO2 on
    seagrass ecosystems will probably be increased
    seagrass biomass and productivity, assuming
    water quality and clarity (low suspended sed-
    iment) are sufficient for photosynthesis to oc-
    cur. Under these conditions, it is probable that
    an increase in total seagrass area will lead to
    more favorable habitat and conditions for asso-
    ciated invertebrate and fish species. However,
    the net effect of ocean acidification on coral reef
    ecosystems will probably be negative as many
    reef-building marine calcifiers will be heavily
    impacted by the combined effects of increasing
    sea-surface temperatures (coral bleaching) and
    decreasing carbonate saturation states of sur-
    face waters in the coming decades (Guinotte
    et al. 2003; Buddemeier et al. 2004). The mag-
    nitude of both ecosystem responses to ocean
    acidification and other environmental changes
    working in synergy is difficult to predict as are

    the net effects on fish abundance and diver-
    sity. Predicting the net effects on fish popula-
    tions is further complicated by the plethora of
    unknowns surrounding the long-term effects of
    increasing CO2 on fish physiology, metabolism,
    and probable range shifts due to ocean
    warming.

    Cold-water Corals and Fishes

    The ecology and species relationships of
    cold-water coral ecosystems are not as ad-
    vanced as the state of knowledge for shallow-
    water coral-reef systems, which is due in large
    part to logistical challenges and the expense
    of operating vessels and submersibles in the
    deep sea. However, cold-water coral ecosys-
    tems are thought to provide important habi-
    tat, feeding grounds, and recruitment/nursery
    functions for many deep-water species, includ-
    ing several commercially important fish species
    (Mortensen 2000; Fossa et al. 2002; Husebo et al.
    2002; Roberts et al. 2006). Many of the species
    relationships are thought to be facultative, but
    nonetheless, high fish densities have been re-
    ported for these structure-forming ecosystems
    (Husebo et al. 2002; Costello et al. 2005; Stone
    2006). Populations of grouper, snapper, and
    amberjack use the Oculina varicosa reefs off
    the Florida coast as feeding and spawning ar-
    eas (Reed 2002), even though their numbers
    have been dramatically reduced by commer-
    cial and recreational fishing in recent decades
    (Koenig et al. 2000). Large aggregations of red-
    fish (Sebastes spp.), ling (Molva molva), and tusk
    (Brosme brosme Ascanius) have been documented
    in the Lophelia pertusa reefs of the North Atlantic
    (Husebo et al. 2002), and strong fish–coral asso-
    ciations exist in the cold-water coral ecosystems
    of the North Pacific (Stone 2006).

    Ocean acidification could have significant
    indirect effects on fishes and other deep-
    sea organisms that rely on cold-water coral
    ecosystems for protection and nutritional re-
    quirements. Roberts and Gage (2003) docu-
    mented over 1300 species living on the Lophelia

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 335

    pertusa reefs in the NE Atlantic. Future depth
    projections for the aragonite saturation hori-
    zons indicate 70% of cold-water scleractinians
    will be in undersaturated waters by the end of
    the century, and significant decreases in cal-
    cification rate could occur well before corals
    experience undersaturated conditions as arag-
    onite saturation state decreases progressively
    over time (Guinotte et al. 2006). Quantifying
    the indirect impacts of ocean acidification on
    coral-associated fishes is not possible due to
    uncertainties surrounding facultative and ob-
    ligate species relationships, but the net effects
    are likely to be negative as cold-water coral
    growth, distribution, and area decrease.

    Plankton

    If reduced calcification decreases a calci-
    fying organism’s fitness or survivorship, then
    some planktonic calcareous species may un-
    dergo shifts in their distributions as the inor-
    ganic carbon chemistry of seawater changes.
    Calcifying species that are CO−2 sensitive could
    potentially be replaced by noncalcifying species
    and/or those species not sensitive to elevated
    pCO2.

    By 2100, surface waters of polar and sub-
    polar regions are projected to become un-
    dersaturated with respect to aragonite (Orr
    et al. 2005). Pteropods are important com-
    ponents of the plankton in high-latitude sys-
    tems, with densities reaching thousands of in-
    dividuals m−3 (e.g., Bathmann et al. 1991;
    Pane et al. 2004). If pteropods require sea-
    water that is supersaturated with respect to
    aragonite, then their habitat would become in-
    creasingly limited, first vertically in the water
    column and then latitudinally, by the shoal-
    ing of the aragonite saturation horizon over
    the next century (Feely et al. 2004; Orr et al.
    2005). If high-latitude surface waters do be-
    come undersaturated with respect to arago-
    nite, pteropods could eventually be eliminated
    from such regions, with consequences to food-
    web dynamics and other ecosystem processes
    (Fabry et al. 2008). In the subarctic Pacific,

    for example, pteropods can be important prey
    for juvenile pink salmon (Oncorhynchus gobuscha),
    as well as chum and sockeye salmon, pollock,
    and other commercially important fishes (Ay-
    din pers. comm.). Armstrong and co-authors
    (2005) reported interannual variability in the
    diet of juvenile pink salmon, with a single
    species of pteropod (Limacina helicina) compris-
    ing 15 to 63% by weight of pink salmon di-
    ets during a 3-year study. Because Pacific pink
    salmon have a short, 2-year life cycle, prey
    quality and abundance during the salmon’s ju-
    venile stage may strongly influence the pink
    salmon’s adult population size and biomass
    (Aydin et al. 2005).

    Jellyfish blooms (scyphomedusae, hydrome-
    dusae, and cubomedusae) have increased over
    the last several decades (Purcell et al. 2007), but
    it is too soon to determine whether such recent
    jellyfish increases will persist or the populations
    will fluctuate with climatic regime shifts, par-
    ticularly those at decadal scales, as has been
    observed previously (Purcell 2005). Attrill and
    colleagues (2007) reported a significant corre-
    lation of jellyfish frequency in the North Sea
    from 1971 to 1995 with decreased pH (from
    8.3 to 8.1) of surface waters. Although the
    causative mechanism is not known, Attrill and
    colleagues (2007) suggest that projected climate
    change and declining ocean pH will increase
    the frequency of jellyfish in the North Sea
    over the next century. Jellyfish are both preda-
    tors and potential competitors of fish and may
    substantially affect pelagic and coastal ecosys-
    tems (Purcell & Arai 2001; Purcell 2005). It is
    important to resolve possible linkages between
    jellyfish blooms and ocean acidification and de-
    termine whether continued changes in the sea-
    water inorganic carbon system will exacerbate
    problematic increases in jellyfish that have been
    associated with climate change, overfishing, eu-
    trophication, and other factors (Purcell et al.
    2007).

    Planktonic ecosystems are complex nonlin-
    ear systems, and the consequences of ocean
    acidification on such ecosystems are largely un-
    known. Substantial changes to species diversity

    336 Annals of the New York Academy of Sciences

    and abundances, food-web dynamics, and
    other fundamental ecological processes could
    occur; however, the interactions and feedbacks
    among the effects of chronic, progressively in-
    creasing ocean acidification and other environ-
    mental variables are difficult to predict. Ecosys-
    tem responses will also depend on the ability of
    biota to adapt to seawater chemistry changes
    that are occurring at rates they have not en-
    countered in their recent evolutionary history
    (Siegenthaler et al. 2005). Future progress will
    likely require integrated approaches involving
    manipulative experiments, field observations,
    and models, particularly at regional scales.

    Summary and Conclusions

    The scientific knowledge base surround-
    ing the biological effects of ocean acidifica-
    tion is in its infancy and the long-term con-
    sequences of changing seawater chemistry on
    marine ecosystems can only be theorized. Most
    is known about the calcification response for
    shallow-water scleractinian corals. Some data
    sets allow the identification of “tipping points”
    or “thresholds” of seawater carbonate chem-
    istry when ocean acidification will cause net
    calcification rates to be less than net dissolu-
    tion rates in coral reef systems (Yates & Halley
    2006; Hoegh-Guldberg et al. 2007). In contrast,
    the potential effects ocean acidification may
    have for the vast majority of marine species
    are not known. Research into the synergistic
    effects of ocean acidification and other human-
    induced environmental changes (e.g., increas-
    ing sea temperatures) on marine food webs
    and the potential transformative effects these
    changes could have on marine ecosystems is
    urgently needed. It is important to have a firm
    understanding of the degree to which ocean
    acidification influences critical physiological
    processes such as respiration, photosynthesis,
    and nutrient dynamics, as these processes are
    important drivers of calcification, ecosystem
    structure, biodiversity, and ultimately ecosys-
    tem health.

    Future ocean acidification research needs in-
    clude increased resources and efforts devoted
    to lab, mesocosm, and in situ experiments, all
    of which will aid in determining the biological
    responses of marine taxa to increased pCO2.
    Mesocosm and in situ experiments may simu-
    late and/or provide more natural conditions
    than single-species lab experiments, but they
    have thus far used abrupt changes in seawater
    chemistry which do not allow for potential ac-
    climation or adaptation by marine organisms.
    There is an additional need for experiments
    on taxa with no commercial value but which
    provide critical habitat and occupy impor-
    tant trophic levels within marine food webs.
    Direct CO2 experiments on commercially im-
    portant species are clearly necessary, but non-
    commercial species play crucial roles in marine
    ecosystems and the life history of most com-
    mercial species. The effects of ocean acid-
    ification on less charismatic species and/or
    species with no economic value should not
    be overlooked. The biological response of ma-
    rine organisms (both commercial and noncom-
    mercial) to ocean acidification will be key to
    making informed policy decisions that con-
    form to sound ecosystem-based management
    principles.

    There is a critical need for well-developed
    spatial and temporal models that give accu-
    rate present day and future estimates of arago-
    nite and calcite saturation states in the coastal
    zones. The shallow continental shelves are
    some of the most biologically productive ar-
    eas in the sea and are home to the majority
    of the world’s fisheries, but accurate carbonate
    saturation state data do not currently exist for
    most coastal regions. Ocean acidification in-
    formation should also be integrated into exist-
    ing ecosystem models, which attempt to predict
    the effects of environmental changes on ma-
    rine populations and ecosystem structure (e.g.,
    Ecopath and Ecosim). Development of these
    tools is essential to making credible predictions
    of future ocean acidification effects on marine
    ecosystems and will aid in guiding management
    decisions.

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 337

    The overwhelming volume of scientific evi-
    dence collated by the IPCC documenting the
    dangers of human-induced climate change, of
    which ocean acidification is only one, should
    end the lingering CO2 emissions reduction de-
    bate. The global CO2 experiment which has
    been under way since the Industrial Revolution
    and the potentially dire consequences this
    uncontrolled experiment poses for marine or-
    ganisms and indeed, all life on Earth, leave no
    doubt that human dependence on fossil fuels
    must end as soon as possible. International col-
    laboration, political will, and large-scale invest-
    ment in clean energy technologies are essen-
    tial to avoiding the most damaging effects of
    human-induced climate change.

    Acknowledgments

    This work was supported in part by MCBI
    grants from the Edwards Mother Earth Foun-
    dation, Marisla Foundation, Moore Family
    Foundation, and Mark and Sharon Bloome.
    Support for VJF was provided in part by
    National Science Foundation grants OCE-
    0551726 and ANT-0538710. We would like
    to thank RW Buddemeier, RA Feely, and an
    anonymous reviewer for constructive inputs on
    an early draft.

    Conflict of Interest

    The authors declare no conflicts of interest.

    References

    Agegian, C.R. 1985. The Biogeochemical Ecology of
    Porolithon gardineri (Foslie). Ph.D. thesis, University
    of Hawaii, Honolulu.

    Armstrong, J.L. et al. 2005. Distribution, size, and interan-
    nual, seasonal and diel food habits of northern Gulf
    of Alaska juvenile pink salmon, Oncorhyncus gorbuscha.
    Deep-Sea Res. II 52: 247–265.

    Attrill, M., J. Wright & M. Edwards. 2007. Climate-
    related increases in jellyfish frequency suggest a more
    gelatinous future for the North Sea. Limnol. Oceanogr.
    52: 480–485.

    Augustin, L. et al. 2004. Eight glacial cycles from an
    Antarctic ice core. Nature 429: 623–628.

    Aydin, K.Y. et al. Linking oceanic foodwebs to coastal
    production and growth rates to Pacific salmon (On-
    corhynchus spp.), using models on three scales. Deep Sea
    Res. II 52: 757–780.

    Bamber, R.N. 1990. The effects of acidic seawater on 3
    species of lamellibranch mollusk. J. Exp. Mar. Biol.
    Ecol. 143: 181–191.

    Bamber, R.N. 1987. The effects of acidic sea water on
    young carpet-shell clams Venerupis decussata (L.)
    (Mollusca: Veneracea). J. Exp. Mar. Biol. Ecol. 108:
    241–260.

    Barcelos e Ramos, J. et al. 2007. Effect of rising atmo-
    spheric carbon dioxide on the marine nitrogen fixer
    Trichodesmium. Global Biogeochem. Cycles 21: np.

    Bathmann, U. et al. 1991. Short-term variations in par-
    ticulate matter sedimentation off Kapp Norvegia,
    Weddell Sea, Antarctica: relation to water mass ad-
    vection, ice cover, plankton biomass and feeding ac-
    tivity. Polar Biol. 11: 185–195.

    Beardall, J. & J.A. Raven. 2004. The potential effects of
    global climate change in microalgal photosynthesis,
    growth and ecology. Phycologia 43: 31–45.

    Beck, M.W. et al. 2001. The identification, conservation,
    and management of estuarine and marine nurseries
    for fish and invertebrates. Bioscience 51: 633–641.

    Beck, M.W. et al. 2003. The role of nearshore ecosystems
    as fish and shellfish nurseries. Issues Ecol. 11: 1–12.

    Beerling, D.J. & R.A. Berner. 2002. Biogeochemical con-
    straints on the Triassic-Jurassic boundary carbon cy-
    cle event. Global Biogeochem. Cycles 16: 101–113.

    Behrenfeld, M.J. et al. 2006. Climate-driven trends in
    contemporary ocean productivity. Nature 444: 752–
    755.

    Berge, J.A. et al. 2006. Effects of increased sea water con-
    centrations of CO2 on growth of the bivalve Mytilus
    edulis L. Chemosphere 62: 681–687.

    Bibby, R. et al. 2007. Ocean acidification disrupts induced
    defences in the intertidal gastropod Littorina littorea.
    Biol. Lett. 3: 699–701.

    Bijma, J., H.J. Spero, & D.W. Lea. 1999. Reassessing
    foraminiferal stable isotope geochemistry: Impact of
    the oceanic carbonate system (experimental results).
    In Use of Proxies in Paleoceanography: Examples from the
    South Atlantic. G. Fischer & G. Wefer, Eds.: 489–512.
    Springer-Verlag. New York.

    Bijma, J., B. Honisch, & R.E. Zeebe. 2002. The im-
    pact of the ocean carbonate chemistry on living
    foraminiferal shell weight: Comment on “Carbon-
    ate ion concentration in glacial-age deep waters of
    the Caribbean Sea” by W.S. Broecker & E. Clark.
    Geochem. Geophys. Geosyst. 3: 1064.

    Borowitzka, L.J. & A.W.D. Larkum 1986. Reef algae.
    Oceanus 29: 49–54.

    338 Annals of the New York Academy of Sciences

    Broecker, W.S. 2003. The Ocean CaCO3 Cycle. In The Oceans
    and Marine Geochemistry. H. Elderfield, Ed.: 1–21.
    Treatise on Geochemistry, Elsevier Pergamon. Ox-
    ford.

    Broecker, W.S. et al. 1979. Fate of fossil fuel carbon dioxide
    and the global carbon budget. Science 206: 409–418.

    Broecker, W.S. & T. Takahashi. 1966. Calcium carbonate
    precipitation on the Bahama Banks. J. Geophys. Res.
    71: 1575–1602.

    Broecker, W.S. & T. Takahashi. 1977. Neutralization of
    fossil fuel CO2 by marine calcium carbonate. In The
    Fate of Fossil Fuel in the Oceans. N.R. Andersen & A.
    Malahoff, Eds.: 213–241. Plenum Press. New York.

    Buddemeier, R.W., J.A. Kleypas & R. Aronson. 2004.
    Coral Reefs and Global Climate Change. Potential
    Contributions of Climate Change to Stresses on
    Coral Reef Ecosystems. Pew Center for Global Climate
    Change. Arlington, VA. 42 pp.

    Buddemeier, R.W. 2007. The future of tropical reefs and coast-
    lines. Presented at the American Association for the Advance-

    ment of Science Annual Meeting. San Francisco, CA, Feb
    16.

    Burkhardt, S. et al. 2001. CO2 and HCO3 uptake in ma-
    rine diatoms acclimated to different CO2 concentra-
    tions. Limnol. Oceanogr. 46: 1378–1391.

    Cairns, S.D. & I.G. Macintyre. 1992. Phylogenetic im-
    plications of calcium carbonate mineralogy in the
    Stylasteridae (Cnidaria:Hydrozoa). Palaios 7: 96–
    107.

    Caldeira, K. & M.R. Rampino. 1993. Aftermath of the
    end-Cretaceous mass extinction: possible biogeo-
    chemical stabilization of the carbon cycle and cli-
    mate. Paleoceanography 8: 515–525.

    Caldeira, K. & M.E. Wickett. 2003. Anthropogenic car-
    bon and ocean pH. Nature 425: 365.

    Caldeira, K. & M.E. Wickett. 2005. Ocean model pre-
    dictions of chemistry changes from carbon dioxide
    emissions to the atmosphere and ocean. J. Geophys.
    Res. 110: np.

    Caldeira, K. et al. 2007. Comment on “Modern-
    age buildup of CO2 and its effects on seawa-
    ter acidity and salinity” by Hugo A. Loaiciga.
    Geophysical Research Letters 34: L18608. doi:
    10.1029/2006GL027288.

    Chisholm, J.R.M. 2000. Calcification by crustose
    coralline algae on the northern Great Barrier Reef,
    Australia. Limnol. Oceanogr. 45: 1476–1484.

    Costello, M. et al. 2005. Role of cold-water Lophelia
    pertusa coral reefs as fish habitat in the NE At-
    lantic. In Cold-water Corals and Ecosystems. A. Freiwald
    & J.M. Roberts, Eds.: 771–805. Springer-Verlag.
    Berlin, Heidelberg.

    Davies, P.J. & J.F. Marshall. 1985. Halimeda bioherms – low
    energy reefs, northern Great Barrier Reef. Proceedings
    of the Fifth International Coral Reef Congress 1: 1–7.

    Delille, B. et al. 2005. Response of primary production
    and calcification to changes of pCO2 during exper-
    imental blooms of the coccolithophorid Emiliania
    huxleyi. Global Biogeochem. Cycles 19: np.

    Diaz-Pulido, G. et al. 2007. Vulnerability of macroalgae of
    the Great Barrier Reef to climate change. In Climate
    Change and the Great Barrier Reef . J.E. Johnson & P.A.
    Marshall, Eds.: 154–192. Great Barrier Reef Marine
    Park Authority and Australian Greenhouse Office,
    Australia.

    Dickson, A.G. & F.J. Millero. 1987. A comparison of the
    equilibrium constants for the dissociation of carbonic
    acid in seawater media. Deep-Sea Res. A 34: 1733–
    1743.

    Doney, S.C. et al. 2007. The impacts of anthropogenic
    nitrogen and sulfur deposition on ocean acidification
    and the inorganic carbon system. Proc. Natl. Acad. Sci.
    104: 14580–14585.

    Dorenbosch, M. et al. 2005. Indo-Pacific seagrass beds and
    mangroves contribute to fish density and diversity on
    adjacent coral reefs. Mar. Ecol. Prog. Ser. 302: 63–76.

    Drew, E.A., & K.M. Abel. 1988. Studies on Halimeda
    I. The distribution and species composition of Hal-
    imeda meadows throughout the Great Barrier Reef
    Province. Coral Reefs 6: 195–205.

    Durako, M.J. 1993. Photosynthetic utilization of CO2 (aq)
    and HCO3 in Thalassia testudinum (Hydrocharita-
    cae). Mar. Biol. 115: 373–380.

    Edmunds, P.J. 2007. Evidence for a decadal-scale decline
    in the growth rates of juvenile scleractinian corals.
    Mar. Ecol. Prog. Ser. 341: 1–13.

    Fabry, V.J. et al. 2008. Impacts of ocean acidification on
    marine fauna and ecosystems processes. J. Mar. Sci.
    In Press.

    Feely, R.A. et al. 1988. Winter-summer variations of calcite
    and aragonite saturation in the northeast Pacific.
    Mar. Chem. 25: 227–241.

    Feely, R.A., & C.T.A. Chen. 1982. The effect of excess
    CO2 on the calculated calcite and aragonite satura-
    tion horizons in the Northeast Pacific. Geophys. Res.
    Lett. 9: 1294–1297.

    Feely, R.A. et al. 2008. Present and future changes in
    seawater chemistry due to ocean acidification. AGU
    Monograph, The Science and Technology of CO2 Seques-
    tration. In Press.

    Feely, R.A. et al. 2004. Impact of anthropogenic CO2
    on the CaCO3 system in the oceans. Science 305:
    362–366.

    Feely, R.A. et al. 1984. Factors influencing the degree of
    saturation of the surface and intermediate waters of
    the North Pacific Ocean with respect to aragonite.
    J. Geophys. Res. 89: 631–640.

    Fine, M. & D. Tchernov. 2007a. Scleractinian coral
    species survive and recover from decalcification.
    Science 315: 1811.

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 339

    Fine, M. & D. Tchernov. 2007b. Response to “Ocean
    acidification and scleractinian corals”. Science 317:
    1032–1033.

    Fossa, J.H., P.B. Mortensen, & D.M. Furevik. 2002. The
    deep-water coral Lophelia pertusa in Norwegian wa-
    ters: distribution and fishery impacts. Hydrobiologia
    13: 1–12.

    Freiwald, A. 2002. Reef-forming cold-water corals. In
    Ocean Margin Systems. G. Wefer, D. Billett, D.B.B.
    Hebbeln Jorgensen, et al., Eds.: 365–385. Springer.
    Heidelberg.

    Freiwald, A. et al. 2004. Coldwater Coral Reefs. UNEP-
    WCMC, Cambridge, UK.

    Freiwald, A. & J.M. Roberts. 2005. Cold-Water Corals and
    Ecosystems. Springer. Heidelberg.

    Gattuso, J.-P. et al. 1998. Effect of calcium carbonate satu-
    ration of seawater on coral calcification. Global Planet.
    Change 18: 37–46.

    Gazeau, F. et al. 2007. Impact of elevated CO2 on
    shellfish calcification. Geophys. Res. Lett. L07603,
    doi:10.1029/2006GL028554.

    Giordano, M., J. Beardall & J.A. Raven. 2005. CO2 con-
    centrating mechanisms in algae: mechanisms, en-
    vironmental modulation, and evolution. Annu. Rev.
    Plant Biol. 56: 99–131.

    Green, M.A. et al. 2004. Dissolution mortality of juvenile
    bivalves in coastal marine deposits. Limnol. Oceanogr.
    49: 727–734.

    Guinotte, J.M., R.W. Buddemeier, & J.A. Kleypas. 2003.
    Future coral reef habitat marginality: temporal and
    spatial effects of climate change in the Pacific basin.
    Coral Reefs 22: 551–58.

    Guinotte, J.M. et al. 2006. Will human induced changes
    in seawater chemistry alter the distribution of deep-
    sea scleractinian corals? Front. Ecol. Environ. 4: 141–
    146.

    Harrington, L. et al. 2005. Synergistic effects of diuron
    and sedimentation on photosynthesis and survival of
    crustose coralline algae. Mar. Pollut. Bull. 51: 415–
    427.

    Heifetz, J. et al. 2005. Corals of the Aleutian Islands. Fish-
    eries Oceanogr. 14: 131–138.

    Hein, M. & K. Sand-Jensen. 1997. CO2 increases oceanic
    primary production. Nature 388: 526.

    Heyward, A.J. & A.P. Negri. 1999. Natural inducers of
    coral larval metamorphosis. Coral Reefs 18: 273–
    279.

    Hillis-Colinvaux, L. 1980. Ecology and taxonomy of Hal-
    imeda: primary producers of coral reefs. Adv. Mar. Biol.
    17: 1–327.

    Hoegh-Guldberg, O. et al. 2007. Coral reefs under rapid
    climate change and ocean acidification. Science 318:
    1737–1742.

    Hinga, K.R. 2002. Effects of pH on coastal marine phy-
    toplankton. Mar.Ecol.Prog.Ser. 238: 281–300.

    Hossain, M.M.M. & S. Ohde. 2006. Calcification of cul-
    tured Porites and Fungia under different aragonite
    saturation states of seawater. Proc. 10th Int.Coral
    Reef Sym. Jpn. Coral Reef Soc. 597–606.

    Houghton, J.T. et al. 2001. Climate Change 2001: The Scientific
    Basis. Cambridge University Press. Cambridge, UK.

    Huesemann, M.H., A.D. Skillman, & E.A. Crecelius.
    2002. The inhibition of marine nitrification by ocean
    disposal of carbon dioxide. Mar. Pollut. Bull. 44: 142–
    148.

    Husebo, A. et al. 2002. Distribution and abundance of fish
    in deep-sea coral habitats. Hydrobiologia 471: 91–99.

    Hutchins, D.A. et al. 2007. CO2 control of Trichodesmium
    N2 fixation, photosynthesis, growth rates, and el-
    emental ratios: implications for past, present, and
    future ocean biogeochemistry. Limnol. Oceanogr. 52:
    1293–1304.

    Invers, O. et al. 2001. Inorganic carbon sources for sea-
    grass photosynthesis: an experimental evaluation of
    bicarbonate use in species inhabiting temperate wa-
    ters. J. Exp. Mar. Biol. Ecol. 265: 203–217.

    Ishimatsu, A. et al. 2005. Physiological effects on fishes in
    a high-CO2 world. J. Geophys. Res. 110: np.

    Ishimatsu, A. et al. 2004. Effects of CO2 on marine fish:
    larvae and adults. J. Oceanogr. 60: 731–741.

    Ishimatsu, A. & J. Kita. 1999. Effects of environmental
    hypercapnia on fish, Jpn. J. Ichthyol. 46: 1–13.

    Kikkawa, T., A. Ishimatsu & J. Kita. 2003. Acute CO2
    tolerance during the early developmental stages of
    four marine teleosts. Environ. Toxicol. 18: 375–382.

    Kikkawa, T., J. Kita, & A. Ishimatsu. 2004. Comparison
    of the lethal effect of CO2 and acidification on red
    sea bream (Pagrus major) during the early develop-
    mental stages. Mar. Pollut. Bull. 48: 108–110.

    Kleypas, J.A. et al. 1999a. Geochemical consequences of
    increased atmospheric carbon dioxide on coral reefs.
    Science 284: 118–120.

    Kleypas, J.A. et al. 1999b. Environmental limits to coral
    reef development: where do we draw the line? Am
    Zool 39: 146–159.

    Kleypas, J.A. et al. 2006. Impacts of Ocean Acidifica-
    tion on Coral Reefs and Other Marine Calcifiers:
    A Guide for Future Research, report of a workshop
    held 18–20 April 2005, St. Petersburg, FL, sponsored
    by NSF, NOAA, and the U.S. Geological Survey, 88
    pp.

    Kleypas, J.A. & C. Langdon. 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 Envi-
    ronment, Indonesian Institute of Sciences, Interna-
    tional Society for Reef Studies, Bali, Indonesia.

    340 Annals of the New York Academy of Sciences

    Koenig, C. et al. 2000. Protection of fish spawning habitat
    for the conservation of warm- temperate reef-fish
    fisheries of shelf-edge reefs of Florida. Bull. Mar. Sci.
    66: 593–616.

    Kuffner, I.B. et al. 2008. Decreased abundance of crus-
    tose coralline algae due to ocean acidification. Nature
    Geoscience In Press.

    Kurihara, H., S. Kato, & A. Ishimatsu. 2007. Effects of
    increased seawater pCO2 on early development of
    the oyster Crassostrea gigas. Aquat. Biol. 1: 91–98.

    Kurihara, H., S. Shimode, & Y. Shirayama. 2004. Sub-
    lethal effects of elevated concentration of CO2 on
    planktonic copepods and sea urchins. J. Oceanogr. 60:
    743–750.

    Kurihara, H. & Y. Shirayama. 2004. Effects of increased
    atmospheric CO2 on sea urchin early development.
    Mar. Ecol. Prog. Ser. 274: 161–169.

    Kuwatani, Y. & T. Nishii. 1969. Effects of decreased pH
    of culture water on the growth of the Japanese pearl
    oyster. Bull. Jap. Soc. Sci. Fish. 35: 342–350.

    Langdon, C. et al. 2000. Effect of calcium carbonate sat-
    uration state on the calcification rate of an experi-
    mental coral reef. Global Biogeochem. Cy. 14: 639–654.

    Langdon, C. et al. 2003. Effect of elevated CO2 on
    the community metabolism of an experimental
    coral reef. Global Biogeochem. Cy. 17: 1011, doi:
    10.1029/2002GB001941.

    Langdon, C. & M.J. Atkinson. 2005. Effect of elevated
    pCO2 on photosynthesis and calcification of corals
    and interactions with seasonal change in tempera-
    ture/irradiance and nutrient enrichment. J. Geophys.
    Res. 110: np.

    Langer, M.R. et al. 2006. Species-specific responses of cal-
    cifying algae to changing seawater carbonate chem-
    istry. Geochem. Geophys. Geosyst. 7: np.

    Leclercq, N., J.-P. Gattuso, & J. Jaubert. 2000. CO2 par-
    tial pressure controls the calcification rate of a coral
    community. Global Change Biol. 6: 329–334.

    Leclercq, N. et al. 2002. Primary production, respiration,
    and calcification of a coral reef mesocosm under
    increased CO2 partial pressure. Limnol. Oceanogr. 47:
    558–564.

    Littler, M.M. & D.S. Littler. 1984. Models of tropical reef
    biogenesis: the contribution of algae. Prog. Phycol. Res.
    3: 323–364.

    Lueker, T.J., A.G. Dickson, & C.D. Keeling. 2000. Ocean
    pCO2 calculated from dissolved inorganic carbon,
    alkalinity, and equations for K1 and K2: validation
    based on laboratory measurements of CO2 in gas
    and seawater at equilibrium. Mar. Chem. 70: 105–
    119.

    Marshall, A.T. & P.L. Clode. 2002. Effect of increased
    calcium concentration in sea water on calcification
    and photosynthesis in the scleractinian coral Galaxea
    fascicularis. J. Exp. Biol. 205: 2107–2113.

    Martin, C.L. & P.D. Tortell. 2006. Bicarbonate trans-
    port and extracellular carbonic anhydrase activity in
    Bering Sea phytoplankton assemblages: Results from
    isotope disequilibrium experiments. Limnol. Oceanogr.
    51: 2111–2121.

    Marubini, F., C. Ferrier-Pagés & J.P. Cuif. 2003. Sup-
    pression of skeletal growth in scleractinian corals by
    decreasing ambient carbonate-ion concentration: a
    crossfamily comparison. Proc. Roy. Soc. Lond. B 270:
    179–184.

    Marubini, F., H. Barnett, C. Langdon & M.J. Atkinson.
    2001. Dependence of calcification on light and car-
    bonate ion concentration for the hermatypic coral
    Porites compressa. Mar. Ecol.-Prog. Ser. 220: 153–162.

    Marubini, F. & B. Thake. 1999. Bicarbonate addition
    promotes coral growth. Limnol. Oceanogr. 44: 716–
    720.

    McKim, J.M. 1977. Evaluation of tests with early life
    stages of fish for predicting long- term toxicity. J.
    Fish. Res. Board Can. 34: 1148–1154.

    Medina, M.A. et al. 2006. Naked corals: skeleton loss in
    Scleractinia. Proc. Natl. Acad. Sci. USA 103: 9096–
    9100.

    Meehl, G.A. et al. 2007. Global Climate Projections. In
    Climate Change 2007: The Physical Science Basis. Contribu-

    tion of Working Group I to the Fourth Assessment Report of the

    Intergovernmental Panel on Climate Change. S. Solomon,
    D. Qin, M. Manning, et al., Eds.: Cambridge Uni-
    versity Press. Cambridge, UK, and New York, NY.

    Mehrbach, C. et al. 1973. Measurement of apparent
    dissociation constants of carbonic acid in seawater
    at atmospheric pressure. Limnol. Oceanogr. 18: 897–
    907.

    Michaelidis, B.C. et al. 2005. Effects of long-term moder-
    ate hypercapnia on acid-base balance and growth
    rate in marine mussels Mytilus galloprovincialis,
    Mar. Ecol. Prog. Ser. 293: 109–118.

    Milliman, J.D. & A.W. Droxler. 1996. Neritic and pelagic
    carbonate sedimentation in the marine environment:
    Ignorance is not bliss. Geol. Rundsch. 85: 496–504.

    Mortensen, P.B. et al. 2001. Distribution, abundance and
    size of Lophelia pertusa coral reefs in mid-Norway in
    relation to seabed characteristics. J. Mar. Biol. Assoc.
    UK. 81: 581–597.

    Mortensen, P.B. 2000. Lophelia pertusa (Scleractinia) in
    Norwegian waters; distribution, growth, and asso-
    ciated fauna. Ph.D. thesis, University of Bergen,
    Bergen, Norway.

    Mucci, A. 1983. The solubility of calcite and aragonite
    in seawater at various salinities, temperatures, and
    one atmosphere total pressure. Am. J. Sci. 283: 780–
    799.

    Mumby, P.J. et al. 2004. Mangroves enhance the biomass
    of coral reef fish communities in the Caribbean.
    Nature 427: 533–536.

    Guinotte & Fabry: Ocean Acidification and Marine Ecosystems 341

    Ohde, S. & M.M.M. Hossain. 2004. Effect of CaCO3
    (aragonite) saturation state of seawater on calcifica-
    tion of Porites coral. Geochem. J. 38: 613–621.

    Ohde, S. & R. Van Woesik. 1999. Carbon dioxide flux
    and metabolic processes of a coral reef. Bull. Mar. Sci.
    65: 559–576.

    Orr, J.C. et al. 2005. Anthropogenic ocean acidification
    over the twenty-first century and its impact on calci-
    fying organisms. Nature 437: 681–686.

    Palacios, S. & R.C. Zimmerman. 2007. Response of eel-
    grass Zostera marina to CO2 enrichment: possible
    impacts of climate change and potential for reme-
    diation of coastal habitats. Mar. Ecol. Prog. Ser. 344:
    1–13.

    Pane, L. et al. 2004. Summer coastal zooplankton biomass
    and copepod community structure near the Italian
    Terra Nova Base (Terra Nova Bay, Ross Sea, Antarc-
    tica). J. Plank. Res. 26: 1479–1488.

    Parrish, J.D. 1989. Fish communities of interacting
    shallow-water habitats in tropical oceanic regions.
    Mar. Ecol. Prog. Ser. 58: 143–160.

    Petit, J.R. et al. 1999. Climate and atmospheric history
    of the past 420000 years from the Vostok ice core,
    Antarctica. Nature 399: 429–436.

    Pollard, D.A. 1984. A review of ecological studies on sea-
    grass fish communities, with particularly reference to
    recent studies in Australia. Aquat. Bot. 18: 3–42.

    Portner, H.O., M. Langenbuch & A. Reipschlager. 2004.
    Biological impacts of elevated ocean CO2 concen-
    trations: lessons from animal physiology and earth
    history. J. Oceanogr. 60: 705–718.

    Purcell, J.E. 2005. Climate effects on formation of jellyfish
    and ctenophore blooms: a review. J. Mar Biol. Ass. UK
    85: 461–476.

    Purcell, J.E. & M.N. Arai. 2001. Interactions of pelagic
    cnidarians and ctenophores with fishes: a review.
    Hydrobiologia 451: 27–44.

    Purcell, J.E., S. Uye & W.-T. Lo. 2007. Anthropogenic
    causes of jellyfish blooms and their direct conse-
    quences for humans: a review. Mar. Ecol. Prog. Ser.
    350: 153–174.

    Reed, J.K. 2002. Deep-water Oculina coral reefs of
    Florida: biology, impacts and management. Hydro-
    biologia 471: 43–55.

    Renegar, D.A. & B.M. Riegl. 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. et al. 2003. Interacting effects of CO2 partial
    pressure and temperature on photosynthesis and cal-
    cification in a scleractinian coral. Global Change Biol.
    9: 1660–1668.

    Riebesell, U. et al. 2000. Reduced calcification of marine
    plankton in response to increased atmospheric CO2.
    Nature 407: 364–367.

    Riebesell, U. et al. 2007. Enhanced biological carbon con-
    sumption in a high CO2 ocean. Nature 450: 545–548.

    Roberts, J.M., J.D. Gage & the ACES party. 2003. As-
    sessing biodiversity associated with cold-water coral
    reefs: Pleasures and pitfalls. Erlanger Geol. Abh. 4: 73.

    Roberts, J.M., A.J. Wheeler, & A. Freiwald. 2006. Reefs
    of the deep: the biology and geology of cold-water
    coral ecosystems. Science 312: 543–547.

    Rogers, A.D. 1999. The biology of Lophelia pertusa (Lin-
    naeus 1758) and other deep- water reef-forming
    corals and impacts from human activities. Int. Rev.
    Hydrobiol. 84: 315–406.

    Roos, A. & W.F. Boron. 1981. Intracellular pH. Physiol.
    Rev. 61: 296–434.

    Rost, B., U. Riebesell & S. Burkhardt. 2003. Carbon ac-
    quisition of bloom-forming marine phytoplankton.
    Limnol. Oceanogr. 48: 55–67.

    Rost, B. & U. Riebesell. 2004. Coccolithophores and
    the biological pump: responses to environmental
    changes. In Coccolithophores—From Molecular Processes
    to Global Impact. H.R. Thierstein & J.R. Young, Eds.:
    76–99. Springer. Berlin Heidelberg.

    Royal Society. 2005. Ocean Acidification Due to Increasing At-
    mospheric Carbon Dioxide. Policy Document 12/05. The
    Royal Society. London, UK.

    Sabine, C.L. et al. 2004. The oceanic sink for anthro-
    pogenic CO2. Science 305: 367–371.

    Schippers, P., M. Lurling & M. Scheffer. 2004. Increase of
    atmospheric CO2 promotes phytoplankton produc-
    tivity. Ecol. Lett. 7: 446–451.

    Schneider, K. & J. Erez. 2006. The effect of carbon-
    ate chemistry on calcification and photosynthesis in
    the hermatypic coral Acropora eurystoma. Limnol.
    Oceanogr. 51: 1284–1293.

    Sciandra, A. et al. 2003. Response of coccolithophorid
    Emiliania huxleyi to elevated partial pressure of CO2
    under nitrogen limitation. Mar. Ecol.-Prog. Ser. 261:
    111–122.

    Sheridan, P. & C. Hays. 2003. Are mangroves nursery
    habitat for transient fishes and decapods? Wetlands
    23: 449–458.

    Siegenthaler, U. et al. 2005. Stable carbon cycle-climate
    relationship during the late Pleistocene. Science 310:
    1313–1317.

    Silverman, J., B. Lazar & J. Erez. 2007. Effect of aragonite
    saturation, temperature, and nutrients on the com-
    munity calcification rate of a coral reef. J. Geophys.
    Res. 112: C05004, doi:10.1029/2006JC003770.

    Skirrow, G. & M. Whitfield. 1975. The effect of in-
    creases in the atmospheric carbon dioxide content
    on the carbonate ion concentration of surface water
    at 25◦C. Limnol. Oceanogr. 20: 103–108.

    Spero, H.J. et al. 1997. Effect of seawater carbonate con-
    centration on foraminiferal carbon and oxygen iso-
    topes. Nature 390: 497–500.

    342 Annals of the New York Academy of Sciences

    Stanley, G.D., Jr. 2006. Photosymbiosis and the evolution
    of modern coral reefs. Science 312: 857–858.

    Stanley, G.D., Jr. 2007. Ocean acidification and sclerac-
    tinian corals. Science 317: 1032–1033.

    Stanley, G.D., Jr. & D.G. Fautin. 2001. The Origin of
    Modern Corals. Science 291: 1913–1914.

    Stone, R.P. 2006. Coral habitat in the Aleutian Islands
    of Alaska: depth distribution, fine-scale species as-
    sociations, and fisheries interactions. Coral Reefs 25:
    229–238.

    Tortell, P.D. & F.M.M. Morel. 2002. Sources of inorganic
    carbon for phytoplankton in the eastern Subtropical
    and Equatorial Pacific Ocean. Limnol. Oceanogr. 47:
    1012–1022.

    Tortell, P.D., J.R. Reinfelder, & F.M.M. Morel. 1997. Ac-
    tive uptake of bicarbonate by diatoms. Nature 390:
    243–244.

    Turley, C.M., J.M. Roberts, & J.M. Guinotte. 2007.
    Corals in deep-water: will the unseen hand of ocean
    acidification destroy cold-water ecosystems? Coral
    Reefs 26: 445–448.

    Veron, J.E. 2008. A Reef in Time: The Great Barrier Reef
    from Beginning to End. Harvard University Press. Cam-
    bridge, MA.

    Whitfield, M. 1975. Future impact of fossil CO2 on sea–
    Reply. Nature 254: 274–275.

    Yates, K.K. & R.B. Halley. 2006. CO3 concentration and
    pCO2 thresholds for calcification and dissolution on
    the Molokai reef flat, Hawaii. Biogeosciences 3: 357–
    369.

    Zachos, J.C. et al. 2005. Rapid acidification of the ocean
    during the Paleocene-Eocene Thermal Maximum.
    Science 308: 1611–1615.

    Zeebe, R.E. & D. Wold-Gladrow. 2001. CO2 in Seawater:
    Equilibrium, Kinetics, Isotopes. Elsevier Oceanography
    Series 65. Amsterdam. pp. 346.

    Zimmerman, R.C. et al. 1997. Impacts of CO2-
    enrichment on productivity and light requirements
    of eelgrass. Plant Physiol. 115: 599–607.

    Zondervan, I. 2007. The effects of light, macronutri-
    ents, trace metals and CO2 on the production of
    calcium carbonate and organic carbon in coccol-
    ithophores – A review. Deep Sea Res. II 54: 521–
    537.

    Zondervan, I. et al. 2001. Decreasing marine biogenic
    calcification: A negative feedback on rising at-
    mospheric pCO2. Global Biogeochem. Cy. 15: 507–
    516.

    MINIREVIE

    W

    EFFECTS OF CLIMATE CHANGE ON GLOBAL SEAWEED COMMUNITIES1

    Christopher D. G. Harley,2 Kathryn M. Anderson, Kyle W. Demes, Jennifer P. Jorve, Rebecca L. Kordas,
    Theraesa A. Coyle

    Department of Zoology and Biodiversity Research Centre, University of British Columbia, 6270 University Blvd, Vancouver,

    British Columbia, V6T1Z4, Canada

    and Michael H. Graham

    Moss Landing Marine Laboratories, 8272 Moss Landing Road, Moss Landing, California, 95039, USA

    Seaweeds are ecologically important primary
    producers, competitors, and ecosystem engineers that
    play a central role in coastal habitats ranging from kelp
    forests to coral reefs. Although seaweeds are known to
    be vulnerable to physical and chemical changes in the
    marine environment, the impacts of ongoing and
    future anthropogenic climate change in seaweed-
    dominated ecosystems remain poorly understood. In
    this review, we describe the ways in which changes in
    the environment directly affect seaweeds in terms of
    their physiology, growth, reproduction, and survival.
    We consider the extent to which seaweed species may
    be able to respond to these changes via adaptation or
    migration. We also examine the extensive reshuffling
    of communities that is occurring as the ecological
    balance between competing species changes, and as
    top-down control by herbivores becomes stronger or
    weaker. Finally, we delve into some of the ecosystem-
    level responses to these changes, including changes in
    primary productivity, diversity, and resilience.
    Although there are several key areas in which
    ecological insight is lacking, we suggest that reasonable
    climate-related hypotheses can be developed and
    tested based on current information. By strategically
    prioritizing research in the areas of complex
    environmental variation, multiple stressor effects,
    evolutionary adaptation, and population, community,
    and ecosystem-level responses, we can rapidly build
    upon our current understanding of seaweed biology
    and climate change ecology to more effectively
    conserve and manage coastal ecosystems.

    Key index words: adaptation; carbon dioxide; climate
    change; community structure; competition; ecophysi-
    ology; ecosystem function; herbivory; marine macro-
    algae; ocean acidification

    Changes in global temperature and ocean chemistry
    associated with increasing greenhouse gas concen-
    trations are forcing widespread shifts in biological
    systems. In response to warming, species ranges are
    shifting toward the poles, up mountainsides, and to
    deeper ocean depths (Parmesan and Yohe 2003,
    Perry et al. 2005). Factors including warming and
    ocean acidification are causing the reorganization
    of local communities as species are added or
    deleted and as interactions among species change
    in importance (Wootton et al. 2008, Harley 2011).
    Because greenhouse gas emission rates continue to
    accelerate, the climatically forced ecological changes
    that have been documented over the past half cen-
    tury will likely pale in comparison to changes in the
    coming decades.
    Global change is, by definition, a global phenom-

    enon, yet some biological systems have received
    more attention than others. Although a great deal
    of research has focused on systems like coral reefs
    and terrestrial forests (e.g., Hoegh-Guldberg et al.
    2007, Aitken et al. 2008), considerably less attention
    has been devoted to seaweed-dominated ecosystems
    (Wernberg et al. 2012). Like corals and trees, sea-
    weeds are key habitat structuring agents that harbor
    incredible biodiversity (Graham 2004, Christie et al.
    2009). Seaweeds form the base of productive food
    webs that include economically valuable species
    (Graham 2004, Norderhaug et al. 2005) and extend
    well beyond the shallow waters in which seaweeds
    dwell (Harrold et al. 1998). Seaweeds are intimately
    linked to human cultural and economic systems via
    the provision of ecosystem goods and services rang-
    ing from food to medicine to storm protection
    (Rönnbäck et al. 2007).
    Here, we describe how global climate change influ-

    ences marine macroalgae and their associated ecosys-
    tems. We begin with the physical and chemical
    changes that are currently at work in the oceans, and
    how these changes may impact seaweed performance
    via changes in stress and resource availability. These
    direct linkages from environment to organism will

    1Received 24 April 2012. Accepted 17 July 2012.
    2Author for correspondence: e-mail harley@zoology.ubc.ca.

    J. Phycol. 48,

    1064

    –1078 (2012)
    © 2012 Phycological Society of America
    DOI: 10.1111/j.1529-8817.2012.01224.x

    1064

    drive species-level responses, including adaptation,
    migration, and extinction. We then consider how
    direct effects of climate change may modify inter-
    specific interactions such as competition, herbivory,
    and disease, and broaden our focus to examine
    changes in whole ecosystem structure and function.
    Finally, we highlight key areas where our understand-
    ing is incomplete, and suggest productive avenues for
    future research.

    ABIOTIC CHANGE IN COASTAL MARINE ENVIRONMENTS

    Rising carbon dioxide concentrations in the atmo-
    sphere and in the oceans are driving a number of
    important physical and chemical changes. These
    include directional, global-scale trends like ocean
    acidification (the shift in ocean chemistry that
    includes reductions in pH and carbonate ion avail-
    ability), warming, and sea-level rise, along with
    regionally specific increases or decreases in wave
    heights, upwelling, terrigenous nutrient runoff, and
    coastal salinity. As these abiotic trends are thor-
    oughly reviewed elsewhere (e.g., Feely et al. 2004,
    IPCC 2007, Rabalais et al. 2009, Wang et al. 2010,
    Zacharioudaki et al. 2011), we pause here only to
    highlight two salient features of this suite of anthro-
    pogenically forced environmental change. First, the
    magnitude of change is remarkable. We have
    already exceeded the maximum CO2 concentration
    experienced in the last 740,000 years (Augustin
    et al. 2004), and will soon exceed the range of CO2
    concentrations experienced in tens of millions of
    years (Pearson and Palmer 2000, IPCC 2007).
    Second, the rate of abiotic change is virtually
    unprecedented. The rise in CO2 concentrations and
    global temperature since the industrial revolution
    are 100–1000 times faster than at any point in the
    past 420,000 years and are still accelerating (Hoegh-
    Guldberg et al. 2007). Corresponding rates of geo-
    chemical change in the oceans currently exceed
    anything recorded in the last 300 million years
    (Hönisch et al. 2012). Both the magnitude and the
    rate of environmental change pose serious
    challenges to marine species that must either
    tolerate or adapt to a new ocean.

    INDIVIDUAL-LEVEL RESPONSES: GAPS IN THE
    ECOPHYSIOLOGICAL FRAMEWORK

    Seaweed survival, growth, and reproduction are
    known to vary with numerous climatically sensitive
    environmental variables including temperature
    (e.g., Lüning and Neushul 1978), desiccation (Davison
    and Pearson 1996, Chu et al. 2012), salinity
    (e.g., Steen 2004a), wave heights (Seymour et al.
    1989, Graham et al. 1997), nutrient supply via
    upwelling and run-off (Lobban and Harrison 1997),
    pH (Kuffner et al. 2008, Martin and Gattuso 2009,
    Diaz-Pulido et al. 2012), and carbon dioxide
    concentration itself (Kroeker et al. 2010). To date,

    our understanding of the relationship between
    environmental change and the performance of indi-
    vidual seaweeds is based on a loose combination of
    mechanistic, physiological research, and phenome-
    nological studies that correlate performance with
    environmental conditions. The seaweed physiologi-
    cal literature is extensive, but much of it predates
    our current understanding of future environmental
    scenarios, and it is not well linked to more ecologi-
    cally oriented studies. Rather, climate change ecolo-
    gists often make use of phenomenological studies to
    make broad-brush predictions for future change.
    Whether due to the lack of information or the lack
    of communication across disciplines, weak or miss-
    ing mechanistic linkages between predicted future
    conditions and seaweed growth, reproduction, and
    survival are problematic. For example, climate
    change will result in novel patterns and combina-
    tions of stress, and a priori predictions regarding
    responses to simultaneous changes in the means
    and variance of multiple environmental stressors are
    difficult to make in the absence of a mechanistic
    understanding of sublethal and lethal stresses in sea-
    weeds. (In this review, we use the term “stress” to
    denote disruptive stress sensu Davison and Pearson
    (1996); stressful conditions are those that adversely
    affect growth via damage and/or resource realloca-
    tion associated with damage prevention and repair).
    Below, we consider what is and what is not known
    about the two most broadly important aspects of
    environmental change, warming and ocean acidifi-
    cation, with a further emphasis on variable impacts
    across different algal life history stages. We then
    detail some of the ways in which an incomplete
    ecophysiological understanding impairs our ability to
    predict seaweed responses to complex environmental
    variation and multiple stressors.
    Thermal ecophysiology. Temperature determines

    the performance of seaweeds, and indeed all organ-
    isms, at the fundamental levels of enzymatic
    processes and metabolic function (reviewed in
    Raven and Geider 1988, Lobban and Harrison
    1997). Seaweeds have evolved biochemical and phys-
    iological adaptations, including variation in the
    identity and concentration of proteins and the prop-
    erties of cell membranes, that enable them to
    optimize their performance with respect to the tem-
    peratures they encounter (Eggert 2012). Although
    seaweeds are generally well adapted to their thermal
    environment, they nevertheless experience tempera-
    tures in nature – particularly during periods of envi-
    ronmental change – that are sufficiently high or low
    to result in disruptive stress in the form of cellular
    and subcellular damage (reviewed in Davison and
    Pearson 1996, Eggert et al. 2012). Such damage
    and any reallocation of resources for protection and
    repair can slow growth, delay development, and lead
    to mortality (Davison and Pearson 1996). In
    response, seaweeds can produce heat shock proteins
    that repair or remove damaged proteins (e.g., Vayda

    CLIMATE CHANGE AND SEAWEEDS 1065

    M
    IN

    IR
    E
    V
    IE

    W

    and Yuan 1994, Lewis et al. 2001). However, protein
    thermal physiology is not well understood in macro-
    algae (Eggert et al. 2012) and the upregulation of
    heat shock protein production is only one of many
    transcriptional changes that occur in seaweeds dur-
    ing periods of thermal stress (Collén et al. 2007,
    Kim et al. 2011). Relevant genomic, transcriptomic,
    and proteomic studies are only just beginning to
    scratch the surface and most links from gene
    expression to organismal performance are far from
    well established.

    As a result of nonstressful conditions at inter-
    mediate temperatures and stress at the extremes,
    the relationship between temperature and most sub-
    cellular, tissue-level, or whole-organism processes is
    described by a hump-shaped thermal performance
    curve. From colder to warmer, these curves gener-
    ally rise exponentially as rates of biochemical reac-
    tions increase, peak at some optimum temperature,
    and then fall rapidly as the biological components
    of the system become less efficient or damaged
    (Kordas et al. 2011, Eggert et al. 2012). When prop-
    erly parameterized across the full-temperature toler-
    ance range of a species, thermal performance
    curves have the potential to predict the physiologi-
    cal effects of any given warming or cooling scenario
    (barring any further acclimatization, adaptation, or
    context-dependent surprises; see below). The effect
    of a small increase in thallus temperature will be
    beneficial when the initial temperature is cooler
    than optimal and detrimental when it is warmer
    than optimal, and the precise change in perfor-
    mance can be predicted from the starting and
    ending temperature values along the curve. Unfor-
    tunately, the shapes of thermal performance curves
    and the positions of their optima are poorly
    described in most seaweeds. Although many physio-
    logical and ecological studies have linked seaweed
    performance to temperature, a substantial fraction
    of these studies do not investigate enough tempera-
    tures across a wide enough range to characterize
    the underlying, nonlinear relationship between the
    two. Furthermore, various physiological parameters
    within an organism differ in the shape and opti-
    mum temperature of their thermal performance
    curves, which limits our ability to use an easily mea-
    sured parameter (e.g., photosynthesis) as a proxy
    for parameters that may be more ecologically
    relevant (e.g., growth and reproduction). Indeed,
    growth rates do tend to peak at lower temperatures
    than photosynthetic rates (Eggert et al. 2012), pre-
    sumably because metabolic rates increase faster than
    photosynthetic rates at higher temperatures. Much
    remains to be learned regarding the thermal depen-
    dence of the key physiological processes that control
    growth, reproduction, and survival across the full
    range of temperatures experienced by an individual
    in its lifetime.
    Ecophysiology of ocean acidification. Carbon dioxide

    concentrations in seaweed habitats are increasing

    with anthropogenic emissions and, in some regions,
    with intensified upwelling of CO2-enriched water
    (Feely et al. 2008). As with terrestrial plants (Long
    et al. 2004), it is tempting to predict that seaweeds
    will benefit from the increase in inorganic carbon
    concentration (Beardall et al. 1998). However, the
    situation in the sea is not so simple. CO2-driven
    effects on photosynthesis and growth depend on
    the degree to which carbon is limiting, which in
    turn varies among habitat types and among taxa.
    Because CO2 diffusion rates are much higher in air
    than in water, seaweeds that are exposed at low tide
    and those with floating canopies at the sea-air inter-
    face have greater access to CO2 (Beardall et al.
    1998). However, aerial exposure does not necessar-
    ily reduce the probability of carbon limitation, as
    exposure at low tide can dramatically reduce rates
    of carbon acquisition (Williams and Dethier 2005)
    and even emersed seaweeds can benefit from
    increasing atmospheric CO2 concentrations (Gao
    et al. 1999, Zou and Gao 2002). Moreover, a strict
    focus on CO2 in air or dissolved in water may be
    misleading as not all species require environmental
    CO2 as a carbon source. Most green and brown
    algae (and many red algae) can also utilize bicar-
    bonate (HCO3

    �) by converting it to CO2 intracellu-
    larly via CO2 concentrating mechanisms (CCMs; see
    Raven et al. 2012 for review). Just as terrestrial C3
    plants are more likely to be CO2-limited and there-
    fore more likely to benefit from elevated CO2 than
    C4 plants (Long et al. 2004), seaweeds lacking
    CCMs are more likely to be carbon-limited and thus
    more likely to benefit from additional CO2(aq). For
    example, experimental addition of CO2 greatly
    increased the growth rate of Lomentaria articulata,
    which cannot use bicarbonate (Kübler et al. 1999),
    but did not enhance photosynthetic rates of species
    with CCMs or of nonbicarbonate using species that
    were not carbon-limited (Cornwall et al. 2012).
    However, species with CCMs did shift away from
    bicarbonate and toward CO2(aq) when CO2 concen-
    trations were high, which may benefit the seaweeds
    by reducing the energetic costs of using CCMs
    (Cornwall et al. 2012). Thus, although there may be
    variation among taxa based on carbon utilization
    strategy, noncalcifying seaweeds as a group will
    likely respond positively to increasing global CO2
    concentrations in general (see Kroeker et al. 2010).
    In addition to providing carbon for photosynthesis,

    anthropogenic CO2 emissions reduce seawater pH
    and the saturation state of calcium carbonate. As
    this increases the cost of calcification and the likeli-
    hood of dissolution, calcifying organisms are partic-
    ularly sensitive to elevated CO2 in seawater. Ocean
    acidification is consistently related to reduced
    growth rates in calcified macroalgae (Kroeker et al.
    2010) and reductions in calcification rate at elevated
    pCO2 have been demonstrated for crustose and
    articulated coralline red algae as well as calcified
    green Halimeda (Gao et al. 1993, Büdenbender

    1066 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    et al. 2011, Price et al. 2011). However, reduced cal-
    cification at higher pCO2 did not emerge as a gen-
    eral pattern in a meta-analysis of multiple seaweed
    studies (Kroeker et al. 2010). This may be because
    the process of calcification, and likewise the effects
    of ocean acidification on calcification, varies among
    seaweeds (Price et al. 2011), and many species are
    able to create microclimates of chemistry favorable
    for calcification regardless of ambient conditions
    (Roleda et al. 2012a). It has therefore been sug-
    gested that the effects of ocean acidification on cal-
    cified species may be manifested as increased
    dissolution rather than reduced production of cal-
    cium carbonate (Roleda et al. 2012a). Reduced pH
    may have important consequences for noncalcifying
    taxa as well (Roleda et al. 2012b), although the
    cumulative effects of climatically realistic, CO2-dri-
    ven pH change on noncalcifying seaweeds remain
    poorly understood.
    Stress and the completion of algal life cycles. Predict-

    ing true individual-level responses to climate change
    in seaweeds is challenging owing to the numerous
    life history stages and transitions upon which envi-
    ronmental change can act (Schiel and Foster 2006).
    Careful consideration of this complexity is impor-
    tant because thermal optima and tolerance limits
    can vary among life history stages within a species
    (e.g., Fain and Murray 1982), and climate effects at
    one life history stage may be magnified or offset by
    impacts (or lack thereof) at other life history stages
    (e.g., Ladah and Zertuche-González 2007).

    To exemplify the degree to which we are ignorant
    of how climate change will impact seaweeds across all
    life history stages, we summarize what is known
    regarding the effects of warming and elevated CO2
    on one particularly well-studied species, the giant
    kelp, Macrocystis pyrifera (Fig. 1). Increased tempera-
    ture is generally thought to have negative effects on

    spore production (Buschmann et al. 2004), germina-
    tion (Buschmann et al. 2004), recruitment (Deysher
    and Dean 1986a, Buschmann et al. 2004), and sporo-
    phyte growth (Rothäusler et al. 2009, 2011) and con-
    text-specific effects on gametogenesis depending on
    the source population and degree of warming (Lüning
    and Neushul 1978, Deysher and Dean 1986b,
    Muñoz et al. 2004). Warming has also been linked
    to mortality of spores, gametophytes, eggs, and
    sporophytes (Ladah and Zertuche-González 2007).
    Much less is known about the effects of increasing
    CO2 concentrations. On the basis of current knowl-
    edge, we can expect positive effects on gametogene-
    sis and variable effects (e.g., positive effect of
    increasing CO2, but negative effect of decreasing
    pH) on germination (Roleda et al. 2012b). Studies
    assessing the potential for interactive temperature
    and CO2 effects are uncommon (see below), and
    nonexistent for M. pyrifera. Thus, even for one of the
    best-studied seaweeds in the world, large knowledge
    gaps greatly hinder our ability to precisely predict
    future changes in population growth and persistence.
    The importance of variability, rates of change, and envi-

    ronmental history. When predicting future ecological
    patterns – and when designing experiments to test
    those predictions – it is tempting to treat environ-
    mental change as a steady shift in mean conditions.
    However, environmental time series are complex (see
    Helmuth et al. 2006 for temperature examples and
    Wootton et al. 2008 for a pH example), and different
    aspects of an environmental signal, including
    extremes, range, and patterns of variability, will have
    different biological consequences. For example,
    seaweed reproduction may only occur if temperatures
    drop below some threshold for a sufficiently long per-
    iod of time, whereas mortality may be more closely
    linked to high temperatures that exceed physiological
    tolerance (Breeman 1988, Wernberg et al. 2011b).

    FIG. 1. Effects of increasing temperature and CO2 on life history processes in Macrocystis pyrifera. Green boxes indicate experimental
    evidence of positive effects, yellow boxes indicate negative effects, hatched boxed indicate both positive and negative (i.e., context-specific)
    effects, and blank boxes represent unquantified responses owing to a lack of published information.

    CLIMATE CHANGE AND SEAWEEDS 1067

    M
    IN
    IR
    E
    V
    IE
    W

    Mortality rates following short exposures to extreme
    temperatures or salinities can be similar to those
    found after longer exposures to less extreme condi-
    tions (Forrest and Blakemore 2006). In an experi-
    ment that manipulated the temporal variation of
    stress, higher variability muted negative impacts of
    stress on some seaweed taxa, but generated negative
    impacts in others (Benedetti-Cecchi et al. 2006).
    Although changes in variability can drive important
    biological changes in other systems (e.g., agricul-
    tural crops and forests; Southworth et al. 2000,
    Giesecke et al. 2010), studies on the effects of differ-
    ent magnitudes and temporal patterns of environmen-
    tal variability, alone or in combination with changes in
    mean conditions, are exceedingly rare for seaweeds.

    Additional aspects of environmental variability
    come into play when one considers that the physio-
    logical limits of individuals and populations are not
    constant (see below). The history of environmental
    variation is often a key predictor of future success,
    as an individual or population that has been
    exposed to stressful conditions in the past may be
    better able to cope with them in the future (Padilla-
    Gamino and Carpenter 2007a). Finally, rapid envi-
    ronmental changes are typically more detrimental
    than slow ones, as rapid change is more likely to
    outpace an organism’s ability to acclimatize or a
    population’s ability to adapt (O’Connor et al.
    2012). We require a better understanding of the
    ecological consequences of the accelerating pace of
    change in the Earth’s climate system to reduce the
    probability of ecological surprises.
    Multiple stressors and nonadditive effects. All of the

    anthropogenically forced changes in the physical
    and chemical environment are occurring simulta-
    neously, and in many cases, the impact of any par-
    ticular stressor on the physiology and performance
    of marine macrophytes will depend upon the pres-
    ence and magnitude of additional limiting or dis-
    ruptive stressors. For example, the importance of –
    and limiting values of – various resources are envi-
    ronmentally dependent, with the degree of light
    limitation at low irradiance varying with tempera-
    ture (e.g., Davison et al. 1991) and the enhance-
    ment of photosynthesis by elevated CO2 varying
    with nutrient availability (Xu et al. 2010). The per-
    cent cover of algal turfs decreased with increasing
    CO2 under ambient nutrients, but the reverse was
    true under elevated nutrients (Russell et al. 2009).
    There are also many interactions among disruptive
    stressors, including temperature, desiccation, pH,
    salinity, and ultraviolet radiation. For example, in
    tropical and warm-temperate crustose coralline
    algae, the negative effect of warmer temperatures
    on bleaching, growth rates, calcification rates, and
    survival were significantly greater under conditions
    of elevated CO2/reduced pH (Anthony et al. 2008,
    Martin and Gattuso 2009, Diaz-Pulido et al. 2012).
    The magnitude and even the direction of UV effects
    depend upon temperature and CO2 (Hoffman et al.

    2003, Swanson and Fox 2007, Gao and Zheng
    2010). Depending on the species and life history
    stage, desiccation has been shown to magnify or
    reduce the effects of high temperature (e.g., Hunt
    and Denny 2008, Chu et al. 2012). As of yet, it is
    difficult to predict when one stressor will increase
    or decrease the effect of another. There are also no
    known biases toward synergistic or antagonistic
    effects; in a meta-analysis of multi-stressor studies on
    Fucus spp., synergistic, additive, and antagonistic out-
    comes were all equally prevalent (Wahl et al. 2011).
    We desperately need to incorporate more ecophysio-
    logical research into a multi-stressor framework to
    improve our understanding of when, where, and why
    important context-dependent outcomes emerge.

    POPULATION AND SPECIES-LEVEL RESPONSES:
    TOLERATE, ADAPT, MOVE, OR DIE

    As described above, environmental change can
    elicit a wide array of responses in individual
    organisms. At the species level, responses to envi-
    ronmental forcing can be distilled down to a small
    set of basic alternatives: (i) persistence without
    acclimatization or adaptation (tolerance), (ii) per-
    sistence with acclimatization or adaptation, (iii)
    persistence enabled by migration to remain within
    some particular climatic niche and (iv) extinction.
    In this section, we devote our discussion to potential
    roles of acclimatization and adaptation in facilitat-
    ing local persistence, to changes in seaweed distribu-
    tional patterns, and to the potential for seaweed
    extirpations.
    Scope for acclimatization and adaptation. There is a

    rich literature on seaweed acclimation (an individ-
    ual-level response to experimental manipulation of
    the environment), acclimatization (an individual-
    level response to natural variation in the environ-
    ment), and local adaptation (a population-level
    response to natural environmental variation) as a
    consequence of variation in temperature, salinity,
    light, and wave forces (Lüning 1990, Lobban and
    Harrison 1997, Eggert et al. 2012). Appropriately
    acclimated/acclimatized individuals or adapted pop-
    ulations may be better able to withstand coming
    environmental change. For example, warm-accli-
    mated Saccharina latissima sporophytes required less
    light to achieve maximum photosynthetic rates and
    were more photosynthetically efficient at high tem-
    peratures (Davison et al. 1991), and warm-accli-
    mated Fucus vesiculosus embryos were more likely to
    survive periods of thermal stress (Li and Brawley
    2004). Although many species can acclimate to envi-
    ronmental changes, some algal populations or
    species may be less able to do so than others. For
    example, some tropical species and subpopulations
    appear to have limited scope for acclimation
    relative to their temperate counterparts, presum-
    ably due to reduced environmental variability in
    tropical habitats (Padilla-Gamino and Carpenter

    1068 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    2007a,b). Regardless, the use of appropriately accli-
    mated/acclimatized individuals is a prerequisite for
    realistic climate change experiments; otherwise,
    short-term measurements may not reflect true
    long-term responses.

    Unlike acclimation, relatively little is known about
    the degree to which evolutionary adaptation may
    “rescue” seaweed species in the face of environmen-
    tal change. The existence of local ecotypes (e.g.,
    Breeman 1988) clearly indicates that adaptation is
    possible, and there is evidence to suggest that sea-
    weeds can evolve and even speciate fairly rapidly (in
    ~400 years in the case of Fucus radicans) when suffi-
    cient selective pressure is applied by the environ-
    ment (Pereyra et al. 2009). However, the degree to
    which most multi-cellular marine species such as
    seaweeds, their competitors, and their consumers
    can adapt over climate change-relevant time scales
    (<100 years) is largely unknown (but see Sunday et al. 2011). Understanding the extent to which spe- cies will acclimatize or adapt to environmental change is crucial for predicting future ecological change. Distributional shifts and the threat of extinction. Because

    environmental conditions directly and indirectly
    influence seaweed distributional patterns at a vari-
    ety of scales (Breeman 1988), changes in the envi-
    ronment will result in changes in seaweed
    distributions. Some of the most readily detectable
    changes are at local (site) scales, where environ-
    mental change can result in shifts in the vertical
    distribution (often termed zonation) of intertidal
    and subtidal seaweeds. Sea-level rise will result in a
    general upward shift of benthic communities en
    masse, although the accompanying changes in the
    relative availability of appropriate substratum types
    and orientations at specific shore levels (e.g., on
    shores with wave cut platforms and cliffs) may drive
    changes in relative algal abundance (Vaselli et al.
    2008). However, zonation patterns are determined
    by far more than just position relative to mean sea
    level. The upper limit of intertidal seaweeds is
    related to thermal and desiccation stress during
    low tide (e.g., Harley 2003), and long-term
    increases in air temperature have resulted in down-
    shore shifts in the upper limit of some species
    (Harley and Paine 2009). The depth range of sub-
    tidal kelps also depends critically on environmental
    factors such as temperature, water motion, and
    water transparency (Graham et al. 2007), and cli-
    mate-related changes in these factors are predicted
    to reduce the depth range of kelp forests (Méléder
    et al. 2010). When the upper and lower depth lim-
    its of a species are set by different agents (e.g.,
    thermal stress, light availability, consumers; Harley
    2003, Graham et al. 2007, Méléder et al. 2010), cli-
    mate change can result in certain species being
    squeezed out of the system entirely (Harley 2011).

    Climate change will drive distributional shifts at
    larger, alongshore scales as well. Increased storm

    frequency could restrict vulnerable species to pro-
    tected shorelines, and changes in salinity may allow
    seaweeds to penetrate further into, or be forced fur-
    ther out of, estuarine embayments and lagoons. The
    most notable large-scale changes, however, are those
    occurring across latitudinal temperature gradients.
    Drastic population declines and even local extinc-
    tions have been documented at the warm (lower lati-
    tude) end of species’ biogeographic ranges during
    periods of warming (e.g., Serisawa et al. 2004).
    Range retraction at low latitudes can be offset by
    expansion into higher latitudes, as in western Europe
    where warm-water species have expanded northward
    (Lima et al. 2007). However, such expansions may
    not be a sustainable escape mechanism for species
    along coastlines with significant geomorphic barri-
    ers, such as the end of a continent. For example,
    poleward migration of seaweed species has been
    observed along the east and west coasts of Australia
    since 1940, but because there is no suitable habitat
    within the range of most species’ dispersal abilities
    further south, continued poleward retreat may
    result in numerous extinctions as species ‘fall off
    the map’ (Wernberg et al. 2011a). Indeed, extinc-
    tions have already been documented for several
    marine macroalgae, although the relative contribu-
    tion of environmental change to these losses
    remains poorly understood (Brodie et al. 2009).

    COMMUNITY-LEVEL RESPONSES: INTERSPECIFIC
    INTERACTIONS AND INDIRECT EFFECTS

    Ecological change in coastal ecosystems reflects
    the combined influence of direct environmental
    impacts on individual species and indirect effects
    mediated by changes in interspecific interactions
    (Harley et al. 2006). We first describe some of the
    ways that competitive, trophic, and symbiotic rela-
    tionships are likely to change in seaweed systems,
    and then discuss the consequences of these changes
    for entire ecosystems in the following section.
    Competitive relationships. Seaweeds compete for

    nutrients, light, and space for attachment, and their
    relative success at acquiring these resources in the
    presence of other photo-autotrophs (or sessile inver-
    tebrates, in the case of space) depends upon both
    resource availability and environmental stress. The
    availability of several resources (e.g., CO2, nitrate,
    ammonium) is changing due to human activities,
    and the effects of changing resource supply will
    depend on the magnitude and direction of these
    changes and the degree to which these resources
    limit algal growth and competitive ability. Increasing
    nitrogen loading tends to favor fast-growing species
    with high nitrogen requirements. In some cases, this
    may lead to competitive dominance by weedy taxa
    (Steen 2004b, Vermeij et al. 2010) and – should
    nutrients trigger a phytoplankton bloom – shading-
    out of benthic seaweeds by phytoplankton (Kava-
    naugh et al. 2009). In other cases, higher nitrogen

    CLIMATE CHANGE AND SEAWEEDS 1069

    M
    IN
    IR
    E
    V
    IE
    W

    merely allows for the persistence of nitrogen-limited
    taxa and thus enhances algal diversity (Bracken and
    Nielsen 2004). Elevated CO2(aq) will differentially
    affect seaweeds depending on their carbon capture
    strategy. The influence of elevated CO2(aq) on sea-
    weeds with carbon concentrating mechanisms, such
    as kelps, is highly light-dependent, and the overall
    effect of rising CO2(aq) on kelp competitive ability
    remains unclear (Hepburn et al. 2011). On the
    other hand, for species that rely on aqueous CO2, like
    turf-forming rhodophytes in New Zealand, elevated
    CO2(aq) should differentially favor their growth,
    which may in turn enhance their competitive ability
    (Hepburn et al. 2011).

    Changes in the severity of environmental stressors
    (e.g., temperature, pH, salinity, wave forces) will
    also affect the outcome of competitive relationships.
    In some cases, environmental extremes remove
    otherwise dominant competitors, allowing subordi-
    nate species to persist (Sousa 1979) and facilitating
    the establishment of non-native taxa (Miller et al.
    2011). Stress need not be lethal to influence the
    outcome of competitive interactions (Davison and
    Pearson 1996). For example, many important com-
    petitors for space are calcified taxa such as crustose
    coralline algae, corals, and mussels, and inhibition
    of growth by reduced pH likely contributes to
    increasing fleshy algal competitive dominance over
    these groups (Wootton et al. 2008, Diaz-Pulido et al.
    2011, Hepburn et al. 2011). The effects of rising
    temperatures may increase or decrease competition,
    and even change competitive interactions into facili-
    tative ones. Elevated temperature increased the
    competitive impacts of Enteromorpha on two species
    of Fucus (Steen 2004b). Conversely, the effects of
    intertidal Ascophyllum nodosum on understory barna-
    cles, like the effects of subtidal Ecklonia radiata on
    E. radiata recruits, shifted from negative (competi-
    tive) to net positive (facilitative) at high tempera-
    tures (Leonard 2000, Wernberg et al. 2010).
    Herbivory. Herbivores are key structuring agents

    in algal communities, influencing everything from
    the survival of individual seaweeds to total algal bio-
    mass and diversity (Lubchenco and Gaines 1981).
    The outcomes of pairwise plant-herbivore interac-
    tions depend on characteristics of both the alga and
    the herbivore, including the palatability of seaweeds,
    the per capita consumption rates of herbivores, and
    the individual and population growth rates and
    overall abundance of both. Abiotic factors associated
    with climate change are known to impact all of
    these attributes.

    The amount of algal tissue that an herbivore can
    or will consume depends on the degree of morpho-
    logical or chemical defense and on aspects of
    nutritional quality such as the C:N ratio (Duffy and
    Hay 1990, Van Alstyne et al. 2009). Elevated temper-
    ature reduced herbivore defenses in F. vesiculosus
    (Weinberger et al. 2011), and changes in nutrient
    availability have been shown to alter algal palatability

    (e.g., Hemmi and Jormalainen 2002). Calcium car-
    bonate in algal tissue is an important anti-herbivore
    defense (Hay et al. 1994) and ocean acidification
    may have dramatic impacts on the palatability of cal-
    cified seaweeds via reduced calcification or
    increased dissolution. Although elevated CO2 would
    be expected to increase C:N ratios in noncalcified
    taxa, the effects of elevated CO2 on seaweed palat-
    ability lags far behind our understanding of such
    effects in phytoplankton and terrestrial plants.
    Climate change will also have direct effects on

    herbivores that will cascade down to primary pro-
    ducers. Several field studies suggest that warming
    sea surface temperatures are associated with
    increases in important herbivore populations and
    concomitant declines in certain algal species (Hart
    and Scheibling 1988, Ling 2008, Hernandez et al.
    2010). Although warming may benefit some grazer
    populations, ocean acidification is likely to be gen-
    erally detrimental to many invertebrate herbivores,
    particularly heavily calcified species such as sea
    urchins and molluscs (Dupont et al. 2010, Crim
    et al. 2011). Volcanic CO2 vent systems provide a
    glimpse into this future; areas of reduced pH near
    CO2 seeps are associated with reductions in urchin
    and shelled gastropod abundance, and the success
    of Padina spp. (despite a reduction of calcium car-
    bonate in the thalli) and of highly palatable Sargas-
    sum vulgar near CO2 vents has been attributed to
    the absence of urchin grazers (Porzio et al. 2011,
    Johnson et al. 2012). The impacts of ocean acidifi-
    cation on other herbivorous taxa, notably crusta-
    ceans and fish, appear to be relatively minor
    (Kroeker et al. 2010).
    Although useful as a starting point, changes to algal

    and invertebrate performance or population size in
    isolation cannot fully predict changes in the impor-
    tance of herbivory. Rather, the overall impact of her-
    bivory depends upon the balance of production and
    consumption of algal tissue. Metabolic theory
    predicts that metabolic rate and scope for activity –
    which in ectothermic herbivores determine the
    demand for and ability to acquire food, respectively,
    – increase more quickly with temperature than algal
    photosynthetic rate and thus primary production
    (O’Connor 2009). As a result of these different tem-
    perature–performance relationships, experimental
    warming increased the relative importance of
    amphipod grazing and decreased algal biomass
    despite generally positive direct effects of warming
    on algal growth (O’Connor 2009). Such rate-depen-
    dent generalizations fall apart, however, when
    abiotic conditions become stressful, and stress differ-
    entially reduces the performance of one or more
    of the interacting species (Kordas et al. 2011).
    One trophic level or the other is often dispropor-
    tionately susceptible to stress associated with
    extremes in temperature, salinity, and wave forces
    (Cubit 1984, Elfwing and Tedengren 2002, Taylor
    and Schiel 2010), making seaweeds relatively safe or

    1070 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    relatively vulnerable to grazing at certain places and
    times. As the environment changes, the times and
    places that seaweeds are most, or least, impacted by
    herbivory will change as well (see e.g., Vinueza et al.
    2006).
    Epibionts, endophytes, and pathogens. Seaweeds live

    in constant association with a variety of microbes,
    fungi, animals, and other algae that live on or in their
    tissues. Of these relationships, the ecological role of
    epibionts is particularly well-studied, with effects on
    seaweed hosts ranging from reduced growth and
    reproduction to increased risk of mechanical break-
    age (e.g., Dantonio 1985). In kelp beds in eastern
    Canada, outbreaks of non-native epiphytic bryozoans
    are triggered by warming events, and these outbreaks
    have led to drastic reductions in the percent cover of
    habitat-forming Saccharina longicruris (Scheibling and
    Gagnon 2009, Saunders et al. 2010). In cases such as
    this, where bryozoan epibionts increase the risk of
    frond breakage (Krumhansl et al. 2011), any local
    increase in storminess may act synergistically with
    warming and infestations by epibionts.

    In contrast to epibionts, the ecology of seaweed
    endophytes and diseases is poorly understood, par-
    ticularly with regard to climate change (Eggert et al.
    2010, Gachon et al. 2010). There is, however,
    mounting evidence that warming will negatively
    impact seaweeds by facilitating bacterial infections
    (Campbell et al. 2011, Case et al. 2011). Departures
    from optimal salinity and irradiance can also make
    seaweeds more susceptible to bacterial disease, as
    evidenced by experiments and observations on
    aquaculture species such as Kappaphycus alvarezii
    (Largo et al. 1995). Although evidence for primarily
    negative pathogen-mediated effects of environmen-
    tal change is slowly accumulating, the generality and
    future magnitude of such negative effects remain
    essentially unknown.

    SHIFTS IN COMMUNITY STRUCTURE

    AND ECOSYSTEM FUNCTION

    Environmental change, coupled with shifts in spe-
    cies interactions and the shuffling of species distri-
    butions, will culminate in potentially far-reaching
    changes in community structure and ecosystem
    function (Harley et al. 2006). Because the responses
    of benthic assemblages are often highly idiosyn-
    cratic, generalizations and specific predictions are
    fraught with uncertainty. While the future states of
    marine ecosystems are far from certain, neither are
    they completely unforeseeable. Several predictions
    and testable hypotheses can be developed around
    our current understanding of seaweed-dominated,
    or potentially seaweed-dominated, ecosystems.

    One system for which specific predictions have
    been made is the rocky intertidal zone in Britain,
    where the effects of climate change have been
    considered in great detail and where relevant long-
    term datasets exist (Hawkins et al. 2008, 2009).

    Wave-protected and semi-exposed British shores are
    typically dominated by large fucoid algae, which are
    dominant competitors for primary space as well as
    ecosystem engineers that provide cool, moist micro-
    habitats for associated species (e.g., Schonbeck and
    Norton 1980, Thompson et al. 1996). Rising air
    temperatures and increasing wave exposure will
    directly reduce fucoid canopies via lethal physiologi-
    cal and hydrodynamic stress (Hawkins et al. 2009).
    The appearance and/or increased abundance of
    warm-water herbivores is expected to further reduce
    algal cover (Mieszkowska et al. 2006, Hawkins et al.
    2009). A more diverse grazer assemblage, coupled
    with the replacement of a structurally complex,
    cold-water barnacle species with a structurally sim-
    ple, warm-water barnacle species, will reduce oppor-
    tunities for fucoid size escapes from microscopic
    stages and thus inhibit the regrowth of the algal
    canopy (Hawkins et al. 2008). The net result is a
    decline in subcanopy habitat and a reduction in
    benthic primary productivity. These changes are
    predicted to reduce the abundance of many inverte-
    brates that rely on cool moist microhabitats and
    decrease invertebrate production in the algal detri-
    tus food web of the strand line; both of these effects
    may negatively impact birds that forage on these
    invertebrate resources (Kendall et al. 2004).
    Like intertidal fucoids, subtidal kelps provide hab-

    itat structure for numerous species, including many
    that are economically important (Graham 2004).
    Kelp forests in Australia, like fucoid communities in
    Britain, are experiencing range expansions and con-
    tractions of both seaweeds and important herbivores
    in association with warming temperatures (Ling
    2008, Wernberg et al. 2011a). In this system, there
    is also evidence that ocean acidification will result
    in important shifts in community structure. Experi-
    mental increases of temperature and CO2 increased
    the biomass of algal turfs (Connell and Russell
    2010). Enhanced cover and biomass of turf-forming
    algae associated with elevated CO2 occurred at the
    expense of coralline crusts, although the magnitude
    of this shift depended on nutrient and light levels
    (Russell et al. 2009, 2011). Increasing dominance of
    turfs in response to rising CO2 may in turn inhibit
    kelp recruitment, which could cause or maintain
    phase shifts from kelps to turfs (Connell and Russell
    2010). However, kelp canopy can, to some degree,
    inhibit the positive effects of elevated CO2 on turfs,
    suggesting that intact kelp forests may be resistant,
    but not resilient, to phase shifts to turf-dominated
    communities (Falkenberg et al. 2012).
    The degree to which results from southwestern

    Australia will generalize to other kelp systems, such
    as those under strong top-down control, is unclear.
    Limited information suggests that elevated CO2 has
    variable but often positive effects on kelps like Nereo-
    cystis luetkeana and M. pyrifera (Thom 1996, Swanson
    and Fox 2007, Roleda et al. 2012b), but negative
    effects on crustose coralline algae (CCA; see above)

    CLIMATE CHANGE AND SEAWEEDS 1071

    M
    IN
    IR
    E
    V
    IE
    W

    and important kelp consumers such as sea urchins
    (Dupont et al. 2010, Reuter et al. 2011). Because
    urchins benefit CCA by preventing overgrowth by
    other seaweeds, and CCA benefit urchins by provid-
    ing settlement cues, models suggest that reductions
    in either taxa may result in a positive feedback loop
    (Baskett and Salomon 2010). In contrast to the neg-
    ative effects of ocean acidification on herbivores,
    the effects of warming may be largely positive for
    herbivores (Hart and Scheibling 1988, Ling 2008,
    Hernandez et al. 2010). A long-term study of warm-
    ing associated with power plant thermal effluent in
    central California has shown that a ~3.5°C increase
    in temperature results in increases in herbivore
    abundance, shifts from cold-water N. luetkeana to
    warmer water M. pyrifera canopies, and a replace-
    ment of understory kelps by foliose red algae
    (Schiel et al. 2004). A reasonable working hypothe-
    sis, therefore, is that kelps in the California Current
    system, particularly in the southern portions of their
    range, may respond positively to the direct and indi-
    rect effects of acidification, but negatively to the
    direct and indirect effects of warming (Fig. 2). The
    relative balance between these opposing forces,
    particularly in systems with complex trophic and
    competitive relationships, remains uncertain.

    Tropical coral reefs are also quite sensitive to
    climate change (Fig. 3). In these systems, corals and
    CCA currently dominate in part because present-day
    conditions are relatively conducive to calcification
    and in part because herbivores benefit the slow-
    growing calcifiers – despite some negative impacts
    of bioerosion – by keeping fleshy macroalgae in
    check (Hoegh-Guldberg et al. 2007, O’Leary and
    McClanahan 2010). However, future changes in
    ocean climate are predicted to destabilize coral reef
    ecosystems, resulting in phase shifts from coral-dom-
    inated reefs to benthic systems dominated by fleshy
    macroalgae (Hoegh-Guldberg et al. 2007, Anthony
    et al. 2011, Diaz-Pulido et al. 2011). Fleshy macro-
    algae are positively affected by increased CO2 (Kuff-
    ner et al. 2008, but see Jokiel et al. 2008), which
    along with elevated nutrients may increase their
    competitive ability. In contrast, reef-building corals
    appear to be in serious trouble due to the influence
    of climatic stressors; warming ocean waters are asso-
    ciated with mass coral bleaching events, increased
    pCO2 decreases coral calcification and growth and
    increases dissolution, and storms cause physical
    damage to weakened reef structures (Hoegh-Guld-
    berg et al. 2007). Other calcified habitat-forming
    reef organisms, such as CCA and the green alga
    Halimeda spp., are also expected to do poorly when
    pH and calcium carbonate saturation drop and tem-
    peratures rise (Kuffner et al. 2008, Price et al. 2011,
    Diaz-Pulido et al. 2012). Indeed, coralline algae are
    relatively rare or absent from both tropical and
    temperate sites with naturally occurring carbon
    dioxide seeps (Hall-Spencer et al. 2008, Fabricius
    et al. 2011, Porzio et al. 2011). CCA are particularly

    important as they are the “cement” that helps hold
    coral reefs together and provide important settle-
    ment surfaces for coral larvae; the loss of these
    crusts is predicted to expedite phase shifts on

    FIG. 2. Future ecological scenarios for temperate kelp forests.
    Solid and dashed arrows represent direct and indirect effects of
    one species on another, respectively (the flow of energy via tro-
    phic interactions is omitted for clarity). Faded icons represent
    functional groups that may still be present but play a strongly
    reduced ecological role. Relative to present-day conditions (upper
    panel), future warming (middle panel) will favor grazers and
    have direct and indirect negative impacts on canopy-forming
    kelps. Future increases in CO2 (lower panel) will have strong neg-
    ative effects on crustose coralline algae and positive effects on
    noncalcified seaweeds both directly via improved growth and indi-
    rectly via reduced consumption by calcified herbivores. The com-
    bined impacts of simultaneous warming and acidification in a
    more realistic climate change scenario remain poorly understood.
    See text for details.

    1072 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    tropical reefs (Diaz-Pulido et al. 2007). The key to
    triggering a phase shift from corals to fleshy macroal-
    gae, however, may rest with the herbivores (Fig. 3).
    In areas where herbivore biomass can be maintained,
    the shift away from coral-dominated systems can be
    delayed, although perhaps not prevented indefinitely
    (Hughes et al. 2003, Hoegh-Guldberg et al. 2007,

    Buddemeier et al. 2011). Should structurally com-
    plex corals be replaced with fleshy macroalgae, a
    considerable loss of biodiversity would result (Hoegh-
    Guldberg et al. 2007).
    Ecosystem shifts, such as those described above,

    may occur rapidly once a system has been pushed
    beyond some threshold or tipping point (Scheffer
    et al. 2001). In some cases, the behavior of the sys-
    tem may not change over a wide range of progres-
    sive impairment (e.g., biomass removal or species
    loss), only to shift suddenly once a threshold is
    crossed (Speidel et al. 2001, Davies et al. 2011).
    Catastrophic phase shifts, as have been observed in
    kelp forests and on coral reefs, are facilitated by
    losses of resilience associated with changes in
    resource supply, food web structure, and distur-
    bance frequency (Folke et al. 2004), all of which are
    altered by CO2-induced environmental change.
    Catastrophic shifts are often difficult to anticipate,
    as relevant environmental thresholds may lie at or
    beyond the range of historical variation, but within
    the range of near-future environmental conditions.
    Should environmental conditions return to below-
    threshold values, recovery may proceed quickly,
    slowly, or not at all (Folke et al. 2004), and the recovery
    trajectory may differ considerably from the original per-
    turbation trajectory (e.g., Baskett and Salomon 2010).

    FUTURE DIRECTIONS: ADDRESSING THE BIG UNKNOWNS

    Although a great deal of progress has been made
    in recent years, there are still significant gaps in our
    understanding which hamper our ability to predict
    the outcomes of global change in seaweed-domi-
    nated systems. Some of the most important areas in
    which we lack a general or even basic understanding
    include (i) the importance of rates, timing, magni-
    tude, and duration of environmental change,
    (ii) non-additive effects of multiple stressors,
    (iii) population-level implications of variable envi-
    ronmental impacts among life-history stages,
    (iv) the scope for population- or species-level adap-
    tation to environmental change and (v) ecological
    responses at the level of communities and ecosys-
    tems, including tipping points and sudden phase
    shifts. With regard to uncertainties in the nature of
    environmental forcing, we require additional ecophys-
    iological and ecomechanical studies – especially ones
    that move beyond single-factor ANOVA designs –
    and further development in the emerging field of
    ecological genomics to identify biological responses
    to key environmental drivers or combinations of
    drivers. Of particular use would be an ecophysiologi-
    cal framework from which the impacts of multiple
    stressors could be predicted a priori (Pörtner and
    Farrell 2008). Once understood, these drivers can
    be incorporated into demographic models to better
    describe and predict changes in population growth
    or decline. Although species-level research on sea-
    weeds, at least with regard to climate change, lags

    FIG. 3. Future ecological scenarios for tropical coral reefs.
    Arrows and shading as in Fig. 2. Relative to the present day
    (upper panel), the combination of warming and ocean acidificat-
    ion will reduce the dominance of calcified taxa such as crustose
    coralline algae and corals (middle and lower panels). However,
    the likelihood of fleshy macroalgae rising to dominance and out-
    competing the calcified taxa depends upon whether they are sup-
    pressed by herbivores (as may happen in a marine protected
    area, middle panel) or not (as may happen on a heavily fished
    reef, lower panel). See text for details.

    CLIMATE CHANGE AND SEAWEEDS 1073

    M
    IN
    IR
    E
    V
    IE
    W

    behind similar work in terrestrial environments
    (e.g., Aitken et al. 2008), there is no reason that
    phycologists could not model a research program
    based on the successes of terrestrial botanists, foresters,
    and agricultural scientists. As for community and
    ecosystem-level change, researchers can make rapid
    progress by focusing on ecological dominants (e.g.,
    kelps) and strong interactors (e.g., sea urchins) as a
    starting point. Individual pieces of the ecological
    puzzle can then be interlinked with mathematical
    models and ground-truthed in areas where environ-
    mental conditions already approximate future pro-
    jections (e.g., volcanic CO2 vents and power plant
    thermal effluent plumes).

    Seaweed beds, coral reefs, and other coastal eco-
    systems provide trillions of dollars of ecosystem
    goods and services every year (Costanza et al. 1997),
    and the degradation of these systems will have far-
    reaching consequences for human societies. Devel-
    oping accurate predictions for the ecological effects
    of climate change in seaweed-dominated systems is
    therefore a high priority, as it will be invaluable for
    effective conservation and management. The climate
    change scenario leading from healthy coral reefs to
    degraded macroalgal beds is an excellent example
    of an ecological prediction that can be used to dic-
    tate management priorities. Although warming and
    ocean acidification are beyond our control in the
    near term, we can manage for coral reef resilience
    by conserving herbivore diversity and abundance
    and reducing nutrient loads (Hoegh-Guldberg et al.
    2007). In some parts of the Caribbean, this strategy
    appears to work in practice; following high tempera-
    ture and hurricane disturbances, coral recovery rates
    were higher in protected areas where algal cover was
    more effectively controlled by herbivores (Mumby
    and Harborne 2010). There is high yet largely
    untapped potential for similarly feasible local-scale
    management options in a wide variety of seaweed-
    dominated coastal ecosystems that are undergoing
    major ecological reorganization in response to
    anthropogenic change (e.g., Russell et al. 2009).
    Identifying the leverage points where conservation
    and management practices are most effective should
    continue to be a major focus of ecological research.

    We thank S. Dudgeon and two anonymous reviewers for their
    constructive criticisms. M. Mach kindly provided the illustra-
    tions. During the writing of this paper, CH was supported in
    part by the Killam Trusts and by the Hakai Network for
    Coastal People, Ecosystems, and Management at Simon
    Fraser University. Funding was provided by the NSF through
    grant OCE 0752523 to M. Graham and C. Harley.

    Aitken, S. N., Yeaman, S., Holliday, J. A., Wang, T. L. & Curtis-
    McLane, S. 2008. Adaptation, migration or extirpation: climate
    change outcomes for tree populations. Evol. Appl. 1:95–111.

    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. Proc. Natl. Acad. Sci.
    USA 105:17442–6.

    Anthony, K. R. N., Maynard, J. A., Diaz-Pulido, G., Mumby, P. J.,
    Marshall, P. A., Cao, L. & Hoegh-Guldberg, O. 2011. Ocean
    acidification and warming will lower coral reef resilience.
    Global Change Biol. 17:1798–808.

    Augustin, L., Barbante, C., Barnes, P. R. F., Barnola, J. M., Bigler, M.,
    Castellano, E., Cattani, O. et al. 2004. Eight glacial cycles from
    an Antarctic ice core. Nature 429:623–8.

    Baskett, M. L. & Salomon, A. K. 2010. Recruitment facilitation
    can drive alternative states on temperate reefs. Ecology
    91:1763–73.

    Beardall, J., Beer, S. & Raven, J. A. 1998. Biodiversity of marine
    plants in an era of climate change: some predictions based
    on physiological performance. Bot. Mar. 41:113–23.

    Benedetti-Cecchi, L., Bertocci, I., Vaselli, S. & Maggi, E. 2006.
    Temporal variance reverses the impact of high mean
    intensity of stress in climate change experiments. Ecology
    87:2489–99.

    Bracken, M. E. S. & Nielsen, K. J. 2004. Diversity of intertidal
    macroalgae increases with nitrogen loading by invertebrates.
    Ecology 85:2828–36.

    Breeman, A. M. 1988. Relative importance of temperature and
    other factors in determining geographic boundaries of
    seaweeds: experimental and phenological evidence. Helgol.
    Wiss. Meeresunters 42:199–241.

    Brodie, J., Andersen, R. A., Kawachi, M. & Millar, A. J. K. 2009.
    Endangered algal species and how to protect them.
    Phycologia 48:423–38.

    Buddemeier, R. W., Lane, D. R. & Martinich, J. A. 2011.
    Modeling regional coral reef responses to global warming
    and changes in ocean chemistry: Caribbean case study. Clim.
    Change 109:375–97.

    Büdenbender, J., Riebesell, U. & Form, A. 2011. Calcification of
    the Arctic coralline red algae Lithothamnion glaciale in
    response to elevated CO2. Mar. Ecol. Prog. Ser. 441:79–87.

    Buschmann, A. H., VaAquez, J., Osorio, P., Reyes, E., Filun, L.,
    Hernandez-Gonzalez, M. C. & Vega, A. 2004. The effect of
    water movement, temperature and salinity on abundance
    and reproductive patterns of Macrocystis spp. (Phaeophyta) at
    different latitudes in Chile. Mar. Biol. 145:849–62.

    Campbell, A. H., Harder, T., Nielsen, S., Kjelleberg, S. &
    Steinberg, P. D. 2011. Climate change and disease: bleaching
    of a chemically defended seaweed. Global Change Biol.
    17:2958–70.

    Case, R. J., Longford, S. R., Campbell, A. H., Low, A., Tujula, N.,
    Steinberg, P. D. & Kjelleberg, S. 2011. Temperature
    induced bacterial virulence and bleaching disease in a
    chemically defended marine macroalga. Environ. Microbiol.
    13:529–37.

    Christie, H., Norderhaug, K. M. & Fredriksen, S. 2009.
    Macrophytes as habitat for fauna. Mar. Ecol. Prog. Ser.
    396:221–33.

    Chu, S. H., Zhang, Q. S., Liu, S. K., Tang, Y. Z., Zhang S. B.,
    Lu Z. C. & Yu Y. Q. 2012. Tolerance of Sargassum thunbergii
    germlings to thermal, osmotic and desiccation stress. Aq. Bot.
    96:1–6.

    Collén, J., Guisle-Marsollier, I., Léger, J. J. & Boyen, C. 2007.
    Response of the transcriptome of the intertidal red seaweed
    Chondrus crispus to controlled and natural stresses. New
    Phytol. 176:45–55.

    Connell, S. D. & Russell, B. D. 2010. The direct effects of
    increasing CO2 and temperature on non-calcifying
    organisms: increasing the potential for phase shifts in kelp
    forests. Proc. R. Soc. Lond. B Biol. Sci. 277:1409–15.

    Cornwall, C. E., Hepburn, C. D., Pritchard D., Currie, K. I.,
    McGraw, C. M., Hunter, K. A. & Hurd, C. L. 2012. Carbon
    use strategies in macroalgae: differential responses to
    lowered pH and implications for ocean acidification.
    J. Phycol. 48:137–44.

    Costanza, R., d’Arge, R., de Groot, R., Farber, S., Grasso, M.,
    Hannon, B., Limburg, K. et al. 1997. The value of the world’s
    ecosystem services and natural capital. Nature 387:253–60.

    Crim, R. N., Sunday, J. M. & Harley, C. D. G. 2011. Elevated
    seawater CO2 concentrations impair larval development and

    1074 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    reduce larval survival in endangered northern
    abalone (Haliotis kamtschatkana). J. Exp. Mar. Biol. Ecol. 400:
    272–7.

    Cubit, J. D. 1984. Herbivory and the seasonal abundance of algae
    on a high intertidal rocky shore. Ecology 65:1904–17.

    Dantonio, C. 1985. Epiphytes on the rocky intertidal red alga
    Rhodomela larix (Turner) C Agardh – negative effects on the
    host and food for herbivores. J. Exp. Mar. Biol. Ecol. 86:
    197–218.

    Davies, T. W., Jenkins, S. R., Kingham, R., Kenworthy, J.,
    Hawkins, S. J. & Hiddink, J. G. 2011. Dominance, biomass
    and extinction resistance determine the consequences of
    biodiversity loss for multiple coastal ecosystem processes.
    PLoS ONE 6:e28362.

    Davison, I. R., Greene, R. M. & Podolak, E. J. 1991. Temperature
    acclimation of respiration and photosynthesis in the brown
    alga Laminaria saccharina. Mar. Biol. 110:449–54.

    Davison, I. R. & Pearson, G. A. 1996. Stress tolerance in intertidal
    seaweeds. J. Phycol. 32:197–211.

    Deysher, L. E. & Dean, T. A. 1986a. In situ recruitment of
    sporophytes of the giant kelp, Macrocystis pyrifera (L) C.A.
    Agardh – effects of physical factors. J. Exp. Mar. Biol. Ecol.
    103:41–63.

    Deysher, L. E. & Dean, T. A. 1986b. Interactive efects of light and
    temperature on sporophyte production in the giant kelp
    Macrocystis pyrifera. Mar. Biol. 93:17–20.

    Diaz-Pulido, G., Anthony, K. R. N., Kline, D. I., Dove, S. &
    Hoegh-Guldberg, O. 2012. Interactions between ocean
    acidification and warming on the mortality and dissolution
    of coralling algae. J. Phycol. 48:32–9.

    Diaz-Pulido, G., Gouezo, M., Tilbrook, B., Dove, S. & Anthony, K.
    R. N. 2011. High CO2 enhances the competitive strength of
    seaweeds over corals. Ecol. Lett. 14:156–62.

    Diaz-Pulido, G., McCook, L. J., Larkum, A. W. D., Lotze, H. K.,
    Raven, J. A., Schaffelke, B., Smith, J. E. & Steneck, R. S.
    2007. Vulnerability of macroalgae of the Great Barrier Reef
    to climate change. In Johnson, J. E. & Marshall, P. A. [Eds.]
    Climate Change and the Great Barrier Reef. Great Barrier Reef
    Marine Park Authority and Australian Greenhouse Office,
    Townsville, pp. 154–92.

    Duffy, J. E. & Hay, M. E. 1990. Seaweed adaptations to herbivory –
    Chemical, structural, and morphological defenses are often
    adjusted to spatial or temporal patterns of attack. Bioscience
    40:368–75.

    Dupont, S., Ortega-Martinez, O. & Thorndyke, M. 2010. Impact
    of near-future ocean acidification on echinoderms.
    Ecotoxicology 19:449–62.

    Eggert, A., Peters, A. F. & Küpper, F. C. 2010. The potential
    impacts of climate change on endophyte infections in kelp
    sporophytes. In Israel, A., Einav, R. & Seckbach, J. [Eds.]
    Seaweeds and Their Role in Globally Changing Environments.
    Springer, New York, pp. 139–54.

    Eggert, A. 2012. Seaweed responses to temperature. In Wiencke,
    C. & Bischof, K. [Eds.] Seaweed Biology. Springer-Verlag,
    Berlin, Germany, pp. 47–66.

    Elfwing, T. & Tedengren, M. 2002. Effects of copper and reduced
    salinity on grazing activity and macroalgae production: a
    short-term study on a mollusc grazer, Trochus maculatus and
    two species of macroalgae in the inner Gulf of Thailand.
    Mar. Biol. 140:913–9.

    Fabricius, K. E., Langdon, C., Uthicke, S., Humphrey, C.,
    Noonan, S., De’ath, G., Okazaki, R., Muehllehner, N., Glas,
    M. S. & Lough, J. M. 2011. Losers and winners in coral reefs
    acclimatized to elevated carbon dioxide concentrations.
    Nature Clim. Change 1:165–9.

    Fain, S. R. & Murray, S. N. 1982. Effects of light and temperature
    on net photosynthesis and dark respiration of gametophytes
    and embryonic sporophytes of Macrocystis pyrifera. J. Phycol.
    18:92–8.

    Falkenberg, L. J., Russell, B. D. & Connell, S. D. 2012. Stability of
    strong species interactions resist the synergistic effects of
    local and global pollution in kelp forests. PLoS ONE 7:
    e33841.

    Feely, R. A., Sabine, C. L., Hernandez-Ayon, J. M., Ianson, D. &
    Hales, B. 2008. Evidence for upwelling of corrosive “acidified”
    water onto the continental shelf. Science 320:1490–2.

    Feely, R. A., Sabine, C. L., Lee, K., Berelson, W., Kleypas, J.,
    Fabry, V. J. & Millero, F. J. 2004. Impact of anthropogenic
    CO2 on the CaCO3 system in the oceans. Science 305:362–6.

    Folke, C., Carpenter, S., Walker, B., Scheffer, M., Elmqvist, T.,
    Gunderson, L. & Holling, C. S. 2004. Regime shifts,
    resilience, and biodiversity in ecosystem management. Annu.
    Rev. Ecol. Syst. 35:557–81.

    Forrest, B. M. & Blakemore, K. A. 2006. Evaluation of treatments
    to reduce the spread of a marine plant pest with aquaculture
    transfers. Aquaculture 257:333–45.

    Gachon, C. M. M., Sime-Ngando, T., Strittmatter, M.,
    Chambouvet, A. & Kim, G. H. 2010. Algal diseases: spotlight
    on a black box. Trends Plant Sci. 15:633–40.

    Gao, K., Aruga, Y., Asada, K., Ishihara, T., Akano, T. & Kiyohara,
    M. 1993. Calcification in the articulated coralline alga
    Corallina pilulifera, with special reference to the effect of
    elevated CO2 concentration. Mar. Biol. 117:129–32.

    Gao, K. S., Ji, Y. & Aruga, Y. 1999. Relationship of CO2
    concentrations to photosynthesis of intertidal macroalgae
    during emersion. Hydrobiologia 399:355–9.

    Gao, K. S. & Zheng, Y. Q. 2010. Combined effects of ocean
    acidification and solar UV radiation on photosynthesis,
    growth, pigmentation and calcification of the coralline alga
    Corallina sessilis (Rhodophyta). Global Change Biol. 16:
    2388–98.

    Giesecke, T., Miller, P. A., Sykes, M. T., Ojala, A. E. K., Seppa, H.
    & Bradshaw, R. H. W. 2010. The effect of past changes in
    inter-annual temperature variability on tree distribution
    limits. J. Biogeogr. 37:1394–405.

    Graham, M. H. 2004. Effects of local deforestation on the
    diversity and structure of Southern California giant kelp
    forest food webs. Ecosystems 7:341–57.

    Graham, M. H., Harrold, C., Lisin, S., Light, K., Watanabe, J. M.
    & Foster, M. S. 1997. Population dynamics of giant kelp
    Macrocystis pyrifera along a wave exposure gradient. Mar. Ecol.
    Prog. Ser. 148:269–79.

    Graham, M. H., Kinlan, B. P., Druehl, L. D., Garske, L. E. &
    Banks, S. 2007. Deep-water refugia as potential hotspots of
    tropical marine diversity and productivity. Proc. Nat. Acad.
    Sci. U.S.A. 104:16576–80.

    Hall-Spencer, J. M., Rodolfo-Metalpa, R., Martin, S., Ransome, E.,
    Fine, M., Turner, S. M., Rowley, S. J., Tedesco, D. & Buia, M. C.
    2008. Volcanic carbon dioxide vents show ecosystem effects of
    ocean acidification. Nature 454:96–9.

    Harley, C. D. G. 2003. Abiotic stress and herbivory interact to set
    range limits across a two-dimensional stress gradient. Ecology
    84:1477–88.

    Harley, C. D. G. 2011. Climate change, keystone predation, and
    biodiversity loss. Science 334:1124–7.

    Harley, C. D. G., Hughes, A. R., Hultgren, K. M., Miner, B. G.,
    Sorte, C. J. B., Thornber, C. S., Rodriguez, L. F., Tomanek,
    L. & Williams, S. L. 2006. The impacts of climate change in
    coastal marine systems. Ecol. Lett. 9:228–41.

    Harley, C. D. G. & Paine, R. T. 2009. Contingencies and compounded
    rare perturbations dictate sudden distributional shifts during
    periods of gradual climate change. Proc. Nat. Acad. Sci. U.S.A.
    106:11172–6.

    Harrold, C., Light, K. & Lisin, S. 1998. Organic enrichment of
    submarine-canyon and continental-shelf benthic communities
    by macroalgal drift imported from nearshore kelp forests.
    Limnol. Oceanogr. 43:669–78.

    Hart, M. W. & Scheibling, R. E. 1988. Heat waves, baby booms,
    and the destruction of kelp beds by sea urchins. Mar. Biol.
    99:167–76.

    Hawkins, S. J., Moore, P. J., Burrows, M. T., Poloczanska, E.,
    Mieszkowska, N., Herbert, R. J. H., Jenkins, S. R., Thompson,
    R. C., Genner, M. J. & Southward, A. J. 2008. Complex
    interactions in a rapidly changing world: responses of rocky
    shore communities to recent climate change. Climate Res.
    37:123–33.

    CLIMATE CHANGE AND SEAWEEDS 1075

    M
    IN
    IR
    E
    V
    IE
    W

    Hawkins, S. J., Sugden, H. E., Mieszkowska, N., Moore, P. J.,
    Poloczanska, E., Leaper, R., Herbert, R. J. H. et al. 2009.
    Consequences of climate-driven biodiversity changes for
    ecosystem functioning of North European rocky shores. Mar.
    Ecol. Prog. Ser. 396:245–59.

    Hay, M. E., Kappel, Q. E. & Fenical, W. 1994. Synergisms in plant
    defenses against herbivores – interactions of chemistry,
    calcification, and plant quality. Ecology 75:1714–26.

    Helmuth, B., Broitman, B. R., Blanchette, C. A., Gilman, S., Halpin,
    P., Harley, C. D. G., O’Donnell, M. J., Hofmann, G. E.,
    Menge, B. & Strickland, D. 2006. Mosaic patterns of thermal
    stress in the rocky intertidal zone: implications for climate
    change. Ecol. Monogr. 76:461–79.

    Hemmi, A. & Jormalainen, V. 2002. Nutrient enhancement
    increases performance of a marine herbivore via quality of
    its food alga. Ecology 83:1052–64.

    Hepburn, C. D., Pritchard, D. W., Cornwall, C. E., McLeod, R. J.,
    Beardall, J., Raven, J. A. & Hurd, C. L. 2011. Diversity of
    carbon use strategies in a kelp forest community: implications
    for a high CO2 ocean. Global Change Biol. 17:2488–97.

    Hernandez, J. C., Clemente, S., Girard, D., Perez-Ruzafa, A. &
    Brito, A. 2010. Effect of temperature on settlement and
    postsettlement survival in a barrens-forming sea urchin. Mar.
    Ecol. Prog. Ser. 413:69–80.

    Hoegh-Guldberg, O., Mumby, P. J., Hooten, A. J., Steneck, R. S.,
    Greenfield, P., Gomez, E., Harvell, C. D. et al. 2007. Coral
    reefs under rapid climate change and ocean acidification.
    Science 318:1737–42.

    Hoffman, J. R., Hansen, L. J. & Klinger, T. 2003. Interactions
    between UV radiation and temperature limit inferences from
    single-factor experiments. J. Phycol. 39:268–72.

    Hönisch, B., Ridgwell, A., Schmidt, D. N., Thomas, E., Gibbs, S. J.,
    Sluijs, A., Zeebe, R. et al. 2012. The geological record of
    ocean acidification. Science 335:1058–63.

    Hughes, T. P., Baird, A. H., Bellwood, D. R., Card, M., Connolly,
    S. R., Folke, C., Grosberg, R. et al. 2003. Climate change,
    human impacts, and the resilience of coral reefs. Science
    301:929–33.

    Hunt, L. J. H. & Denny, M. W. 2008. Desiccation protection and
    disruption: a trade-off for an intertidal marine alga. J. Phycol.
    44:1164–70.

    IPCC 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.
    Cambridge University Press, Cambridge, UK.

    Johnson, V. R., Russell, B. D., Fabricius, K. E., Brownlee, C. &
    Hall-Spencer, J. M. 2012. Temperate and tropical brown
    macroalae thrive, despite decalcification, along natural CO2
    gradients. Global Change Biol. 18:2792–803.

    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 27:473–83.

    Kavanaugh, M. T., Nielsen, K. J., Chan, F. T., Menge, B. A.,
    Letelier, R. M. & Goodrich, L. M. 2009. Experimental
    assessment of the effects of shade on an intertidal kelp: do
    phytoplankton blooms inhibit growth of open-coast
    macroalgae? Limnol. Oceanogr. 54:276–88.

    Kendall, M. A., Burrows, M. T., Southward, A. J. & Hawkins, S. J.
    2004. Predicting the effects of marine climate change on the
    invertebrate prey of the birds of rocky shores. Ibis 146:40–7.

    Kim, E., Park, H. S., Jung, Y., Choi, D. W., Jeong, W. J., Hwang,
    M. S., Park, E. J. & Gong, Y. G. 2011. Identification of the
    high-temperature response genes from Porphyra seriata
    (Rhodophyta) expression sequence tags and enhancement
    of heat tolerance of Chlamydomonas (Chlorophyta) by
    expression of the Porphyra HTR2 gene. J. Phycol. 47:821–8.

    Kordas, R. L., Harley, C. D. G. & O’Connor, M. I. 2011.
    Community ecology in a warming world: the influence of
    temperature on interspecific interactions in marine systems.
    J. Exp. Mar. Biol. Ecol. 400:218–26.

    Kroeker, K. J., Kordas, R. L., Crim, R. N. & Singh, G. G. 2010.
    Meta-analysis reveals negative yet variable effects of

    ocean acidification on marine organisms. Ecol. Lett. 13:
    1419–34.

    Krumhansl, K. A., Lee, J. M. & Scheibling, R. E. 2011. Grazing
    damage and encrustation by an invasive bryozoan reduce the
    ability of kelps to withstand breakage by waves. J. Exp. Mar.
    Biol. Ecol. 407:12–8.

    Kübler, J. E., Johnston, A. M. & Raven, J. A. 1999. The effects of
    reduced and elevated CO2 and O2 on the seaweed
    Lomentaria articulata. Plant, Cell Environ. 22:1303–10.

    Kuffner I. B., Andersson, A. J., Jokiel, P. L., Rodgers, K. S. &
    Mackenzie, F. T. 2008. Decreased abundance of crustose
    coralline algae due to ocean acidification. Nature Geosci.
    1:114–7.

    Ladah, L. B. & Zertuche-González, J. A. 2007. Survival of
    microscopic stages of a perennial kelp (Macrocystis pyrifera)
    from the center and the southern extreme of its range in
    the Northern Hemisphere after exposure to simulated El
    Nino stress. Mar. Biol. 152:677–86.

    Largo, D. B., Fukami, K. & Nishijima, T. 1995. Occasional
    pathogenic bacteria promoting ice-ice disease in the
    carrageenan-producing red algae Kappaphycus alvarezii and
    Eucheuma denticulatum (Solieriaceae, Gigartinales,
    Rhodophyta). J. Appl. Phycol. 7:545–54.

    Leonard, G. H. 2000. Latitudinal variation in species interactions:
    a test in the New England rocky intertidal zone. Ecology
    81:1015–30.

    Lewis, S., Donkin, M. E. & Depledge, M. H. 2001. Hsp70
    expression in Enteromorpha intestinalis (Chlorophyta) exposed
    to environmental stressors. Aquat. Toxicol. 51:277–91.

    Li, R. & Brawley, S. H. 2004. Improved survival under heat stress
    in intertidal embryos (Fucus spp.) simultaneously exposed to
    hypersalinity and the effect of parental thermal history. Mar.
    Biol. 144:205–13.

    Lima, F. P., Ribeiro, P. A., Queiroz, N., Hawkins, S. J. & Santos,
    A. M. 2007. Do distributional shifts of northern and
    southern species of algae match the warming pattern? Global
    Change Biol. 13:2592–604.

    Ling, S. D. 2008. Range expansion of a habitat-modifying species
    leads to loss of taxonomic diversity: a new and impoverished
    reef state. Oecologia 156:883–94.

    Lobban, C. S. & Harrison, P. J. 1997. Seaweed Ecology and
    Physiology. Cambridge University Press, New York.

    Long, S. P., Ainsworth, E. A., Rogers, A. & Ort, D. R. 2004. Rising
    atmospheric carbon dioxide: plants face the future. Ann.
    Rev. Plant Biol. 55:591–628.

    Lubchenco, J. & Gaines, S. D. 1981. A unified approach to
    marine plant-herbivore interactions. I. Populations and
    communities. Ann. Rev. Ecol. Syst. 12:405–37.

    Lüning, K. 1990. Seaweeds – Their Environment, Biogeography, and
    Ecophysiology. John Wiley & Sons Inc., New York.

    Lüning, K. & Neushul, M. 1978. Light and temperature
    demands for growth and reproduction of Laminarian
    gametophytes in southern and central California. Mar. Biol.
    45:297–309.

    Martin, S. & Gattuso, J. P. 2009. Response of Mediterranean
    coralline algae to ocean acidification and elevated
    temperature. Global Change Biol. 15:2089–100.

    Méléder, V., Populus, J., Guillaumont, B., Perrot, T. & Mouquet,
    P. 2010. Predictive modelling of seabed habitats: case study
    of subtidal kelp forests on the coast of Brittany. France. Mar.
    Biol. 157:1525–41.

    Mieszkowska, N., Kendall, M. A., Hawkins, S. J., Leaper, R.,
    Williamson, P., Hardman-Mountford, N. J. & Southward, A.
    J. 2006. Changes in the range of some common rocky shore
    species in Britain – a response to climate change?
    Hydrobiologia 555:241–51.

    Miller, K. A., Aguilar-Rosas, L. E. & Pedroche, F. F. 2011. A
    review of non-native seaweeds from California, USA and Baja
    California, Mexico. Hidrobiologica 21:365–79.

    Mumby, P. J. & Harborne, A. R. 2010. Marine reserves enhance
    the recovery of corals on caribbean reefs. PLoS ONE 5:e8657.

    Muñoz, V., Hernández-González, M. C., Buschmann, A. H.,
    Graham, M. H. & Vásquez, J. A. 2004. Variability in per

    1076 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    capita oogonia and sporophyte production from giant kelp
    gametophytes (Macrocystis pyrifera, Phaeophyceae). Rev. Chil.
    Hist. Nat. 77:639–47.

    Norderhaug, K. N., Christie, H., Fossa, J. H. & Fredriksen, S.
    2005. Fish-macrofauna interactions in a kelp (Laminaria
    hyperborea) forest. J. Mar.Biol. Assoc. UK 85:1279–86.

    O’Connor, M. I. 2009. Warming strengthens an herbivore-plant
    interaction. Ecology 90:388–98.

    O’Connor, M. I., Selig, E. R., Pinsky, M. L. & Altermatt, F. 2012.
    Toward a conceptual synthesis for climate change responses.
    Global Ecol. Biogeogr. 21:693–703.

    O’Leary, J. K. & McClanahan, T. R. 2010. Trophic cascades result
    in large-scale coralline algae loss through differential grazer
    effects. Ecology 91:3584–97.

    Padilla-Gamino, J. L. & Carpenter, R. C. 2007a. Seasonal
    acclimatization of Asparagopsis taxiformis (Rhodophyta) from
    different biogeographic regions. Limnol. Oceanogr. 52:833–42.

    Padilla-Gamino, J. L. & Carpenter, R. C. 2007b. Thermal
    ecophysiology of Laurencia pacifica and Laurencia nidifica
    (Ceramiales, Rhodophyta) from tropical and warm-
    temperate regions. J. Phycol. 43:686–92.

    Parmesan, C. & Yohe, G. 2003. A globally coherent fingerprint of
    climate change impacts across natural systems. Nature 421:
    37–42.

    Pearson, P. N. & Palmer, M. R. 2000. Atmospheric carbon
    dioxide concentrations over the past 60 million years. Nature
    406:695–9.

    Pereyra, R. T., Bergstrom, L., Kautsky, L. & Johannesson, K. 2009.
    Rapid speciation in a newly opened postglacial marine
    environment, the Baltic Sea. BMC Evol. Biol. 9:70.

    Perry, A. L., Low, P. J., Ellis, J. R. & Reynolds, J. D. 2005. Climate
    change and distribution shifts in marine fishes. Science
    308:1912–5.

    Pörtner, H. O. & Farrell, A. P. 2008. Physiology and climate
    change. Science 322:690–2.

    Porzio, L., Buia, M. C. & Hall-Spencer, J. M. 2011. Effects of
    ocean acidification on macroalgal communities. J. Exp. Mar.
    Biol. Ecol. 400:278–87.

    Price, N. N., Hamilton, S. L., Tootell, J. S. & Smith, J. E. 2011.
    Species-specific consequences of ocean acidification for the
    calcareous tropical green algae Halimeda. Mar. Ecol. Prog. Ser.
    440:67–78.

    Rabalais, N. N., Turner, R. E., Diaz, R. J. & Justic, D. 2009. Global
    change and eutrophication of coastal waters. ICES J. Mar. Sci.
    66:1528–37.

    Raven, J. A. & Geider, R. J. 1988. Temperature and algal growth.
    New Phytol. 110:441–61.

    Raven, J. A., Giordano, M., Beardall, J. & Maberly, S. C. 2012.
    Algal evolution in relation to atmospheric CO2: carboxylases,
    carbon-concentrating mechanisms and carbon oxidation
    cycles. Phil. Trans. R. Soc. B: Biol. Sci. 367:493–507.

    Reuter, K. E., Lotterhos, K. E., Crim, R. N., Thompson, C. A. &
    Harley, C. D. G. 2011. Elevated pCO2 increases sperm
    limitation and risk of polyspermy in the red sea urchin
    Strongylocentrotus franciscanus. Global Change Biol. 17:163–71.

    Roleda, M. Y., Boyd, P. W. & Hurd, C. L. 2012a. Before
    ocean acidification: calcifier chemistry lessons. J. Phycol.
    48:840–3.

    Roleda, M. Y., Morris, J. N., McGraw, C. M. & Hurd, C. L. 2012b.
    Ocean acidification and seaweed reproduction: increased
    CO2 ameliorates the negative effect of lowered pH on
    meiospore germination in the giant kelp Macrocystis pyrifera
    (Laminariales, Phaeophyceae). Global Change Biol. 18:854–64.

    Rönnbäck, P., Kautsky, N., Pihl, L., Troell, M., Söderqvist, T. &
    Wennhage, H. 2007. Ecosystem goods and services from
    Swedish coastal habitats: identification, valuation, and
    implications of ecosystem shifts. Ambio 36:534–44.

    Rothäusler, E., Gomez, I., Hinojosa, I. A., Karsten, U., Tala, F. &
    Thiel, M. 2009. Effect of temperature and grazing on growth
    and reproduction of floating Macrocystis spp. (Phaeophyceae)
    along a latitudinal gradient. J. Phycol. 45:547–59.

    Rothäusler, E., Gomez, I., Karsten, U., Tala, F. & Thiel, M. 2011.
    Physiological acclimation of floating Macrocystis pyrifera to

    temperature and irradiance ensures long-term persistence at
    the sea surface at mid-latitudes. J. Exp. Mar. Biol. Ecol. 405:
    33–41.

    Russell, B. D., Passarelli, C. A. & Connell, S. D. 2011. Forecasted
    CO2 modifies the influence of light in shaping subtidal
    habitat. J. Phycol. 47:744–52.

    Russell, B. D., Thompson, J. A. I., Falkenberg, L. J. & Connell, S. D.
    2009. Synergistic effects of climate change and local stressors:
    CO2 and nutrient-driven change in subtidal rocky habitats.
    Global Change Biol. 15:2153–62.

    Saunders, M. I., Metaxas, A. & Filgueira, R. 2010. Implications of
    warming temperatures for population outbreaks of a
    nonindigenous species (Membranipora membranacea, Bryozoa)
    in rocky subtidal ecosystems. Limnol. Oceanogr. 55:1627–42.

    Scheffer, M., Carpenter, S., Foley, J. A., Folke, C. & Walker, B.
    2001. Catastrophic shifts in ecosystems. Nature 413:591–6.

    Scheibling, R. E. & Gagnon, P. 2009. Temperature-mediated
    outbreak dynamics of the invasive bryozoan Membranipora
    membranacea in Nova Scotian kelp beds. Mar. Ecol. Prog. Ser.
    390:1–13.

    Schiel, D. R. & Foster, M. S. 2006. The Population Biology of Large
    Brown Seaweeds: Ecological Consequences of Multiphase Life
    Histories in Dynamic Coastal Environments. Annu. Rev. Ecol.
    Evol. Syst., Annual Reviews, Palo Alto, California, pp. 343–72.

    Schiel, D. R., Steinbeck, J. R. & Foster, M. S. 2004. Ten years of
    induced ocean warming causes comprehensive changes in
    marine benthic communities. Ecology 85:1833–9.

    Schonbeck, M. W. & Norton, T. A. 1980. Factors controlling the
    lower limits of fucoid algae on the shore. J. Exp. Mar. Biol.
    Ecol. 43:131–50.

    Serisawa, Y., Imoto, Z., Ishikawa, T. & Ohno, M. 2004. Decline of
    the Ecklonia cava population associated with increased
    seawater temperatures in Tosa Bay, southern Japan. Fish. Sci.
    70:189–91.

    Seymour, R. J., Tegner, M. J., Dayton, P. K. & Parnell, P. E. 1989.
    Storm wave-induced mortality of giant kelp, Macrocystis
    pyrifera, in southern California. Estuar. Coast. Shelf Sci. 28:
    277–92.

    Sousa, W. P. 1979. Disturbance in marine intertidal boulder
    fields: the nonequilibrium maintenance of species diversity.
    Ecology 60:1225–39.

    Southworth, J., Randolph, J. C., Habeck, M., Doering, O. C.,
    Pfeifer, R. A., Rao, D. G. & Johnston, J. J. 2000.
    Consequences of future climate change and changing
    climate variability on maize yields in the midwestern United
    States. Agric. Ecosyst. Environ. 82:139–58.

    Speidel, M., Harley, C. D. G. & Wonham, M. J. 2001. Recovery of
    the brown alga Fucus gardneri following a range of removal
    intensities. Aq. Bot. 71:273–80.

    Steen, H. 2004a. Effects of reduced salinity on reproduction and
    germling development in Sargassum muticum (Phaeophyceae,
    Fucales). Eur. J. Phycol. 39:293–9.

    Steen, H. 2004b. Interspecific competition between Enteromorpha
    (Ulvales: Chlorophyceae) and Fucus (Fucales: Phaeophyceae)
    germlings: effects of nutrient concentration, temperature,
    and settlement density. Mar. Ecol. Prog. Ser. 278:89–101.

    Sunday, J. M., Crim, R. N., Harley, C. D. G. & Hart, M. W. 2011.
    Quantifying rates of evolutionary adaptation in response to
    ocean acidification. PLoS ONE 6:e22881.

    Swanson, A. K. & Fox, C. H. 2007. Altered kelp (Laminariales)
    phlorotannins and growth under elevated carbon dioxide
    and ultraviolet-B treatments can influence associated
    intertidal food webs. Global Change Biol. 13:1696–709.

    Taylor, D. I. & Schiel, D. R. 2010. Algal populations controlled by
    fish herbivory across a wave exposure gradient on southern
    temperate shores. Ecology 91:201–11.

    Thom, R. M. 1996. CO2-enrichment effects on eelgrass (Zostera
    marina L) and bull kelp (Nereocystis luetkeana (Mert) P & R).
    Water Air Soil Poll. 88:383–91.

    Thompson, R. C., Wilson, B. J., Tobin, M. L., Hill, A. S. &
    Hawkins, S. J. 1996. Biologically generated habitat provision
    and diversity of rocky shore organisms at a hierarchy of
    spatial scales. J. Exp. Mar. Biol. Ecol. 202:73–84.

    CLIMATE CHANGE AND SEAWEEDS 1077

    M
    IN
    IR
    E
    V
    IE
    W

    Van Alstyne, K. L., Pelletreau, K. N. & Kirby, A. 2009. Nutritional
    preferences override chemical defenses in determining food
    choice by a generalist herbivore, Littorina sitkana. J. Exp. Mar.
    Biol. Ecol. 379:85–91.

    Vaselli, S., Bertocci, I., Maggi, E. & Benedetti-Cecchi, L. 2008.
    Assessing the consequences of sea level rise: effects of
    changes in the slope of the substratum on sessile
    assemblages of rocky seashores. Mar. Ecol. Prog. Ser. 368:9–22.

    Vayda, M. E. & Yuan, M. L. 1994. The heat-shock response of an
    Antarctic alga is evident at 5°C. Plant Mol. Biol. 24:229–33.

    Vermeij, M. J. A., van Moorselaar, I., Engelhard, S., Hornlein, C.,
    Vonk, S. M. & Visser, P. M. 2010. The effects of nutrient
    enrichment and herbivore abundance on the ability of turf
    algae to overgrow coral in the Caribbean. PLoS ONE 5:
    e14312.

    Vinueza, L. R., Branch, G. M., Branch, M. L. & Bustamante, R. H.
    2006. Top-down herbivory and bottom-up El Niño effects on
    Galapagos rocky-shore communities. Ecol. Monogr. 76:111–31.

    Wahl, M., Jormalainen, V., Eriksson, B. K., Coyer, J. A., Molis, M.,
    Schubert, H., Dethier, M. et al. 2011. Stress ecology in Fucus:
    abiotic, biotic, and genetic interactions. In Lesser, M. [Ed.]
    Advances in Marine Biology, Vol 59. Elsevier Academic Press
    Inc, San Diego, California, pp. 37–105.

    Wang, M. Y., Overland, J. E. & Bond, N. A. 2010. Climate projections
    for selected large marine ecosystems. J. Mar. Syst. 79:258–66.

    Weinberger, F., Rohde, S., Oschmann, Y., Shahnaz, L., Dobretsov, S.
    & Wahl, M. 2011. Effects of limitation stress and of disruptive
    stress on induced antigrazing defense in the bladder wrack
    Fucus vesiculosus. Mar. Ecol. Prog. Ser. 427:83–94.

    Wernberg, T., Russell, B. D., Thomsen, M. S., Gurgel, C. F. D.,
    Bradshaw, C. J. A., Poloczanska, E. S. & Connell, S. D. 2011a.

    Seaweed communities in retreat from ocean warming.
    Current Biol. 21:1828–32.

    Wernberg, T., Smale, D. A. & Thomsen, M. S. 2012. A decade of
    climate change experiments on marine organisms: procedures,
    patterns and problems. Global Change Biol. 18:1491–8.

    Wernberg, T., Thomsen, M. S., Tuya, F. & Kendrick, G. A. 2011b.
    Biogenic habitat structure of seaweeds change along a
    latitudinal gradient in ocean temperature. J. Exp. Mar. Bio.
    Ecol. 400:264–71.

    Wernberg, T., Thomsen, M. S., Tuya, F., Kendrick, G. A., Staehr,
    P. A. & Toohey, B. D. 2010. Decreasing resilience of kelp
    beds along a latitudinal temperature gradient: potential
    implications for a warmer future. Ecol. Lett. 13:685–94.

    Williams, S. L. & Dethier, M. N. 2005. High and dry: variation in
    net photosynthesis of the intertidal seaweed Fucus gardneri.
    Ecology 86:2373–9.

    Wootton, J. T., Pfister, C. A. & Forester, J. D. 2008. Dynamic
    patterns and ecological impacts of declining ocean pH in a
    high-resolution multi-year dataset. Proc. Nat. Acad. Sci. U.S.A.
    105:18848–53.

    Xu, Z. G., Zou, D. H. & Gao, K. S. 2010. Effects of elevated CO2 and
    phosphorus supply on growth, photosynthesis and nutrient
    uptake in the marine macroalga Gracilaria lemaneiformis
    (Rhodophyta). Bot. Mar. 53:123–9.

    Zacharioudaki, A., Pan, S. Q., Simmonds, D., Magar, V. & Reeve,
    D. E. 2011. Future wave climate over the west-European shelf
    seas. Ocean Dynam. 61:807–27.

    Zou, D. & Gao, K. 2002. Effects of desiccation and CO2
    concentrations on emersed photosynthesis in Porphyra
    haitanensis (Bangiales, Rhodophyta), a species farmed in
    China. Eur. J. Phycol. 37:587–92.

    1078 CHRISTOPHER D. G. HARLEY ET AL.

    M
    IN
    IR
    E
    V
    IE
    W

    High summer temperatures amplify functional differences between
    coral- and algae-dominated reef

    communities

    FLORIAN ROTH ,1,2,3,10 NILS RäDECKER ,1,4,5 SUSANA CARVALHO ,1 CARLOS M. DUARTE ,1,6 VINCENT
    SADERNE ,1 ANDREA ANTON ,1,6 LUIS SILVA ,1 MARIA LI. CALLEJA ,1,7 XOSÉ ANXELU G. MORÁN ,1

    CHRISTIAN R. VOOLSTRA ,1,4 BENJAMIN KüRTEN ,1,8 BURTON H. JONES ,1 AND CHRISTIAN WILD 9

    1Red Sea Research Center, King Abdullah University of Science and Technology (KAUST), Thuwal 23955 Saudi Arabia
    2Baltic Sea Centre, Stockholm University, Stockholm 10691 Sweden

    3Faculty of Biological and Environmental Sciences, Tvärminne Zoological Station, University of Helsinki, Helsinki 00014 Finland
    4Department of Biology, University of Konstanz, Konstanz 78457 Germany

    5Laboratory for Biological Geochemistry, School of Architecture, Civil and Environmental Engineering, École Polytechnique Fédérale
    de Lausanne (EPFL), Lausanne 1015 Switzerland

    6Computational Biology Research Center, King Abdullah University of Science and Technology (KAUST), Thuwal 23955 Saudi
    Arabia

    7Department of Climate Geochemistry, Max Planck Institute for Chemistry (MPIC), Mainz 55128 Germany
    8Project Management Jülich, Jülich Research Centre GmbH, Rostock 52425 Germany

    9Marine Ecology, Faculty of Biology and Chemistry, University of Bremen, Bremen 28359 Germany

    Citation: Roth, F., N. Rädecker, S. Carvalho, C. M. Duarte, V. Saderne, A. Anton, L. Silva, M. L. L. Call-
    eja, X. A. G. Morán, C. R. Voolstra, B. Kürten, B. H. Jones, and C. Wild. 2021. High summer tempera-
    tures amplify functional differences between coral- and algae-dominated reef communities. Ecology 10

    2

    (2):e03226. 10.1002/ecy.3226

    Abstract. Shifts from coral to algal dominance are expected to increase in tropical coral
    reefs as a result of anthropogenic disturbances. The consequences for key ecosystem functions
    such as primary productivity, calcification, and nutrient recycling are poorly understood, par-
    ticularly under changing environmental conditions. We used a novel in situ incubation
    approach to compare functions of coral- and algae-dominated communities in the central Red
    Sea bimonthly over an entire year. In situ gross and net community primary productivity, calci-
    fication, dissolved organic carbon fluxes, dissolved inorganic nitrogen fluxes, and their respec-
    tive activation energies were quantified to describe the effects of seasonal changes. Overall,
    coral-dominated communities exhibited 30% lower net productivity and 10 times higher calcifi-
    cation than algae-dominated communities. Estimated activation energies indicated a higher
    thermal sensitivity of coral-dominated communities. In these communities, net productivity
    and calcification were negatively correlated with temperature (>40% and >65% reduction,
    respectively, with +5°C increase from winter to summer), whereas carbon losses via respiration
    and dissolved organic carbon release more than doubled at higher temperatures. In contrast,
    algae-dominated communities doubled net productivity in summer, while calcification and dis-
    solved organic carbon fluxes were unaffected. These results suggest pronounced changes in
    community functioning associated with coral-algal phase shifts. Algae-dominated communities
    may outcompete coral-dominated communities because of their higher productivity and car-
    bon retention to support fast biomass accumulation while compromising the formation of
    important reef framework structures. Higher temperatures likely amplify these functional dif-
    ferences, indicating a high vulnerability of ecosystem functions of coral-dominated communi-
    ties to temperatures even below coral bleaching thresholds. Our results suggest that ocean
    warming may not only cause but also amplify coral–algal phase shifts in coral reefs.

    Key words: activation energy; biogeochemical cycling; climate change; community budget; ecosystem
    functioning; regime shifts.

    INTRODUCTION

    Community shifts and the ongoing loss of biodiversity
    (Brondizio et al. 2019) are altering the productivity and
    biogeochemistry of many ecosystems globally

    (Middleton and Grace 2004, Hooper et al. 2012, Naeem
    et al. 2012). These changes compound with local and
    global environmental perturbations, which can acceler-
    ate the alteration of essential ecosystem processes (Bal-
    vanera et al. 2006, Stachowicz et al. 2007). Thermal
    stress caused by climate change is, thereby, likely to exhi-
    bit the most substantial impact (Stillman 2019).
    Tropical coral reefs are hotspots of biodiversity that

    provide various ecosystem services that are supported by

    Manuscript received 17 March 2020; revised 6 July 2020;
    accepted 24 August 2020. Corresponding Editor: Richard B.
    Aronson.

    10 E-mail: florian.roth@su.se

    Article e03226; page 1

    Ecology, 102(2), 2021, e03226
    © 2020 The Authors. Ecology published by Wiley Periodicals LLC on behalf of Ecological Society of America
    This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any
    medium, provided the original work is properly cited.

    https://orcid.org/0000-0003-4004-5863

    https://orcid.org/0000-0003-4004-5863

    https://orcid.org/0000-0003-4004-5863

    https://orcid.org/0000-0002-2387-8567

    https://orcid.org/0000-0002-2387-8567

    https://orcid.org/0000-0002-2387-8567

    https://orcid.org/0000-0003-1300-1953

    https://orcid.org/0000-0003-1300-1953

    https://orcid.org/0000-0003-1300-1953

    https://orcid.org/0000-0002-1213-1361

    https://orcid.org/0000-0002-1213-1361

    https://orcid.org/0000-0002-1213-1361

    https://orcid.org/0000-0003-3968-2718

    https://orcid.org/0000-0003-3968-2718

    https://orcid.org/0000-0003-3968-2718

    https://orcid.org/0000-0002-4104-2966

    https://orcid.org/0000-0002-4104-2966

    https://orcid.org/0000-0002-4104-2966

    https://orcid.org/0000-0001-8434-6559

    https://orcid.org/0000-0001-8434-6559

    https://orcid.org/0000-0001-8434-6559

    https://orcid.org/0000-0002-5992-2013

    https://orcid.org/0000-0002-5992-2013

    https://orcid.org/0000-0002-5992-2013

    https://orcid.org/0000-0002-9823-5339

    https://orcid.org/0000-0002-9823-5339

    https://orcid.org/0000-0002-9823-5339

    https://orcid.org/0000-0003-4555-379

    5

    https://orcid.org/0000-0003-4555-3795

    https://orcid.org/0000-0003-4555-3795

    https://orcid.org/0000-0003-0328-7847

    https://orcid.org/0000-0003-0328-7847

    https://orcid.org/0000-0003-0328-7847

    https://orcid.org/0000-0002-9599-1593

    https://orcid.org/0000-0002-9599-1593

    https://orcid.org/0000-0002-9599-1593

    https://orcid.org/0000-0001-9637-6536

    https://orcid.org/0000-0001-9637-6536

    https://orcid.org/0000-0001-9637-6536

    info:doi/10.1002/ecy.3226

    mailto:

    http://creativecommons.org/licenses/by/4.0/

    http://crossmark.crossref.org/dialog/?doi=10.1002%2Fecy.3226&domain=pdf&date_stamp=2020-12-27

    one or more metabolic or biogeochemical functions
    (e.g., primary production, calcification, organic matter
    fluxes, and nutrient cycling; Moberg and Folke 1999).
    Many of these processes are primarily driven by sclerac-
    tinian corals, the “ecosystem engineers” of tropical reefs
    (Wild et al. 2011). However, the combination of global
    and local anthropogenic stressors has caused extensive
    coral mortality and subsequent shifts from complex
    coral-dominated communities to simplified communities
    with a predominance of filamentous turf- and macroal-
    gae in many reefs around the world (Done 1992, Bell-
    wood et al. 2004, Hughes et al. 2007, Graham et al.
    2015). Although coral–algal phase shifts are increasingly
    observed globally, the consequences for reef ecosystem
    functions such as productivity, calcification, and nutri-
    ent cycling are poorly understood. Laboratory and
    mesocosm studies indicate that reef algae, particularly
    the widespread filamentous turfs, are metabolically very
    different from corals, and generally display significantly
    higher primary production rates (Rix et al. 2015, Cardini
    et al. 2016). At the same time, the fraction of the photo-
    synthetically fixed carbon (C) being exuded into the
    environment is generally more labile (Nelson et al.
    2013). At the community level, these differences may
    result in changes in the carbonate chemistry of seawater
    (McMahon et al. 2013, Bernstein et al. 2016), disrupted
    trophic structures (Johnson et al. 1995, Hempson et al.
    2018), or increased microbial loads on algae-dominated
    reefs worldwide (Jessen et al. 2013, Haas et al. 2016).
    Divergent responses to changing environmental condi-

    tions may amplify ecosystem functions of corals and
    algae differently. As such, changing temperature regimes
    and recurrent heatwaves, which are increasing in fre-
    quency and magnitude (Frölicher et al. 2018, Oliver
    et al. 2018), can have detrimental effects on tropical
    coral reef taxa (Lough et al. 2018, Hughes et al. 2019).
    In corals, sublethal heat stress during summer can

    compromise primary production and calcification (Rey-
    naud et al. 2003, Anthony et al. 2008), thereby altering
    the release of organic and inorganic products (Niggl
    et al. 2009, Piggot et al. 2009). In contrast, benthic turf-
    and macroalgae may be less sensitive to heat (Koch et al.
    2013), showing increased productivity and net growth
    with rising temperature (Bender et al. 2014). Likewise,
    temperature-related productivity optima and mortality
    thresholds of algae are often well above those of corals
    (Anton et al. 2020). Similarly, the abundance of reef
    algae can increase seasonally, especially during the sum-
    mer months (Lirman and Biber 2000, Diaz-Pulido and
    Garzón-Ferreira 2002, Ateweberhan et al. 2006).
    However, few studies investigated the effects of coral–-

    algal phase shifts on community metabolism, particu-
    larly in situ. This paucity of information probably
    reflects the logistical challenges of quantifying the func-
    tions of structurally complex communities in their natu-
    ral environment (Roth et al. 2019). Currently, most data
    describing ecosystem functions are derived from labora-
    tory (e.g., Cardini et al. 2016) and mesocosm (e.g.,

    Langdon et al. 2003, Bellworthy and Fine 2018,
    Edmunds et al. 2020) studies using either single organ-
    isms or simplified reconstructed communities to predict
    in situ changes at the community scale. Although these
    approaches provide valuable mechanistic insights and
    permit a tight control of environmental conditions dur-
    ing the experiments, they can only approximate natural
    conditions. However, primary production, calcification,
    and organic matter recycling critically depend on local
    environmental conditions, biodiversity, and system
    heterogeneity (Baird et al. 2007). In addition, large parts
    of the energy and nutrient pool are remineralized by
    microbial communities or cryptic fauna within the reef
    matrix (Richter and Wunsch 1999, de Jongh and Van
    Duyl 2004, Maldonado et al. 2012), all of which are gen-
    erally not considered in ex situ experimental setups.
    Concordantly, Roth et al. (2019) highlighted in a com-
    parison between laboratory-based single-organism and
    in situ incubations that ex situ measurements that are
    scaled up to average-constructed communities can over-
    estimate community-wide net primary production and
    underestimate respiration and gross photosynthesis by
    20–90%. Hence, laboratory experiments can only pro-
    vide a glimpse of the complex environmental dynamics
    (e.g., seasonality) that shape the ecological processes of
    reef communities (Damgaard 2019).
    To overcome these experimental constraints, we used

    a novel in situ approach that allowed the quantification
    of major metabolic and biogeochemical pathways (Roth
    et al. 2019) of co-occurring natural coral- and algae-
    dominated reef communities in the Red Sea. With a total
    of 112 light and dark in situ incubations, we measured
    rates of community production (i.e., net community pro-
    duction [NCP], community respiration [CR], and gross
    primary production [GPP]), net community calcification
    (NCC), net dissolved organic carbon (DOC), and dis-
    solved inorganic nitrogen (DIN) fluxes bimonthly for an
    entire year. In addition, we quantified the thermal-de-
    pendence of the functioning of benthic communities by
    applying principles of the metabolic theory of ecology
    (MTE; Sibly et al. 2012). We quantified the temperature
    sensitivity of metabolic processes using the activation
    energy (Ea), as the slope or rate of change in the rise and
    falling phases of a thermal performance curve before
    and after achieving the optimal temperature. Although
    activation energies are commonly assessed at the organ-
    ism level (Garcı́a et al. 2018, Savva et al. 2018, Anton
    et al. 2020), they also provide useful insights regarding
    the sensitivity of community metabolism to warming
    (Follstad Shah et al. 2017, Morán et al. 2017, Padfield
    et al. 2017).
    Thus, we (1) directly compare the magnitudes and

    directions of key functions of coral-dominated and
    phase-shifted algae-dominated reef communities, (2

    )

    derive their functional responses to environmental
    changes induced by seasonality, and (3) describe their
    thermal sensitivity to seasonally variable temperature
    changes.

    Article e03226; page 2 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    MATERIALS AND METHODS

    Study site and environmental conditions

    This experiment was carried out at Abu Shosha reef
    located in the central Red Sea on the west coast of
    Saudi Arabia (22°18’16.3’’ N; 39°02’57.7’’ E) from
    January 2017 until January 2018. Key environmental
    variables were monitored at the sampling site and were
    previously reported in Roth et al. (2018). Water temper-
    ature was measured continuously (logging interval =
    30 min) for the whole study period with Onset HOBO
    temperature/light data loggers (accuracy: �0.2°C)
    deployed at the seafloor. Salinity was measured at each
    day of sampling with a WTW TetraCon® conductivity
    cell (accuracy: �0.5% of value). Light availability was
    measured continuously (logging interval = 1 min) on
    three full days per month with the above-mentioned
    Onset HOBO data logger. Light readings were con-
    verted from lux to photosynthetically active radiation
    (PAR; μmol quanta�m−2�s−1; 400–700 nm wavelengths)
    by intercalibration and conversion as outlined in Roth
    et al. (2018), and values are presented as daytime
    means. Seawater samples for the determination of dis-
    solved nitrate (NO�3 ), nitrite (NO


    2 ), ammonium

    (NHþ4 ), phosphate (PO
    3�
    4 ), and monomeric silicate (Si

    (OH)4) were taken in triplicates each month from 1 m
    above the seafloor. Details for sampling and analysis
    can be found in Appendix S1: Section S1. The sum of
    NO�3 , NO


    2 , and NH

    þ
    4 is termed “dissolved inorganic

    nitrogen” (DIN) henceforth.

    Benthic communities selected for in situ incubations

    Abu Shosha reef is characterized by a heterogeneous
    mosaic of patches of coral- and algae-dominated com-
    munities. Thus, this site allows for the quantification of
    the functionality of both communities under identical
    environmental conditions. Nnatural coral- and algae-
    dominated reef communities surrounded by sand were
    haphazardly selected at the study site at 5 m water depth
    within an area of 50 × 50 m. The communities had to
    fulfill the following characteristics to qualify as “suitable
    candidates” for later incubations: (1) Coral-dominated
    communities were defined by having >40% coral cover
    but <10% algae cover; (2) algae-dominated communities were defined by having >40% algae cover but <10% coral cover; (3) all communities had to fit into the incu- bation chambers (max. diameter 50 cm, max. height 39 cm). Among all suitable candidates in the study area, four coral-dominated and four algae-dominated com- munities were chosen randomly. These eight communi- ties were revisited each month of sampling. The community composition at the level of major

    functional groups was assessed for each community
    three times during the study period (i.e., in the begin-
    ning, after 6 months, and at the end of the experiments).
    Details on the assessment and statistical evaluation of

    the community composition can be found in Appendix
    S1: Section S1. Two-factor permutational multivariate
    analysis of variance (PERMANOVA) indicated that dif-
    ferences between communities grouped according to
    coral and algal dominance were significant (P = 0.001;
    visualization in Appendix S1: Fig. S1). As no significant
    changes over time in the relative benthic cover were
    detected (Appendix S1: Table S2),the benthic commu-
    nity composition of each treatment was averaged over
    all replicates and survey points (Fig. 1a).
    Algae-dominated communities were considered

    “phase-shifted,” as complex structures and the occur-
    rence of coral rubble indicated that branching corals
    were present previously. The co-occurrence of coral and
    algal reef communities within small spatial ranges was
    reported as “mosaic-dynamics” before (e.g., Edmunds
    2002, Tkachenko et al. 2007) and may be explained by a
    combination of local processes and historical effects,
    such as previous stress events or adaptation (Done et al.
    1991, Bythell et al. 2000, Edmunds 2002).
    Cryptic habitats can encompass about 60–75% of the

    total surface area of a reef (Richter and Wunsch 1999,
    Richter et al. 2001), but organisms living in cracks and
    crevices within the communities’ matrices could not be
    assessed by our conventional benthic surveys. These
    organisms (e.g., sponges, bryozoans, and tunicates),
    however, metabolize organic matter in the order of
    15–30% of the gross production of a coral reef (reviewed
    in de Jongh and Van Duyl 2004), driving community res-
    piration and other biogeochemical fluxes assessed in this
    study. As our benthic incubation chambers jointly cap-
    tured the metabolism of all members of the communities
    (i.e., from the visible surface and cryptic habitats), we
    refrained from assigning measured metabolic activities
    to functional groups on the visible surface only, as the
    inferred contribution would be highly biased. Thus,
    measurements presented in this study represent commu-
    nity-wide processes that include all compartments of the
    reef benthos and the surrounding water.

    In situ incubations and quantification of community
    functions

    In situ incubations with benthic chambers were per-
    formed according to the protocol described in Roth
    et al. (2019). In brief, chambers were constructed from
    polymethyl methacrylate (PMMA) cylinders with
    removable gastight lids of the same material. All cham-
    bers were equipped with individual water circulation
    pumps with adjustable flow control, autonomous
    recording dissolved oxygen (DO), and temperature sen-
    sors (HOBO U26; temperature corrected and salinity
    adjusted), and two sampling ports for discrete water
    samples. Incubations were carried out on three consecu-
    tive days in January, March, May, July, September, and
    November 2017, and in January 2018. Generally, on day
    1, divers deployed four chambers on coral-dominated,
    and four chambers on algae-dominated reef

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 3

    communities (Fig. 1a). The chambers were positioned
    carefully and left in place with open tops (no lids) until
    the next morning. On the second day, incubations
    started at around 09:00 a.m. by tightly securing the lids
    and closing all sampling ports during natural daylight
    conditions. The exact incubation start and end time was
    recorded for each chamber. Incubations ran for approxi-
    mately 2 h. The chambers were left in place with open
    tops for a second set of incubation on the following day.
    On day 3, benthic communities were incubated at “simu-
    lated” darkness during the same period used for incuba-
    tions the previous day. The procedure followed the same
    as on day two; however, all chambers were covered with
    thick black PVC covers. Any light penetration into the
    chambers was prevented, as validated by control read-
    ings of Onset HOBO temperature/light data loggers
    within chambers. Between deployments, all materials
    were rinsed with freshwater, washed with 4% hydrochlo-
    ric acid (HCl), and subsequently rinsed with deionized

    water for reliable water chemistry measurements that
    included sensitive DOC samples.
    Discrete water samples for dissolved inorganic carbon

    (DIC), total alkalinity (TA), DOC, and DIN were with-
    drawn from the sampling ports with acid-washed syr-
    inges at the beginning and the end of each incubation
    (details for the analysis of these samples can be found in
    Appendix S1: Section S1). Changes in seawater chem-
    istry between start and end of incubations were used to
    calculate rates of NCP, CR, GPP (calculated as GPP =
    NCP + |CR|), NCC, and fluxes of DOC, and DIN. All
    rates and fluxes were extrapolated to incubation water
    volume (in L) and normalized to incubation duration (in
    h) and the planar reef area (in m2) of the enclosed ben-
    thic community adapted after Roth et al. (2019).
    Productivity and respiration rates (NCP and CR; in

    mmol C�m−2�h−1) were calculated by changes in DIC
    concentrations, taking into account calcification and
    dissolution rates according to an adapted protocol by

    (a)

    (b)

    (c)

    (

    (

    (%)

    FIG 1. Experimental setup, relative benthic cover, and key environmental variables at the study site. (a) Relative benthic cover of
    functional groups in the studied coral- and algae-dominated reef communities, and exemplary pictures of the incubation chambers
    on the respective substrates. Details on the community composition of each replicate and time point can be found in Appendix S1:
    Section S1 and Fig. S1. (b) Photosynthetically active radiation (PAR, in μmol photons�m−2�s−1), dissolved inorganic nitrogen (DIN,
    in μM), and aragonite saturation state (Ωarag) at the study site from January 2017 until January 2018. Circles represent values of dis-
    crete samples for DIN and Ωarag, and daytime averages of three separate days per month for PAR; lines represent the smoothed
    trend through the means. (c) Plot of seawater temperature at experimental site from January 2017 until January 2018. Each dot rep-
    resents one measurement at 30-min intervals. Dashed horizontal line depicts the mean maximum annual temperature modeled for
    the region from 1982 to 2015, taken from Chaidez et al. (2017). [Color figure can be viewed at wileyonlinelibrary.com]

    Article e03226; page 4 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    www.wileyonlinelibrary.com

    Albright et al. (2013). Rates of NCP and CR based on
    DIC fluxes were compared to rates based on oxygen
    fluxes from continuous measurements with DO sensors.
    No discrepancy between C and oxygen measurements
    was detected (r = 0.99, P < 0.0001, n = 112). The calcu- lated photosynthetic (1.05 � 0.02) and respiratory (0.97 � 0.02) quotients agree with those obtained by various authors elsewhere, typically ∼1 mol of oxygen produced for 1 mol C fixed, and vice versa (e.g., Gattuso et al. 1999b, Atkinson and Falter 2003). NCC (in mmol CaCO3�m−2�h−1) was calculated by

    concentration differences in TA, which are primarily
    caused by calcification and dissolution of CaCO3,
    whereby TA is reduced (increased) by two molar equiva-
    lents for every mole of CaCO3 produced (dissolved)
    (Zeebe and Wolf-Gladrow 2001). Nutrients fluxes (i.e.,
    NO�3 , NH

    þ
    4 , PO

    3�
    4 , and SO

    2�
    4 ) that cause a change in TA

    unrelated to calcification and dissolution were
    accounted for according to Zeebe and Wolf-Gladrow
    (2001) and Wolf-Gladrow et al. (2007). DOC (in mmol
    C�m−2�h−1) and DIN (in µmol N�m−2�h−1) fluxes were
    calculated from concentration differences between start
    and endpoints. Any temperature corrections that were
    necessary for seawater chemistry calculations were
    achieved by temperature readings from individual tem-
    perature loggers within each chamber.
    Although measurements presented in this study only

    relate to small benthic communities, most studies cur-
    rently work with individual reef organisms (e.g., Anton
    et al. 2020) or reconstructed communities (e.g.,
    Edmunds et al. 2020) to derive community functions.
    Thus, the results presented here are among the best
    approximations for the quantification of community-
    wide (biogeochemical) ecosystem functions of
    untouched, natural coral reef communities in situ (but
    see Haas et al. 2013, Van Heuven et al. 2018).

    Data analysis

    Statistical analyses were performed using JMP©
    Pro14 (SAS Institute) statistic software. Environmental
    variables and response parameters from incubations
    were grouped into spring (March–May), summer
    (June–September), fall (October–November), and winter
    (December–February) for statistical analysis. For the
    seasonal comparison and to derive GPP/CR ratios,
    hourly rates from light and dark incubations were used
    to calculate daily net fluxes. We acknowledge that there
    is a chance for a slight over- or underestimation because
    of associated changes in environmental conditions (e.g.,
    light) during the course of the day. Thus, to minimize
    the error associated with extrapolating hourly rates, we
    chose a time window for daylight incubations (from
    around 09:00 a.m. to 11:00 a.m.) that is closest to day-
    time average irradiation and excludes the “ramping up”
    phase in the early morning hours and extreme values
    that can occur during midday. As all incubations during
    all sampling periods were conducted at the very same

    time of the day, incubations are comparable across com-
    munity types and time points.
    The full seawater carbonate system parameters were

    derived for each sampling period from measured salinity,
    temperature, nutrients, TA, and DIC data using the
    CO2SYS Microsoft Excel Macro by Pierrot et al. (2006)
    and the R package Seacarb (Lavigne and Gattuso 2013)
    (Appendix S1: Table S3). Environmental variables were
    tested for differences with two-tailed t-tests. Response
    parameters from incubations (GPP, NCP, CR, NCC,
    DOC, and DIN) were assessed by linear mixed models
    (LMMs) to test for differences in the respective response
    parameters with ‘treatment’ (coral- vs. algae-dominated)
    and ‘season’ (spring, summer, fall, and winter) as fixed
    factors, and the sampling dates (date) within seasons
    and the replicates of the communities (community ID)
    as random factors. Tukey’s Honest Significant Differ-
    ence (HSD) test was used for pairwise comparisons if
    significant interactions (treatment * season) were found.
    Detailed statistical results, including significant post hoc
    comparisons, are presented in Appendix S1: Table S4.
    The relationships between response (e.g., metabolic
    functions) and explanatory variables (e.g., environmen-
    tal variables) were assessed by linear regression models.
    The thermal sensitivity of the metabolic processes

    (GPP, NCP, CR, NCC, DOC, and DIN) was explored
    by calculating the activation energy (Ea) based on
    Arrhenius equations (Sibly et al. 2012) within the sea-
    sonal thermal regime (25.0–32.8°C). The activation ener-
    gies (Ea in eV) were estimated by fitting a linear
    regression equation between the natural logarithm of the
    metabolic rates and the reciprocal of temperature (1/kT),
    where k is the Boltzmann’s constant (8.62 × 10−5 eV/K)
    and T is the water temperature (K). To deal with obser-
    vations ≤0 on log-transformed data (e.g., NCC, DOC,
    and DIN rates), a constant was added (i.e., ln(rate + 1 −
    min value(rate)) to shift all values above zero (Legendre
    and Legendre 2012). The alternative of excluding ≤0 val-
    ues was discarded because these measurements are an
    important part of the biological processes under investi-
    gation (Canavero et al. 2018).

    RESULTS

    Environmental conditions

    Monthly monitored environmental variables at the
    study site exhibited strong seasonal patterns (Fig. 1b,c,
    Appendix S1: Table S2). The average seawater tempera-
    ture ranged from 25.8 � 0.2°C in winter to
    32.3 � 0.1°C in summer (Fig. 1c). Simultaneously, aver-
    age daytime PAR intensities at 5 m water depth
    increased from 130 � 2 μmol photons�m−2�s−1 to
    465 � 14 μmol photons�m−2�s−1 (Fig. 1b). Seawater
    DIN concentrations were lowest in spring and winter
    (0.46 � 0.02 and 0.66 � 0.04 μM DIN, respectively)
    and significantly higher in summer and fall (1.03 � 0.06
    and 1.08 � 0.13 μM DIN, respectively, Fig. 1b).

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 5

    Functions of coral- and algae-dominated reef communities

    The rates observed along the various deployments
    (Fig. 2) were used to calculate average values over the
    whole study period (Table 1). Average NCP was 30%
    and CR 50% lower in coral- as compared to algae-domi-
    nated communities (NCP, mean � SE: 26.7 � 1.2 and
    36.9 � 1.7 mmol C�m−2�h−1, respectively; CR:
    −10.9 � 0.7 and −20.9 � 1.4 mmol C�m−2�h−1, respec-
    tively; Fig. 2a,b). Integrated over 1 d, these differences
    yielded a 40% lower GPP of coral- compared to algae-
    dominated communities, with GPP/CR ratios of
    2.4 � 0.1 and 2.0 � 0.2, respectively (Table 1).
    NCC in the light was sixfold higher in coral- com-

    pared to algae-dominated communities, averaging
    7.9 � 0.7 mmol and 1.3 � 0.2 mmol CaCO3�m−2�h−1,
    respectively. In the dark, coral-dominated communities
    displayed an NCC of 3.0 � 0.3 mmol CaCO3�m−2�h−1,
    whereas algae-dominated communities exhibited an
    NCC of −0.5 � 0.2 mmol CaCO3�m−2�h−1 (represent-
    ing net CaCO3 dissolution). Integrating the hourly rates
    over 24 h, corals showed a 10-fold higher NCC com-
    pared to algae-dominated reef communities (Table 1).
    Coral-dominated communities were net sources of

    DOC during both light and dark incubations, with aver-
    age fluxes of 0.57 � 0.07 and 0.62 � 0.11 mmol C�m−2-
    �h−1, respectively (Table 1). In contrast, algae-dominated
    communities released similar amounts of DOC as corals
    in the light (0.70 � 0.08 mmol C�m−2�h−1) but were net
    sinks of DOC in the dark (−0.37 � 0.05 mmol C�m−2-
    �h−1). When integrated over 24 h, net DOC fluxes in
    coral-dominated communities were 3.5-fold higher than
    those in algae-dominated communities (Table 1).
    Both coral- and algae-dominated communities were net

    sources of DIN (1.62 � 0.21 and 1.66 � 0.23
    mmol N�m−2�d−1, respectively; Table 1), with no signifi-
    cant differences between treatments. There were, however,
    significant differences between light and dark incubations:
    algae-dominated communities released three times more
    DIN in the dark as compared to light conditions
    (109.9 � 14.1 μmol N�m−2�h−1 and 28.5 � 8.7, respec-
    tively). In contrast, coral-dominated communities released
    DIN at consistent rates during light and in dark incuba-
    tions (65.7 � 11.6 and 69.7 � 8.9 μmol N�m−2�h−1).

    Temporal variability of reef functions

    Both C and N fluxes showed temporal variations;
    however, significant differences in the magnitude and
    directions occurred between coral- and algae-dominated
    reef communities (Fig. 2, Table 1; detailed statistics in
    Appendix S1: Table S4).
    Daily-integrated GPP in coral-dominated communi-

    ties remained stable throughout the year; however, CR
    increased by >60% from winter to summer, resulting in
    40% lower NCP (Table 1). GPP in algae-dominated
    communities doubled from winter to summer, resulting
    in significantly increased NCP that peaked at

    >500 mmol C�m−2�d−1 in summer. In both community
    types, variations in NCP were significantly correlated
    with seawater temperature (Fig. 3a). NCP of coral-

    Dark incubationsLight incubations

    –8

    0

    0

    80

    160

    2

    40

    –5

    0
    5

    10

    Ja
    nu

    ar
    y

    M
    ar

    ch
    M

    ay
    Ju

    ly
    S

    ep
    te

    m
    be

    r
    N

    ov
    em

    be
    r

    Ja
    nu
    ar
    y
    Ja
    nu
    ar
    y
    M
    ar
    ch
    M
    ay
    Ju
    ly
    S
    ep
    te
    m
    be
    r
    N
    ov
    em
    be
    r
    Ja
    nu
    ar
    y

    N
    C

    C

    (
    m

    m
    o

    l C
    a
    C

    O
    m

    ·
    h

    )
    D

    O
    C

    (
    m
    m
    o

    l C
    m

    ·h
    )

    D
    IN

    (
    μ
    m

    o
    l N

    m
    ·
    h

    )

    N

    C
    P

    a
    n
    d

    C
    R

    (
    m
    m
    o
    l C
    m
    ·
    h
    )

    20

    0
    20
    40

    Coral-dominated
    Algae-dominated

    (b)

    (c)

    (d)

    (e) (f)

    (g) (h)

    (a)

    Months

    0.7

    0
    0.7

    1.

    4

    FIG 2. Hourly biogeochemical fluxes during light (left panels)
    and dark (right panels) incubations of coral- and algae-dominated
    reef communities of the central Red Sea. Presented are all data
    from bimonthly incubations of (a, b) community metabolism (di-
    vided into net community production (NCP) and community res-
    piration (CR), (c, d) net community calcification (NCC), (e, f) net
    dissolved organic carbon (DOC) fluxes, and (g, h) net dissolved
    inorganic nitrogen (DIN) fluxes. Dashed lines connect the
    bimonthly means. Shaded areas connect the standard error around
    the means. [Color figure can be viewed at wileyonlinelibrary.com]

    Article e03226; page 6 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    www.wileyonlinelibrary.com

    dominated communities exhibited a negative (r = −0.81,
    P < 0.0001, n = 26), and NCP of algae-dominated com- munities exhibited a positive (r = 0.83, P < 0.0001, n = 26) relationship with increasing temperature. Hence, coral-dominated communities showed apparent negative Ea values for GPP and NCP, indicating them to be in the falling phase of the performance curve (i.e., past the opti- mum temperature), while apparent Ea for CR was posi- tive, indicating an opposite trend (Appendix S1: Fig. S3). In contrast, algae-dominated communities had positive Ea values for GPP, NCP, and, CR (Table 2), indicating that these communities remained within the rising phase of the performance curve throughout the study period (including the summer months; Appendix S1: Fig. S3). NCC in coral-dominated communities peaked in

    spring, fall, and winter, with no differences between
    these seasons, but dropped sharply by >60% in summer
    (Table 1). Contrary, NCC in algae-dominated communi-
    ties was consistently low throughout the year (Fig. 2c,d;
    Table 1). NCC in coral-dominated communities corre-
    lated negatively with increasing water temperatures
    (r = −0.62, P = 0.0005, n = 26; Fig. 3b), negatively with
    Ωarag (r = −0.36, P = 0.0070, n = 54) (Fig. 4a), and pos-
    itively with NCP (r = 0.73, P < 0.0001, n = 54; Fig. 4b). NCC of algae-dominated communities did neither corre- late significantly with Ωarag, NCP, nor water tempera- ture. The corresponding apparent Ea values are presented in Table 2.

    Net DOC fluxes per day in coral-dominated commu-
    nities were lowest in winter and doubled in summer
    (Table 1), owing to both increases in DOC releases dur-
    ing dark and light incubations. While generally one
    order of magnitude lower, net DOC fluxes integrated per
    day remained stable in algae-dominated communities
    throughout most of the year but increased towards sum-
    mer (Table 1). Both community types showed a positive
    correlation between net DOC fluxes and increasing tem-
    perature (Fig. 3c), which was also reflected in positive
    Ea values (Table 2, Appendix S1: Fig. S3).
    DIN fluxes in both coral- and algae-dominated com-

    munities were lowest in spring and twofold higher during
    the rest of the year (Table 1). DIN fluctuations did not
    correspond to changes in seawater temperature (Fig. 4d)
    but showed a significant positive relationship (r = 0.57,
    P = 0.0017, n = 26 for corals; r = 0.39, P = 0.0385,
    n = 26 for algae) with increasing DIN concentrations of
    the ambient seawater (Appendix S1: Fig. S2).

    DISCUSSION

    Phase shifts from corals to algae may drastically change
    reef community functions

    The carbon and carbonate pathways are well
    described in the literature for undisturbed coral reef
    communities, both in direction and magnitude

    TABLE 1. Seasonal gross primary production (GPP), net community production (NCP), community respiration (CR), the ratio of
    GPP and CR (GPP/CR), net community calcification (NCC), net dissolved organic carbon (DOC) fluxes, and net dissolved
    inorganic nitrogen (DIN) fluxes of coral- and algae-dominated reef communities of the central Red Sea.

    GPP (mmol
    C⋅m⁻2⋅d⁻1)

    NCP (mmol
    C⋅m⁻2⋅d⁻1)

    CR (mmol
    C⋅m⁻2⋅d⁻1) GPP/CR

    NCC (mmol
    CaCO₃⋅m⁻2⋅d⁻1)

    DOC (mmol
    C⋅m⁻2⋅d⁻1)

    DIN (mmol
    N⋅m⁻2⋅d⁻1)

    Annual mean
    Coral 581 � 19 320 � 15 −261 � 17 2.4 � 0.1 131 � 10 14.3 � 1.8 1.6 � 0.2
    Algae 946 � 51 443 � 21 −503 � 33 2.0 � 0.2 10 � 3 4.0 � 0.9 1.7 � 0.2
    P <0.0001* <0.0001* <0.0001* <0.0001* <0.0001* <0.00001* 0.9351

    Spring
    Coral 638 � 21 353 � 12 −285 � 25 2.3 � 0.2 162 � 11 14.1 � 3.6 0.7 � 0.4
    Algae 974 � 41 439 � 22 −535 � 27 1.8 � 0.0 7 � 6 1.6 � 1.6 1.3 � 0.4
    ∣t∣ <0.0001* 0.0833 <0.0001* 0.0134* <0.0001* 0.0246* 0.9573

    Summer
    Coral 575 � 45 231 � 24 −344 � 26 1.7 � 0.1 61 � 11 18.9 � 3.9 2.2 � 0.2
    Algae 1,224 � 13 553 � 12 −671 � 9 1.8 � 0.0 13 � 3 7.1 � 1.9 1.9 � 0.4
    ∣t∣ <0.0001* <0.0001* <0.0001* 0.9310 0.0024 0.0432* 0.9979

    Fall
    Coral 447 � 44 292 � 25 −154 � 20 2.9 � 0.1 176 � 7 16.2 � 1.7 2.3 � 0.7
    Algae 1,053 � 94 470 � 68 −583 � 43 1.8 � 0.1 15 � 11 2.3 � 3.0 2.3 � 0.7
    ∣t∣ <0.0001* 0.0017* <0.0001* <.00001* <0.0001* 0.0111* 1.0000

    Winter
    Coral 597 � 20 390 � 10 −207 � 15 2.9 � 0.1 147 � 10 9.1 � 3.0 1.6 � 0.3
    Algae 586 � 30 324 � 17 −262 � 18 2.3 � 0.1 8 � 4 4.1 � 0.4 1.5 � 0.5
    ∣t∣ 1.0000 0.3351 0.6166 0.0005 <0.0001* 0.1388 1.0000

    Notes: Values represent averages from all incubations during the respective season � SE. Significant differences between treat-
    ments were assessed by linear mixed models (LMMs), and the differences between treatments and seasons by Tukey’s Honest Sig-
    nificant Difference (HSD; Appendix S1: Table S4) test. Asterisks highlight significant P values (P or |t| < 0.05).

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 7

    (Appendix S1: Table S5). Concomitant to previous find-
    ings, coral-dominated communities in our experiments
    displayed: (1) low net community production despite

    high gross productivity, implying that biomass accumu-
    lates slowly (Gattuso et al. 1998); (2) net DOC fluxes,
    where the accumulation of exudates outpaces its con-
    sumption (Nelson et al. 2013, Quinlan et al. 2018), rep-
    resenting an organic C loss that further reduces C
    available to support net benthic biomass accretion, and
    promotes an efficient transfer of C via benthic-pelagic
    coupling (Wild et al. 2004); and (3) high rates of net
    community calcification, consistent with the accretion of
    carbonate structures typical for tropical coral reefs (Gat-
    tuso et al. 1998, 1999b, Atkinson and Falter 2003). Yet,
    our seasonal comparative in situ approach revealed that
    functions of reef ecosystems change with shifts from a
    coral- to an algae-dominated benthic community.
    Specifically, algae-dominated communities displayed a

    higher organic C metabolism, along with residual net
    calcification. Significantly higher GPP and >40% ele-
    vated NCP in algae-dominated communities indicate a
    greater potential for autotrophic biomass accumulation
    per planar square meter of reef (Fong and Paul 2011,
    Kelly et al. 2017), despite a lower GPP/CR ratio
    (Table 1). The lower GPP/CR ratio in algae communi-
    ties compared to coral communities seems counterintu-
    itive considering the dominant organisms (algae as
    autotrophs compared to mixotrophic corals); however,
    algae-dominated reef structures host numerous

    N
    C

    C
    (
    m

    m
    o
    l C
    a
    C
    O
    m

    ·
    d

    )

    26 28 30 32

    (b)
    0
    80
    160

    240 ·d
    )

    26 28 30 32 32

    (c)
    0
    20

    40
    N

    C
    P
    (
    m
    m
    o
    l C
    m
    ·
    d
    )
    26 28 30 32

    (a) ·d
    )

    26 28 30

    (d)
    0
    2
    4

    Coral-dominated Algae-dominated

    200

    400

    600

    FIG 3. Relationship of temperature with (a) net community production (NCP), (b) net community calcification (NCC), (c) dis-
    solved organic carbon (DOC) fluxes, and (d) dissolved inorganic nitrogen (DIN) fluxes in coral- and algae-dominated reef commu-
    nities. Rates represent the net flux integrated over a 24-h day (assuming 12 h of light and 12 h of dark). Solid lines represent the
    linear regressions; shaded areas in transparent colors represent the 95% confidence intervals. [Color figure can be viewed at wileyon
    linelibrary.com]

    TABLE 2. Apparent activation energies (Ea, eV) for coral- and algae-dominated reef communities as the slope of the Arrhenius
    relationship between the natural logarithm of specific metabolic rates and the inverted temperature (1/kT).

    GPP NCP CR NCC DOC DIN

    Coral Algae Coral Algae Coral Algae Coral Algae Coral Algae Coral Algae

    Ea (eV) −0.09 0.75 −0.58 0.55 0.49 0.93 −0.82 0.65 0.73 0.19 0.12 0.13
    r2 0.03 0.71 0.54 0.59 0.26 0.66 0.30 0.10 0.28 0.01 0.02 0.02
    P 0.3861 <0.0001 <0.0001 <0.0001 0.0065 <0.0001 0.0023 0.1073 0.0038 0.5619 0.5126 0.4439

    Notes: GPP = gross primary production (mmol C�m−2�d−1); NCP = net community production (mmol C�m−2�d−1); CR = com-
    munity respiration (mmol C�m−2�d−1); NCC = net community calcification = (mmol CaCO3�m−2�d−1); DOC = dissolved organic
    carbon fluxes (mmol C�m−2�d−1); DIN = dissolved inorganic nitrogen fluxes (mmol N�m−2�d−1); r2 = square of correlation coeffi-
    cient.

    –5
    0
    5
    10

    –30 0 30 60

    NCP (mmol C m ·h )

    N
    C
    C
    (
    m
    m
    o
    l C
    a
    C
    O
    m
    ·h
    )

    – 5

    0
    5
    10

    3 4 5 6

    N
    C
    C
    (
    m
    m
    o
    l C
    a
    C
    O
    m

    ·h
    )

    arag

    (a) (b)

    Coral-dominated Algae-dominated

    FIG 4. Relationship between (a) net community calcifica-
    tion (NCC) and aragonite saturation state (Ωarag), and (b) NCC
    and net community production (NCP) in coral- and algae-dom-
    inated reef communities. Closed circles indicate measurements
    from light incubations, and open circles indicate dark incuba-
    tions. Solid lines represent the linear regressions; shaded areas
    in transparent colors represent the 95% confidence intervals.
    [Color figure can be viewed at wileyonlinelibrary.com]

    Article e03226; page 8 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    www.wileyonlinelibrary.com

    www.wileyonlinelibrary.com

    www.wileyonlinelibrary.com

    heterotrophs that rely on the high algal biomass produc-
    tion, fueling community-wide respiration by direct her-
    bivory (Klumpp and McKinnon 1989, Russ 2003) and
    indirect detritivory by invertebrates (Kramer et al.
    2013). Likewise, sponges and other filter feeders are
    commonly associated with degraded reef habitats (Abele
    and Patton 1976), feeding on DOC or algal debris (Rix
    et al. 2017, 2018). In addition, heterotrophic bacteria
    within algal communities remineralize labile DOC
    released by algae (Nelson et al. 2011, Haas et al. 2013).
    Thereby, the consumption of algal-derived C (i.e., the C
    retention within the system) can shorten the average
    trophic path length, and reduce the average trophic level
    of the second-order consumers (reviewed in Johnson
    et al. 1995). In support of these trophic interactions, we
    observed an apparent consumption of DOC in the dark
    in algae-dominated communities, limiting the net DOC
    flux integrated over 24 h (i.e., release in the light bal-
    anced by consumption in the dark). Our results indicate
    that algae-associated organisms readily remineralize
    DOC, concurring with demonstrated DOC depletion in
    algae-dominated shallow reefs elsewhere (Nelson et al.
    2011, Haas et al. 2016).
    Along with alterations of the organic C cycle, calcifi-

    cation was reduced within algae-dominated communi-
    ties. The slope of the relationship between NCP and
    NCC is commonly used as an indicator of reef health,
    indicative for the relative proportion of calcifying to
    noncalcifying organisms in a benthic community (Alb-
    right et al. 2013, Takeshita et al. 2016). We observed a
    slope of 0.18 in coral- and 0.03 in algae-dominated com-
    munities, and the slope averages 0.22 based on 52 reefs
    around the world (Gattuso et al. 1999a). The significant
    difference highlights the shift from calcifying corals to
    noncalcifying organisms and a decoupling of the organic
    carbon (production vs. respiration) and carbonate (calci-
    fication vs. dissolution) cycles in algae-dominated com-
    munities (McMahon et al. 2019).
    Unraveling nitrogen pathways in coral reefs is crucial

    to understand how high productivity is supported
    despite low ambient nutrient concentrations (D’Elia and
    Wiebe 1990, Szmant 2002, Atkinson and Falter 2003).
    Both coral- and algae-dominated communities were net
    sources of DIN over the study period with no significant
    differences between community types. The flux rates
    generally fell within the published range of in situ mea-
    surements (−4 to 5 mmol N as NOx�m−2�d−1; reviewed
    in Atkinson and Falter 2003). They are in stark contrast,
    however, with the expectation that net autotrophic com-
    munities would act as sinks for dissolved inorganic nutri-
    ents, as generally measured in single organism
    incubations (e.g., Den Haan et al. 2016). Although a the-
    oretical N requirement of 34–83 mmol N�m−2�d−1 to
    support NCP can be expected (stoichiometric calcula-
    tions with NCP ranging from 230 to 550 mmol C�m−2-
    �d−1; assuming a C/N ratio of 6.6; see Redfield 1958), the
    effects of assimilation were likely masked by concurrent
    community-wide processes that produce DIN (Gruber

    et al. 2019). For example, cavities within Red Sea reefs
    can be considerable sources of DIN (>20 mmol N�m−2-
    �d−1) as sponges and other filter feeders utilize dissolved
    organic matter (Richter et al. 2001). Also, microbial
    communities can consume and transform organic N
    compounds (Yahel et al. 2003, Moulton et al. 2016, Pfis-
    ter and Altabet 2019), potentially increasing the commu-
    nity-wide DIN release into the environment. Other
    pathways, such as N2 fixation (Cardini et al. 2016) or
    heterotrophic feeding on particulates (Ribes et al. 2003,
    Houlbrèque and Ferrier-Pagès 2009) are additional N
    sources that potentially limit/mask the N uptake from
    DIN.
    Overall, algae-dominated communities displayed a

    higher potential of biomass accumulation or export (i.e.,
    high NCP), associated with a higher total amount of C
    available (i.e., high GPP) to the ecosystem. The high
    NCP of algae communities facilitates rapid lateral and
    vegetative overgrowth of bare substrates (Diaz-Pulido
    and Garzón-Ferreira 2002, Roth et al. 2018). At the
    same time, limited reef accretion (i.e., low NCC) within
    algal habitats may compromise the topographic com-
    plexity of phase-shifted reef communities (Wild et al.
    2011), limit the recruitment of corals (Harrington et al.
    2004, Roth et al. 2017, 2018), and increase reef erosion
    (Adey 1978).

    High temperatures during summer amplify functional
    differences between coral- and algae-dominated

    communities

    Our data highlight that functions related to the carbon
    and carbonate cycles of coral- and algae-dominated
    communities are strongly but inversely affected by tem-
    perature (summarized in Fig. 5), with implications for
    their response to warming.
    The results from activation energies show a higher

    sensitivity to thermal stress of coral-dominated com-
    pared to algae-dominated communities during summer
    (Table 2, Appendix S1: Fig. S3). Particularly, apparent
    activation energies for NCP and NCC of coral-domi-
    nated communities (Ea = −0.58 and −0.82 eV; respec-
    tively) were in the falling phase of the performance
    curves, and thus, past the optimum temperature to peak
    rates. Thereby, the community metabolism of corals is
    pushed toward carbon losses because of high CR relative
    to GPP in summer. Likewise, the activation energy of
    CR was positive, as respiration continued to increase
    with temperature through the seasonal thermal range.
    Both the lower ratio of GPP to CR (GPP/CR) and
    reduced NCP integrated over diel cycles indicate that
    coral-dominated communities shifted toward a more
    heterotrophic state with increasing temperature. It is
    thus apparent that summer temperatures exceeded the
    metabolic optima for coral-dominated communities,
    which was previously suggested for individuals of Pocil-
    lopora verrucosa (Sawall et al. 2015, Roik et al. 2016,
    Anton et al. 2020) and Stylophora pistillata (Anton et al.

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 9

    2020) in the central Red Sea. This contrasts with many
    reef locations worldwide, where primary production
    maxima are typically observed during the warmest
    months of the year (e.g., Scheufen et al. 2017). At the
    same time, coral-dominated communities displayed
    enhanced rates of net DOC fluxes with warming (Ea =
    0.73 eV), which can be attributed to an increased
    release of cellular matter and/or mucoid exudates during
    thermal stress in corals (Niggl et al. 2009, Scheufen et al.
    2017). Although mucus released during higher tempera-
    tures may help to protect corals against pathogens (Glasl
    et al. 2016) or high UV radiation (Gleason and Welling-
    ton 1993), it poses an increased loss of organic C that
    can be used for community growth and/or export at con-
    stant GPP rates (Fig. 5). Along with this trend, NCC

    dropped by >50% from the annual mean during sum-
    mer, with most of this decline realized as waters warmed
    from 30 to 32°C. Overall, decreased NCC strongly corre-
    lated with decreased NCP as temperatures increased (as
    revealed by high negative apparent activation energies of
    NCC, Ea = −0.82 eV), with a temperature threshold at
    around 30.5°C (Appendix S1: Fig. S3). This thermal
    threshold is near the reported thermal optimum of Pocil-
    lopora verrucosa and Stylophora pistillata for gross pri-
    mary production (29.9 and 31.9°C, respectively; Anton
    et al. 2020) in the central Red Sea, indicating a strong
    thermal sensitivity of coral-dominated communities
    soon after the thermal optimum is exceeded. Accord-
    ingly, in the temperature range above the optimum, the
    rate of calcification decreased despite increased Ωarag in

    FIG 5. In situ community metabolism of natural coral- and algae-dominated reef communities in the central Red Sea, Saudi
    Arabia. Schematic was derived from all data available in the given lower (blue) and upper (red) temperature ranges. Organic carbon
    pathways refer to photosynthesis, respiration, and dissolved organic carbon (DOC) fluxes, and the inorganic carbon pathway refers
    to the formation and dissolution of calcium carbonate. The thicknesses of the bars scale with the actual flux measurements from
    in situ incubations. [Color figure can be viewed at wileyonlinelibrary.com]

    Article e03226; page 10 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    www.wileyonlinelibrary.com

    summer (Fig. 4a; Silverman et al. 2007). Physiological
    factors can also strongly affect the biomineralization
    process. As calcification mainly depends on the photo-
    synthetic efficiency of the endosymbionts within corals
    (Gattuso et al. 1999a, Allemand et al. 2004), a lower
    NCC may occur for thermally stressed corals, limiting
    reef accretion and stabilization (Jokiel and Coles 1977,
    De’ath et al. 2009).
    In contrast, positive activation energies for algae-

    dominated communities (Ea = 0.93, 0.75, and 0.55 eV
    for CR, GPP, and NCP; respectively) indicate that these
    communities benefit from higher temperatures and that
    thermal optima were not reached in summer. In fact,
    some macroalgae species (Halimeda tuna) from the cen-
    tral Red Sea have a reported thermal optimum (31.7°C)
    for gross primary production (Anton et al. 2020) that is
    close to the maximum summer temperature recorded
    during our incubations (32.5°C). The increases in GPP
    and CR along a thermal gradient in algae-dominated
    communities highlight a higher turnover of organic C.
    However, increases in C fixation outweighed increases in
    respiratory C consumption, resulting in higher NCP in
    the summer. In contrast to previous reports (e.g., Barron
    et al. 2014), DOC fluxes in algae-dominated communi-
    ties showed only a weak temperature dependence. How-
    ever, although the overall net DOC fluxes remained
    relatively stable, differences in net production during the
    light and net consumption in the dark amplified with
    temperature (Fig. 3), limiting losses of organic C
    through this process. As a result, more organic C was
    retained within algae-dominated communities in sum-
    mer, supporting biomass accumulation and export in the
    community (Fig. 5).

    Implications for reef ecosystem functioning under global
    change

    Theoretical studies have provided a sound under-
    standing of the relationship between biodiversity loss
    and ecosystem functions in tropical coral reefs (reviewed
    in Hughes et al. 2017). However, considerable knowledge
    gaps remain, in particular, on how metabolic and bio-
    geochemical processes differ between coral- and algae-
    dominated communities, and how these respond to sea-
    sonal fluctuations in environmental conditions. As algal
    cover is expected to increase in coral reefs, our long-term
    in situ experiments reveal how these novel communities
    in general, and how thermal stress in particular, may
    alter pivotal ecological functions of future reefs. Our
    data show that fundamental metabolic and biogeochem-
    ical characteristics of coral-dominated communities are
    disturbed by shifts from coral to algal dominance and
    may, thereby, compromise the future stability and resili-
    ence of coral reef biota. These responses may be further
    compounded by differential thermal responses between
    coral and algae species (e.g., Anton et al. 2020).
    The sensitivity of corals and their symbionts to rising

    temperatures has been documented extensively (Hoegh-

    Guldberg 1999). Thermal anomalies exceeding 1–2°C
    above the mean summer maximum temperature can
    compromise the symbiosis (e.g., Weeks et al. 2008), lead-
    ing to coral bleaching and reduced coral survival (Baird
    and Marshall 2002, Baker et al. 2008). However, our
    study did not record temperatures exceeding the local
    mean summer maxima reported for the region (see
    Fig. 1c; Chaidez et al. 2017) and, likewise, no apparent
    signs of coral bleaching were observed. Nevertheless,
    growth of Pocillopora verrucosa and Stylophora pistillata
    (Anton et al. 2020) and calcification of Pocillopora verru-
    cosa (Roik et al. 2018) are already reduced under current
    summer conditions in the Red Sea, as also highlighted
    by the present study. For the warmer part of the year,
    algae-dominated communities have, thus, a metabolic
    advantage over coral-dominated communities because
    they maintain high NCP. Importantly, the consequences
    of global warming may manifest not only in terms of
    higher-than-normal temperatures but also in a longer-
    than-normal duration of the seasonal peak temperatures
    (Fitt et al. 2001). Under such conditions, if coral mortal-
    ity events occur, algae may quickly spread and shift coral
    ecosystems more rapidly towards systems that are domi-
    nated by algae (McManus et al. 2019), as already
    reported on some reefs in the southern Red Sea follow-
    ing the coral bleaching event in 2015 (Anton et al. 2020).
    The frequency and intensity of climate-driven stress

    events on coral reefs will inevitably aggravate in the near
    future. Our results suggest that the anticipated increase
    in the spatial footprint of algae-dominated communities
    would exacerbate the magnitude of the functional
    changes described here. Ocean warming likely enhances
    the competitive advantage of algae- over coral-domi-
    nated communities (Anton et al. 2020), thus promoting
    a positive feedback loop of reef degradation. Similar
    effects of warming are likely to be operational in other
    temperature-sensitive and calcifying communities. As
    these organisms are central to the formation of reef
    ecosystems, critical changes in the biodiversity and func-
    tioning may be witnessed. Thus, appropriate manage-
    ment practices designed to limit the proliferation of
    algae are needed for maintaining reefs dominated by
    corals and the important ecosystem services they sup-
    port.

    ACKNOWLEDGMENTS

    We are grateful to the personnel from the King Abdullah
    University of Science and Technology (KAUST) Coastal and
    Marine Resources Core (CMOR) Laboratory for logistical sup-
    port. The authors would also like to acknowledge Rodrigo Vil-
    lalobos and João Cúrdia, who helped during fieldwork.
    Figure 5 was produced by Xavier Pita, scientific illustrator at
    KAUST. We would like to thank the two anonymous reviewers
    and the editor for their helpful suggestions and comments. The
    research was supported by KAUST baseline funding to BHJ
    and by grant Wi 2677/9-1 from the German Research Founda-
    tion (DFG) to CW. Author contributions: FR, CW, and SC
    conceptualized and designed research. FR, LS, MLC, and VS
    performed research. FR, NR, VS, AA, LS, BK, and MLC

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 11

    analyzed data. CMD, XAGM, CRV, and BHJ contributed to
    research materials, logistics and to interpreting data. FR wrote
    original draft of the manuscript with support by CW. All
    authors read and approved the final manuscript.

    LITERATURE CITED

    Abele, L. G., and W. K. Patton. 1976. The size of coral heads
    and the community biology of associated decapod crus-
    taceans. Journal of Biogeography 3:35.

    Adey, W. H. 1978. Coral reef morphogensis: a multidimensional
    model. Science 202(4370): 831–837.

    Albright, R., C. Langdon, and K. R. N. Anthony. 2013.
    Dynamics of seawater carbonate chemistry, production, and
    calcification of a coral reef flat, Central Great Barrier Reef.
    Biogeosciences Discussions 10:7641–7676.

    Allemand, D., C. Ferrier-Pagès, P. Furla, F. Houlbrèque, S.
    Puverel, S. Reynaud, É. Tambutté, S. Tambutté, and D. Zoc-
    cola. 2004. Biomineralisation in reef-building corals: from
    molecular mechanisms to environmental control. Comptes
    Rendus Palevol 3:453–467.

    Anthony, K. R. N., D. I. Kline, G. Diaz-Pulido, S. Dove, and O.
    Hoegh-Guldberg. 2008. Ocean acidification causes bleaching
    and productivity loss in coral reef builders. Proceedings of the
    National Academy of Sciences of the United States of Amer-
    ica 105:17442–17446.

    Anton, A., J. L. Randle, F. C. Garcia, S. Rossbach, J. I. Ellis, M.
    Weinzierl, and C. M. Duarte. 2020. Differential thermal toler-
    ance between algae and corals may trigger the proliferation of
    algae in coral reefs. Global Change Biology 26:4316–4327.

    Ateweberhan, M., J. H. Bruggemann, and A. M. Breeman.
    2006. Effects of extreme seasonality on community structure
    and functional group dynamics of coral reef algae in the
    southern Red Sea (Eritrea). Coral Reefs 25:391–406.

    Atkinson, M. J., and J. L. Falter. 2003. Coral reefs. in K. Black
    and G. Shimmieldeditors. Pages 40–64 Biogeochemistry of
    Marine Systems, Boca Raton, Florida: CRC Press.

    Baird, A. H., and P. A. Marshall. 2002. Mortality, growth and
    reproduction in scleractinian corals following bleaching on
    the Great Barrier Reef. Marine Ecology Progress Series
    237:133–141.

    Baird, D., S. Brown, L. Lagadic, M. Liess, L. Maltby, M. Mor-
    eira-Santos, R. Schulz, and G. Scott. 2007. In situ–based
    effects measures: determining the ecological relevance of mea-
    sured responses. Integrated Environmental Assessment and
    Management 3:259–267.

    Baker, A. C., P. W. Glynn, and B. Riegl. 2008. Climate change
    and coral reef bleaching: an ecological assessment of long-
    term impacts, recovery trends and future outlook. Estuarine,
    Coastal and Shelf Science 80:435–471.

    Balvanera, P., A. B. Pfisterer, N. Buchmann, J. S. He, T. Naka-
    shizuka, D. Raffaelli, and B. Schmid. 2006. Quantifying the
    evidence for biodiversity effects on ecosystem functioning
    and services. Ecology Letters 9:1146–1156.

    Barron, C., E. T. Apostolaki, and C. M. Duarte. 2014. Dis-
    solved organic carbon fluxes by seagrass meadows and
    macroalgal beds. Frontiers in Marine Science 1:42.

    Bellwood, D. R., T. P. Hughes, C. Folke, and M. Nyström.
    2004. Confronting the coral reef crisis. Nature 429:827–833.

    Bellworthy, J., and M. Fine. 2018. The Red Sea Simulator: A
    high-precision climate change mesocosm with automated
    monitoring for the long-term study of coral reef organisms.
    Limnology and Oceanography: Methods 16:367–375.

    Bender, D., G. Diaz-Pulido, and S. Dove. 2014. Warming and
    acidification promote cyanobacterial dominance in turf algal
    assemblages. Marine Ecology Progress Series 517:271–284.

    Bernstein, W. N., K. A. Hughen, C. Langdon, D. C. McCorkle,
    and S. J. Lentz. 2016. Environmental controls on daytime net
    community calcification on a Red Sea reef flat. Coral Reefs
    35:697–711.

    Brondizio, E., J. Settele, S. Diaz, and T. Ngo. 2019. Global
    assessment report on biodiversity and ecosystem services of
    the Intergovernmental Science—Policy Platform on Biodiver-
    sity and Ecosystem Services. IPBES Secretariat, Bonn, Ger-
    many.

    Bythell, J. C., Z. M. Hillis-Starr, and C. S. Rogers. 2000. Local
    variability but landscape stability in coral reef communities
    following repeated hurricane impacts. Marine Ecology Pro-
    gress Series 204:93–100.

    Canavero, A., M. Arim, F. Pérez, F. M. Jaksic, and P. A. Mar-
    quet. 2018. A metabolic view of amphibian local community
    structure: the role of activation energy. Ecography
    41:388–400.

    Cardini, U., V. N. Bednarz, N. van Hoytema, A. Rovere, M. S.
    Naumann, M. M. D. Al-Rshaidat, and C. Wild. 2016. Budget
    of primary production and dinitrogen fixation in a highly sea-
    sonal Red Sea coral reef. Ecosystems 19:771–785.

    Chaidez, V., D. Dreano, S. Agusti, C. M. Duarte, and I. Hoteit.
    2017. Decadal trends in Red Sea maximum surface tempera-
    ture. Scientific Reports 7:1–8.

    D’Elia, C. F., and W. J. Wiebe. 1990. Biogeochemical nutrient
    cycles in coral reef ecosystems. Pages 49–74 in Z. Dubinsky,
    editor. Ecosystems of the world 25: coral reefs. Elsevier, New
    York, New York, USA.

    Damgaard, C. 2019. A critique of the space-for-time substitu-
    tion practice in community ecology. Trends in Ecology and
    Evolution 34:416–421.

    de Jongh, F., and F. C. Van Duyl. 2004. The impact of suspen-
    sion feeders in cryptic habitats on the trophodynamics of
    coral reefs. Thesis. http://www.science.uva.nl/ZMA/Inverteb
    rates/Coel/scirep/Cryptic_Reef_Habitats

    De’ath, G., J. M. Lough, and K. E. Fabricius. 2009. Declining
    coral calcification on the Great Barrier Reef. Science
    323:116–119.

    Den Haan, J., et al. 2016. Nitrogen and phosphorus uptake
    rates of different species from a coral reef community after a
    nutrient pulse. Scientific Reports 6:1–13.

    Diaz-Pulido, G., and J. Garzón-Ferreira. 2002. Seasonality in
    algal assemblages on upwelling-influenced coral reefs in the
    Colombian Caribbean. Botanica Marina 45:284–292.

    Done, T. J. 1992. Phase shifts in coral reef communities and
    their ecological significance. in V. Jaccarini and E. Martens
    editors. Pages 121–132 Book: The ecology of mangrove and
    related ecosystems, Dordrecht, Netherlands: Springer Inter-
    national Publishing.

    Done, T. J., P. K. Dayton, A. E. Dayton, and R. Steger. 1991.
    Regional and local variability in recovery of shallow coral
    communities: Moorea, French Polynesia and central Great
    Barrier Reef. Coral Reefs 9:183–192.

    Edmunds, P. J. 2002. Long-term dynamics of coral reefs in St.
    John, US Virgin Islands. Coral Reefs 21:357–367.

    Edmunds, P. J., S. S. Doo, and R. C. Carpenter. 2020. Year-long
    effects of high pCO2 on the community structure of a tropical
    fore reef assembled in outdoor flumes. ICES Journal of Mar-
    ine Science 77:1055–1065.

    Fitt, W., B. Brown, M. Warner, and R. Dunne. 2001. Coral
    bleaching: interpretation of thermal tolerance limits and ther-
    mal thresholds in tropical corals. Coral Reefs 20:51–65.

    Follstad Shah, J. J., et al. 2017. Global synthesis of the tempera-
    ture sensitivity of leaf litter breakdown in streams and rivers.
    Global Change Biology 23:3064–3075.

    Fong, P., and V. J. Paul. 2011. Coral reef algae. in Z. Dubinsky
    and N Stambler editors. Pages 241–272 Book: Coral reefs: an

    Article e03226; page 12 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    http://www.science.uva.nl/ZMA/Invertebrates/Coel/scirep/Cryptic_Reef_Habitats

    http://www.science.uva.nl/ZMA/Invertebrates/Coel/scirep/Cryptic_Reef_Habitats

    ecosystem in transition, Dordrecht, Netherlands: . Springer
    International Publishing.

    Frölicher, T. L., E. M. Fischer, and N. Gruber. 2018. Marine
    heatwaves under global warming. Nature 560:360–364.

    Garcı́a, F. C., E. Bestion, R. Warfield, and G. Yvon-Durocher.
    2018. Changes in temperature alter the relationship between
    biodiversity and ecosystem functioning. Proceedings of the
    National Academy of Sciences of the United States of Amer-
    ica 115:10989–10994.

    Gattuso, J. P., D. Allemand, and M. Frankignoulle. 1999a. Pho-
    tosynthesis and calcification at cellular, organismal and com-
    munity levels in coral reefs: a review on interactions and
    control by carbonate chemistry. American Zoologist
    39:160–183.

    Gattuso, J. P., M. Frankignoulle, and S. V. Smith. 1999b. Mea-
    surement of community metabolism and significance in the
    coral reef CO2 source-sink debate. Proceedings of the
    National Academy of Sciences of the United States of Amer-
    ica 96:13017–13022.

    Gattuso, J. P., M. Frankignoulle, and R. Wollast. 1998. Carbon
    and carbonate metabolism in coastal aquatic ecosystems.
    Annual Review of Ecology and Systematics 29:405–434.

    Glasl, B., G. J. Herndl, and P. R. Frade. 2016. The microbiome
    of coral surface mucus has a key role in mediating holobiont
    health and survival upon disturbance. ISME Journal
    10:2280–2292.

    Gleason, D. F., and G. M. Wellington. 1993. Ultraviolet radia-
    tion and coral bleaching. Nature 365:836–838.

    Graham, N. A. J., S. Jennings, M. A. MacNeil, D. Mouillot,
    and S. K. Wilson. 2015. Predicting climate-driven regime
    shifts versus rebound potential in coral reefs. Nature
    518:1–17.

    Gruber, R. K., R. J. Lowe, and J. L. Falter. 2019. Tidal and sea-
    sonal forcing of dissolved nutrient fluxes in reef communities.
    Biogeosciences 16 1921–1935. https://doi.org/10.5194/bg-16-
    1921-2019.

    Haas, A. F., et al. 2016. Global microbialization of coral reefs.
    Nature Microbiology 1: 16042.

    Haas, A. F., C. E. Nelson, F. Rohwer, L. Wegley-Kelly, S. D.
    Quistad, C. A. Carlson, J. J. Leichter, M. Hatay, and J. E.
    Smith. 2013. Influence of coral and algal exudates on micro-
    bially mediated reef metabolism. PeerJ 1:e108.

    Harrington, L., K. Fabricius, G. D’eath, A. Negri, L. I. H.
    Arrington, K. A. F. Abricius, and G. L. D. E. Ath. 2004.
    Recognition and selection of settlement substrata determine
    post-settlement survival in corals. Ecology 85:3428–3437.

    Hempson, T. N., N. A. J. Graham, M. A. Macneil, A. S. Hoey,
    and S. K. Wilson. 2018. Ecosystem regime shifts disrupt
    trophic structure. Ecological Applications 28:191–200.

    Hoegh-Guldberg, O. 1999. Climate change, coral bleaching and
    the future of the world’s coral reefs. Marine and Freshwater
    Research 50:839–866.

    Hooper, D. U., E. C. Adair, B. J. Cardinale, J. E. K. Byrnes, B.
    A. Hungate, K. L. Matulich, A. Gonzalez, J. E. Duffy, L.
    Gamfeldt, and M. I. Connor. 2012. A global synthesis reveals
    biodiversity loss as a major driver of ecosystem change. Nat-
    ure 486:105–108.

    Houlbrèque, F., and C. Ferrier-Pagès. 2009. Heterotrophy in
    tropical scleractinian corals. Biological Reviews 84:1–17.

    Hughes, T. P., et al. 2017. Coral reefs in the Anthropocene. Nat-
    ure 546:82–90.

    Hughes, T. P., et al. 2019. Ecological memory modifies the
    cumulative impact of recurrent climate extremes. Nature Cli-
    mate Change 9:40–43.

    Hughes, T. P., M. J. Rodrigues, D. R. Bellwood, D. Ceccarelli,
    O. Hoegh-Guldberg, L. McCook, N. Moltschaniwskyj, M. S.
    Pratchett, R. S. Steneck, and B. Willis. 2007. Phase shifts,

    herbivory, and the resilience of coral reefs to climate change.
    Current Biology 17:360–365.

    Jessen, C., J. F. Villa Lizcano, T. Bayer, C. Roder, M. Aranda,
    C. Wild, and C. R. Voolstra. 2013. In-situ effects of eutrophi-
    cation and overfishing on physiology and bacterial diversity
    of the red sea coral Acropora hemprichii. PLoS One 8:e62091.

    Johnson, C., D. Klumpp, J. Field, and R. Bradbury. 1995. Car-
    bon flux on coral reefs: effects of large shifts in community
    structure. Marine Ecology Progress Series 126:123–143.

    Jokiel, P. L., and S. L. Coles. 1977. Effects of temperature on
    the mortality and growth of Hawaiian reef corals. Marine
    Biology 43:201–208.

    Kelly, E. L. A., Y. Eynaud, I. D. Williams, R. T. Sparks, M. L. Dai-
    ler, S. A. Sandin, and J. E. Smith. 2017. A budget of algal pro-
    duction and consumption by herbivorous fish in an herbivore
    fisheries management area, Maui. Hawaii. Ecosphere 8:e01899.

    Klumpp, W., and A. McKinnon. 1989. Community structure,
    biomass and productivity of epilithic algal communities on
    the Great Barrier Reef: dynamics at different spatial scales.
    Marine Ecology Progress Series 86:77–89.

    Koch, M., G. Bowes, C. Ross, and X. H. Zhang. 2013. Climate
    change and ocean acidification effects on seagrasses and mar-
    ine macroalgae. Global Change Biology 19:103–132.

    Kramer, M. J., O. Bellwood, and D. R. Bellwood. 2013. The
    trophic importance of algal turfs for coral reef fishes: the
    crustacean link. Coral Reefs 32:575–583.

    Langdon, C., W. S. Broecker, D. E. Hammond, E. Glenn, K.
    Fitzsimmons, S. G. Nelson, T.-H. Peng, I. Hajdas, and G.
    Bonani. 2003. Effect of elevated CO2 on the community
    metabolism of an experimental coral reef. Global Biogeo-
    chemical Cycles 17:1011.

    Lavigne, H., and J. P. Gattuso. 2013. Package “seacarb”: seawa-
    ter carbonate chemistry with R, v. 2.4. 8. http://www2.uaem.
    mx/r-mirror/web/packages/seacarb/seacarb

    Legendre, P., and L. Legendre. 2012. Numerical ecology. Third
    edition. Elsevier, Amsterdam, The Netherlands.

    Lirman, D., and P. Biber. 2000. Seasonal dynamics of macroal-
    gal communities of the northern Florida reef tract. Botanica
    Marina 43:305–314.

    Lough, J. M., K. D. Anderson, and T. P. Hughes. 2018. Increas-
    ing thermal stress for tropical coral reefs: 1871–2017. Scien-
    tific Reports 8:6079.

    Maldonado, M., M. Ribes, and F. C. van Duyl. 2012. Nutrient
    fluxes through sponges. Biology, budgets, and ecological
    implications. Advances in Marine Biology 62:113–182.

    McMahon, A., I. R. Santos, T. Cyronak, and B. D. Eyre. 2013.
    Hysteresis between coral reef calcification and the seawater
    aragonite saturation state. Geophysical Research Letters
    40:4675–4679.

    McMahon, A., I. R. Santos, K. G. Schulz, A. Scott, J. Silver-
    man, K. L. Davis, and D. T. Maher. 2019. Coral reef calcifica-
    tion and production after the 2016 bleaching event at Lizard
    Island, Great Barrier reef. Journal of Geophysical Research
    Oceans 124:4003–4016.

    McManus, L. C., V. V. Vasconcelos, S. A. Levin, D. M. Thomp-
    son, J. A. Kleypas, F. S. Castruccio, E. N. Curchitser, and J.
    R. Watson. 2019. Extreme temperature events will drive coral
    decline in the Coral Triangle. Global Change Biology
    26:2120–2133.

    Middleton, B., and J. Grace. 2004. Biodiversity and ecosystem
    functioning: synthesis and perspectives. Restoration Ecology
    12:611–612.

    Moberg, F. F., and C. Folke. 1999. Ecological goods and ser-
    vices of coral reef ecosystems. Ecological Economics
    29:215–233.

    Morán, X. A. G., J. M. Gasol, M. C. Pernice, J. Mangot, R.
    Massana, E. Lara, D. Vaqué, and C. M. Duarte. 2017.

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 13

    https://doi.org/10.5194/bg-16-1921-2019

    https://doi.org/10.5194/bg-16-1921-2019

    http://www2.uaem.mx/r-mirror/web/packages/seacarb/seacarb

    http://www2.uaem.mx/r-mirror/web/packages/seacarb/seacarb

    Temperature regulation of marine heterotrophic prokaryotes
    increases latitudinally as a breach between bottom-up and
    top-down controls. Global Change Biology 23:3956–3964.

    Moulton, O. M., M. A. Altabet, J. M. Beman, L. A. Deegan, J.
    Lloret, M. K. Lyons, J. A. Nelson, and C. A. Pfister. 2016.
    Microbial associations with macrobiota in coastal ecosys-
    tems: patterns and implications for nitrogen cycling. Frontiers
    in Ecology and the Environment 14:200–208.

    Naeem, S., J. E. Duffy, and E. Zavaleta. 2012. The functions of
    biological diversity in an age of extinction. Science
    336:1401–1406.

    Nelson, C. E., A. L. Alldredge, E. A. McCliment, L. A.
    Amaral-Zettler, and C. A. Carlson. 2011. Depleted dissolved
    organic carbon and distinct bacterial communities in the
    water column of a rapid-flushing coral reef ecosystem. ISME
    Journal 5:1374–1387.

    Nelson, C. E., S. J. Goldberg, L. Wegley Kelly, A. F. Haas, J. E.
    Smith, F. Rohwer, and C. A. Carlson. 2013. Coral and
    macroalgal exudates vary in neutral sugar composition and
    differentially enrich reef bacterioplankton lineages. ISME
    Journal 7:962–979.

    Niggl, W., M. Glas, C. Laforsch, Mayr, C., and Wild, C. 2009.
    First evidence of coral bleaching stimulating organic matter
    release by reef corals. Pages 905–910inB. Riegl, editor. Pro-
    ceedings of the 11th International Coral Reef Symposium.
    https://www.researchgate.net/profile/Martin_Gutbrod2/publi
    cation/257233311_First_evidence_of_coral_bleaching_stimu
    lating_organic_matter_release_by_reef_corals/links/56938980
    08aec14fa55e9224/First-evidence-of-coral-bleaching-stimula
    ting-organic-matter-release-by-reef-corals .

    Oliver, E. C. J., et al. 2018. Longer and more frequent marine
    heatwaves over the past century. Nature Communications
    9:1–12.

    Padfield, D., C. Lowe, A. Buckling, R. Ffrench-Constant, S.
    Jennings, F. Shelley, J. S. Ólafsson, and G. Yvon-Durocher.
    2017. Metabolic compensation constrains the temperature
    dependence of gross primary production. Ecology Letters
    20:1250–1260.

    Pfister, C. A., and M. A. Altabet. 2019. Enhanced microbial
    nitrogen transformations in association with macrobiota
    from the rocky intertidal. Biogeosciences 16:193–206.

    Pierrot, D., E. Lewis, and D. W. R. Wallace. 2006. MS Excel
    program developed for CO2 system calculations. Oak Ridge,
    Tennessee, USA. https://www.ncei.noaa.gov/access/ocean-ca
    rbon-data-system/oceans/CO2SYS/co2rprt.html.

    Piggot, A. M., B. W. Fouke, M. Sivaguru, R. A. Sanford, and
    H. R. Gaskins. 2009. Change in zooxanthellae and mucocyte
    tissue density as an adaptive response to environmental stress
    by the coral, Montastraea annularis. Marine Biology
    156:2379–2389.

    Quinlan, Z. A., K. Remple, M. D. Fox, N. J. Silbiger, T. A. Oli-
    ver, H. M. Putnam, L. Wegley Kelly, C. A. Carlson, M. J.
    Donahue, and C. E. Nelson. 2018. Fluorescent organic exu-
    dates of corals and algae in tropical reefs are compositionally
    distinct and increase with nutrient enrichment. Limnology
    and Oceanography Letters 3:331–340.

    Redfield, A. C. 1958. The biological control of chemical factors
    in the environment. American Scientist 46:230A–A221.

    Reynaud, S., N. Leclercq, S. Romaine-Lioud, C. Ferrier-Pagès,
    J. Jaubert, and J. P. Gattuso. 2003. Interacting effects of CO2
    partial pressure and temperature on photosynthesis and calci-
    fication in a scleractinian coral. Global Change Biology
    9:1660–1668.

    Ribes, M., R. Coma, M. Atkinson, and R. Kinzie. 2003. Parti-
    cle removal by coral reef communities: picoplankton is a
    major source of nitrogen. Marine Ecology Progress Series
    257:13–23.

    Richter, C., and M. Wunsch. 1999. Cavity-dwelling suspension
    feeders in coral reefs—A new link in reef trophodynamics.
    Marine Ecology Progress Series 188:105–116.

    Richter, C., M. Wunsch, M. Rasheed, I. Kötter, and M. I. Bad-
    ran. 2001. Endoscopic exploration of Red Sea coral reefs
    reveals dense populations of cavity-dwelling sponges. Nature
    413:726–730.

    Rix, L., V. N. Bednarz, U. Cardini, N. Van Hoytema, F. A. Al-
    Horani, C. Wild, and M. S. Naumann. 2015. Seasonality in
    dinitrogen fixation and primary productivity by coral reef
    framework substrates from the northern Red Sea. Marine
    Ecology Progress Series 533:79–92.

    Rix, L., J. M. De Goeij, D. Van Oevelen, U. Struck, F. A. Al-
    Horani, C. Wild, and M. S. Naumann. 2018. Reef sponges
    facilitate the transfer of coral-derived organic matter to their
    associated fauna via the sponge loop. Marine Ecology Pro-
    gress Series 589:85–96.

    Rix, L., et al. 2017. Differential recycling of coral and algal dis-
    solved organic matter via the sponge loop. Functional Ecol-
    ogy 31:778–789.

    Roik, A., C. Roder, T. Röthig, and C. R. Voolstra. 2016. Spatial
    and seasonal reef calcification in corals and calcareous crusts
    in the central Red Sea. Coral Reefs 35:681–693.

    Roik, A., T. Röthig, C. Pogoreutz, V. Saderne, and C. R. Vool-
    stra. 2018. Coral reef carbonate budgets and ecological dri-
    vers in the central Red Sea—A naturally high temperature
    and high total alkalinity environment. Biogeosciences
    15:6277–6296.

    Roth, F., F. Saalmann, T. Thomson, D. J. Coker, R. Villalobos,
    B. H. Jones, C. Wild, and S. Carvalho. 2018. Coral reef degra-
    dation affects the potential for reef recovery after disturbance.
    Marine Environmental Research 142:48–58.

    Roth, F., I. Stuhldreier, C. Sánchez-Noguera, S. Carvalho, and
    C. Wild. 2017. Simulated overfishing and natural eutrophica-
    tion promote the relative success of a non-indigenous ascid-
    ian in coral reefs at the pacific coast of Costa Rica. Aquatic
    Invasions 12:435–446.

    Roth, F., C. Wild, S. Carvalho, N. Rädecker, C. R. Voolstra, B.
    Kürten, H. Anlauf, Y. C. El-Khaled, R. Carolan, and B. H.
    Jones. 2019. An in situ approach for measuring biogeochemi-
    cal fluxes in structurally complex benthic communities. Meth-
    ods in Ecology and Evolution 10:712–725.

    Russ, G. R. 2003. Grazer biomass correlates more strongly with
    production than with biomass of algal turfs on a coral reef.
    Coral Reefs 22:63–67.

    Savva, I., S. Bennett, G. Roca, G. Jordà, and N. Marbà. 2018.
    Thermal tolerance of Mediterranean marine macrophytes:
    Vulnerability to global warming. Ecology and Evolution
    8:12032–12043.

    Sawall, Y., A. Al-Sofyani, S. Hohn, E. Banguera-Hinestroza, C.
    R. Voolstra, and M. Wahl. 2015. Extensive phenotypic plas-
    ticity of a Red Sea coral over a strong latitudinal temperature
    gradient suggests limited acclimatization potential to warm-
    ing. Scientific Reports 5:8940.

    Scheufen, T., W. E. Krämer, R. Iglesias-Prieto, and S. Enrı́quez.
    2017. Seasonal variation modulates coral sensibility to heat-
    stress and explains annual changes in coral productivity. Sci-
    entific Reports 7:4937.

    Sibly, R., J. Brown, and A. Kodric-Brown. 2012. Metabolic
    ecology: a scaling approach. Wiley-Blackwell: Oxford, UK
    ISBN: 978-1-119-96851-1.

    Silverman, J., B. Lazar, and J. Erez. 2007. Effect of aragonite
    saturation, temperature, and nutrients on the community cal-
    cification rate of a coral reef. Journal of Geophysical
    Research: Oceans 112:1–14.

    Stachowicz, J. J., J. F. Bruno, and J. E. Duffy. 2007. Understand-
    ing the effects of marine biodiversity on communities and

    Article e03226; page 14 FLORIAN ROTH ET AL. Ecology, Vol. 102, No. 2

    https://www.researchgate.net/profile/Martin_Gutbrod2/publication/257233311_First_evidence_of_coral_bleaching_stimulating_organic_matter_release_by_reef_corals/links/5693898008aec14fa55e9224/First-evidence-of-coral-bleaching-stimulating-organic-matter-release-by-reef-corals

    https://www.researchgate.net/profile/Martin_Gutbrod2/publication/257233311_First_evidence_of_coral_bleaching_stimulating_organic_matter_release_by_reef_corals/links/5693898008aec14fa55e9224/First-evidence-of-coral-bleaching-stimulating-organic-matter-release-by-reef-corals

    https://www.researchgate.net/profile/Martin_Gutbrod2/publication/257233311_First_evidence_of_coral_bleaching_stimulating_organic_matter_release_by_reef_corals/links/5693898008aec14fa55e9224/First-evidence-of-coral-bleaching-stimulating-organic-matter-release-by-reef-corals

    https://www.researchgate.net/profile/Martin_Gutbrod2/publication/257233311_First_evidence_of_coral_bleaching_stimulating_organic_matter_release_by_reef_corals/links/5693898008aec14fa55e9224/First-evidence-of-coral-bleaching-stimulating-organic-matter-release-by-reef-corals

    https://www.researchgate.net/profile/Martin_Gutbrod2/publication/257233311_First_evidence_of_coral_bleaching_stimulating_organic_matter_release_by_reef_corals/links/5693898008aec14fa55e9224/First-evidence-of-coral-bleaching-stimulating-organic-matter-release-by-reef-corals

    https://www.ncei.noaa.gov/access/ocean-carbon-data-system/oceans/CO2SYS/co2rprt.html

    https://www.ncei.noaa.gov/access/ocean-carbon-data-system/oceans/CO2SYS/co2rprt.html

    ecosystems. Annual Review of Ecology, Evolution, and Sys-
    tematics 38:739–766.

    Stillman, J. H. 2019. Heat waves, the new normal: summertime
    temperature extremes will impact animals, ecosystems, and
    human communities. Physiology 34:86–100.

    Szmant, A. M. 2002. Nutrient enrichment on coral reefs: is
    it a major cause of coral reef decline? Estuaries 25:
    743–766.

    Takeshita, Y., W. Mcgillis, E. M. Briggs, A. L. Carter, E. M.
    Donham, T. R. Martz, N. N. Price, and J. E. Smith. 2016.
    Assessment of net community production and calcification of
    a coral reef using a boundary layer approach. Journal of Geo-
    physical Research: Oceans 121:5655–5671.

    Tkachenko, K. S., B. J. Wu, L. S. Fang, and T. Y. Fan. 2007.
    Dynamics of a coral reef community after mass mortality of
    branching Acropora corals and an outbreak of anemones.
    Marine Biology 151:185–194.

    Van Heuven, S. M. A. C., A. E. Webb, D. M. de Bakker, E.
    Meesters, F. C. Van Duyl, G.-J.-J. Reichart, and L. J. De
    Nooijer. 2018. In-situ incubation of a coral patch for commu-
    nity-scale assessment of metabolic and chemical processes on
    a reef slope. PeerJ 2018:e5966.

    Weeks, S. J., K. R. N. Anthony, A. Bakun, G. C. Feldman, and
    O.-H. Guldberg. 2008. Improved predictions of coral bleach-
    ing using seasonal baselines and higher spatial resolution.
    Limnology and Oceanography 53:1369–1375.

    Wild, C., et al. 2011. Climate change impedes scleractinian cor-
    als as primary reef ecosystem engineers. Marine and Freshwa-
    ter Research 62:205–215.

    Wild, C., M. Huettel, A. Klueter, S. G. Kremb, M. Y. M.
    Rasheed, and B. B. Jørgensen. 2004. Coral mucus functions
    as an energy carrier and particle trap in the reef ecosystem.
    Nature 428:66–70.

    Wolf-Gladrow, D. A., R. E. Zeebe, C. Klaas, A. Körtzinger,
    and A. G. Dickson. 2007. Total alkalinity: The explicit con-
    servative expression and its application to biogeochemical
    processes. Marine Chemistry 106:287–300.

    Yahel, G., J. H. Sharp, D. Marie, C. Häse, and A. Genin. 2003.
    In situ feeding and element removal in the symbiont-bearing
    sponge Theonella swinhoei: Bulk DOC is the major source for
    carbon. Limnology and Oceanography 48:141–149.

    Zeebe, R. E., and D. A. Wolf-Gladrow. 2001. CO2 in seawater :
    equilibrium, kinetics, isotopes. First editionAmsterdam,
    Netherlands: Elsevier Science. ISBN: 9780080929903

    SUPPORTING INFORMATION

    Additional supporting information may be found in the online version of this article at http://onlinelibrary.wiley.com/doi/
    10.1002/ecy.3226/suppinfo

    February 2021 FUNCTIONING OF CORAL REEF COMMUNITIES Article e03226; page 15

    http://onlinelibrary.wiley.com/doi/10.1002/ecy.3226/suppinfo

    http://onlinelibrary.wiley.com/doi/10.1002/ecy.3226/suppinfo

    Journal of Experimental Marine Biology and Ecology 534 (2021) 15147

    7

    Available online 6 November 2020
    0022-0981/© 2020 Elsevier B.V. All rights reserved.

  • Coral-macroalgal competition under ocean warming and acidification
  • Lena Rölfer a, b, *, 1, Hauke Reuter a, b, Sebastian C.A. Ferse a, b, Andreas Kubicek c, Sophie Dove c,
    Ove Hoegh-Guldberg c, Dorothea Bender-Champ c

    a Leibniz Centre for Tropical Marine Research (ZMT), Fahrenheitstraße 6, D-28359 Bremen, Germany
    b Faculty of Biology & Chemistry (FB2), University of Bremen, D-28359 Bremen, Germany
    c Global Change Institute and School for Biological Sciences, University of Queensland, 4072 Brisbane, Australia

    A R T I C L E I N F O

    Keywords:
    Coral-macroalgal interaction
    Ocean acidification
    Representative concentration pathways
    Porites lobata
    Chlorodesmis fastigiata
    Climate change

    A B S T R A C T

    Competition between corals and macroalgae is frequently observed on reefs with the outcome of these in-
    teractions affecting the relative abundance of reef organisms and therefore reef health. Anthropogenic activities
    have resulted in increased atmospheric CO2 levels and a subsequent rise in ocean temperatures. In addition to
    increasing water temperature, elevated CO2 levels are leading to a decrease in oceanic pH (ocean acidification).
    These two changes have the potential to alter ecological processes within the oceans, including the outcome of
    competitive coral-macroalgal interactions. In our study, we explored the combined effect of temperature increase
    and ocean acidification on the competition between the coral Porites lobata and on the Great Barrier Reef
    abundant macroalga Chlorodesmis fastigiata. A temperature increase of +1 ◦C above present temperatures and
    CO2 increase of +85 ppm were used to simulate a low end emission scenario for the mid- to late 21st century,
    according to the Representative Concentration Pathway 2.6 (RCP2.6). Our results revealed that the net photo-
    synthesis of P. lobata decreased when it was in contact with C. fastigiata under ambient conditions, and that dark
    respiration increased under RCP2.6 conditions. The Photosynthesis to Respiration (P:R) ratios of corals as they
    interacted with macroalgal competitors were not significantly different between scenarios. Dark calcification
    rates of corals under RCP2.6 conditions, however, were negative and significantly decreased compared to
    ambient conditions. Light calcification rates were negatively affected by the interaction of macroalgal contact in
    the RCP2.6 scenario, compared to algal mimics and to coral under ambient conditions. Chlorophyll a, and protein
    content increased in the RCP2.6 scenario, but were not influenced by contact with the macroalga. We conclude
    that the coral host was negatively affected by RCP2.6 conditions, whereas the productivity of its symbionts
    (zooxanthellae) was enhanced. While a negative effect of the macroalga (C. fastigiata) on the coral (P. lobata) was
    observed for the P:R ratio under control conditions, it was not enhanced under RCP2.6 conditions.

    1. Introduction

    Macroalgae are important organisms on coral reefs, contributing
    significantly to primary production (Gattuso et al., 1998) and nitrogen
    fixation (Heil et al., 2004). On a healthy reef, corals generally pre-
    dominate the benthic community and are generally competitively su-
    perior to macroalgae (Chadwick and Morrow, 2011). However, in recent
    years reef ecosystems experienced dramatic declines in coral cover due
    to anthropogenic impacts such as global climate change, ocean acidifi-
    cation, eutrophication, sedimentation and overfishing as well as disease

    outbreaks (Hoegh-Guldberg et al., 2007; Hughes et al., 2010, 2007).
    Between 2014 and 2017 a 36 month global heatwave led to multiple
    bleaching events on coral reefs and on the Great Barrier Reef to a loss of
    shallow water corals of 22–30% with even 50% in the northern parts
    (Hughes et al., 2017; Eakin et al., 2018).

    The free space on the reef created by high coral mortality can be
    taken up by other sessile benthic organisms such as macroalgae, cor-
    allimorpharians and sponges (Aronson and Precht, 2001; Norström
    et al., 2009). Competition between benthic, sessile organisms is one of
    the main factors shaping the community composition on reefs (Dayton,

    Abbreviations: RCP2.6, Representative Concentration Pathway 2.6; Pnet, net oxygen production (net photosynthesis); Rdark, dark respiration rate; P:R, Photo-
    synthesis per Respiration ratio.

    * Corresponding author at: Leibniz Centre for Tropical Marine Research (ZMT), Fahrenheitstraße 6, D-28359 Bremen, Germany.
    E-mail address: lena@roelfer.de (L. Rölfer).

    1 Present address: Climate Service Center Germany (GERICS), Helmholtz-Zentrum Geesthacht, Fischertwiete 1, D-20095 Hamburg, Germany.

    Contents lists available at ScienceDirect

    Journal of Experimental Marine Biology and Ecology

    journal homepage: www.elsevier.com/locate/jembe

    https://doi.org/10.1016/j.jembe.2020.151477
    Received 4 June 2020; Received in revised form 2 October 2020; Accepted 29 October 2020

    mailto:lena@roelfer.de

    www.sciencedirect.com/science/journal/00220981

    https://www.elsevier.com/locate/jembe

    https://doi.org/10.1016/j.jembe.2020.151477

    https://doi.org/10.1016/j.jembe.2020.151477

    https://doi.org/10.1016/j.jembe.2020.151477

    http://crossmark.crossref.org/dialog/?doi=10.1016/j.jembe.2020.151477&domain=pdf

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    2

    1971). Macroalgae are competing for space in different ways, including
    physical abrasion of coral tissue (Coyer et al., 1993), shading (Hughes,
    1994) or allelochemicals (i.e. harmful chemicals) to induce bleaching or
    death in corals (Longo and Hay, 2017; McCook, 2001; Nugues et al.,
    2004). Furthermore, macroalgal exudates can lead to microbe-induced
    mortality of adjacent corals (Clements et al., 2020; Smith et al., 2006).
    Corals that are weakened by anthropogenic impacts are not able to
    invest energy in spatial competition as their energy is needed for various
    maintenance functions (Diaz-Pulido et al., 2011; Foster et al., 2008;
    Rinkevich and Loya, 1985). Additionally, changes in the reef environ-
    ment such as high nutrient availability (eutrophication) and overfishing
    of herbivores can result in enhanced growth rates of macroalgae as well
    as their release from predation pressure (Hughes et al., 2007, 1987;
    Shenkar et al., 2008). As a consequence, macroalgae become the
    stronger competitor and proliferate over the reef environment (Done,
    1992; McCook, 1999), which may lead to a phase shift from a coral to an
    algal dominated state (Anton et al., 2020; Done, 1992; Hughes et al.,
    2010; Norström et al., 2009).

    Ocean warming and acidification, combined with disturbances such
    as overfishing and nutrient enrichment, have a high potential to also
    decrease resilience of coral reefs (Anthony et al., 2011; Dove et al.,
    2013) and may change the outcome of competition (Chadwick and
    Morrow, 2011; Diaz-Pulido et al., 2011; Hoegh-Guldberg et al., 2007).
    Thus, on many reefs worldwide macroalgae are the winners in the
    competition for space on coral reefs (Gardner, 2003; McCook, 1999;
    Mumby et al., 2013; Scheffer et al., 2001).

    Corals are particularly vulnerable to ocean acidification (Pörtner
    et al., 2019), resulting in a significant reduction in calcification rate
    through a decreased aragonite saturation, which controls coral calcifi-
    cation (Doney et al., 2009; Kleypas and Langdon, 2006; Langdon, 2002).
    With an increase of dissolved CO2, however, productivity may be
    enhanced, as CO2 is often a limiting factor in the marine realm.
    Enhanced productivity under elevated CO2 levels has been shown for
    both, macroalgae (Gao et al., 1993, 1991; Wu et al., 2008) and
    zooxanthellae (Al-Moghrabi et al., 1996; Leggat et al., 1999).

    Under current conditions, corals are already at their thermal limits,
    and with temperatures continuing to rise, corals will be pushed more
    frequently beyond their thermal tolerance threshold, as oceans warm
    (Hoegh-Guldberg, 1999; Hoegh-Guldberg et al., 2007; Anton et al.,
    2020). This was last observed in the longest and most severe coral
    bleaching event 2014–2017, when global monthly sea surface temper-
    ature maxima increased to 30–31 ◦C (Lough et al., 2018), which led to a
    mass mortality of shallow water corals (Heron et al., 2016; Hughes et al.,
    2017).

    Investigations about coral-algal interactions have strongly increased
    during the last decade, monitoring interactions both in situ and in tank
    experiments (Bender et al., 2012; Birrell et al., 2008; Brown et al., 2019;
    Del Monaco et al., 2017; Diaz-Pulido et al., 2010; Diaz-Pulido and
    Barrón, 2020; Jompa and McCook, 2003; McCook, 2001; Rasher et al.,
    2011; Rasher and Hay, 2010; Vieira et al., 2016)However, information
    on the impacts of global and climate change on ecological interactions
    are still underrepresented in the literatura.

    In order to explore the interaction of the macroalgae and corals
    under climate change, we used the macroalgae Chlorodesmis fastigiata (C.
    Agardh) S.C.Ducker and the coral Porites lobata (Dana, 1846) from
    Heron Island on the southern Great Barrier Reef. Both species are
    abundant and interactions between each are frequently observed
    (Jompa and McCook, 2003). C. fastigiata, a siphonous green macroalga,
    has been shown to have mostly negative impacts on corals, such as
    inducing bleaching (Bonaldo and Hay, 2014), decreased photosynthesis
    (Rasher et al., 2011; Rasher and Hay, 2010), polyp retraction by abra-
    sion (Jompa and McCook, 2003), reduced tissue recovery (Bender et al.,
    2012), and reduced coral settlement (Birrell et al., 2008). P. lobata is a
    massive, colonial coral from the order Scleractinia and constitutes one of
    the most important and occurring reef-building corals on Pacific coral
    reefs (Budd, 1986).

    The combined effect of elevated CO2 and temperature on the inter-
    action between P. lobata and C. fastigiata was tested against ambient
    seawater as control, to provide further insights into coral-algal in-
    teractions under changing environmental conditions. Temperature of
    the treatment water was increased by 1 ◦C (compared to today; +2 ◦C
    compared to preindustrial times) and the CO2-concentration by
    +85 ppm according to the RCP2.6 scenario for the mid- to late 21st
    century (IPCC, 2013). We hypothesized that the coral would be nega-
    tively affected by both, the interaction with macroalgae and RCP2.6
    conditions, and that the interaction with live algae would induce
    stronger effects than mimics because of biological/chemical in addition
    to mechanical effects (e.g. shading or abrasion).

    2. Materials and methods

    2.1. Study site and collection of corals and macroalgae

    This study was done between October and December 2016 (late
    spring/Australian summer) at Heron Island Research Station (HIRS),
    located in the southern section of the Great Barrier Reef. Organisms
    were collected with permission from the Great Barrier Reef Marine Park
    Authority (permit number G16/38942.1). Colonies of P. lobata
    (approximately 30 cm in diameter) were collected at the reef flat of
    Heron Reef (four colonies, 23◦26′S, 151◦54′E) and the adjacent Wistari
    Reef (two colonies, 23◦27′S, 151◦52′E). From the colonies, 90 coral
    cores (further referred to as corals) of 5 cm diameter were drilled using a
    core saw, cut to a height of 1.5 cm and put into tanks with constant flow-
    through of seawater for a recovery time of 5 days.

    Thirty specimens of C. fastigiata of approximately the same size as
    coral cores were collected with a small amount of substrate (Bonaldo
    and Hay, 2014) at the reef flat of Heron Reef. Care was taken not to
    injure holdfast or other tissue to avoid leaking of chemical compounds.
    Macroalgal substrate was cleaned of crabs and other organisms that
    lived and fed on the macroalgae were removed. Macroalgae were put
    into tanks with corals, with care not to touch the corals (or other mac-
    roalgae). After the recovery time of five days, seawater in all tanks was
    slowly switched to treatment water over another five days in order to let
    corals and algae acclimatize to the new conditions. Macroalgae mimics
    were made from 13 cm long pieces of fibre rope (0.7 cm in diameter),
    which were bent once in the middle, tied at the base and separated into
    its fibres (Edgar and Klumpp, 2003), and pieces of substrate were
    attached to the bottom. After the acclimatization time, mimics and
    specimens of C. fastigiata were tied to coral cores using rubber bands
    (without touching coral tissue) (Bonaldo and Hay, 2014) to avoid
    mimics from floating and to assure that macroalgae stayed with the
    same coral for the time of the experiment.

    2.2. Experimental set-up, maintenance and monitoring

    To assess the effect of elevated temperature and ocean acidification
    on a coral-algal interaction, organisms were subjected to ambient con-
    ditions and the Intergovernmental Panel on Climate Change (IPCC)
    scenario RCP2.6 conditions (Dove et al., 2013). Data collected at a
    reference site in the Wistari Channel (adjacent to Heron Reef) was used
    as a baseline for ambient temperature and CO2. Sea water was pumped
    from the reef flat to a holding tower at Heron Island Research Station
    and redistributed into two sumps, in which CO2 and temperature
    treatment conditions were established (as described by Dove et al.,
    2013). From these sumps the experimental tanks were supplied with the
    respective treatment water. Temperature and pH feedback sensors in
    experimental tanks were connected to a system controller, which then
    adjusted the conditions in the sumps at intervals of two hours to guar-
    antee exact diurnal and seasonal conditions within the experimental
    tanks (Fig. 1). See Dove et al. (2013) for details. Both water treatments
    (ambient and RCP2.6) were applied to nine tanks each (18 tanks in
    total). Per treatment condition, three tanks each contained five

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    3

    replicates of either a) only corals, b) corals with interacting macroalgae,
    or c) corals with macroalgal mimics (Fig. 2), summing up to 15 coral
    replicates per contact treatment. Organisms were kept under treatment
    conditions for a period of 22 days and physiological measurements were
    performed subsequently.

    Experimental tanks (60*20*38 cm; ~ 35 l) containing corals and
    macroalgae were covered with Lee-filters (Old Steel Blue, 725) on the
    sides and lids to imitate light intensity at a water depth of 3–6 m on the
    reef flat. They were exposed to natural sun irradiance with a constant
    flow-through of seawater (0.8–1 l*min− 1). Pumps (Clearpond Infiniti
    800) were installed to agitate seawater in order to avoid the forming of a
    boundary layer. Tanks were cleaned from fouling organisms every three
    days, with macroalgal mimics being washed in freshwater to remove
    biofouling, and settled substrate was carefully cleaned off corals with
    toothbrushes. At the same time, tanks were rotated to avoid confounding
    of light and temperature differences.

    To monitor the abiotic conditions within tanks (Table 1), we
    installed four light loggers (PAR Sensor, Odyssey, Dataflow Systems, New
    Zealand) and five temperature loggers (HOBO Data Loggers, Onset) as
    well as four pH-probes (Mettler Toledo, Port Melbourne, Victoria,

    Australia, InPro4501VP X connected to an Aquatronica Aquarium
    Controller ACQ110). pH was measured on the scale pHsw and the pH
    probes were calibrated every second to third day by a two point cali-
    bration. The loggers and probes were randomly swapped between tanks
    every three days to monitor every tank throughout the experiment.
    Temperature and pCO2 contents of sumps were recorded additionally
    (Table 1). pCO2 was measured in the sumps (logged continuously every
    3 min) and calculated in CO2SYS (developed by E. Lewis and W.R.
    Wallace) based on twice daily alkalinity and salinity sampling, and
    continuous temperature and pH monitoring at 10 min intervals (see
    Dove et al. (2013) for more detail). The total alkalinity of each tank was
    sampled once a week at midday and midnight (Table 1) and measured
    using a Mettler Toledo titrating system (T50) by Gran titration after
    Dickson et al. (2003) using the method with a precision of ±3 μmol Kg− 1
    or better as described in Kline et al. (2012). RCP2.6 conditions of ~1 ◦C
    increase and a CO2 level of +85 ppm compared to ambient conditions
    could be maintained during the experimental period.

    Fig. 1. Schematic of sumps and wet table set
    up with partial pressure of CO2(pCO2)/
    temperature control system. Seawater from
    the tower (light blue) is pumped into sumps
    (dark blue, red), treatment conditions are
    applied and tanks (n = 3 per interaction
    treatment) on wet table are connected to
    sumps. System controller (grey) with feed-
    back loop (dotted line) to adjust conditions
    in a two hour interval. AMB: ambient
    seawater; RCP2.6: treatment seawater with
    increased temperature and pCO2. (For
    interpretation of the references to colour in
    this figure legend, the reader is referred to
    the web version of this article.)

    Fig. 2. Pictures of interaction treatments: a) coral P. lobata, b) coral P. lobata with interacting alga C. fastigiata, and c) coral P. lobata with algal mimic (made out of
    plastic fibre rope). Ambient and RCP2.6 water treatments were applied to each 3 tanks with 5 replicates each per interaction treatment.

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    4

    2.3. Measurement of respiration and calcification rates

    In order to investigate treatment effects on the metabolism of corals
    and macroalgae, the net oxygen production (Pnet) and dark respiration
    rates (Rdark) were measured following the methodology of Crawley et al.
    (2010) at the end of the experiment (i.e. after 22 days in treatment
    conditions). Organisms (ncoral = 12 per treatment, nmacroalgae = 6 for
    RCP2.6 and 8 for ambient) were dark adapted for 45 min prior to dark
    respiration measurements. For Pnet measurements the light intensity was
    adjusted by modifying the distance of a metal halide lamp (Ocean Light
    T5 MH combo 150 W, with 2 × 24 Ocean Blue Actinic, Aqua-Medic of
    North America, LLC) from the specimens. The average light intensity
    within the experimental tanks matched that measured at midday of the
    previous week (694.5 μmol photons m− 2 s− 1 ≈ h = 32.5 cm). Oxygen
    flux was recorded using high-precision optical oxygen sensors (optodes)
    connected to a logging system (Oxy-10, PreSens, Germany) for 12 and
    20 min for macroalgae and corals, respectively, and was normalized to
    the surface area of corals (see below) and fresh weight of macroalgae. All
    measurements were conducted in the respective treatment water (tam-
    bient = 25.9 ◦C, tRCP2.6 = 26.9 ◦C) calculated as a mean for the previous
    week. The photosynthesis per respiration (P:R) ratio was calculated by
    dividing Pnet by -Rdark.

    The light and dark calcification rates were measured in addition to
    the photosynthetic/respiratory rates of the corals. To do this, alkalinity
    samples (n = 5 per treatment) were taken from respirometry incubations
    and the changes in the calcification of corals measured using the alka-
    linity anomaly technique using bulk water samples as blanks. Alkalinity
    samples were stored in a fridge at 4 ◦C, and within eight weeks after
    collection (Huang et al., 2012), were measured using a Mettler Toledo
    titrating system (T50) by Gran titration after Dickson et al. (2003) with a
    precision of ±3 μmol Kg− 1 or better using the method described in Kline
    et al. (2012). The calcification rate was then calculated from the total
    alkalinity after Zundelevich et al. (2007), corrected by 10− 3 for unit
    conversion and divided by two as molar amount of dissolved CaCO3
    equals only half of observed AT increase (Chisholm and Gattuso, 1991).
    Calcification rates were normalized to time and surface area and
    expressed in μg CaCO3 cm− 2 h− 1.

    2.4. Growth measurements

    To examine possible differences in growth rates of organisms be-
    tween treatments, buoyant weight of corals (Jokiel and Maragos, 1978)
    (n = 15 per treatment) and fresh weight of macroalgae (n = 15 per
    treatment) were measured before the tank period and after respirometry
    measurements were conducted (i.e after 26 days). Fresh weight of
    macroalgae was calculated by subtracting the weight of the substrate
    (measured by buoyant weight) from the total weight. To minimize in-
    accuracy of measurements, macroalgae were blotted with paper tissue
    prior to weighing. Weight differences were calculated as percentage
    change of buoyant weight and fresh weight for corals and macroalgae,
    respectively. After weighing, all samples were frozen at − 20 ◦C pending
    further analysis.

    2.5. Tissue analysis of corals

    Tissue samples were taken from same corals used for respirometry
    measurements (n = 12 per treatment), by removal with a seawater jet
    (Johannes and Wiebe, 1970). Samples were poured into a Falcon tube
    and then vortexed and centrifuged at 4500 rpm for 5 min at 4 ◦C. The
    supernatant was poured into a 2 ml Eppendorf tube for protein analysis.
    To dilute the pellet left in the tube, 10 ml of filtered seawater was added
    and vortexed. From this dilution, 1 ml was pipetted into two tubes for
    zooxanthellae count and chlorophyll a analysis.

    The population density of zooxanthellae was measured using a
    Neubauer hemocytometer (Neubauer-improved, Marienfeld GmbH),
    counting cells within three squares of four replicate grids. Chlorophyll a
    was measured by adding 2 ml of 100% acetone to the 1 ml pellet dilu-
    tion. Samples were sonicated in an ice bath for 10 min to extract the
    pigments and then centrifuged at 4500 rpm for 5 min at 4 ◦C to separate
    the pellet from the pigment solution. The supernatant was poured into a
    new Falcon tube and frozen and the extraction was repeated twice until
    supernatant was clear. After the extraction, tubes with supernatants
    were centrifuged again to remove debris (Hellebust and Craigie, 1978).
    Samples were measured in a spectrophotometer (SpectrostarNano) at
    wavelengths of 663 nm and 645 nm and blanks of acetone were
    measured after every ten samples and subtracted directly. The amount of
    chlorophyll a was then calculated after Arnon (1949) and expressed in g
    l− 1. As a quantitative indicator for the thickness of the tissue, the protein
    content was measured. Subsamples were measured as triplicates in the
    spectrophotometer at wavelengths of 235 nm and 280 nm. Protein
    content was calculated after Whitaker and Granum (1980) and
    expressed in g l− 1. To normalize tissue properties, respirometry mea-
    surements and calcification rates, the corals’ surface area was measured
    using the paraffin wax technique (Stimson and Kinzie III, 1991).

    2.6. Data analysis

    Before statistical analysis of data was performed, all variables were
    tested for a possible tank effect by using the lmne package in R Studio,
    which compares Gaussian linear and nonlinear mixed-effect models. The
    test was performed for models with and without tank as a random factor.
    Since the tank factor was non-significant for all variables (p > 0.25)
    (results in supplementary table S2), specimens were used as replicates
    (Underwood, 1997), hence increasing the power of the analysis.

    All response data of corals were tested using a two-factor analysis of
    variance (ANOVA) with “scenario” (ambient; RCP2.6) and “contact”
    (coral; coral-algal interaction; coral with algal mimic) as fixed factors,
    including the interaction term. Response data of macroalgae were tested
    using a one-way ANOVA with “scenario” as factor. To account for
    multiple comparisons of physiological parameters for P. lobata a Bon-
    ferroni correction was applied reducing the α-level of significance to
    0.005. When significant effects of factors occurred, ANOVAs were fol-
    lowed by a Tukey multiple comparisons test to identify significant
    groups. Data were tested for homogeneity of variance (visual inspection

    Table 1
    Summary of values of water chemistry data for scenario conditions.

    Ambient RCP2.6

    Temperature [◦C] tanks 25.95 ± 0.65 26.88 ± 0.75
    pHSW tanks 8.07 ± 0.03 7.96 ± 0.03
    Temperature [◦C] sumps 25.38 ± 0.44 26.29 ± 0.57
    pCO2 [ppm] sumps 465.27 ± 53.48 550.58 ± 49.52
    Total alkalinity day [μmol Eq. L− 1]

    Day 2289.65 ± 12.23 2287.18 ± 11.86
    Night 2269.11 ± 26.18 2271.64 ± 25.18

    pCO2 [μppm] CO2Sys
    Day 268.17 ± 30.32 282.13 ± 25.32
    Night 311.26 ± 26.52 331.45 ± 21.68

    HCO3
    − [μmol Kg SW− 1]

    Day 1706.37 ± 43.74 1723.38 ± 36.09
    Night 1744.8 ± 35.13 1767.53 ± 31.80

    CO3
    2− [μmol Kg SW− 1]

    Day 234.64 ± 15.85 226.79 ± 11.60
    Night 210.61 ± 10.87 202.61 ± 7.98

    ΩAragonite
    Day 3.66 ± 0.25 3.54 ± 0.18
    Night 3.28 ± 0.17 3.16 ± 0.12

    Temperature, pH, pCO2 values are means over experimental period, continu-
    ously measured over the experimental period (22 days). High standard deviation
    is due to daily variability. Total alkalinity, pCO2, HCO3

    − , CO3 and ΩAragonite of
    treatments are given as means over experimental period, measured once a week
    at midday and midnight in tanks and were estimated using CO2SYS software. Kg
    SW, kilogram of seawater.

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    5

    of residuals vs. fitted values), and normality of residuals was tested using
    Shapiro-Wilk normality test. Non-normally distributed data were log or
    power transformed to correct for right- or left skew, respectively. Sta-
    tistical analysis of data was performed using R Studio Version 1.0.143 (R
    Core Team, 2015) and results were expressed as boxplots using the ggplot
    package. Some samples of chlorophyll a and protein content of corals
    were lost throughout the analysis, reducing the degrees of freedom in
    the analysis.

    3. Results

    3.1. Corals

    Coral growth, reported as percentage change in buoyant weight, was
    not significantly affected by the RCP2.6 or macroalgae treatment, or a
    combination of both. However, slightly negative growth was observed in
    the RCP2.6 coral treatment with an average of − 0.34% change in
    buoyant weight (see supplementary TableS1).

    The technique used for the respirometry measurements allowed the
    quantification of calcification rates under treatment conditions in the
    presence and absence of light for a particular point in time. The response
    of light calcification was dominated by an interaction of “scenario” and
    “contact”. In the absence of macroalgae or mimics, corals showed
    significantly higher light calcification under ambient conditions with
    48.7 ± 10.3 CaCO3 [μg h− 1 cm− 2] as compared to 15.1 ± 5.9 CaCO3
    [μg h− 1 cm− 2] in the RCP2.6 scenario (Table 2, Fig. 3a). Corals in the
    ambient treatment also performed significantly different compared to
    corals in interaction with macroalgae in the RCP2.6 treatment. Dark
    calcification was significantly greater in the ambient scenario, with

    negative dark calcification in the RCP2.6 scenario for all contact treat-
    ments (Table 2, Fig. 3b).

    Pnet was significantly influenced by “contact” (Table 2, supplemen-
    tary material Fig.S1A) and lower in the coral-algal interaction treatment
    compared to only corals, while there was no difference of both treat-
    ments to the mimics treatment. Rdark was significantly affected by
    “scenario” (Table 2, supplementary material Fig.S1B), with higher
    respiration under RCP2.6 compared to ambient conditions. For the P:R
    ratio there was a highly significant effect of “contact” (Fig. 3c, Table 2),
    with differences among all treatments. P:R ratio was highest for the
    mimics treatment, followed by the coral treatment and lowest for the
    coral-algal interaction treatment. While the P:R ratio was higher under
    ambient conditions for only corals and corals with mimics compared to
    RCP2.6 scenario conditions, there was an opposite trend for the coral-
    algal interaction. However, the interaction term of “scenario” and
    “contact” was not significant (Anova, p = 0.03, Table 2).

    Chlorophyll a (Fig. 4d and Protein content (Fig. 4a) were signifi-
    cantly different among “scenarios” (Table 2). The mean value of chlo-
    rophyll a content in the coral-algal interaction under RCP2.6 conditions
    was more than two times higher than for the interaction under ambient
    conditions (Fig. 4d). There was a slight trend of a higher zooxanthellae
    population density in the RCP2.6 compared to the ambient scenario,
    irrespective of the contact treatment (Table 2, Fig. 4b).

    3.2. Macroalgae

    The macroalga at the centre of this study was very sensitive, and
    began to die after eight days of exposure to treatment conditions.
    Macroalgae that died were replaced once by ‘back-up’ macroalgae,
    which were acclimated and kept in additional tanks with ambient and
    RCP2.6 conditions, respectively. However, due to permit limitations no
    more macroalgae could be replaced after that. Over the remainder of the
    experiment another 14 macroalgae died, some of which were ‘back-up’
    macroalgae, summing up to a total of 30 dead macroalgae (n = 13 in
    ambient, n = 17 in RCP2.6). Only 16 macroalgae survived throughout
    the whole experimental time (n = 9 in ambient, n = 7 in RCP2.6),
    reducing the degrees of freedom in the analysis.

    Due to differences in size of macroalgae, growth was expressed as
    percentage change in fresh weight. Growth of individuals was positive
    and negative in both of the scenarios, with negative values resulting
    from a loss of filaments, explaining a high standard deviation. However,
    total change in fresh weight was positive in both treatments, with no
    significant difference between treatments (Table 3). There was also no
    significant difference in Rnet between the scenarios (Table 3). Rdark was
    slightly higher in the RCP2.6 compared to the ambient scenario,
    resulting in a non-significant difference of P:R ratios between scenarios
    (Table 3).

    4. Discussion

    We investigated the important issues as to how ecological competi-
    tion may vary under climate change. To do so, we investigated the in-
    teractions between the coral P. lobata and the potential competitor, the
    fleshy alga C.fastigiate, under low rates of future ocean warming and
    acidification. Corals and macroalgae were exposed to a temperature
    increase of +1 ◦C and a CO2 increase of +85 ppm above ambient, which
    is close to the RCP2.6 scenario of the IPCC at mid- to late 21st century
    (IPCC, 2013). This corresponds to CO2 levels expected if action is taken
    globally in accordance with the Paris Agreement and refers to the best-
    case scenario. We found that both, interaction with macroalgae and
    the combined effect of temperature and CO2 affected the coral, whereas
    no significant impact of treatment conditions was detected for the
    macroalgae.

    Table 2
    ANOVA output of different variables for P. lobata with bold values indicating
    significant effects on the variable.

    Variable Source of variation df F p

    Bouyant weight Scenario 1 0.279 0.599
    Contact 2 1.633 0.202
    Scenario x Contact 2 1.587 0.211
    Residuals 81

    Light calcification Scenario 1 4.354 0.048
    Contact 2 2.878 0.076
    Scenario x Contact 2 7.480 0.003
    Residuals 24

    Dark calcification Scenario 1 36.792 <0.001 Contact 2 5.007 0.013 Scenario x Contact 2 2.843 0.078 Residuals 24

    Zooxanthellae Scenario 1 7.031 0.010
    Contact 2 0.821 0.445
    Scenario x Contact 2 0.034 0.967
    Residuals 66

    Chlorophyll a Scenario 1 13.909 <0.001 Contact 2 0.068 0.934 Scenario x Contact 2 3.434 0.039 Residuals 54

    Protein Scenario 1 12.448 <0.001 Contact 2 0.848 0.433 Scenario x Contact 2 2.228 0.117 Residuals 60

    Pnet Scenario 1 2.390 0.127
    Contact 2 5.893 0.004
    Scenario x Contact 2 3.485 0.036
    Residuals 66

    Rdark Scenario 1 8.749 0.004
    Contact 2 3.631 0.032
    Scenario x Contact 2 0.156 0.856
    Residuals 66

    P:R Scenario 1 2.899 0.093
    Contact 2 13.466 <0.001 Scenario x Contact 2 3.693 0.030 Residuals 66

    df = degrees of freedom; F = F-value; p = p-value (significance <0.005).

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    6

    Fig. 3. Calcification rates of P. lobata under a light (n = 5), and b dark conditions (n = 5), measured during respirometry incubations. Letters indicate significant
    differences between interactions (a) or among scenarios (b).

    Fig. 4. a Protein content (n = 11), b Zooxanthellae density (n = 12), c Photosynthesis/Respiration (P:R) ratio (n = 12), d Chlorophyll a content (n = 10), protein
    content (n = 11) for P. lobata. Letters indicate significant differences among scenarios (a,d) or contact (c).

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477
    7

    4.1. Calcification under low rates of warming and acidification

    The calcification of massive species such as P. lobata can be slow, as
    compared to faster growing corals such as branching species (Lough
    et al., 1999). In this study, percentage change in buoyant weight
    (~deposited calcium carbonate) was close to zero in all treatments. It is
    very likely, however, that the period of our tank experiment (22 days)
    was too short, as well as the increase of pCO2 too small, to lead to a
    detectable effect in buoyant weight. Anthony et al. (2008) measured
    growth of P. lobata over a period of eight weeks and reported slightly
    reduced growth rate at 520–700 ppm and a ~ 40% decrease at
    1000–1300 ppm. Diaz-Pulido et al. (2011) found that the linear exten-
    sion of the fast growing coral Acropora intermedia, measured over eight
    weeks, was also strongly negatively affected by CO2 treatments, but
    showed no significant difference between treatments with or without
    competition with the seaweed Lobophora papenfussii.

    While a change in buoyant weight was not detected, measurements
    of dark calcification rates were significantly decreased under conditions
    similar to RCP2.6 for all interaction treatments, as well as for the coral
    treatment under light conditions. This is in agreement with other studies
    on coral of the genus Porites, including P. lobata (Anthony et al., 2008);
    P. lutea (Ohde and Hossain, 2004); P. compressa (Marubini et al., 2003)
    and supports the sensitivity of corals to elevated CO2 (Kleypas and
    Langdon, 2006). In the present study, negative calcification (i.e. decal-
    cification) was observed in all treatments under RCP2.6 conditions in
    the dark. CaCO3 dissolution even exceeded light calcification in the coral
    and coral-algae interaction treatment which would lead to negative
    growth rates in these treatments if measured over a longer period. While
    the amount of energy available from photosynthesis (P:R) was stable
    among scenario conditions in the coral-algal interaction, dark calcifi-
    cation was reduced under RCP2.6 conditions. This suggests that re-
    sources were used for processes other than calcification that demanded
    higher energy expenditure under RCP2.6 conditions in the dark while in
    contact with the alga.

    Calcification under illuminated conditions in the RCP2.6 scenario
    was significantly reduced in the interaction with the macroalgae
    compared to algal mimics and to coral under ambient conditions. Those
    results suggest that macroalgal mimics benefitted the light calcification
    of corals through shading by reducing irradiance and therefore light
    stress (Anthony et al., 2008), while this positive effect could not be
    detected for live algae.

    4.2. Photosynthesis and respiration: evidence of a CO2 fertilization effect?

    Contrary to the results for the calcification rates, we found that
    RCP2.6 conditions significantly increased chlorophyll a and protein
    content. The increase in chlorophyll a coincides with an increased net
    photosynthesis under RCP2.6 conditions and might be explained by a
    ‘CO2 fertilization effect’ due to the greater availability of CO2 to
    photosynthesize. An increase in chlorophyll an under future conditions
    was also found in the branching corals Stylophora pistillata under raised
    temperature (Reynaud et al., 2003) and Acropora formosa under elevated
    CO2 (Crawley et al., 2010). Crawley et al. (2010) used an increase of

    >200 ppm, but interestingly, a CO2 increase of only +85 ppm in this
    study was sufficient to lead to an increase in chlorophyll a. This is
    comparable to another study, which found an increase in productivity at
    an intermediate CO2 scenario (520–705 ppm), while the positive effect
    was mitigated at high CO2 (1010–1350 ppm) (Anthony et al., 2008). The
    positive effect of CO2 on chlorophyll a found in this study facilitates the
    slight increase of zooxanthellae and therefore protein content, which is
    an indicator for the nutritional condition of the coral (Ferrier-Pagès
    et al., 2003). However, the positive effect of CO2 on the chlorophyll a
    content could be mitigated when CO2 concentrations reach a higher
    level, as more energy is needed to maintain base functions of the coral
    host. This negative effect may be further enhanced by other anthropo-
    genic stressors, which weaken the competitive strength of corals over
    macroalgae (Diaz-Pulido et al., 2011; Foster et al., 2008).

    While a decrease in photosynthesis of corals in contact with various
    macroalgae is documented (Rasher et al., 2011; Rasher and Hay, 2010),
    the interacting effects of CO2 and temperature on the interaction have
    scarcely been considered yet. Our study showed that there was a sig-
    nificant decreased P:R ratio of corals in interaction with macroalgae
    compared to no contact and the mimics treatment irrespective of the
    temperature/CO2 regime the corals were under. RCP2.6 conditions had
    a lesser negative effect that was only visible as a trend in the coral only
    and mimics treatment. C. fastigiata caused a significant decrease of
    photosynthetic efficiency in the corals Montipora digitata, Acropora mil-
    lepora and Pocillopora damicornis under ambient conditions, which was
    more severe compared to the effect of seven other common macroalgae
    (Longo and Hay, 2017; Rasher et al., 2011). In our study, however,
    corals were not actually in contact with macroalgae during the physio-
    logical measurements, because each species’ metabolic rate was
    measured separately. Negative carry-over effects of macroalgae on the
    coral photosynthesis and respiration found in our study might be even
    more enhanced if measured whilst in contact.

    4.3. Physical impacts of competitors

    Impacts of macroalgae can harm corals by various mechanisms
    including shading and abrasion (McCook et al., 2001) as well as
    biochemical reactions, e.g. the induced bleaching due to harmful
    chemicals (Longo and Hay, 2017; McCook, 2001; Nugues et al., 2004).
    Indeed, C. fastigiata has been shown to produce allelochemicals that can
    suppress photosynthesis (Rasher et al., 2011) and cause bleaching
    (Rasher and Hay, 2010). A study by Del Monaco et al. (2017) investi-
    gated the impact of allelochemical extracts from C. fastigiata on corals
    over the same time scale as our project and Diaz-Pulido and Barrón
    (2020) tested the release of dissolved organic carbon, which can pro-
    mote bacterial metabolism on corals surface and subsequent mortality,
    under future CO2 conditions. Both studies conclude that the effects of
    C. fastigiata were not more harmful to corals under future climatic
    conditions (Del Monaco et al., 2017; Diaz-Pulido and Barrón, 2020).This
    is in agreement with our results, which show an effect of C. fastigiata on
    the P:R ratio, which however was not enhanced under RCP2.6
    conditions.

    A recent study by Brown et al. (2019) suggests that coral-algal in-
    teractions are temporally variable across seasons. Photosynthetic rates
    of the coral A. intermedia in contact with H. heteromorpha were reduced
    in winter and increased in summer, while calcification rates in summer
    reduced in contact with the algae. Even though photosynthetic activity
    was increased in contact with the algae, negative effects of a high-end
    ocean acidification and warming scenario (RCP8.5) reduced overall
    performance of corals (Brown et al., 2019), which is comparable to the
    results of our study.

    The impacts of competitors can vary strongly among coral-algal in-
    teractions. C. fastigiata is a siphonous macroalga and therefore lacks
    discrete cell walls (Rasher et al., 2011). For that reason, handling such as
    collection or tank cleaning might have had a significant impact on the
    health of the organisms, resulting in a high mortality rate of the

    Table 3
    ANOVA output of different variables for C. fastigiata.

    Variable Source of variation df F p

    Fresh weight Scenario 1 1.470 0.244
    Residuals 15

    Pnet Scenario 1 1.852 0.197
    Residuals 13

    Rdark Scenario 1 0.336 0.572
    Residuals 13

    P:R Scenario 1 0.341 0.569
    Residuals 13

    df = degrees of freedom; F = F-value; p = p-value (significance >0.05).

    L. Rölfer et al.

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    8

    macroalgae. Additionally, cycles of periodic loss and reappearance are
    known to occur in C. fastigiata, but the timing is unknown (Jompa and
    McCook, 2003). As the mortality rate was higher in the RCP2.6 scenario,
    temperature and CO2 increase might have affected algal health and led
    to a die-off. However, measurements are hardly comparable between
    treatments, as time in treatment differed between minimum 13 days and
    maximum 22 days, due to the mortality. Furthermore, only 9 corals in
    the ambient and 7 corals in the RCP2.6 treatment had an algal partner at
    the end of the study. Despite the early die off of macroalgae, we still
    measured their effects on corals, which, however, might have been more
    visible in the absence of algal mortality, especially in the RCP2.6 sce-
    nario, where macroalgae died relatively early in the experiment.

    5. Conclusions

    P. lobata has a ‘massive ‘coral morphology, and has previously been
    shown to be less affected by the interaction with competing macroalgae
    compared to other coral species. The negative impact of C. fastigiata was
    only visible in the decrease of the P:R ratio, but the study shows no
    enhanced impact under RCP2.6 conditions. The energy budget of the
    coral in this study, however, was very likely negatively influenced by
    RCP2.6 scenario conditions. Calcification, which is directly linked to the
    aragonite saturation, was probably negatively affected by the increase in
    CO2 as shown in earlier studies. We hypothesize that the productivity of
    zooxanthellae might be enhanced under the RCP2.6 scenario due to
    elevated CO2 availability (CO2 fertilization effect), leading to an in-
    crease of chlorophyll a. The coral host, however, was rather stressed,
    resulting in higher respiration and decreased calcification.

    C. fastigiata is known for its strong allelopathy, but also very sensitive
    under experimental conditions. While the impacts of the algae on the
    coral were small, a temperature and CO2 increase of more than 1 ◦C and
    85 ppm respectively over longer periods, whether due to global warming
    or warm water periods (e.g. El Niño), might have significant impacts on
    coral-macroalgal interactions. Hence, further studies with less sensitive
    macroalgae are needed to investigate the likelihood of interaction shifts
    for P. lobata under future climatic regimes.

    Declaration of Competing Interest

    The authors declare that they have no known competing financial
    interests or personal relationships that could have appeared to influence
    the work reported in this paper.

    Acknowledgements

    We would like to thank the University of Queensland and the Leibniz
    Centre for Tropical Marine Research (ZMT) for funding. LR was sup-
    ported by the ZMT ‘Student Research Grant’. DB-C and OHG were
    supported by the Australian Research Council Laureate Fellowship
    program. Thanks to the research assistants at Heron Island Research
    Station for technical support.

    Appendix A. Supplementary data

    Supplementary data to this article can be found online at https://doi.
    org/10.1016/j.jembe.2020.151477.

    References

    Al-Moghrabi, S., Goiran, C., Allemand, D., Speziale, N., Jaubert, J., 1996. Inorganic
    carbon uptake for photosynthesis by the symbiotic coral-dinoflagellate association II.
    Mechanisms for bicarbonate uptake. J. Exp. Mar. Biol. Ecol. 199, 227–248. https://
    doi.org/10.1016/0022-0981(95)00202-2.

    Anthony, Kline, D.I., Diaz-Pulido, G., Dove, S., Hoegh-Guldberg, O., 2008. Ocean
    acidification causes bleaching and productivity loss in coral reef builders. Proc. Natl.
    Acad. Sci. 105, 17442–17446. https://doi.org/10.1073/pnas.0804478105.

    Anthony, K.R., Maynard, J.A., Diaz-Pulido, G., Mumby, P.J., Marshall, P.A., Cao, L.,
    Hoegh-Guldberg, O., 2011. Ocean acidification and warming will lower coral reef
    resilience. Glob. Chang. Biol. 17, 1798–1808.

    Anton, A., Randle, J.L., Garcia, F.C., Rossbach, S., Ellis, J.I., Weinzierl, M., Duarte, C.M.,
    2020. Differential thermal tolerance between algae and corals may trigger the
    proliferation of algae in coral reefs. Glob. Chang. Biol. 26, 4316–4327. https://doi.
    org/10.1111/gcb.15141.

    Arnon, D.I., 1949. Copper enzymes in isolated chloroplasts. Polyphenoloxidase in Beta
    vulgaris. Plant Physiol. 24 (1).

    Aronson, R.B., Precht, W.F., 2001. White-band disease and the changing face of
    Caribbean coral reefs. In: Porter, J.W. (Ed.), The Ecology and Etiology of Newly
    Emerging Marine Diseases. Springer Netherlands, Dordrecht, pp. 25–38. https://doi.
    org/10.1007/978-94-017-3284-0_2.

    Bender, D., Diaz-Pulido, G., Dove, S., 2012. Effects of macroalgae on corals recovering
    from disturbance. J. Exp. Mar. Biol. Ecol. 429, 15–19. https://doi.org/10.1016/j.
    jembe.2012.06.014.

    Birrell, C.L., McCook, L.J., Willis, B.L., Diaz-Pulido, G.A., 2008. Effects of benthic algae
    on the replenishment of corals and the implications for the resilience of coral reefs.
    Ocean. Mar. Biol. Annu. Rev. 46, 25–63.

    Bonaldo, R.M., Hay, M.E., 2014. Seaweed-coral interactions: variance in seaweed
    allelopathy, coral susceptibility, and potential effects on coral resilience. PLoS One 9,
    e85786.

    Budd, A.F., 1986. Neogene Paleontology in the Northern Dominican Republic: 3. The
    Family Poritidae (Anthozoa: Scleractinia).

    Brown, K.T., Bender-Champ, D., Kenyon, T.M., Rémond, C., Hoegh-Guldberg, O.,
    Dove, S., 2019. Temporal effects of ocean warming and acidification on coral–algal
    competition. Coral Reefs 38, 297–309. https://doi.org/10.1007/s00338-019-01775-
    y.

    Chadwick, N.E., Morrow, K.M., 2011. Competition among sessile organisms on coral
    reefs. In: Dubinsky, Z., Stambler, N. (Eds.), Coral Reefs: An Ecosystem in Transition.
    Springer Netherlands, Dordrecht, pp. 347–371. https://doi.org/10.1007/978-94-
    007-0114-4_20.

    Chisholm, J.R., Gattuso, J.-P., 1991. Validation of the alkalinity anomaly technique for
    investigating calcification of photosynthesis in coral reef communities. Limnol.
    Oceanogr. 36, 1232–1239.

    Clements, C.S., Burns, A.S., Stewart, F.J., Hay, M.E., 2020. Seaweed-coral competition in
    the field: effects on coral growth, photosynthesis and microbiomes require direct
    contact. Proc. Biol. Sci. 287, 20200366 https://doi.org/10.1098/rspb.2020.0366.

    Coyer, J.A., Ambrose, R.F., Engle, J.M., Carroll, J.C., 1993. Interactions between corals
    and algae on a temperate zone rocky reef: mediation by sea urchins. J. Exp. Mar.
    Biol. Ecol. 167, 21–37.

    Crawley, A., Kline, D.I., Dunn, S., Anthony, K., Dove, S., 2010. The effect of ocean
    acidification on symbiont photorespiration and productivity in Acropora formosa.
    Glob. Chang. Biol. 16, 851–863. https://doi.org/10.1111/j.1365-2486.2009.01943.
    x.

    Dayton, P.K., 1971. Competition, disturbance, and community organization: the
    provision and subsequent utilization of space in a rocky intertidal community. Ecol.
    Monogr. 41, 351–389. https://doi.org/10.2307/1948498.

    Del Monaco, C., Hay, M.E., Gartrell, P., Mumby, P.J., Diaz-Pulido, G., 2017. Effects of
    ocean acidification on the potency of macroalgal allelopathy to a common coral. Sci.
    Rep. 7, 41053 https://doi.org/10.1038/srep41053.

    Diaz-Pulido, G., Barrón, C., 2020. CO2 enrichment stimulates dissolved organic carbon
    release in coral reef macroalgae. J. Phycol. 56, 1039–1052. https://doi.org/
    10.1111/jpy.13002.

    Diaz-Pulido, G., Harii, S., McCook, L.J., Hoegh-Guldberg, O., 2010. The impact of benthic
    algae on the settlement of a reef-building coral. Coral Reefs 29, 203–208. https://
    doi.org/10.1007/s00338-009-0573-x.

    Diaz-Pulido, G., Gouezo, M., Tilbrook, B., Dove, S., Anthony, K.R.N., 2011. High CO2
    enhances the competitive strength of seaweeds over corals: coral-algal competition
    and high CO2. Ecol. Lett. 14, 156–162. https://doi.org/10.1111/j.1461-
    0248.2010.01565.x.

    Dickson, A.G., Afghan, J.D., Anderson, G.C., 2003. Reference materials for oceanic CO2
    analysis: a method for the certification of total alkalinity. Mar. Chem. 80, 185–197.

    Done, T.J., 1992. Phase shifts in coral reef communities and their ecological significance.
    Hydrobiologia 247, 121–132. https://doi.org/10.1007/BF00008211.

    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. https://doi.org/10.1146/annurev.
    marine.010908.163834.

    Dove, S.G., Kline, D.I., Pantos, O., Angly, F.E., Tyson, G.W., Hoegh-Guldberg, O., 2013.
    Future reef decalcification under a business-as-usual CO2 emission scenario. Proc.
    Natl. Acad. Sci. 110, 15342–15347. https://doi.org/10.1073/pnas.1302701110.

    Eakin, C.M., Liu, G., Gomez, A.M., De la Couri, J.L., Heron, S.F., Skirving, W.J., Geiger, E.
    F., Marsh, B.L., Tirak, K.V., Strong, A.E., 2018. Unprecedented three years of global
    coral bleaching 2014–17. Sidebar 3.1. [in State of the Climate in 2017]. Bull. Am.
    Meteorol. Soc. 99 (8), S74–S75.

    Edgar, G.J., Klumpp, D.W., 2003. Consistencies over regional scales in assemblages of
    mobile epifauna associated with natural and artificial plants of different shape.
    Aquat. Bot. 75, 275–291.

    Ferrier-Pagès, C., Witting, J., Tambutté, E., Sebens, K.P., 2003. Effect of natural
    zooplankton feeding on the tissue and skeletal growth of the scleractinian coral
    Stylophora pistillata. Coral Reefs 22, 229–240. https://doi.org/10.1007/s00338-
    003-0312-7.

    Foster, N., Box, S., Mumby, P., 2008. Competitive effects of macroalgae on the fecundity
    of the reef-building coral Montastraea annularis. Mar. Ecol. Prog. Ser. 367, 143–152.
    https://doi.org/10.3354/meps07594.

    L. Rölfer et al.

    https://doi.org/10.1016/j.jembe.2020.151477

    https://doi.org/10.1016/j.jembe.2020.151477

    https://doi.org/10.1016/0022-0981(95)00202-2

    https://doi.org/10.1016/0022-0981(95)00202-2

    https://doi.org/10.1073/pnas.0804478105

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0015

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0015

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0015

    https://doi.org/10.1111/gcb.15141

    https://doi.org/10.1111/gcb.15141

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0025

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0025

    https://doi.org/10.1007/978-94-017-3284-0_2

    https://doi.org/10.1007/978-94-017-3284-0_2

    https://doi.org/10.1016/j.jembe.2012.06.014

    https://doi.org/10.1016/j.jembe.2012.06.014

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0040

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0040

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0040

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0045

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0045

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0045

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0050

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0050

    https://doi.org/10.1007/s00338-019-01775-y

    https://doi.org/10.1007/s00338-019-01775-y

    https://doi.org/10.1007/978-94-007-0114-4_20

    https://doi.org/10.1007/978-94-007-0114-4_20

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0060

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0060

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0060

    https://doi.org/10.1098/rspb.2020.0366

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0070

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0070

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0070

    https://doi.org/10.1111/j.1365-2486.2009.01943.x

    https://doi.org/10.1111/j.1365-2486.2009.01943.x

    https://doi.org/10.2307/1948498

    https://doi.org/10.1038/srep41053

    https://doi.org/10.1111/jpy.13002

    https://doi.org/10.1111/jpy.13002

    https://doi.org/10.1007/s00338-009-0573-x

    https://doi.org/10.1007/s00338-009-0573-x

    https://doi.org/10.1111/j.1461-0248.2010.01565.x

    https://doi.org/10.1111/j.1461-0248.2010.01565.x

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0110

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0110

    https://doi.org/10.1007/BF00008211

    https://doi.org/10.1146/annurev.marine.010908.163834

    https://doi.org/10.1146/annurev.marine.010908.163834

    https://doi.org/10.1073/pnas.1302701110

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0130

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0130

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0130

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0130

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0135

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0135

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0135

    https://doi.org/10.1007/s00338-003-0312-7

    https://doi.org/10.1007/s00338-003-0312-7

    https://doi.org/10.3354/meps07594

    Journal of Experimental Marine Biology and Ecology 534 (2021) 151477

    9

    Gao, K., Aruga, Y., Asada, K., Ishihara, T., Akano, T., Kiyohara, M., 1991. Enhanced
    growth of the red algaPorphyra yezoensis Ueda in high CO2 concentrations. J. Appl.
    Phycol. 3, 355–362.

    Gao, K., Aruga, Y., Asada, K., Kiyohara, M., 1993. Influence of enhanced CO2 on growth
    and photosynthesis of the red algae Gracilaria sp. and G. chilensis. J. Appl. Phycol. 5,
    563–571. https://doi.org/10.1007/BF02184635.

    Gardner, T.A., 2003. Long-term region-wide declines in Caribbean corals. Science 301,
    958–960. https://doi.org/10.1126/science.1086050.

    Gattuso, J.-P., Frankignoulle, M., Wollast, R., 1998. Carbon and carbonate metabolism in
    coastal aquatic ecosystems. Annu. Rev. Ecol. Syst. 29, 405–434.

    Heil, C.A., Chaston, K., Jones, A., Bird, P., Longstaff, B., Costanzo, S., Dennison, W.C.,
    2004. Benthic microalgae in coral reef sediments of the southern great barrier reef,
    Australia. Coral Reefs 23, 336–343. https://doi.org/10.1007/s00338-004-0390-1.

    Hellebust, J.A., Craigie, J.S., 1978. Handbook of Phycological Methods Physiological and
    Biological Methods. Cambridge University Press Cambridge, London.

    Heron, S.F., Maynard, J.A., van Hooidonk, R., Eakin, C.M., 2016. Warming trends and
    bleaching stress of the World’s coral reefs 1985–2012. Sci. Rep. 6 https://doi.org/
    10.1038/srep38402.

    Hoegh-Guldberg, O., 1999. Climate change, coral bleaching and the future of the World’s
    coral reefs. Mar. Freshw. Res. 50, 839. https://doi.org/10.1071/MF99078.

    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., et al., 2007. Coral reefs under
    rapid climate change and ocean acidification. Science 318, 1737–1742.

    Huang, W.-J., Wang, Y., Cai, W., 2012. Assessment of sample storage techniques for total
    alkalinity and dissolved inorganic carbon in seawater. Limnol. Oceanogr. Methods
    10. https://doi.org/10.4319/lom.2012.10.711.

    Hughes, T.P., 1994. Catastrophes, phase shifts, and large-scale degradation of a
    Caribbean coral reef. Science 265, 1547–1551. https://doi.org/10.1126/
    science.265.5178.1547.

    Hughes, T.P., Reed, D.C., Boyle, M.-J., 1987. Herbivory on coral reefs: community
    structure following mass mortalities of sea urchins. J. Exp. Mar. Biol. Ecol. 113,
    39–59. https://doi.org/10.1016/0022-0981(87)90081-5.

    Hughes, T.P., Rodrigues, M.J., Bellwood, D.R., Ceccarelli, D., Hoegh-Guldberg, O.,
    McCook, L., Moltschaniwskyj, N., Pratchett, M.S., Steneck, R.S., Willis, B., 2007.
    Phase shifts, herbivory, and the resilience of coral reefs to climate change. Curr. Biol.
    17, 360–365. https://doi.org/10.1016/j.cub.2006.12.049.

    Hughes, T.P., Graham, N.A.J., Jackson, J.B.C., Mumby, P.J., Steneck, R.S., 2010. Rising
    to the challenge of sustaining coral reef resilience. Trends Ecol. Evol. 25, 633–642.
    https://doi.org/10.1016/j.tree.2010.07.011.

    Hughes, T.P., Kerry, J.T., Álvarez-Noriega, M., Álvarez-Romero, J.G., Anderson, K.D.,
    Baird, A.H., Babcock, R.C., Beger, M., Bellwood, D.R., Berkelmans, R., Bridge, T.C.,
    Butler, I.R., Byrne, M., Cantin, N.E., Comeau, S., Connolly, S.R., Cumming, G.S.,
    Dalton, S.J., Diaz-Pulido, G., Eakin, C.M., Figueira, W.F., Gilmour, J.P., Harrison, H.
    B., Heron, S.F., Hoey, A.S., Hobbs, J.-P.A., Hoogenboom, M.O., Kennedy, E.V.,
    Kuo, C., Lough, J.M., Lowe, R.J., Liu, G., McCulloch, M.T., Malcolm, H.A.,
    McWilliam, M.J., Pandolfi, J.M., Pears, R.J., Pratchett, M.S., Schoepf, V.,
    Simpson, T., Skirving, W.J., Sommer, B., Torda, G., Wachenfeld, D.R., Willis, B.L.,
    Wilson, S.K., 2017. Global warming and recurrent mass bleaching of corals. Nature
    543, 373–377. https://doi.org/10.1038/nature21707.

    IPCC, 2013. Climate Change 2013 – The Physical Science Basis, Climate Change 2013:
    The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment
    Report of the Intergovernmental Panel on Climate Change. Cambridge University
    Press, Cambridge, United Kingdom and New York, NY, USA. https://doi.org/
    10.1017/cbo9781107415324.

    Johannes, R.E., Wiebe, W.J., 1970. Method for determination of coral tissue biomass and
    composition. Limnol. Oceanogr. 15, 822–824.

    Jokiel, P.L., Maragos, J.E., 1978. Coral growth: Buoyant weight technique. In:
    Stoddart, D.R., Johannes, R.E. (Eds.), Coral Reefs: Research Methods, Monographs
    on Oceanographic Methodology. UNESCO, Paris, pp. 529–542.

    Jompa, J., McCook, L.J., 2003. Contrasting effects of turf algae on corals: massive Porites
    spp. are unaffected by mixed-species turfs, but killed by the red alga Anotrichium
    tenue. Mar. Ecol. Prog. Ser. 258, 79–86.

    Kleypas, J.A., Langdon, C., 2006. Coral reefs and changing seawater carbonate
    chemistry. In: Phinney, J.T., Hoegh-Guldberg, O., Kleypas, J., Skirving, W.,
    Strong, A. (Eds.), Coastal and Estuarine Studies. American Geophysical Union,
    Washington, D. C, pp. 73–110. https://doi.org/10.1029/61CE06.

    Kline, D.I., Teneva, L., Schneider, K., Miard, T., Chai, A., Marker, M., Headley, K.,
    Opdyke, B., Nash, M., Valetich, M., Caves, J.K., Russell, B.D., Connell, S.D.,
    Kirkwood, B.J., Brewer, P., Peltzer, E., Silverman, J., Caldeira, K., Dunbar, R.B.,
    Koseff, J.R., Monismith, S.G., Mitchell, B.G., Dove, S., Hoegh-Guldberg, O., 2012.
    A short-term in situ CO2 enrichment experiment on Heron Island (GBR). Sci. Rep. 2
    https://doi.org/10.1038/srep00413.

    Langdon, C., 2002. Review of experimental evidence for effects of CO2 on calcification of
    reef builders. In: Proc. 9th Int. Coral Reef Sym, pp. 1091–1098.

    Leggat, W., Badger, M.R., Yellowlees, D., 1999. Evidence for an inorganic carbon-
    concentrating mechanism in the symbiotic Dinoflagellate Symbiodinium sp. Plant
    Physiol. 121, 1247–1255. https://doi.org/10.1104/pp.121.4.1247.

    Longo, G.O., Hay, M.E., 2017. Seaweed allelopathy to corals: are active compounds on,
    or in, seaweeds? Coral Reefs 36, 247–253. https://doi.org/10.1007/s00338-016-
    1526-9.

    Lough, J.M., Barnes, D.J., Devereux, M.J., Tobin, B.J., Tobin, S., 1999. Variability in
    growth characteristics of massive Porites on the great barrier reef. CRC Reef Res.
    Cent. Tech. Rep. 95.

    Lough, J.M., Anderson, K.D., Hughes, T.P., 2018. Increasing thermal stress for tropical
    coral reefs: 1871–2017. Sci. Rep. 8, 6079. https://doi.org/10.1038/s41598-018-
    24530-9.

    Marubini, F., Ferrier-Pages, C., Cuif, J.-P., 2003. Suppression of skeletal growth in
    scleractinian corals by decreasing ambient carbonate-ion concentration: a cross-
    family comparison. Proc. R. Soc. Lond. B Biol. Sci. 270, 179–184.

    McCook, L.J., 1999. Macroalgae, nutrients and phase shifts on coral reefs: scientific
    issues and management consequences for the great barrier reef. Coral Reefs 18,
    357–367.

    McCook, L., 2001. Competition between corals and algal turfs along a gradient of
    terrestrial influence in the nearshore central great barrier reef. Coral Reefs 19,
    419–425. https://doi.org/10.1007/s003380000119.

    McCook, L., Jompa, J., Diaz-Pulido, G., 2001. Competition between corals and algae on
    coral reefs: a review of evidence and mechanisms. Coral Reefs 19, 400–417. https://
    doi.org/10.1007/s003380000129.

    Mumby, P.J., Steneck, R.S., Hastings, A., 2013. Evidence for and against the existence of
    alternate attractors on coral reefs. Oikos 122, 481–491. https://doi.org/10.1111/
    j.1600-0706.2012.00262.x.

    Norström, A., Nyström, M., Lokrantz, J., Folke, C., 2009. Alternative states on coral reefs:
    beyond coral–macroalgal phase shifts. Mar. Ecol. Prog. Ser. 376, 295–306. https://
    doi.org/10.3354/meps07815.

    Nugues, M.M., Smith, G.W., Hooidonk, R.J., Seabra, M.I., Bak, R.P.M., 2004. Algal
    contact as a trigger for coral disease: algal contact triggers coral disease. Ecol. Lett. 7,
    919–923. https://doi.org/10.1111/j.1461-0248.2004.00651.x.

    Ohde, S., Hossain, M.M.M., 2004. Effect of CaCO3 (aragonite) saturation state of
    seawater on calcification of Porites coral. Geochem. J. 38, 613–621.

    Pörtner, H.O., Roberts, D.C., Masson-Delmotte, V., Zhai, P., Tignor, M., Poloczanska, E.,
    Mintebeck, K., Nicolai, M., Okem, A., Petzold, J., Rama, B., Weyer, N., 2019. IPCC
    The Ocean and Cryosphere in a Changing Climate Summary for Policmakers. IPCC
    Spec. Rep. Ocean Cryosph. a Chang. Clim. SPM-1-SPM-42.

    R Core Team, 2015. R: A Language and Environment for Statistical Computing. R
    Foundation for Statistical Computing, Vienna, Austria.

    Rasher, D.B., Hay, M.E., 2010. Chemically rich seaweeds poison corals when not
    controlled by herbivores. Proc. Natl. Acad. Sci. 107, 9683–9688. https://doi.org/
    10.1073/pnas.0912095107.

    Rasher, D.B., Stout, E.P., Engel, S., Kubanek, J., Hay, M.E., 2011. Macroalgal terpenes
    function as allelopathic agents against reef corals. Proc. Natl. Acad. Sci. 108,
    17726–17731. https://doi.org/10.1073/pnas.1108628108.

    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 scleractinian coral. Glob. Chang. Biol. 9, 1660–1668. https://
    doi.org/10.1046/j.1365-2486.2003.00678.x.

    Rinkevich, B., Loya, Y., 1985. Intraspecific competition in a reef coral: effects on growth
    and reproduction. Oecologia 66, 100–105. https://doi.org/10.1007/BF00378559.

    Scheffer, M., Carpenter, S., Foley, J.A., Folke, C., Walker, B., 2001. Catastrophic shifts in
    ecosystems. Nature 413, 591–596. https://doi.org/10.1038/35098000.

    Shenkar, N., Bronstein, O., Loya, Y., 2008. Population dynamics of a coral reef ascidian
    in a deteriorating environment. Mar. Ecol. Prog. Ser. 367, 163–171. https://doi.org/
    10.3354/meps07579.

    Smith, J.E., Shaw, M., Edwards, R.A., Obura, D., Pantos, O., Sala, E., Sandin, S.A.,
    Smriga, S., Hatay, M., Rohwer, F.L., 2006. Indirect effects of algae on coral: Algae-
    mediated, microbe-induced coral mortality. Ecol. Lett. 9, 835–845. https://doi.org/
    10.1111/j.1461-0248.2006.00937.x.

    Stimson, J., Kinzie III, R.A., 1991. The temporal pattern and rate of release of
    zooxanthellae from the reef coral Pocillopora damicornis (Linnaeus) under nitrogen-
    enrichment and control conditions. J. Exp. Mar. Biol. Ecol. 153, 63–74.

    Underwood, A.J., 1997. Experiments in Ecology: Their Logical Design and Interpretation
    Using Analysis of Variance. Cambridge University Press.

    Vieira, C., Thomas, O.P., Culioli, G., Genta-Jouve, G., Houlbreque, F., Gaubert, J., De
    Clerck, O., Payri, C.E., 2016. Allelopathic interactions between the brown algal
    genus Lobophora (Dictyotales, Phaeophyceae) and scleractinian corals. Sci. Rep. 6,
    18637.

    Whitaker, J.R., Granum, P.E., 1980. An absolute method for protein determination based
    on difference in absorbance at 235 and 280 nm. Anal. Biochem. 109, 156–159.

    Wu, H., Zou, D., Gao, K., 2008. Impacts of increased atmospheric CO2 concentration on
    photosynthesis and growth of micro- and macro-algae. Sci China C Life Sci 51,
    1144–1150. https://doi.org/10.1007/s11427-008-0142-5.

    Zundelevich, A., Lazar, B., Ilan, M., 2007. Chemical versus mechanical bioerosion of
    coral reefs by boring sponges-lessons from Pione cf. vastifica. J. Exp. Biol. 210,
    91–96.

    L. Rölfer et al.

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0150

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0150

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0150

    https://doi.org/10.1007/BF02184635

    https://doi.org/10.1126/science.1086050

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0165

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0165

    https://doi.org/10.1007/s00338-004-0390-1

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0175

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0175

    https://doi.org/10.1038/srep38402

    https://doi.org/10.1038/srep38402

    https://doi.org/10.1071/MF99078

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0190

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0190

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0190

    https://doi.org/10.4319/lom.2012.10.711

    https://doi.org/10.1126/science.265.5178.1547

    https://doi.org/10.1126/science.265.5178.1547

    https://doi.org/10.1016/0022-0981(87)90081-5

    https://doi.org/10.1016/j.cub.2006.12.049

    https://doi.org/10.1016/j.tree.2010.07.011

    https://doi.org/10.1038/nature21707

    https://doi.org/10.1017/cbo9781107415324

    https://doi.org/10.1017/cbo9781107415324

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0225

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0225

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0230

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0230

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0230

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0235

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0235

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0235

    https://doi.org/10.1029/61CE06

    https://doi.org/10.1038/srep00413

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0250

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0250

    https://doi.org/10.1104/pp.121.4.1247

    https://doi.org/10.1007/s00338-016-1526-9

    https://doi.org/10.1007/s00338-016-1526-9

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0265

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0265

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0265

    https://doi.org/10.1038/s41598-018-24530-9

    https://doi.org/10.1038/s41598-018-24530-9

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0275

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0275

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0275

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0280

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0280

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0280

    https://doi.org/10.1007/s003380000119

    https://doi.org/10.1007/s003380000129

    https://doi.org/10.1007/s003380000129

    https://doi.org/10.1111/j.1600-0706.2012.00262.x

    https://doi.org/10.1111/j.1600-0706.2012.00262.x

    https://doi.org/10.3354/meps07815

    https://doi.org/10.3354/meps07815

    https://doi.org/10.1111/j.1461-0248.2004.00651.x

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0310

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0310

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0315

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0315

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0315

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0315

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0320

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0320

    https://doi.org/10.1073/pnas.0912095107

    https://doi.org/10.1073/pnas.0912095107

    https://doi.org/10.1073/pnas.1108628108

    https://doi.org/10.1046/j.1365-2486.2003.00678.x

    https://doi.org/10.1046/j.1365-2486.2003.00678.x

    https://doi.org/10.1007/BF00378559

    https://doi.org/10.1038/35098000

    https://doi.org/10.3354/meps07579

    https://doi.org/10.3354/meps07579

    https://doi.org/10.1111/j.1461-0248.2006.00937.x

    https://doi.org/10.1111/j.1461-0248.2006.00937.x

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0355

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0355

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0355

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0360

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0360

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0365

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0365

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0365

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0365

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0370

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0370

    https://doi.org/10.1007/s11427-008-0142-5

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0380

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0380

    http://refhub.elsevier.com/S0022-0981(20)30246-X/rf0380

      Coral-macroalgal competition under ocean warming and acidification
      1 Introduction
      2 Materials and methods
      2.1 Study site and collection of corals and macroalgae
      2.2 Experimental set-up, maintenance and monitoring
      2.3 Measurement of respiration and calcification rates
      2.4 Growth measurements
      2.5 Tissue analysis of corals
      2.6 Data analysis
      3 Results
      3.1 Corals
      3.2 Macroalgae
      4 Discussion
      4.1 Calcification under low rates of warming and acidification
      4.2 Photosynthesis and respiration: evidence of a CO2 fertilization effect?
      4.3 Physical impacts of competitors
      5 Conclusions
      Declaration of Competing Interest
      Acknowledgements
      Appendix A Supplementary data
      References

    Journal of Experimental Marine Biology and Ecology 535 (2021) 15148

    9

    Available online 13 November 2020
    0022-0981/© 2020 Elsevier B.V. All rights reserved.

    Irradiance, photosynthesis and elevated pCO2 effects on net calcification in
    tropical reef macroalgae

    C. McNicholl , M.S. Koch *

    Florida Atlantic University, Boca Raton, FL 33431, USA

    A R T I C L E I N F O

    Keywords:
    Coral reef
    Dissolution
    pH
    Climate change
    Ocean acidification

    A B S T R A C T

    Calcifying tropical macroalgae produce sediment, build three-dimensional habitats, and provide substrate for
    invertebrate larvae on reefs. Thus, lower calcification rates under declining pH and increasing ocean pCO2, or
    ocean acidification, is a concern. In the present study, calcification rates were examined experimentally under
    predicted end-of-the-century seawater pCO2 (1116 μatm) and pH (7.67) compared to ambient controls (pCO2
    409 μatm; pH 8.04). Nine reef macroalgae with diverse calcification locations, calcium carbonate structure,
    photophysiology, and site-specific irradiance were examined under light and dark conditions. Species included
    five from a high light patch reef on the Florida Keys Reef Tract (FKRT) and four species from low light reef walls
    on Little Cayman Island (LCI). Experiments on FKRT and LCI species were conducted at 500 and 50 μmol photons
    m− 2 s− 1 in situ irradiance, respectively. Calcification rates independent of photosystem-II (PSII) were also
    investigated for FKRT species. The most consistent negative effect of elevated pCO2 on calcification rates in the
    tropical macroalgae examined occurred in the dark. Most species (89%) had net calcification rates of zero or net
    dissolution in the dark at low pH. Species from the FKRT that sustained positive net calcification rates in the light
    at low pH also maintained ~30% of their net calcification rates without PSII at ambient pH. However, calcifi-
    cation rates in the light independent of PSII were not sustained at low pH. Regardless of these low pH effects,
    most FKRT species daily net calcification rates, integrating light/dark rates over a 24h period, were not signif-
    icantly different between low and ambient pH. This was due to a 10-fold lower dark, compared to light, calci-
    fication rate, and a strong correspondence between calcification and photosynthetic rates. Interestingly, low-light
    species sustained calcification rates on par with high-light species without high rates of photosynthesis. Low-light
    species’ morphology and physiology that promote high calcification rates at ambient pH, may increase their
    vulnerability to low pH. Our data indicate that the negative effect of elevated pCO2 and low pH on tropical
    macroalgae at the organismal level is their impact on dark net calcification, probably enhanced dissolution.
    However, elevated pCO2 and low pH effects on macroalgae daily calcification rates are greatest in species with
    lower net calcification rates in the light. Thus, macroalgae able to maintain high calcification rates in the light
    (high and low irradiance) at low pH, and/or sustain strong biotic control with high [H+] in the bulk seawater, are
    expected to dominate under global change.

    1. Introduction

    Marine calcifier persistence and sustained calcification rates remain
    uncertain under future predictions of ocean acidification. Since the in-
    dustrial revolution, global ocean pH has decreased 0.1 pH units, and a
    further 0.3–0.4 reduction is predicted to occur by the year 2100 due to
    anthropogenic CO2 emissions (Gehlen et al. 2014; Hartin et al. 2016). A
    global decrease in ocean pH affects calcification rates in marine organ-
    isms, such as coral, shellfish, phytoplankton, and ecologically important

    macroalgae (Andersson et al. 2009; Fabry et al. 2008; Hoegh-Guldberg
    et al. 2007; Koch et al. 2013; Orr et al. 2005; Ries et al. 2009). Dimin-
    ished calcification rates of tropical reef macroalgae is a concern because
    of their ecological role in carbonate sediment production, building of 3-
    dimensional reef habitat structure, and providing substrate for inverte-
    brate larval settlement (Adey 1998; Nelson 2009). Many studies have
    examined the effects of elevated partial pressure of CO2 (pCO2) that
    lowers seawater pH on macroalgal calcification, but often with con-
    flicting results (Hofmann et al. 2014; Koch et al. 2013; Nelson 2009;

    * Corresponding author.
    E-mail addresses: cmcnicholl2015@fau.edu (C. McNicholl), mkoch@fau.edu (M.S. Koch).

    Contents lists available at ScienceDirect

    Journal of Experimental Marine Biology and Ecology

    journal homepage: www.elsevier.com/locate/jembe

    https://doi.org/10.1016/j.jembe.2020.151489
    Received 19 November 2019; Received in revised form 22 October 2020; Accepted 2 November 2020

    mailto:cmcnicholl2015@fau.edu

    mailto:mkoch@fau.edu

    www.sciencedirect.com/science/journal/00220981

    https://www.elsevier.com/locate/jembe

    https://doi.org/10.1016/j.jembe.2020.151489

    https://doi.org/10.1016/j.jembe.2020.151489

    https://doi.org/10.1016/j.jembe.2020.151489

    http://crossmark.crossref.org/dialog/?doi=10.1016/j.jembe.2020.151489&domain=pdf

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    2

    Porzio et al. 2011). Discrepancies in the literature may depend on
    species-specific calcification mechanisms, photophysiology, location of
    calcification site, and calcium carbonate (CaCO3) crystal form and
    mineral content (reviewed in Basso 2012; Hofmann and Bischof 2014;
    Koch et al. 2013).

    In marine macroalgae, CaCO3 precipitation occurs in areas that are
    isolated or semi-isolated from bulk seawater where the saturation state
    (ΩCaCO3) can be elevated to promote calcification (Borowitzka and
    Larkum 1987). Calcification typically occurs in the cell walls of Rho-
    dophyta (red macroalgae), such as crustose coralline algae (CCA), and
    other Rhodophyta families (e.g., Peyssonneliacea) (Adey et al. 2013;
    Basso 2012). In the Chlorophyta (green macroalgae), calcification oc-
    curs in sheaths surounding filaments or specific compartments con-
    nected to external seawater by diffusive channels (Borowitzka and
    Larkum 1987, 1976). CCA are thought to be the most sensitive to
    declining pH and elevated pCO2 due to the proximity of their calcifying
    sites to overlying bulk seawater and the high magnesium-calcite content
    of their crystals. Magnesium concentrations in CaCO3 lattice have a
    positive relationship with temperature and higher concentrations of Mg
    result in a relatively more soluble polymorph of CaCO3 (Kamenos et al.
    2009; Kamenos and Law 2010; Mccoy and Kamenos 2015). A number of
    studies have shown negative effects of elevated pCO2 and low pH on
    CCA calcification (Anthony et al. 2008; Basso 2012; Comeau et al. 2019;
    Diaz-Pulido et al. 2014; Gao et al. 1993; Kato et al. 2014; Noisette et al.
    2013); however, other studies imply CCA resistance (Comeau et al.
    2018, 2017, 2013; Cornwall et al. 2017; Dutra et al. 2015; Ries et al.
    2009). Chlorophytes, and some rhodophytes, have an aragonite poly-
    morph of CaCO3, which is less soluble than the high magnesium poly-
    morph found in CCA (Borowitzka and Larkum 1987). Precipitating a less
    soluble CaCO3 polymorph within semi-isolated compartments may be
    advantageous to resist elevated pCO2 and low pH (Comeau et al. 2013;
    Peach et al. 2017b, 2016; Ries 2011; Vogel et al. 2015a), yet there are a
    number of studies showing lower net calcification under declining pH
    and elevated pCO2 conditions (Meyer et al. 2016; Price et al. 2011). In
    addition to the potential effects of morphology and polymorphs,
    photosynthesis affects calcification in marine macroalgae (Diaz-Pulido
    et al. 2007; Hofmann and Bischof 2014; Koch et al. 2013; Porzio et al.
    2011; Raven and Hurd 2012).

    Photosynthesis has been shown to increase calcification in marine
    macroalgae (Borowitzka and Larkum 1987; De Beer and Larkum 2001;
    Gao et al. 1993; Koch et al. 2013; Pentecost 1978; Semesi et al. 2009;
    Wizemann et al. 2014), but the influence of increasing pCO2 and [H

    +] on
    the coupling of these two processes is only recently being disentangled
    (Brown et al. 2019; Comeau et al. 2018; Hofmann et al. 2016; McNicholl
    et al. 2020, 2019). A majority of marine macroalgae use carbon
    concentrating mechanisms (CCMs) to saturate RuBisCO with CO2 for
    photosynthesis (Raven and Hurd 2012). In the process of HCO

    3

    − uptake,
    HCO3

    − dehydrogenation to CO2 and OH
    − , catalyzed by external carbonic

    anhydrase (CAext), can neutralize H
    + and raise the macroalgal surface

    pH. Immediate (seconds) light-triggered pH increase in macro-and
    micro-algal surfaces detected with microsensors, combined with
    photosynthetic inhibitors, provides evidence that photosynthesis and
    light are major drivers of pH control at the seawater-cell surface inter-
    face (Chrachri et al. 2018; Cornwall et al. 2015, 2013; De Beer and
    Larkum 2001; Hofmann et al. 2016; McNicholl et al. 2019). Presence
    and maintenance of a high thalli surface pH may support calcification
    under declining pH and elevated pCO2 (Cornwall et al. 2014; Hofmann
    et al. 2016; McNicholl et al. 2019). In addition to photosynthesis, light-
    triggered H+ transport pumps independent of photosystem II (PSII) have
    been identified in several species of macroalgae and may facilitate
    calcification (De Beer and Larkum 2001; Hofmann et al. 2016; McNi-
    choll et al. 2019). Electron microscopy of epithallial cells show in-
    vaginations that have been postulated to promote proton pumping in
    coralline algae during decalcification/recalcification to support thalli
    growth (Pueschel et al., 2005). Thus, photosynthesis and active proton
    pumping will likely play an important role for continued calcification

    (or dissolution) processes.
    While calcification continues to occur in some marine macroalgae in

    the dark, rates are typically reduced or become net negative (Chisholm
    2000; El Haïkali et al. 2004; Vogel et al. 2015b), and may rely on
    accumulated energy stored during periods of irradiance (Mccoy and
    Kamenos 2015). Dark dissolution is likely driven by lower pH at calci-
    fication sites (Borowitzka and Larkum 1976; Wizemann et al. 2014).
    Amplification of this effect may occur with lower seawater carbonate ion
    concentrations [CO3

    2− ] and lower ΩCaCO3, as well as increased seawater
    [H+] and pCO2 in the bulk seawater (Comeau et al. 2012; McNicholl
    et al. 2020). A buildup of external CO2 or H

    + in the dark may prevent
    removing H+ from the calcifying space against a higher [H+] concen-
    tration in seawater (Cyronak et al. 2015; Jokiel 2011), although some
    biotic control and dissolution itself can buffer pH of the bulk seawater at
    the surface in the dark. Internal cellular acid-base regulation may also
    become difficult in the dark when the electrochemical gradient proton
    motive force reverses from passive diffusive efflux of H+ (pHsw:pHcell
    8.2:7.2) to requiring active H+ transport (pHsw:pHcell 7.8:7.2), as
    exemplified in calcifying phytoplankton with proton channels (Taylor
    et al. 2012). Higher energetic demands for proton pumps or lack of H+

    regulation could shift net calcification to net dissolution at night even
    for those groups, such as the CCA, with a known potential to biotically
    control calcification/decalcification as part of their life history (Pueschel
    et al., 2005).

    The objective of this study was to determine the effects of a 0.4 pH
    decline (pH 7.7) from the current ambient pH (8.1), as predicted for
    2100 (Gehlen et al. 2014; Hartin et al. 2016), on net calcification rates of
    nine tropical reef macroalgae. Species examined were characterized by
    diverse calcification locations, CaCO3 content, polymorph, CCMs, and
    site-specific irradiance levels. This included five species from a high-
    light patch reef on the Florida Keys Reef Tract (FKRT) and four spe-
    cies from under reef wall ledges with low light on Little Cayman Island
    (LCI). Calcification rates independent of photosystem-II (PSII) were also
    investigated for FKRT species. All calcification experiments were con-
    ducted in the light and dark. We hypothesized that algae from high-light
    patch reefs would continue to calcify at elevated pCO2 due to their high
    potential to raise surface pH through photosynthesis (McNicholl et al.
    2019) and calcifying space pH (Cornwall et al. 2017). Low-light species
    were predicted to have low photosynthetic rates; therefore, calcification
    rates would be reduced at elevated pCO2. Further, we postulated that
    elevated pCO2 would not lower net calcification rates in FKRT species
    shown to exhibit light-triggered thalli H+ regulation independent of PSII
    based on microsensor studies (McNicholl et al. 2019). We proposed that
    net calcification rates would be lower in the dark and this decline would
    be amplified at low pH.

    2. Materials & methods

    2.1. Calcified macroalgae collection and site parameters

    2.1.1. Florida keys reef tract (high light environment)
    Intact individuals of five dominant calcifying algae, including 3

    Rhodophyta (Neogoniolithon strictum, Jania adhaerens, CCA) and 2
    Chlorophyta (Halimeda scabra, Udotea luna) were collected (between
    March – June 2018) from a shallow Looe Key patch reef (~3–5 m)
    (24.62055◦ N, 81.37078◦ W) on the FKRT (Fig. 1a, Fig. 2a-e). Neo-
    goniolithon strictum and J. adhaerens were collected by hand from the
    benthos, while H. scabra and U. luna were collected by inserting a dive
    knife into the surrounding carbonate sediment and lifting out the thalli
    with rhizoids intact. CCA-covered carbonate rubble (Fig. 2c) was also
    collected.

    Field light levels were measured during collections at midday
    (~11:00–13:00) just above the benthos with a 4π spherical PAR quan-
    tum sensor (LI-193, LI-COR Inc.). The shallow patch reef site had high-
    irradiance (~800 μmol photo m− 2 s− 1) based on average midday mea-
    surements during collections. Experiments with high-light species from

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489
    3

    the FKRT were conducted under a full-spectrum LED light (Kessil,
    A360W E-Series Tuna Sun) at 500 μmol photons m− 2 s− 1 (LI-COR
    Quantum 4 π Light Sensor). While this was lower than mid-day field
    irradiance at the benthos, 500 μmol photons m− 2 s− 1 PAR approximated
    light levels that saturated photosynthesis in these species across a range
    of pH (Zweng et al. 2018).

    Seawater in situ chemistry was measured or calculated for each
    macroalgal collection (n = 3) (Table 1a). Site pH (Orion A211,
    8302BNUMD; calibrated with NBS standards, Thermo Fisher Scienti-
    fic®), conductivity (salinity = 35.5) and temperature (28.5 ± 1.1 ◦C)
    (YSI 650 MDS) were determined in the field. Water samples (n = 3;
    60 mL) were collected and total alkalinity determined within 48 h or
    fixed with HgCl2 (0.02%) and measured within eight weeks. For clarity
    and continuity, measurements for pH are reported on the NBS scale since
    correction to total scale using the TRIS buffer was not available for ex-
    periments conducted on LCI. Total alkalinity (TA) was measured by
    open-cell titration (Metrohm Titrando® 888) with 0.01 N HCl. Certified
    reference material (CRM Batch #156; Dickson Lab, Scripps Institute of

    Oceanography) was also run for each batch of TA samples. The CRM
    offset was used to correct TA readings from each batch. This way the
    data had no systematic bias from the true value. A certified standard
    TRIS buffer was used to calibrate pH for total alkalinity analysis (TRIS
    buffer, Dickson Lab, Scripps Institute of Oceanography: pH = 8.21,
    mV = − 64.9, 25 ◦C and 35 salinity). TA measurements were performed
    in triplicate unless the initial two measurements were within
    ±5 μmol kg− 1 of each other. Total alkalinity, temperature, conductivity,
    and pHNBS data were used to calculate DIC speciation (CO2SYS, Pierrot
    et al. 2006 with; K1, K2 from Mehrbach et al., 1973 refit by Dickson and
    Millero, 1987) for each experiment.

    Following collections, algal samples were immediately transported
    (4 h) in aerated coolers to Florida Atlantic University (FAU) in Boca
    Raton, FL. Samples were separated by species and acclimated for 48 h in
    9 L aquaria held within a large mesocosm tank (~500 L) over which 2
    (100 cm length x 10 cm width) full-irradiance spectrum fixtures
    (BuildMyLED Inc.) were hung. To provide all 9 aquaria with a similar
    light field, the fixtures were hung ~0.5 m above the tank and provided

    Fig. 1. Collection sites at (a, top panel) a shallow (3–5 m) high irradiance patch reef of the Florida Keys Reef Tract (Looe Key Reef; 24.62055◦ N, 81.37078◦ W) and
    (b, middle panel) a deeper (20 m) low irradiance wall reef on Little Cayman Island (Rockbottom Reef; 19.7025◦N, 80.05694◦ W). The (c) experimental setup for total
    alkalinity anomaly incubations where macroalgae were raised above the bottom of the beakers on a perforated disk (see insert) to allow for continued, slow stirring
    below the macroalga. Irradiance on the reefs was measured using a 4 π sensor; on the deep wall reefs (d) an underwater data logger was deployed.

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    4

    Fig. 2. Species from the (a-e) shallow (3–5 m) high
    irradiance patch reef of the Florida Keys Reef Tract and
    (f-i) deeper (20 m) low irradiance wall reef on Little
    Cayman Island used to examine the effects of elevated
    pCO2 and low pH predicted for 2100 on net calcifica-
    tion rates in light and dark experiments. Species from
    the FKRT (a) Neogoniolithon strictum, (b) Jania adhae-
    rens, (c) CCA, (d) Halimeda scabra, (e) Udotea luna and
    LCI Peyssonneliaceae (f) Peypink, (g) Peyred, (h) Hal-
    imeda copiosa and (i) Halimeda goreauii.

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    5

    ~250 μmol photons m− 2 s− 1 to each aquarium on a 12:12 light:dark
    cycle. All aquaria were semi-immersed (~80%) in the mesocosm tank
    with water maintained at 29 ◦C, the average temperature for the FKRT
    during summer (Kuffner et al. 2015). Water in the aquaria was aerated
    and circulated via small submersible pumps and replenished (75%)
    every 2 d with seawater from the flow-through seawater system at the
    FAU marine lab (coastal Atlantic Ocean). Experiments were completed
    within 5 d of collections.

    2.1.2. Little Cayman Island reef (low light environment)
    Calcified macroalgae were collected from a low-light reef wall

    (Fig. 1b) on LCI and experiments conducted at the Little Cayman
    Research Center (LCRC) (July 2018). Four species, including 2 Rhodo-
    phyta prostrate lobed crustose algae (Family Peyssonneliaceae; Peyred
    and Peypink) and 2 Chlorophyta (Halimeda goreauii and Halimeda copiosa)
    were collected under ledges and crevices along the upper reaches of a
    reef wall (Rockbottom ~20 m; 19.7025◦N, 80.05694◦ W; Fig. 2 f-i). The
    Halimeda species hung under ledges (Fig. 1b, Fig. 2h-i) and were
    collected from the holdfast removing the loosely attached filaments.
    Peyssonneliaceae samples (Fig. 2 f-g) were carefully removed with a
    small chisel (20 cm) where the crustose lobes were attached to the
    substrate. While macroalgae were being collected, light was measured
    (LI-COR Quantum 4 π Light Sensor) in situ under ledges (Fig. 1d), col-
    lecting data during midday (~11:00–13:00) for ~5 min with s− 1 interval
    (RBRsolo3 single channel data logger). The range of light under ledges at
    the collection site was ~5 to 50 μmol photons m− 2 s− 1. Macroalgae were
    kept in aerated seawater from the reef at low light and immediately
    transportation to LCRC (< 2 h). At LCRC, algae were separated into 4 L aquaria, kept aerated in the shade (~ 50 μmol photons m− 2 s− 1) with natural sunlight and experiments run within 24 h. Seawater was collected and immediately analyzed for pH and salinity upon returning to the lab, and samples were fixed with HgCl2 (0.02%) for total alkalinity analysis within 8 weeks.

    2.2. Elevated pCO2 experiments

    Net calcification rates were determined for macroalgae in seawater

    adjusted to pH (7.7) predicted for 2100 (Gehlen et al. 2014; Hartin et al.
    2016) and ambient controls (8.1). Final pH and pCO2 treatment levels
    attained (see results) were determined from the initial and end mea-
    surement of each light (between 0800 and 1500) and dark (between
    1900 and 0300) experiment. Calcification rates of FKRT and LCI species
    were determined using the TAA technique. The TAA technique for
    determining calcification rates in marine calcifiers is based on the
    changes in seawater TA over time. The TAA target was 3–10 times the
    accuracy of the method (~10 μmol kg− 1) and within 10% of TA (Lang-
    don et al. 2010). Based on this protocol, experimental incubation time
    for high-light species from the FKRT (n = 3–5) was 1–3 h in the light,
    depending on species, and 4 h in the dark. Incubation time for low-light
    species from LCI (n = 4) was 4.5 h in the light and 5.5 h in the dark.
    Blank runs with seawater were also conducted. Experiments were con-
    ducted in glass beakers (150–250 mL) covered with parafilm and
    secured with rubber bands to reduce atmospheric gas exchange (Fig. 1c)
    according to Chisholm and Gattuso (1991). Individual thalli were sus-
    pended approximately 2 cm above the bottom of the beaker on a
    perforated disk with a stirbar underneath and flow created with a stir-
    plate (Fig. 1c insert).

    Experimental seawater was filtered (0.45 μM) and brought to tem-
    perature (29 ◦C) in a waterbath before assigning treatment. Low pH
    treatment was obtained by bubbling seawater with pure CO2 prior to
    incubations. Initial and post-incubation pH, O2 (Orion A329), and
    temperature (Orion A211, 8302BNUMD) were recorded. Algae were
    acclimated to experimental seawater for 15 min. A relatively short
    acclimation time was needed to keep carbonate chemistry and △pH to a
    minimum before incubations. Low-light experiments with LCI species
    were conducted using the same experimental setup and protocols with
    the exception that these species were incubated at 50 μmol photons
    m− 2 s− 1.

    Ambient and low pH treatment runs were randomized so ~50% of
    the runs were low pH treatments, followed by ambient controls, and the
    inverse for the other 50%. In this experiment, repeated measures were
    used (i.e., treatments were applied sequentially to the same individual
    thalli). When similar TAA experiments were run during 45Ca experi-
    ments (McNicholl et al. 2020), similar results were found for TAA in

    Table 1
    Summary of (a) average seawater carbonate chemistry during incubations averaged across experiments in light/dark and ambient (A) or low (L) pH. In situ seawater
    conditions (n = 3) from collection sites Florida Keys Reef Tract (FKRT) and Little Cayman Island (LCI) shown. Carbonate chemistry parameters were calculated using
    CO2SYS (Pierrot et al. 2006) applying experimental seawater temperature (29 ◦C) and salinity (35.5), and (b) change (△) in seawater pH, total alkalinity (TA) and
    oxygen (O2) during incubations in ambient and low pH experiments in the light and dark averaged across all experiments (n = 38). Means ± SD. Details of carbonate
    chemistry and △pH, △TA and △O2 are presented for all experiments by species and treatments in supplemental tables (Table S1, S2).
    Summary of (a) average seawater carbonate chemistry during incubations averaged across experiments in light/dark and ambient (A) or low (L) pH. In situ seawater
    conditions (n = 3) from collection sites on the Florida Keys Reef Tract (FKRT) and Little Cayman Island (LCI) reefs are shown. Carbonate chemistry parameters were
    calculated using CO2SYS (Pierrot et al. 2006) applying experimental seawater temperature (29 ◦C) and salinity (35.5).The (b) change (△) in seawater pH, total
    alkalinity (TA) and oxygen (O2) in the light and dark are averaged across all experiments (n = 38). Means ± SD. Details of carbonate chemistry and △pH, △TA and
    △O2 are presented for all experiments by species and treatments in supplemental tables (Table S1, S2).

    (a) pH pCO2
    (μatm)

    HCO3

    (μmol kg− 1)
    CO3

    2−

    (μmol kg− 1)
    TA

    (μmol kg− 1)
    ΩCa ΩArag

    Experimental
    Light – A

    Light – L
    8.07 ± 0.06
    7.71 ± 0.05

    367 ± 64
    992 ± 140

    1665 ± 76
    1981 ± 58

    248 ± 23
    130 ± 12

    2283 ± 29
    2302 ± 39

    6.0 ± 0.6
    3.1 ± 0.3

    4.0 ± 0.4
    2.1 ± 0.2

    Dark – A
    Dark – L

    8.00 ± 0.03
    7.62 ± 0.03

    451 ± 34
    1239 ± 112

    1760 ± 31
    2062 ± 29

    222 ± 11
    111 ± 8

    2305 ± 31
    2329 ± 31

    5.4 ± 0.3
    2.7 ± 0.2

    3.6 ± 0.2
    1.8 ± 0.1

    In Situ
    FKRT

    LCI
    8.10 ± 0.02
    8.05 ± 0.03

    343 ± 11
    402 ± 37

    1662 ± 17
    1752 ± 39

    259 ± 4
    243 ± 13

    2292 ± 31
    2348 ± 24

    6.2 ± 0.1
    5.9 ± 0.3

    4.2 ± 0.1
    3.9 ± 0.2

    (b) Ambient pH Experiments Low pH Experiments

    △ pH
    △ TA

    (μmol kg− 1)
    △ O2

    (mg L− 1)
    △ pH

    △ TA
    (μmol kg− 1)

    △ O2
    (mg L− 1)

    Experimental
    Light <0.01 ± 0.09 − 70 ± 41 1.60 ± 0.93 0.09 ± 0.09 − 43 ± 50 1.82 ± 0.90 Dark − 0.11 ± 0.03 − 17 ± 12 − 0.42 ± 0.33 − 0.08 ± 0.03 22 ± 16 − 0.51 ± 0.31

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    6

    FKRT species using completely independent thalli for each run.

    2.3. PSII inhibitor experiments

    To examine light-dark PSII-independent calcification rates, an in-
    hibitor, 3-(3,4-Dichlorophenyl)-1,1-dimethylurea (DCMU), was added
    to the seawater according to Hofmann et al. (2016), after De Beer and
    Larkum (2001) and Borowitzka and Larkum (1976). DCMU was also
    highly effective at arresting photosynthesis in microsensor studies,
    eliminating O2 flux from the diffusive boundary layer (100 μm) at the
    thalli surface (McNicholl et al. 2019). A stock solution of 0.05 M DCMU
    was added to reach a concentration of 4 μM. Algae were left in DCMU-
    amended seawater until no change in O2 production was detected
    upon illumination (~30 min). Incubations were conducted (as above) at
    both pH levels and light/dark conditions with DCMU. DCMU experi-
    ments were only performed on species from the FKRT species due to time
    constraints on LCI.

    2.4. Calculation of net calcification and chemistry analysis

    Net calcification (Gnet) rates in μmol CaCO3 g dwt− 1 h− 1 were
    calculated from TA changes (△TA) during the experiments based on the
    following equation (Eq. 1):

    Gnet = − 0.5ρw
    ∆TA*v
    Wa*t

    (1)

    where: ρw is seawater density (kg L− 1) and ∆TA (μmol kg− 1) is the final
    TA minus the initial TA, v is the volume of seawater (L), Wa is the algal
    dry weight (g) and t is the incubation time in hours. Calcification rates
    were normalized to dry weight (60 ◦C) for FKRT species with the
    exception of CCA that had a complex 3-D structure. CCA was normalized
    to g of a flexible 2-D surface (foil) according to Marsh (1970). Calcifi-
    cation rates of Halimeda spp. from LCI were normalized to dry weight as
    above and to cm− 2 of thalli surface for LCI Peyssonneliaceae. Daily
    calcification rates were calculated by combining light (LGnet) and dark
    (DGnet) net calcification rates using a 12:12 light:dark cycle, as a first-
    order estimate of rates over 24 h (Eq. 2).

    Daily Gnet = (LGnet*12) + (DGnet*12) (2)

    2.5. Photosynthesis and respiration

    Photosynthesis and respiration rates were determined from initial
    and final dissolved O2 measurements (optical probe) using data from
    incubation runs in the light and dark, respectively. The O2 flux rates
    were normalized to dry mass (g) or surface area, as described above, and
    adjusted to seawater volume and time.

    2.6. Statistical analysis

    A repeated measures two-way ANOVA was performed (SigmaPlot
    v13.0, Systat Software Inc.) to compare calcification rates across light/
    dark at ambient and low pH treatments for experiments with and
    without DCMU. The assumptions of normality and homogeneity of
    variance were tested using the Shapiro-Wilkes and Brown-Forsythe tests,
    respectively. The assumption of sphericity in repeated measures was
    tested using the Mauchly’s test. Differences amongst means were
    established using the Holm-Sidak post-hoc test. The effect of pH treat-
    ments on the calculated daily calcification rates were determined using a
    t-test. Regression analysis was used to establish the relationship between
    net photosynthesis and calcification (SigmaPlot v13.0, Systat Software
    Inc.). Significance levels were established at p < 0.05 unless otherwise stated.

    3. Results

    3.1. Carbonate chemistry and △pH, △TA and △O2

    Carbonate chemistry, treatment and in situ pH and pCO2, △pH,
    TA and △O2 are summarized in Table 1. All data by species and

    treatments are presented in supplemental tables (Tables S1, S2). The
    resulting average pH and pCO2 across experiments and treatments was
    7.67 and 1116 μatm, respectively (Table 1). The average ambient pH
    and pCO2 for controls were 8.04 and 409 μatm, respectively (Table 1).
    The averages of the initial and end pCO2 during all experiments was
    approximately 3-fold higher in the elevated pCO2 treatments compared
    to controls, resulting in a pH of 7.71 for the low pH treatment compared
    to 8.07 for ambient pH (Table 1a). The dark experiments were ~ 0.08 pH
    lower than in the light due to metabolic differences, but the values were
    within the pH variance found in the light experiments (Table 1a). The
    concentrations of CO3

    2− and saturation state of CaCO3 declined ~50% in
    the high pCO2 treatments but remained above CaCO3 saturation (Ω > 1)
    for all experiments (Table 1a). The △pH and △O2 across experiments
    indicate only modest changes in chemistry occurred throughout the
    experiments (Table 1b). The △pH was on average ± 0.03 to ±0.09
    showing that our treatments were close to those applied at the initiation
    of the experiment. This is also indicated by a small change in metabolic
    O2, which was similar in the light for both high and low pH treatments
    (Table 1b). Average TA changes in the light were at least 3 times our
    accuracy for CRM TA (± 10 μmol kg− 1) and remained significantly lower
    than 10% of seawater TA. Further, the TA data for all the experiments
    highlight our general result that in the light △ TA was negative (indi-
    cating net calcification occurred) (Table 1b).

    3.2. High light species (FKRT) calcification

    Greater net calcification rates were found in the light compared to
    the dark for all species (Fig. 3) based on 2-way ANOVA main effects. Net
    calcification rates were similar at low and ambient pH for N. strictum
    (Holm-Sidak, t = 1.38 p = 0.214) and J. adhaerens (Holm-Sidak, t = 1.99
    p = 0.085) in the light (Fig. 3a,b). Halimeda scabra net calcification rates
    increased 16% at low pH in the light (Fig. 3d; Holm-Sidak, t = 2.54,
    p = 0.035), indicating a positive response to elevated pCO2. In contrast,
    U. luna (Holm-Sidak, t = 3.07 p = 0.018) and CCA (Holm-Sidak, t = 4.56
    p = 0.028) net calcification rates significantly declined at low pH rela-
    tive to controls in the light with CCA eliciting the strongest negative
    effect (Fig. 3c,e).

    Only N. strictum and J. adhaerens showed significant positive net
    calcification rates in the dark and only at ambient pH (Fig. 3a,b). Dark
    calcification rates for these two species were only a tenth the rates
    observed in the light. Even at ambient pH, U. luna exhibited net disso-
    lution in the dark (Fig. 3c), while H. scabra and CCA had net calcification
    rates approximating zero (Fig. 3d,e). Neogoniolithon strictum,
    J. adhaerens, and H. scabra (Holm-Sidak, t = 2.73 p = 0.032, t = 2.34
    p = 0.050, t = 2.71 p = 0.027, respectively) had significantly lower net
    calcification rates in the dark at low pH compared to ambient pH
    (Fig. 3a,b,d). CCA had 87% less net calcification on average at low pH
    compared to ambient pH in the dark, although not significant (Holm-
    Sidak, t = 1.59 p = 0.226).

    The two species that were negatively affected by low pH in the light,
    U. luna and CCA, had lower daily calcification rates at low pH relative to
    ambient controls (Fig. 3c,e; t-test, t = 2.65 p = 0.038, t = 5.27 p = 0.006,
    respectively). In contrast, the three species with no significant negative
    low pH effect on net calcification in the light, N. strictum (t-test, t = 1.21
    p = 0.262), J. adhaerens (t-test, t = 1.16 p = 0.278), and H. scabra (t-test,
    t = 0.03 p = 0.977), had similar daily calcification rates in low and
    ambient pH treatments (Fig. 3a,b,d). Halimeda scabra’s significantly
    lower net calcification rate in the dark at low pH was offset by increased
    calcification rates in the light at low pH, resulting in no difference in
    daily calcification rates (Fig. 3d).

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    7

    3.2.1. Photosynthesis and calcification
    Net photosynthesis and respiration were not different between the

    low and ambient pH for any species from the FKRT or LCI (Table 2).
    Species from the FKRT that exhibited a broad range in photosynthetic
    rates amongst individuals exhibited a strong correlation between net
    photosynthesis and net calcification rates (Fig. 4). Linear relationships
    between net photosynthesis and calcification rates for N. strictum and
    H. scabra were stronger at low pH (R2 = 0.96; R2 = 0.94) compared to
    ambient pH (R2 = 0.68; R2 = 0.64), respectively (Fig. 4a,c), while for
    J. adhaerens the relationships were relatively similar (R2 = 0.87,
    R2 = 0.98; Fig. 4b). The other two species did not have a broad range in
    photosynthetic and calcification rates, thus no relationship could be
    established.

    3.2.2. Inhibitor experiments (FKRT)
    The photosynthesis inhibitor (DCMU) arrested O2 flux in the light for

    all species, providing confidence that calcification in the presence of
    DCMU was non-PSII light-dependent calcification (Table 2). Calcifica-
    tion was stimulated in the light without PSII in three FKRT species,

    including N. strictum, J. adhaerens, and H. scabra (Fig. 5a,b,d). These
    species maintained a relatively high percentage (22 to 34%) of the
    calcification rates attained in the light without PSII inhibition (Fig. 3a,b,
    d). This light-triggered non-PSII net calcification was significantly
    greater than calcification rates measured in the dark at ambient pH
    (Holm-Sidak, t = 3.58 p = 0.012, t = 3.98 p = 0.004, t = 3.00 p = 0.032,
    N. strictum, J. adhaerens, and H. scabra, respectively). However, non-PSII
    light-dependent calcification rates were not sustained at low pH.
    Calcification rates at low pH without PSII were similar in the light and
    dark (Holm-Sidak, t = 0.43 p = 0.681, t = 1.84 p = 0.104, t = 1.01
    p = 0.351, N. strictum, J. adhaerens, and H. scabra, respectively). No
    significant light-triggered calcification with DCMU was observed for
    U. luna (Two-way ANOVA, F1,4 = 1.55 p = 0.281) or CCA (Two-way
    ANOVA, F1,4 = 0.065 p = 0.812). For all species examined, no net posi-
    tive calcification rates were measured in the presence of DCMU at low
    pH in the dark (Fig. 5).

    Fig. 3. Net calcification rates (n = 3–5) of Florida Keys Reef Tract macroalgal species (a = Neogoniolithon strictum, b = Jania adhaerens, c = Udotea luna, d = Halimeda
    scabra, e = CCA) in the light (500 μmol photons m− 2 s− 1) and dark at ambient and low pH. Totals are daily net calcification rates calculated by combining light and
    dark calcification rates at the respective treatment pH and normalized to 24 h. Means ± standard errors are shown. Different lowercase letters represent significant
    differences from a two-way ANOVA with Light x pCO2 treatments and a post-hoc Holm-Sidak to compare between means (P < 0.05). *CCA calcification rates are normalized to g of flexible 2-D surface due to high 3-D complexity. Differences in daily calcification rates and CCA within light treatments were determined by t-tests; daily differences shown with capital letters (P < 0.05).

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    8

    3.3. Low light species (LCI) calcification

    Although the low-irradiance Halimeda species from LCI reefs were
    incubated at 10-fold lower irradiance and had 10 times lower photo-
    synthetic rates compared to FKRT Halimeda species (Table 2), their
    calcification rates in the light (Fig. 6a,b) were similar (Fig. 3d). At
    ambient pH, net calcification rates were significantly higher in the light
    relative to the dark (Fig. 6; Holm-Sidak, t = 4.92 p = 0.003, t = 3.28
    p = 0.038, t = 3.37 p = 0.023, H. goreauii, H. copiosa, Peyr, respectively)
    with the exception of Peyp, that approached significance (Fig. 6d; Holm-
    Sidak, t = 2.59 p = 0.073). Lower dark calcification rates compared to
    the light were consistent with results from the FKRT species (Fig. 3).
    However, in contrast to Halimeda from the FKRT, relatively high positive
    net calcification rates were maintained in the dark by Halimeda species
    from LCI reefs, H. goreauii (39%) and H. copiosa (22%) (Fig. 6a,b). These
    dark calcification rates are more than twice those of N. strictum and J.
    adhaerens from the FKRT (Fig. 3a,b). Calcification rates at low pH in the
    light decreased by 49% and 28% for H. goreauii and H. copiosa, respec-
    tively (Fig. 6a,b), but was only significant for H. goreauii (Holm-Sidak,
    t = 3.19 p = 0.021), likely due to high variance in H. copiosa (Holm-
    Sidak, t = 1.78 p = 0.149). Although both H. goreauii and H. copiosa had
    on average ~ 50% lower daily calcification rates at low pH in the light,
    the differences only approached significance for H. goreauii (t-test,
    t = 2.15 p = 0.0748) and was not significant for H. copiosa (t-test,
    t = 1.18 p = 0.281). Peyred from LCI only exhibited positive net calcifi-
    cation rates in the light at ambient pH and showed net dissolution at low
    pH in the light and dark (Fig. 6c). Although similar trends were
    observed, low pH only significantly reduced Peypink (Holm-Sidak,
    t = 3.17 p = 0.047) net calcification rates in the dark, but not Peyred
    (Holm-Sidak, t = 0.800 p = 0.460). Daily calcification rates were

    significantly lower at low compared ambient pH for Peypink (t-test,
    t = 2.93 p = 0.026) and approached significance for Peyred (t-test,
    t = 2.30 p = 0.061) (Fig. 6c,d).

    4. Discussion

    The most consistent negative effect of low pH and elevated pCO2 on
    calcification rates in tropical macroalgae examined from FKRT and LCI
    occurred in the dark, albeit effects in the light controlled daily calcifi-
    cation rates. Most of the species examined (89%) had calcification rates
    of zero or net dissolution in the dark under 2100 predictions for pH and
    ocean carbonate chemistry. Other experimental and field studies also
    indicate negative effects of elevated pCO2 and low pH on dark calcifi-
    cation rates. Dark calcification rates of Halimeda opuntia decreased
    167% at a low pH (~7.8) tropical CO2 seep site compared to non-seep
    adjacent control areas (Vogel et al. 2015a). These data are comparable
    to the 171% decrease in dark calcification rates observed at low pH for
    the three Halimeda species in this study. A temperate coralline alga
    (Lithothamnion glaciale) with net positive calcification rates in the dark at
    ambient pH, also exhibited net dissolution when exposed to low pH in

    Table 2
    Net photosynthesis and respiration rates during the light and dark calcification
    incubations at ambient and low pH and with and without a photosystem II in-
    hibitor (DCMU). All data are normalized to gram dry weight per hour with the
    exception of CCA from the Florida Reef Tract and Peyssonneliaceae from Little
    Cayman Island (see below). Mean +/− SE (n = 5–3).

    Oxygen (μmol O2 g dwt− 1 or cm− 2 h− 1)*

    Net Photosynthesis Respiration

    Amb pH Low pH Amb pH Low pH

    Florida Reef
    Tract (No
    DCMU)

    N. strictum 8.66 ± 1.66 8.83 ± 1.42 -0.76 ± 0.08 -0.96 ± 0.17
    J. adhaerens 15.47 ± 4.13 19.29 ± 4.09 -2.39 ± 0.49 -2.97 ± 0.70
    U. luna 12.85 ± 1.51 12.54 ± 0.95 -2.08 ± 0.34 -2.10 ± 0.48
    H. scabra 8.12 ± 1.95 8.40 ± 2.12 -1.30 ± 0.22 -1.36 ± 0.45
    *CCA 37.24 ± 13.08 48.30 ± 13.59 -6.28 ± 1.84 -8.62 ± 2.79

    Florida Reef
    Tract (+
    DCMU)

    N. strictum -1.09 ± 0.42 -1.75 ± 0.44 -1.29 ± 0.21 -1.37 ± 0.26
    J. adhaerens -2.63 ± 1.08 -3.81 ± 0.76 -2.92 ± 0.50 -4.13 ± 0.70
    U. luna -2.70 ± 0.88 -3.56 ± 0.24 -3.31 ± 0.71 -3.39 ± 0.99
    H. scabra -2.51 ± 0.73 -2.94 ± 0.52 -2.45 ± 0.71 -2.31 ± 0.53
    *CCA -11.76 ± 2.64 0.02 ± 0.35 -14.40 ± 4.62 -11.19 ± 1.27

    Little
    Cayman
    Island (No
    DCMU)

    H. goreauii 1.37 ± 0.98 2.55 ± 0.84 -0.49 ± 0.19 -0.33 ± 0.31
    H. copiosa 3.34 ± 0.94 3.26 ± 1.00 -0.53 ± 0.32 -0.92 ± 0.32
    *Peyred 0.19 ± 0.01 0.19 ± 0.04 -0.05 ± 0.01 -0.05 ± 0.03
    *Peypink 0.12 ± 0.03 0.17 ± 0.03 -0.01 ± 0.01 -0.06 ± 0.02

    * CCA from the Florida Reef Tract normalized to g of flexible 2-D surface due
    to high 3-D complexity and Peyssonneliaceae; Peyred and Peypink from Little
    Cayman Island to cm− 2 of thalli surface.

    Fig. 4. The linear relationship between net photosynthetic and calcification
    rates (n = 5) in three Florida Reef Tract species (a = Neogoniolithon strictum,
    b = Jania adhaerens, c = Halimeda scabra). Equations and R2 for linear re-
    gressions are shown for ambient pH (circles with solid line) and low pH (tri-
    angles with dashed line) treatments.`

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489
    9

    the dark (Kamenos et al. 2013). Reduced net calcification by 164% in
    macroalgae (H. opuntia) from the Great Barrier Reef at low pH in the
    dark was interpreted as a negative amplifying effect of low pH (Vogel
    et al. 2015b). The impact of low pH and elevated pCO2 on preferentially
    nighttime net calcification (Kamenos et al. 2013; Venn et al. 2019; Vogel
    et al. 2015b; this study) necessitates a greater understanding of these
    mechanisms.

    The constraints on net calcification rates in the dark at low pH is
    likely attributable to greater rates of dissolution. This conjecture is
    reasonable given McNicholl et al. (2020) found in 45Ca experiments
    either no significant difference between gross calcification rates in the
    dark between low (7.7) and ambient (8.1) pH, or a shift from positive
    gross calcification rates to net dissolution at low pH in the dark, in a
    majority of the FKRT species examined herein (N. strictum, J. adhaerens,
    H. scabra, U. luna). This was the case, even though net calcification rates
    significantly declined at low pH in the dark in all four of the FKRT
    species examined (McNicholl et al. 2020). McNicholl et al. (2020) also
    established a strong relationship (R2 = 0.82) between increasing TA and

    loss of 45Ca from pre-labelled 45CaCO3 thalli only in the low pH treat-
    ment, suggesting dissolution. These data indicate that the ability to form
    new CaCO3 is not the primary factor constraining net calcification rates
    in the dark at low pH for the majority of FKRT species examined herein.
    Further, respiration rates cannot account for the increased dissolution at
    low pH in darkness. None of the species examined exhibited greater
    respiration rates in the dark at low pH. Comeau et al. (2016) also found
    respiration rates to be insensitive to elevated pCO2 in 6 coral and 6
    macroalgal species from reefs in Moorea (French Polynesia). Other
    macroalgal studies support the conclusion that respiration rates do not
    increase in response to low pH conditions in the dark (Kamenos et al.
    2013; Martin et al. 2013a; Semesi et al. 2009; Zou et al. 2011; Zweng
    et al. 2018). A general amplifying effect of low pH on dissolution in the
    dark was observed in both high- and low-light macroalgae in the present
    study, regardless of species-specific photophysiology, calcification
    location, taxonomy, or thalli CaCO3 polymorph, suggesting a funda-
    mental relationship that necessitates further research.

    In contrast to results in the dark, low pH had no significant negative

    Fig. 5. Net calcification rates (n = 3–5) of Florida Reef Tract species (a = Neogoniolithon strictum, b = Jania adhaerens, c = Udotea luna, d = Halimeda scabra, e = CCA)
    with the photosynthesis inhibitor DCMU in the light (500 μmol photons m− 2 s− 1) and dark at ambient and low pH. Means ± standard errors are shown. Different
    lowercase letters represent significant differences from a two-way ANOVA with Light x pCO2 treatments (P < 0.05). *CCA calcification rates are normalized to g of flexible 2-D surface due to high 3-D complexity.

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    10

    effect on FKRT species net calcification rates under high irradiance
    (500 μmol photons m− 2 s− 1) except for U. luna. Neogoniolithon strictum
    and J. adhaerens maintained 87% of their calcification rates and
    H. scabra increased net calcification rates 16% at high light and low pH.
    Two other Halimeda species (H. digitate and H. opuntia) from CO2 seep
    sites in Indonesia at low pH (7.7) were also found to increase calcifi-
    cation rates in the light (131% and 41%, respectively) relative to adja-
    cent ambient pH control sites (Vogel et al. 2015a). The FKRT species that
    maintained calcification rates under low pH and elevated pCO2 in the
    light had a strong positive relationship between net photosynthesis and
    calcification at ambient (R2 0.64 to 0.98) and low pH (R2 0.87 to 0.96).
    While the importance of photosynthesis in macroalgal calcification has
    long been appreciated (Borowitzka 1981; Koch et al. 2013; Pentecost
    1978), how this relationship is sustained under low pH has not been
    resolved. Further, in some species, as was shown for H. heteromorpha, net
    photosynthetic rates do not always positively correspond to rates of
    calcification at low pH (Brown et al. 2019). Photosynthesis has been
    shown to elevate pH at the macroalgal thalli surface under low pH and
    elevated pCO2 (Cornwall et al. 2015; Hofmann et al. 2016; McNicholl
    et al. 2019). This photosynthetically-driven increase in surface pH can
    mitigate the negative effects of bulk seawater acidification on net
    calcification (Cornwall et al. 2014). Photosynthesis also elevates pH
    within the calcifying space of corals and macroalgae, even under low
    external bulk seawater pH (Comeau et al. 2018; Cornwall et al. 2017;
    Venn et al. 2019).

    In addition to photosynthesis, light-dependent proton pumps inde-
    pendent of PSII may be important for macroalgal calcification. Three
    species from the FKRT (H. scabra, N. strictum, and J. adhaerens) main-
    tained 22% to 34% of their calcification rates in the light independent of
    PSII. These same three species were observed to control thalli surface H+

    light/dark dynamics independent of PSII (McNicholl et al. 2019). Thus,
    active pH regulation independent of PSII may be linked to calcification

    in macroalgae. Calcification rates in the light without PSII were not
    sustained at low pH in the present study, even though H+ dynamics
    seemingly continue at low pH (McNicholl et al. 2020). This was possibly
    due to unsustainable H+ transport requirements, acid-base disfunction,
    and/or changes in the electrochemical gradients of H+ across the plas-
    malemma. Thus, we suggest photosynthesis and light-triggered proton
    pumps may promote calcification in the light at ambient pH, but proton
    pumps become overwhelmed under elevated [H+] at low pH. Under this
    scenario, calcification would be more dependent on high rates of
    photosynthesis as ocean pH declines.

    Although the FKRT species were reliant on photosynthesis to facili-
    tate high calcification rates, Halimeda species growing on reef walls
    maintained similarly high calcification rates at 10-fold lower irradiance.
    The distinct morphology and physiology of the three Halimeda species
    examined in this study may explain the divergent responses to low pH.
    H. scabra from the FKRT that sustained calcification in the light at low
    pH has relatively high organic:inorganic carbon ratios (Peach et al.
    2017b; Vroom et al. 2003) indicating a high photosynthetic capacity. In
    contrast, species with lower organic:inorganic carbon ratios, H. goreauii
    and H. copiosa, did not compensate for low pH via photosynthesis while
    growing in low light (50 μmol photons m− 2 s− 1). Low-light adapted
    Halimeda species likely have an alternative strategy to promote high
    rates of calcification. One hypothesis is that short diffusive pathways
    (<5 μm, Peach et al. 2017a) that connect their calcifying space to external bulk seawater allow for efficient export of H+, a byproduct of calcification that can lower internal pH and limit further calcification. The path-length may also facilitate diffusive uptake of Ca2+ and CO3

    2−

    into the calcifying space in support of calcification. It has been shown by
    Peach et al. (2017a) that the shorter the diffusive path length the greater
    the %CaCO3 in Halimeda species from LCI, including the species exam-
    ined herein. A short path-length morphology in H. goreauii and
    H. copiosa, combined with a low respiration rate due to a low organic:

    Fig. 6. Net calcification rates (n = 4) of Little Cayman Island species (two chlorophyte Halimeda species: a = Halimeda goreauii, b = Halimeda copiosa, and two
    rhodophyte species from the Peyssonneliaceae family: c = Peyr, d = Peyp) in low light (50 μmol photons m− 2 s− 1) and dark at ambient and low pH. Totals are daily net
    calcification rates calculated by combining light and dark calcification rates at respective pH and normalized to 24 h. Means ± standard errors are shown. Different
    lowercase letters represent significant differences from a two-way ANOVA with Light x pCO2 treatments and post-hoc analysis with Holm-Sidak (P < 0.05). Peys- sonneliaceae calcification rates were normalized to surface area of thalli lobes (cm2). Significant differences in daily calcification rates were determined by t-tests and shown with capital letters (P < 0.05, unless otherwise shown).

    C. McNicholl and M.S. Koch

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    11

    inorganic ratio, likely accounts for high calcification rates measured for
    LCI Halimeda at ambient pH, regardless of low photosynthetic rates. Low
    light adaptation, and potentially a greater dependence on proton pumps
    associated with biotic control, was shown by H. goreauii’s and
    H. copiosa’s ability to maintain 39% and 22% of their calcification rates
    in the dark at ambient pH. However, low-light adapted morphology and
    physiology that allow Halimeda species to be effective and dominant
    calcifiers on deep reefs (Littler et al. 1985; Vroom et al. 2003), may also
    enhance their vulnerability to low pH. For example, low-irradiance
    Halimeda species are likely more dependent on proton pumps for calci-
    fication. We observed a significant loss of non-PSII light-induced calci-
    fication in FKRT species at low pH, potentially an indicator of a decline
    in biotic control of calcification. While we (McNicholl et al. 2019) and
    others (De Beer and Larkum 2001) observed Halimeda species to possess
    proton pumps, the loss of proton pump function in regards to calcifi-
    cation has not been examined in low-light species at low pH.

    Two species from the high light FKRT site (U. luna and CCA) also had
    significant declines in net calcification in the light at low pH. Net
    dissolution occurred in these two species in the dark under both ambient
    and low pH. The apparent lower resistance to dissolution and lower pH
    in the light may correspond to the proximity of these species’ calcifi-
    cation sites to bulk seawater. Udotea luna calcification occurs within
    external sheaths along thalli filaments that are directly exposed to bulk
    seawater (Bohm 1978). Udotea luna dark dissolution rates were also
    found to be high in both ambient and low pH treatments in 45Ca ex-
    periments (McNicholl et al. 2020). Further, microsensor experiments
    (McNicholl et al. 2019) singled out U. luna amongst FKRT species as
    having the least ability to raise pH at the thalli surface in response to
    light at low pH, leading to the suggestion that it has weak biotic control.
    Meyer et al. (2016) showed Udotea flabellum to also exhibit net disso-
    lution in the dark under ambient pH. Further, U. flabellum had a 36%
    lower net calcification rate under low pH in the light, relative to con-
    trols. CCA calcification sites are also proximate to seawater because of
    its prostrate form. This group is often recognized as being highly
    vulnerable to declining pH and elevated pCO2 due to a high‑magnesium
    CaCO3 structure (Ries 2011). Secondary calcification also occurs be-
    tween filaments in coralline macroalgae which are more exposed to bulk
    seawater than cell wall calcification, and thus less resistant to increased
    [H+] (Cornwall et al. 2017; Hofmann et al. 2012). While we observed a
    negative low pH effect on calcification rates of prostrate encrusting
    rhodophytes in these short-term incubations, Peyssonneliaceae (Dutra
    et al. 2015) and CCA (Kamenos et al. 2016; Martin et al. 2013b) have
    exhibited more robust responses to elevated pCO2 in longer-term
    studies. Kamenos et al. (2013) also detected molecular-level changes
    in CCA carbonate minerals that were exposed to abrupt, but not slow,
    treatments of low pH (7.77). These data correspond to recent results
    demonstrating the ability of CCA to acclimate to lower pH over several
    generations (Cornwall et al. 2020). Thus, mechanisms and species-
    specific resistance/vulnerabilities to future changes in pH and carbon-
    ate chemistry need further examination in both short and long-term
    studies.

    Based on our research, we propose that negative responses to
    elevated pCO2 and low pH at the organismal level for calcifying tropical
    macroalgae are primarily associated with effects on net calcification in
    the dark for high-light species. This is likely attributable to greater
    dissolution in the dark at low pH. Even with greater dissolution in the
    dark at low pH, daily net calcification rates can be unaffected by low pH
    because of high net calcification rates in the light and low overall
    calcification rates in the dark. Low pH and elevated pCO2 effects on daily
    calcification rates appear to be greatest in species that exhibit declines in
    net calcification rates in the light. Further, our inhibition experiments
    lead us to suggest that PSII-independent calcification mechanisms may
    become overwhelmed at low pH with greater [H+] in the bulk seawater.
    Thus, low-light species’ morphology and strategy that evolved to sustain
    calcification without high rates of photosynthesis might make them
    more vulnerable to greater [H+] in the bulk seawater. Light-driven

    processes, including photosynthesis and/or H+ control, will be essen-
    tial to sustain or enhance daytime calcification to offset nighttime
    dissolution and maintain a net positive daily calcification rate. Thus,
    macroalgae able to maintain high calcification rates in the light (high
    and low irradiance) at low pH, and/or sustain strong biotic control with
    high [H+] in the bulk seawater, are expected to dominate under global
    change.

    Supplementary data to this article can be found online at https://doi.
    org/10.1016/j.jembe.2020.151489.

    CRediT authorship contribution statement

    C. McNicholl: Investigation, Formal analysis, Writing – original
    draft, Visualization, Writing – review & editing, Validation, Software,
    Methodology, Conceptualization. M.S. Koch: Writing – original draft,
    Visualization, Writing – review & editing, Project administration, Su-
    pervision, Resources, Validation, Methodology, Conceptualization.

    Declaration of Competing Interest

    The authors declare that they have no known competing financial
    interests or personal relationships that could have appeared to influence
    the work reported in this paper.

    Acknowledgements

    This research was funded by the National Science Foundation Ocean
    Acidification Program-CRI-OA Grant #1416376. The authors would like
    to thank Chris Johnson, Kimberly McFarlane, and the undergraduate
    students that assisted in the field and lab. Dr. Carrie Manfrino is
    recognized for her support in the field, and the Cayman Island Marine
    Conservation Board and Department of the Environment for permitting
    our LCI research. We also appreciate the anonymous reviewers and
    editors that significantly improved the manuscript.

    References

    Adey, W.H., 1998. Coral reefs: algal structured and mediated ecosystems in shallow,
    turbulent, alkaline waters. J. Phycol. 34, 393–406.

    Adey, W.H., Halfar, J., Williams, B., 2013. The coralline genus Clathromorphum Foslie
    emend. Adey. Smithson. Contrib. Mar. Sci. 40, 1–41.

    Andersson, A.J., Kuffner, I.B., Mackenzie, F.T., Jokiel, P.L., Rodgers, K.S., Tan, A., 2009.
    Net loss of CaCO3 from a subtropical calcifying community due to seawater
    acidification: mesocosm-scale experimental evidence. Biogeosciences 6, 1811–1823.
    https://doi.org/10.5194/bg-6-1811-2009.

    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. Proc. Natl.
    Acad. Sci. 105, 17442–17446. https://doi.org/10.1073/pnas.0804478105.

    Basso, D., 2012. Carbonate production by calcareous red algae and global change.
    Geodiversitas 34, 13–33. https://doi.org/10.5252/g2012n1a2.

    Bohm, L., 1978. Application of the 45Ca tracer method for determination of calcification
    rates in calcareous algae: effect of calcium exchange and differential saturation of
    algal calcium pools. Mar. Biol. 47, 9–14. https://doi.org/10.1007/bf00397013.

    Borowitzka, M.A., 1981. Photosynthesis and calcification in the articulated coralline red
    algae Amphiroa anceps and A. foliacea. Mar. Biol. 62, 17–23. https://doi.org/
    10.1007/BF00396947.

    Borowitzka, M.A., Larkum, A.W.D., 1976. Calcification in the green alga Halimeda: IV.
    The action of metabolic inhibitors on photosynthesis and calcification. J. Exp. Bot.
    27, 894–907. https://doi.org/10.1093/jxb/27.5.894.

    Borowitzka, M.A., Larkum, A.W.D., 1987. Calcification in algae: mechanisms and the
    role of metabolism. CRC. Crit. Rev. Plant Sci. 6, 1–45. https://doi.org/10.1080/
    07352688709382246.

    Brown, K.T., Bender-Champ, D., Kenyon, T.M., Rémond, C., Hoegh-Guldberg, O.,
    Dove, S., 2019. Temporal effects of ocean warming and acidification on coral–algal
    competition. Coral Reefs 38, 297–309. https://doi.org/10.1007/s00338-019-01775-
    y.

    Chisholm, J.R.M., 2000. Calcification by crustose coralline algae on the northern great
    barrier reef. Australia. Limnol. Oceanogr. 45, 1476–1484. https://doi.org/10.2307/
    2670432.

    Chisholm, J.R.M., Gattuso, J., 1991. Validation of the alkalinity anomaly technique for
    investigating calcification of photosynthesis in coral reef communities. Limnol.
    Oceanogr. 36, 1232–1239. https://doi.org/10.4319/lo.1991.36.6.1232.

    Chrachri, A., Hopkinson, B.M., Flynn, K., Brownlee, C., Wheeler, G.L., 2018. Dynamic
    changes in carbonate chemistry in the microenvironment around single marine

    C. McNicholl and M.S. Koch

    https://doi.org/10.1016/j.jembe.2020.151489

    https://doi.org/10.1016/j.jembe.2020.151489

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0005

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0005

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0010

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0010

    https://doi.org/10.5194/bg-6-1811-2009

    https://doi.org/10.1073/pnas.0804478105

    https://doi.org/10.5252/g2012n1a2

    https://doi.org/10.1007/bf003970

    13

    https://doi.org/10.1007/BF00396947

    https://doi.org/10.1007/BF00396947

    https://doi.org/10.1093/jxb/27.5.894

    https://doi.org/10.1080/07352688709382246

    https://doi.org/10.1080/07352688709382246

    https://doi.org/10.1007/s00338-019-01775-y

    https://doi.org/10.1007/s00338-019-01775-y

    https://doi.org/10.2307/2670432

    https://doi.org/10.2307/2670432

    https://doi.org/10.4319/lo.1991.36.6.1232

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489

    12

    phytoplankton cells. Nat. Commun. 1–12. https://doi.org/10.1038/s41467-017-
    02426-y.

    Comeau, S., Carpenter, R.C., Edmunds, P.J., 2012. Coral reef calcifiers buffer their
    response to ocean acidification using both bicarbonate and carbonate. Proc. R. Soc. B
    280, 1–8. https://doi.org/10.1098/rspb.2012.2374.

    Comeau, S., Edmunds, P.J., Spindel, N.B., Carpenter, R.C., 2013. The responses of eight
    coral reef calcifiers to increasing partial pressure of CO2 do not exhibit a tipping
    point. Limnol. Oceanogr. 58, 388–398. https://doi.org/10.4319/lo.2013.58.1.0388.

    Comeau, S., Carpenter, R., Edmunds, P., 2016. The effects of pCO2 on photosynthesis and
    respiration of tropical scleractinian corals and calcified algae. ICES J. Mar. Sci.
    https://doi.org/10.1093/icesjms/fsv267.

    Comeau, S., Cornwall, C.E., McCulloch, M.T., 2017. Decoupling between the response of
    coral calcifying fluid pH and calcification to ocean acidification. Sci. Rep. 7, 7573.
    https://doi.org/10.1038/s41598-017-08003-z.

    Comeau, S., Cornwall, C.E., DeCarlo, T.M., Krieger, E., McCulloch, M.T., 2018. Similar
    controls on calcification under ocean acidification across unrelated coral reef taxa.
    Glob. Chang. Biol. 24, 1–12. https://doi.org/10.1111/gcb.14379.

    Comeau, S., Cornwall, C.E., DeCarlo, T.M., Doo, S.S., Carpenter, R.C., McCulloch, M.T.,
    2019. Resistance to ocean acidification in coral reef taxa is not gained by
    acclimatization. Nat. Clim. Chang. 9, 477–483. https://doi.org/10.1038/s41558-
    019-0486-9.

    Cornwall, C.E., Hepburn, C.D., Pilditch, C.A., Hurd, C.L., 2013. Concentration boundary
    layers around complex assemblages of macroalgae: implications for the effects of
    ocean acidification on understory coralline algae. Limnol. Oceanogr. 58, 121–130.
    https://doi.org/10.4319/lo.2013.58.1.0121.

    Cornwall, C.E., Boyd, P.W., McGraw, C.M., Hepburn, C.D., Pilditch, C.A., Morris, J.N.,
    Smith, A.M., Hurd, C.L., 2014. Diffusion boundary layers ameliorate the negative
    effects of ocean acidification on the temperate coralline macroalga Arthrocardia
    corymbosa. PLoS One 9, 1–9. https://doi.org/10.1371/journal.pone.0097235.

    Cornwall, C.E., Pilditch, C.A., Hepburn, C.D., Hurd, C.L., 2015. Canopy macroalgae
    influence understorey corallines’ metabolic control of near-surface pH and oxygen
    concentration. Mar. Ecol. Prog. Ser. 525, 81–95. https://doi.org/10.3354/
    meps11190.

    Cornwall, C.E., Comeau, S., McCulloch, M.T., 2017. Coralline algae elevate pH at the site
    of calcification under ocean acidification. Glob. Chang. Biol. 23, 4245–4256. https://
    doi.org/10.1111/gcb.13673.

    Cornwall, C.E., Comeau, S., DeCarlo, T.M., Larcombe, E., Moore, B., Giltrow, K.,
    Puerzer, F., D’Alexis, Q., McCulloch, M.T., 2020. A coralline alga gains tolerance to
    ocean acidification over multiple generations of exposure. Nat. Clim. Chang. 10,
    143–146. https://doi.org/10.1038/s41558-019-0681-8.

    Cyronak, T., Schulz, K.G., Jokiel, P.L., 2015. The omega myth : what really drives lower
    calcification rates in an acidifying ocean. ICES J. Mar. Sci. 1–5. https://doi.org/
    10.1093/icesjms/fsv075.

    De Beer, D., Larkum, A.W.D., 2001. Photosynthesis and calcification in the calcifying
    algae Halimeda discoidea studied with microsensors. Plant Cell Environ. 24,
    1209–1217. https://doi.org/10.1046/j.1365-3040.2001.00772.x.

    Diaz-Pulido, G., McCook, L.J., Larkum, A.D., Lotze, H.K., Raven, J.A., Schaffelke, B.,
    Smith, J.E., Steneck, R.S., 2007. Vulnerability of Macroalgae of the Great Barrier
    Reef to Climate Change, in: Climate Change and the Great Barrier Reef: A
    Vulnerability Assessment.

    Diaz-Pulido, G., Nash, M.C., Anthony, K.R.N., Bender, D., Opdyke, B.N., Reyes-Nivia, C.,
    Troitzsch, U., 2014. Greenhouse conditions induce mineralogical changes and
    dolomite accumulation in coralline algae on tropical reefs. Nat. Commun. 5, 1–9.
    https://doi.org/10.1038/ncomms4310.

    Dutra, E., Koch, M., Peach, K.E., Manfrino, C., 2015. Tropical crustose coralline algal
    individual and community responses to elevated pCO2 under high and low
    irradiance. ICES J. Mar. Sci. 707–721.

    El Haïkali, B., Bensoussan, N., Romano, J.C., Bousquet, V., 2004. Estimation of
    photosynthesis and calcification rates of Corallina elongata Ellis and Solander, 1786,
    by measurements of dissolved oxygen, pH and total alkalinity. Sci. Mar. 68, 45–56.
    https://doi.org/10.3989/scimar.2004.68n145.

    Fabry, V.J., Seibel, B., Feely, R., Orr, J., 2008. Impacts of ocean acidification on marine
    fauna and ecosystem processes. Int. Counc. Explor. Sea 414–432. https://doi.org/
    10.1093/icesjms/fsn048.

    Gao, K., Aruga, Y., Asada, K., Ishihara, T., Akano, T., Kiyohara, M., 1993. Calcification in
    the articulated coralline alga Corallina pilulifera, with special reference to the effect
    of elevated CO2 concentration. Mar. Biol. 117, 129–132. https://doi.org/10.1007/
    BF00346434.

    Gehlen, M., Séférian, R., Jones, D.O.B., Roy, T., Roth, R., Barry, J., Bopp, L., Doney, S.C.,
    Dunne, J.P., Heinze, C., Joos, F., Orr, J.C., Resplandy, L., Segschneider, J.,
    Tjiputra, J., 2014. Projected pH reductions by 2100 might put deep North Atlantic
    biodiversity at risk. Biogeosciences 11, 6955–6967. https://doi.org/10.5194/bg-11-
    6955-2014.

    Hartin, C.A., Bond-Lamberty, B., Patel, P., Mundra, A., 2016. Ocean acidification over
    the next three centuries using a simple global climate carbon-cycle model:
    projections and sensitivities. Biogeosciences 13, 4329–4342. https://doi.org/
    10.5194/bg-13-4329-2016.

    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.,
    Iglesiaia-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. https://doi.org/10.1126/science.1152509.

    Hofmann, L.C., Bischof, K., 2014. Ocean acidification effects on calcifying macroalgae.
    Aquat. Biol. 22, 261–279. https://doi.org/10.3354/ab00581.

    Hofmann, L.C., Yildiz, G., Hanelt, D., Bischof, K., 2012. Physiological responses of the
    calcifying rhodophyte, Corallina officinalis (L.), to future CO2 levels. Mar. Biol. 159,
    783–792. https://doi.org/10.1007/s00227-011-1854-9.

    Hofmann, L.C., Heiden, J., Bischof, K., Teichberg, M., 2014. Nutrient availability affects
    the response of the calcifying chlorophyte Halimeda opuntia (L.) J.V. Lamouroux to
    low pH. Planta 239, 231–242. https://doi.org/10.1007/s00425-013-1982-1.

    Hofmann, L.C., Koch, M., De Beer, D., 2016. Biotic control of surface pH and evidence of
    light-induced H+ pumping and Ca2+ -H+ exchange in a tropical crustose coralline
    alga. PLoS One 11, 1–24. https://doi.org/10.1371/journal.pone.0159057.

    Jokiel, P.L., 2011. Ocean acidification and control of reef coral calcification by boundary
    layer limitation of proton flux. Bull. Mar. Sci. 87, 639–657. https://doi.org/
    10.5343/bms.2010.1107.

    Kamenos, N.A., Law, A., 2010. Temperature controls on coralline algal skeletal growth.
    J. Phycol. 46, 331–335. https://doi.org/10.1111/j.1529-8817.2009.00780.x.

    Kamenos, N.A., Cusack, M., Huthwelker, T., Lagarde, P., Scheibling, R.E., 2009. Mg-
    lattice associations in red coralline algae. Geochim. Cosmochim. Acta 73,
    1901–1907. https://doi.org/10.1016/j.gca.2009.01.010.

    Kamenos, N.A., Burdett, H.L., Aloisio, E., Findlay, H.S., Martin, S., Longbone, C.,
    Dunn, J., Widdicombe, S., Calosi, P., 2013. Coralline algal structure is more sensitive
    to rate, rather than the magnitude, of ocean acidification. Glob. Chang. Biol. 19,
    3621–3628. https://doi.org/10.1111/gcb.12351.

    Kamenos, N.A., Perna, G., Gambi, M.C., Micheli, F., Kroeker, K.J., 2016. Coralline algae
    in a naturally acidified ecosystem persist by maintaining control of skeletal
    mineralogy and size. Proc. R. Soc. B Biol. Sci. 283, 20161159 https://doi.org/
    10.1098/rspb.2016.1159.

    Kato, A., Hikami, M., Kumagai, N.H., Suzuki, A., Nojiri, Y., Sakai, K., 2014. Negative
    effects of ocean acidification on two crustose coralline species using genetically
    homogeneous samples. Mar. Environ. Res. 94, 1–6. https://doi.org/10.1016/j.
    marenvres.2013.10.010.

    Koch, M., Bowes, G., Ross, C., Zhang, X.-H., 2013. Climate change and ocean
    acidification effects on seagrasses and marine macroalgae. Glob. Chang. Biol. 19
    https://doi.org/10.1111/j.1365-2486.2012.02791.x.

    Kuffner, I.B., Lidz, B.H., Hudson, J.H., Anderson, J.S., 2015. A century of ocean warming
    on Florida keys coral reefs: historic in situ observations. Estuar. Coasts 38,
    1085–1096. https://doi.org/10.1007/s12237-014-9875-5.

    Langdon, C., Gattuso, J., Andersson, A., 2010. Part 3: measurements of calcification and
    dissolution of benthic organisms and communities. In: Riebesell, U., Fabry, V.J.,
    Hansson, L. (Eds.), Guide to Best Practices for Ocean Acidification Research and Data
    Reporting. European Commission, Office CDMA 03/115, B-1049, Brussels,
    pp. 213–232.

    Littler, M.M., Littler, D.S., Blair, S.M., Norris, J.N., 1985. Deepest known plant life
    discovered on an uncharted seamount. Science 227, 57–59. https://doi.org/
    10.1126/science.227.4682.57.

    Marsh, J.A., 1970. Primary productivity of reef-building calcareous red algae. Ecology
    51, 255–263.

    Martin, S., Charnoz, A., Gattuso, J.P., 2013a. Photosynthesis, respiration and
    calcification in the Mediterranean crustose coralline alga Lithophyllum cabiochae
    (Corallinales, Rhodophyta). Eur. J. Phycol. 48, 163–172. https://doi.org/10.1080/
    09670262.2013.786790.

    Martin, S., Cohu, S., Vignot, C., Zimmerman, G., Gattuso, J.P., 2013b. One-year
    experiment on the physiological response of the Mediterranean crustose coralline
    alga, Lithophyllum cabiochae, to elevated pCO2 and temperature. Ecol. Evol. 3,
    676–693. https://doi.org/10.1002/ece3.475.

    Mccoy, S.J., Kamenos, N.A., 2015. Coralline algae (Rhodophyta) in a changing world:
    integrating ecological, physiological, and geochemical responses to global change.
    J. Phycol. 51, 6–24. https://doi.org/10.1111/jpy.12262.

    McNicholl, C., Koch, M.S., Hofmann, L.C., 2019. Photosynthesis and light-dependent
    proton pumps increase boundary layer pH in tropical macroalgae: a proposed
    mechanism to sustain calcification under ocean acidification. J. Exp. Mar. Bio. Ecol.
    521 (1–12).

    McNicholl, C., Koch, M.S., Swarzenski, P.W., Oberhaensli, F.R., Taylor, A., Batista, M.G.,
    Metian, M., 2020. Ocean acidification effects on calcification and dissolution in
    tropical reef macroalgae. Coral Reefs. https://doi.org/10.1007/s00338-020-01991-
    x.

    Meyer, F.W., Schubert, N., Diele, K., Teichberg, M., Wild, C., Enriquez, S., 2016. Effect of
    inorganic and organic carbon enrichments (DIC and DOC) on the photosynthesis and
    calcification rates of two calcifying green algae from a Caribbean reef lagoon. PLoS
    One 11. https://doi.org/10.1371/journal.pone.0160268.

    Nelson, W.A., 2009. Calcified macroalgae–critical to coastal ecosystems and vulnerable
    to change: a review. Mar. Freshw. Res. 60, 787–801. https://doi.org/10.1071/
    MF08335.

    Noisette, F., Duong, G., Six, C., Davoult, D., Martin, S., 2013. Effects of elevated pCO2 on
    the metabolism of a temperate rhodolith Lithothamnion corallioides grown under
    different temperatures. J. Phycol. 49, 746–757. https://doi.org/10.1111/jpy.12085.

    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.,
    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. https://
    doi.org/10.1038/nature04095.

    Peach, K.E., Koch, M.S., Blackwelder, P.L., 2016. Effects of elevated pCO2 and irradiance
    on growth, photosynthesis and calcification in Halimeda discoidea. Mar. Ecol. Prog.
    Ser. 544, 143–158. https://doi.org/10.3354/meps11591.

    C. McNicholl and M.S. Koch

    https://doi.org/10.1038/s41467-017-02426-y

    https://doi.org/10.1038/s41467-017-02426-y

    https://doi.org/10.1098/rspb.2012.2374

    https://doi.org/10.4319/lo.2013.58.1.0388

    https://doi.org/10.1093/icesjms/fsv267

    https://doi.org/10.1038/s41598-017-08003-z

    https://doi.org/10.1111/gcb.14379

    https://doi.org/10.1038/s41558-019-0486-9

    https://doi.org/10.1038/s41558-019-0486-9

    https://doi.org/10.4319/lo.2013.58.1.0121

    https://doi.org/10.1371/journal.pone.0097235

    https://doi.org/10.3354/meps11190

    https://doi.org/10.3354/meps11190

    https://doi.org/10.1111/gcb.13673

    https://doi.org/10.1111/gcb.13673

    https://doi.org/10.1038/s41558-019-0681-8

    https://doi.org/10.1093/icesjms/fsv075

    https://doi.org/10.1093/icesjms/fsv075

    https://doi.org/10.1046/j.1365-3040.2001.00772.x

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0135

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0135

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0135

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0135

    https://doi.org/10.1038/ncomms4310

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0145

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0145

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0145

    https://doi.org/10.3989/scimar.2004.68n145

    https://doi.org/10.1093/icesjms/fsn048

    https://doi.org/10.1093/icesjms/fsn048

    https://doi.org/10.1007/BF00346434

    https://doi.org/10.1007/BF00346434

    https://doi.org/10.5194/bg-11-6955-2014

    https://doi.org/10.5194/bg-11-6955-2014

    https://doi.org/10.5194/bg-13-4329-2016

    https://doi.org/10.5194/bg-13-4329-2016

    https://doi.org/10.1126/science.1152509

    https://doi.org/10.3354/ab00581

    https://doi.org/10.1007/s00227-011-1854-9

    https://doi.org/10.1007/s00425-013-1982-1

    https://doi.org/10.1371/journal.pone.0159057

    https://doi.org/10.5343/bms.2010.1107

    https://doi.org/10.5343/bms.2010.1107

    https://doi.org/10.1111/j.1529-8817.2009.00780.x

    https://doi.org/10.1016/j.gca.2009.01.010

    https://doi.org/10.1111/gcb.12351

    https://doi.org/10.1098/rspb.2016.1159

    https://doi.org/10.1098/rspb.2016.1159

    https://doi.org/10.1016/j.marenvres.2013.10.010

    https://doi.org/10.1016/j.marenvres.2013.10.010

    https://doi.org/10.1111/j.1365-2486.2012.02791.x

    https://doi.org/10.1007/s12237-014-9875-5

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0240

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0240

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0240

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0240

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0240

    https://doi.org/10.1126/science.227.4682.57

    https://doi.org/10.1126/science.227.4682.57

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0250

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0250

    https://doi.org/10.1080/09670262.2013.786790

    https://doi.org/10.1080/09670262.2013.786790

    https://doi.org/10.1002/ece3.475

    https://doi.org/10.1111/jpy.12262

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0270

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0270

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0270

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0270

    https://doi.org/10.1007/s00338-020-01991-x

    https://doi.org/10.1007/s00338-020-01991-x

    https://doi.org/10.1371/journal.pone.0160268

    https://doi.org/10.1071/MF08335

    https://doi.org/10.1071/MF08335

    https://doi.org/10.1111/jpy.12085

    https://doi.org/10.1038/nature04095

    https://doi.org/10.1038/nature04095

    https://doi.org/10.3354/meps11591

    Journal of Experimental Marine Biology and Ecology 535 (2021) 151489
    13

    Peach, K.E., Koch, M.S., Blackwelder, P.L., Guerrero-Given, D., Kamasawa, N., 2017a.
    Primary utricle structure of six Halimeda species and potential relevance for ocean
    acidification tolerance. Bot. Mar. 60, 1–11. https://doi.org/10.1515/bot-2016-0055.

    Peach, K.E., Koch, M.S., Blackwelder, P.L., Manfrino, C., 2017b. Calcification and
    photophysiology responses to elevated pCO2 in six Halimeda species from contrasting
    irradiance environments on little Cayman island reefs. J. Exp. Mar. Bio. Ecol. 486,
    114–126. https://doi.org/10.1016/j.jembe.2016.09.008.

    Pentecost, A., 1978. Calcification and photosynthesis in Corallina officinalis l. using the
    14CO2 method. Br. Phycol. J. 13, 383–390. https://doi.org/10.1080/
    00071617800650431.

    Pierrot, D., Lewis, E., Wallace, D., 2006. MS Excel Program Developed for CO2 Systems
    Calculations: ORNL/CDIAC 105a.

    Porzio, L., Buia, M.C., Hall-Spencer, J.M., 2011. Effects of ocean acidification on
    macroalgal communities. J. Exp. Mar. Bio. Ecol. 400, 278–287. https://doi.org/
    10.1016/j.jembe.2011.02.011.

    Price, N.N., Hamilton, S.L., Tootell, J.S., Smith, J.E., 2011. Species-specific consequences
    of ocean acidification for the calcareous tropical green algae Halimeda. Mar. Ecol.
    Prog. Ser. 440, 67–78. https://doi.org/10.3354/meps09309.

    Pueschel, C.M., Judson, B.L., Wegeberg, S., 2005. Decalcification during epithallial cell
    turnover in Jania adhaerens (Corallinales, Rhodophyta). Phycologia 44, 156–162.
    https://doi.org/10.2216/0031-8884(2005)44[156:DDECTI]2.0.CO;2.

    Raven, J.A., Hurd, C.L., 2012. Ecophysiology of photosynthesis in macroalgae.
    Photosynth. Res. 113, 105–125. https://doi.org/10.1007/s11120-012-9768-z.

    Ries, J.B., 2011. Skeletal mineralogy in a high-CO2 world. J. Exp. Mar. Bio. Ecol. 403,
    54–64. https://doi.org/10.1016/j.jembe.2011.04.006.

    Ries, J.B., Cohen, A.L., McCorkle, D.C., 2009. Marine calcifiers exhibit mixed responses
    to CO2-induced ocean acidification. Geology 37, 1131–1134. https://doi.org/
    10.1130/G30210A.1.

    Semesi, I.S., Kangwe, J., Björk, M., 2009. Alterations in seawater pH and CO2 affect
    calcification and photosynthesis in the tropical coralline alga, Hydrolithon sp.

    (Rhodophyta). Estuar. Coast. Shelf Sci. 84, 337–341. https://doi.org/10.1016/j.
    ecss.2009.03.038.

    Taylor, A.R., Brownlee, C., Wheeler, G.L., 2012. Proton channels in algae: reasons to be
    excited. Trends Plant Sci. 17, 675–684. https://doi.org/10.1016/j.
    tplants.2012.06.009.

    Venn, A.A., Tambutté, E., Caminiti-Segonds, N., Techer, N., Allemand, D., Tambutté, S.,
    2019. Effects of light and darkness on pH regulation in three coral species exposed to
    seawater acidification. Sci. Rep. 9, 1–12. https://doi.org/10.1038/s41598-018-
    38168-0.

    Vogel, N., Fabricius, K.S., Strahl, J., Noonan, S., Wild, C., Uthicke, S., 2015a. Calcareous
    green alga Halimeda tolerates ocean acidification conditions at tropical carbon
    dioxide seeps. Limnol. Oceanogr. 60, 263–275. https://doi.org/10.1002/lno.10021.

    Vogel, N., Meyer, F.W., Wild, C., Uthicke, S., 2015b. Decreased light availability can
    amplify negative impacts of ocean acidification on calcifying coral reef organisms.
    Mar. Ecol. Prog. Ser. 521, 49–61. https://doi.org/10.3354/meps11088.

    Vroom, P.S., Smith, C.M., Coyer, J.A., Walters, L.J., Hunter, C.L., Beach, K.S., Smith, J.E.,
    2003. Field biology of Halimeda tuna (Bryopsidales, Chlorophyta) across a depth
    gradient: comparative growth, survivorship, recruitment, and reproduction.
    Hydrobiologia 501, 149–166. https://doi.org/10.1023/A:1026287816324.

    Wizemann, A., Meyer, F.W., Westphal, H., 2014. A new model for the calcification of the
    green macro-alga Halimeda opuntia (Lamouroux). Coral Reefs 33, 951–964. https://
    doi.org/10.1007/s00338-014-1183-9.

    Zou, D., Gao, K., Luo, H., 2011. Short- and long-term effects of elevated CO2 on
    photosynthesis and respiration in the marine macroalga Hizikia Fusiformis
    (Sargassaceae, Phaeophyta) grown at low and high N supplies. J. Phycol. 47, 87–97.
    https://doi.org/10.1111/j.1529-8817.2010.00929.x.

    Zweng, R.C., Koch, M.S., Bowes, G., 2018. The role of irradiance and C-use strategies in
    tropical macroalgae photosynthetic response to ocean acidification. Sci. Rep. 8,
    1–11. https://doi.org/10.1038/s41598-018-27333-0.

    C. McNicholl and M.S. Koch

    https://doi.org/10.1515/bot-2016-0055

    https://doi.org/10.1016/j.jembe.2016.09.008

    https://doi.org/10.1080/00071617800650431

    https://doi.org/10.1080/00071617800650431

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0320

    http://refhub.elsevier.com/S0022-0981(19)30489-7/rf0320

    https://doi.org/10.1016/j.jembe.2011.02.011

    https://doi.org/10.1016/j.jembe.2011.02.011

    https://doi.org/10.3354/meps09309

    https://doi.org/10.2216/0031-8884(2005)44[156:DDECTI]2.0.CO;2

    https://doi.org/10.1007/s11120-012-9768-z

    https://doi.org/10.1016/j.jembe.2011.04.006

    https://doi.org/10.1130/G30210A.1

    https://doi.org/10.1130/G30210A.1

    https://doi.org/10.1016/j.ecss.2009.03.038

    https://doi.org/10.1016/j.ecss.2009.03.038

    https://doi.org/10.1016/j.tplants.2012.06.009

    https://doi.org/10.1016/j.tplants.2012.06.009

    https://doi.org/10.1038/s41598-018-38168-0

    https://doi.org/10.1038/s41598-018-38168-0

    https://doi.org/10.1002/lno.10021

    https://doi.org/10.3354/meps11088

    https://doi.org/10.1023/A:1026287816324

    https://doi.org/10.1007/s00338-014-1183-9

    https://doi.org/10.1007/s00338-014-1183-9

    https://doi.org/10.1111/j.1529-8817.2010.00929.x

    https://doi.org/10.1038/s41598-018-27333-0

    • Irradiance, photosynthesis and elevated pCO2 effects on net calcification in tropical reef macroalgae
    • 1 Introduction
      2 Materials & methods
      2.1 Calcified macroalgae collection and site parameters
      2.1.1 Florida keys reef tract (high light environment)
      2.1.2 Little Cayman Island reef (low light environment)
      2.2 Elevated pCO2 experiments
      2.3 PSII inhibitor experiments
      2.4 Calculation of net calcification and chemistry analysis
      2.5 Photosynthesis and respiration
      2.6 Statistical analysis
      3 Results
      3.1 Carbonate chemistry and △pH, △TA and △O2
      3.2 High light species (FKRT) calcification
      3.2.1 Photosynthesis and calcification
      3.2.2 Inhibitor experiments (FKRT)
      3.3 Low light species (LCI) calcification
      4 Discussion
      CRediT authorship contribution statement
      Declaration of Competing Interest
      Acknowledgements
      References

    Glob Change Biol. 2020;00:1–14. wileyonlinelibrary.com/journal/gcb  |  1© 2020 John Wiley & Sons Ltd

    Received: 28 July 2020  |  Accepted: 13 November 2020
    DOI: 10.1111/gcb.15455

    P R I M A R Y R E S E A R C H A R T I C L E

    Ocean acidification locks algal communities in a species-poor
    early successional stage

    Ben P. Harvey1  | Koetsu Kon1  | Sylvain Agostini1  | Shigeki Wada1  |
    Jason M. Hall-Spencer1,2

    1Shimoda Marine Research Center,
    University of Tsukuba, Shizuoka, Japan
    2Marine Biology and Ecology Research
    Centre, University of Plymouth, Plymouth,
    UK

    Correspondence
    Ben P. Harvey, Shimoda Marine Research
    Center, University of Tsukuba, 5-10-1
    Shimoda, Shizuoka 415-0025, Japan.
    Email: ben.harvey@shimoda.tsukuba.ac.jp

    Funding information
    Japan Society for the Promotion
    of Science, Grant/Award Number:
    17K17622; Ministry of Environment,
    Government of Japan, Grant/Award
    Number: 4RF-1701; University of Tsukuba

    Abstract
    Long-term exposure to CO2-enriched waters can considerably alter marine biological
    community development, often resulting in simplified systems dominated by turf
    algae that possess reduced biodiversity and low ecological complexity. Current un-
    derstanding of the underlying processes by which ocean acidification alters biologi-
    cal community development and stability remains limited, making the management
    of such shifts problematic. Here, we deployed recruitment tiles in reference (pHT
    8.137 ± 0.056 SD) and CO2-enriched conditions (pHT 7.788 ± 0.105 SD) at a volcanic
    CO2 seep in Japan to assess the underlying processes and patterns of algal commu-
    nity development. We assessed (i) algal community succession in two different sea-
    sons (Cooler months: January–July, and warmer months: July–January), (ii) the effects
    of initial community composition on subsequent community succession (by recipro-
    cally transplanting preestablished communities for a further 6 months), and (iii) the
    community production of resulting communities, to assess how their functioning was
    altered (following 12 months recruitment). Settlement tiles became dominated by
    turf algae under CO2-enrichment and had lower biomass, diversity and complexity, a
    pattern consistent across seasons. This locked the community in a species-poor early
    successional stage. In terms of community functioning, the elevated pCO2 commu-
    nity had greater net community production, but this did not result in increased algal
    community cover, biomass, biodiversity or structural complexity. Taken together, this
    shows that both new and established communities become simplified by rising CO2
    levels. Our transplant of preestablished communities from enriched CO2 to refer-
    ence conditions demonstrated their high resilience, since they became indistinguish-
    able from communities maintained entirely in reference conditions. This shows that
    meaningful reductions in pCO2 can enable the recovery of algal communities. By
    understanding the ecological processes responsible for driving shifts in community
    composition, we can better assess how communities are likely to be altered by ocean
    acidification.

    K E Y W O R D S
    CO2 seeps, community dynamics, competition, ecosystem function, global change ecology,
    inhibition, turf algae

    www.wileyonlinelibrary.com/journal/gcb

    mailto:

    https://orcid.org/0000-0002-4971-1634

    https://orcid.org/0000-0003-1379-0702

    https://orcid.org/0000-0001-9040-9296

    https://orcid.org/0000-0001-6893-7498

    https://orcid.org/0000-0002-6915-2518

    mailto:ben.harvey@shimoda.tsukuba.ac.jp

    http://crossmark.crossref.org/dialog/?doi=10.1111%2Fgcb.15455&domain=pdf&date_stamp=2021-01-10

    2  |    HARVEY Et Al.

    1   |   I N T R O D U C T I O N

    The oceanic uptake of anthropogenic carbon dioxide emissions is
    a global environmental issue termed ocean acidification. The ef-
    fects of ocean acidification are detrimental to a wide range of ma-
    rine organisms (Harvey et al., 2013; Kroeker, Kordas, et al., 2013),
    and this affects ecosystem functioning and the goods and services
    that people derive from marine resources (Gattuso et al., 2015;
    Hall-Spencer & Harvey, 2019). To better understand the effects
    of ocean acidification, there has been an effort in recent years to
    move beyond aquarium-based experiments on single species to-
    wards in-situ experiments (e.g. Albright et al., 2016, 2018; Brown
    et al., 2016), long-term mesocosm observations (e.g. Algueró-Muñiz
    et al., 2017; Moulin et al., 2015), and studies using natural CO2 seeps
    (e.g. Agostini et al., 2018; Fabricius et al., 2011; Hall-Spencer et al.,
    2008; Milazzo et al., 2014). These approaches have shown that long-
    term exposure to ocean acidification conditions projected for the
    end of the century fundamentally alters the composition of marine
    biological communities, usually resulting in simplified systems with
    reduced biodiversity and less ecological complexity (Agostini et al.,
    2018; Sunday et al., 2017; Vizzini et al., 2017). Many of these studies
    have been observation-based, and so an understanding of the un-
    derlying processes responsible for driving these patterns in commu-
    nity development remains limited. To help assess the future effects
    of ocean acidification, it would be useful to better understand how
    community development processes are affected by rising levels of
    seawater CO2 (Gaylord et al., 2015), and how such changes will influ-
    ence their associated ecosystem functioning.

    Ecological theory suggests that the successional trajectories of
    ‘disturbed’ marine subtidal communities will be primarily driven by
    physical stresses, competition for resources through the mecha-
    nisms of ‘facilitation’ and ‘inhibition’ (Connell & Slatyer, 1977) and
    the strength of associated bottom-up and top-down interactions
    (Gruner et al., 2008; Jenkins et al., 1999). One of the difficulties in
    predicting how community development will be affected by ocean
    acidification is that the changes in carbonate chemistry can simul-
    taneously act as both resource and stressor (Connell et al., 2013,
    2018; Milazzo et al., 2019). It provides a bottom-up resource to
    primary producers by enhancing the availability of bicarbonate and
    CO2 (Connell et al., 2013; Koch et al., 2013), but also acts as a phys-
    ical stressor to many organisms (including calcified primary produc-
    ers) via negative effects on their physiology (Harvey et al., 2013;
    Kroeker, Kordas, et al., 2013). Subsequently, marine communities are
    expected to be re-organized by the effects of ocean acidification.
    Ocean acidification alters the initial successional trajectories of algal
    communities, which lead to dominance by fleshy algae over calcified
    algae in acidified conditions projected for the end of the century in
    both temperate and tropical settings (Crook et al., 2016; Kroeker
    et al., 2012). Enriched CO2 alters competitive interactions, acting as
    a physical stressor to calcified macroalgae whereas turf algae can
    use the additional carbon to boost growth, which allows turf algae to
    attain dominance (also see Connell et al., 2018). Fast-growing oppor-
    tunistic (r-selected) turf algal species are usually suppressed beneath

    macroalgal canopies on temperate reefs (Johnson & Mann, 1988)
    and by top-down control of grazers in coral reefs (Hughes et al.,
    2007). In the absence of strong competition or compensatory pro-
    cesses (e.g. Connell et al., 2018; Ghedini & Connell, 2016; Ghedini
    et al., 2015), turf species can become dominant thereby changing
    the ecosystem state.

    Under present-day conditions, it has been suggested that despite
    bottom-up control of primary production being pervasive, top-down
    control by consumers has a stronger influence on the trajectories of
    algal community succession (Gruner et al., 2008; Hillebrand et al.,
    2007). For example, intense grazing by sea urchins and herbivo-
    rous fish can prevent kelp forest growth resulting in ‘urchin barrens’
    dominated by crustose coralline algae (Kelly et al., 2016; Ling et al.,
    2015). Ocean acidification is expected to reduce bottom-up control
    on those species which are carbon-limited, as long as sufficient nu-
    trients are available (Celis-Plá et al., 2015; Gordillo et al., 2003; Li
    et al., 2012). Top-down control by benthic invertebrates in acidified
    conditions may also diminish, given that at CO2 seeps the abun-
    dance and size of many marine fauna are reduced (Garilli et al., 2015;
    Harvey et al., 2016, 2018), with such examples as the observed num-
    ber of sea urchin feeding halos being reduced in a CO2 seep (Kroeker
    et al., 2013). Fish communities include a greater proportion of her-
    bivorous fish within acidified conditions (during the period of peak
    macroalgae biomass; Cattano et al., 2020), and so it may be possible
    in some systems for fish to maintain top-down control (Baggini et al.,
    2015). Taken together, any strong reductions in bottom-up and/or
    top-down control are likely to alter community successional tra-
    jectories and allow r-selected opportunistic species to outcompete
    other species and dominate under ocean acidification.

    Seasonality is an important aspect of shallow-water ecosystems,
    and yet the consequences of seasonally induced environmental fluc-
    tuations have rarely been considered in ocean acidification studies
    (Baggini et al., 2014; Godbold & Solan, 2013). Algal communities in
    temperate and warm temperate ecosystems experience large seasonal
    changes in environmental conditions (Figure 1), which result in con-
    siderable temporal shifts with a period of high recruitment and peak
    biomass typically occurring in late spring. Thus the responses of algal
    communities to ocean acidification will likely be strongly influenced by
    seasonality (Baggini et al., 2014). In the Northern Pacific Ocean, this
    is further complicated by the occurrence of typhoons, which typically
    occur between July and October in Japan. Typhoons act as a substan-
    tial physical disturbance that affects benthic community structure and
    habitat complexity, such as through the removal of corals (Done, 1992),
    macroalgae (Cattano et al., 2020) and seagrass cover (Wilson et al.,
    2020), and can indirectly change the community function of associated
    species (e.g. fish; Cattano et al., 2020).

    Observations at natural CO2 seeps worldwide provide a good
    understanding of how long-term ocean acidification simplifies the
    composition of climax communities (Foo et al., 2018; González-
    Delgado & Hernández, 2018; Hall-Spencer & Harvey, 2019), yet it
    remains unclear whether these simplified communities develop due
    to altered successional trajectory, stunted community development
    (via successional inhibition) or are driven by reduced bottom-up and/

        |  3HARVEY Et Al.

    or top-down control. To address these gaps, we deployed recruit-
    ment tiles in reference (~350 μatm pCO2) and acidified (~900 μatm
    pCO2) conditions using a natural CO2 seep area as an analogue for
    end of the century pCO2 conditions (the representative concentra-
    tion pathway (RCP) 8.5 scenario, 851 to 1370 μatm; IPCC, 2013) to
    assess the early to mid-successional trajectories of algal communities
    in two different seasons (cooler months January to July, and warmer
    months July to January). The study was carried out over these two
    time periods to investigate whether the effects of ocean acidification
    on community development are temporally consistent. Following this
    we carried out a reciprocal transplant of some of those established
    communities, in order to assess the effects of initial community com-
    position on subsequent community succession in the reference and
    acidified conditions. Finally, we assessed the community production
    of these reciprocally transplanted communities (including the asso-
    ciated sessile invertebrate communities which contribute in terms of
    respiration), in order to determine how any changes in community
    composition will alter their ecosystem functioning.

    2   |   M AT E R I A L S A N D M E T H O D S

    2.1  |  Experimental design

    To investigate our core question of how ocean acidification in-
    fluences early community succession of algal communities, ex-
    periments using recruitment tiles were carried out using an
    acidified area of the Shikine Island CO2 seep, Japan (34°19′9ʺN,
    139°12′18ʺE), and a nearby reference pCO2 area in an adjacent bay

    (~600 m away by the shortest route). Both the reference and acidi-
    fied locations (hereafter ‘350 μatm’ and ‘900 μatm’, respectively)
    have had their carbonate chemistry and biology well character-
    ized previously (Agostini et al., 2015, 2018; Cattano et al., 2020;
    Harvey et al., 2018, 2019; Kerfahi et al., 2020; Witkowski et al.,
    2019), and we present 2 months of additional original pHT (Figure
    S1) and temperature data collected at the ‘900 μatm’ location with
    a Durafet sensor (SeaFET, Sea-Bird Scientific) using the same ap-
    proach as Agostini et al. (2018). Salinity was measured concurrently
    using Hobo conductivity loggers (U24-002-C), and discrete sam-
    ples for total alkalinity were collected throughout the study period,
    with total alkalinity measured using an auto-titrator (916 Ti-Touch,
    Metrohm). In summary, the ‘350 μatm’ location had a mean pHT of
    8.137 ± 0.056 (SD) and the ‘900 μatm’ location had a mean pHT of
    7.781 ± 0.105 (SD), and the mean carbonate chemistry of the two
    locations is presented in Table 1. Long-term temperature data were
    recorded over a 1-year period by deploying a temperature logger
    (HOBO Pendant Temperature/Light 64K Data Logger) at ~6 m
    depth in each site. The ‘900 μatm’ elevated pCO2 location repre-
    sents an end of the century projection for reductions in pH (the RCP
    8.5 scenario; IPCC, 2013), and was not confounded by differences in
    temperature, salinity, dissolved oxygen, total alkalinity, nutrients or
    depth relative to reference sites used for comparison (Agostini et al.,
    2015, 2018; Harvey et al., 2019). Our basalt recruitment tiles were
    130 × 130 × 15 mm and were secured using individual anchor bolts
    (8.5 mm width, 70 mm length) drilled into rock by SCUBA divers at
    ~6 m depth (Nemo Underwater Drill). The tiles at each location were
    deployed haphazardly across a c. 400 m2 area (with at least 5 m be-
    tween individual tiles), fixed to upward-facing substrata.

    F I G U R E 1  Conceptual representation
    of the recruitment tile treatments. (a)
    Tiles were deployed for 6 months during
    the ‘Cold Period’ or ‘Warm Period’ in
    either reference pCO2 (350 μatm; blue
    line) or acidified conditions (900 μatm;
    red line). (b) Tiles from the ‘Warm Period’
    were then used as part of a reciprocal
    transplant either being transplanted into
    350 or 900 μatm conditions for a further
    6 months

    4  |    HARVEY Et Al.

    2.1.1  |  Seasonal experiment

    For the first experiment, five recruitment tiles were individually de-
    ployed in each location (350 and 900 μatm) during the cooler months
    of January 2017–July 2017 (hereafter termed ‘Cold Period’), with
    eight recruitment tiles deployed in each location (350 and 900 μatm)
    during the warmer months of July 2017–January 2018 (hereafter
    termed ‘Warm Period’). Mean seawater temperature (±SD) during
    the ‘Cold Period’ was 18.14 ± 1.81°C at 350 μatm and 18.07 ± 1.63°C
    at 900 μatm, and during the ‘Warm Period’ was 22.86 ± 2.97°C at
    350 μatm and 22.67 ± 2.83°C at 900 μatm. See Figure 1a for a con-
    ceptual overview of the experimental design.

    2.1.2  |  Reciprocal experiment

    For the second experiment, tiles from the ‘Warm Period’ of the
    seasonal experiment were used (each tile had an algal community
    following 6 months recruitment). Sixteen tiles from the seasonal ex-
    periment were reciprocally transplanted into the 350 and 900 μatm
    locations for a further 6 months to assess the effects of initial com-
    munity composition on subsequent community succession in refer-
    ence and acidified conditions (four tiles in each combination, see
    Figure 1b).

    2.2  |  Community analysis

    For both the seasonal experiment and the reciprocal transplanta-
    tion experiment, following the 6 month experimental period each
    tile was brought into the laboratory and photographed (Nikon
    D7200, Nikon). For each 12 month tile, two photos were taken to
    image the upperstorey community, and (after removal of the up-
    perstorey community by hand) the understorey community. These
    data were then combined for analysis. Community composition
    was assessed using ImageJ (Abràmoff et al., 2004) by overlaying
    64 points on a grid, and recording the abundance of each algal
    functional group underlying that point. Functional groups used
    followed Steneck and Dethier (1994), with algal grouping being

    based on their morphology, thallus size and complexity (micro-
    algae, filamentous algae, foliose algae, corticated foliose algae,
    corticated macrophytes, leathery macrophytes, articulated cal-
    careous algae and crustose algae). The habitat complexity of each
    tile was determined by combining the abundance of each algal
    functional group with a rank (0–5) based on the biogenic habitat
    complexity provided by that functional group. The ranking was
    based on categories assigned by Steneck and Dethier (1994): mi-
    croalgae = 1, filamentous algae = 2, foliose algae = 3, corticated
    foliose algae = 3.5, corticated macrophytes = 4, leathery macro-
    phytes = 5, articulated calcareous algae = 4 and crustose algae = 3.
    The habitat complexity score was then normalized to between 0
    and 1.

    2.3  |  Community production

    Community production and respiration of individual tiles were as-
    sessed by measuring, with an Orion 4-Star pH and dissolved oxy-
    gen meter (Thermo Scientific), the changes in dissolved oxygen
    concentrations during an incubation within a 2.5 L seawater con-
    tainer (15 cm wide × 20 cm length × 10 cm height) in a tempera-
    ture-controlled water bath. Magnetic stirrers (M-1 Controller and
    MS101A Stirrer, AS-One) were used to continuously mix the sea-
    water within each container throughout measurements. Seawater
    for each treatment (pHNBS 8.05 ± 0.01 SD vs. pHNBS 7.83 ± 0.01
    SD) was collected from the same location off-shore (oceanic pHNBS
    8.05) and treatments of pHNBS 7.83 ± 0.01 SD were acquired by
    bubbling with pure CO2 (Fukurow pH Controller; Aqua Geek).
    Community production and respiration were measured over a
    150 min period; first determining oxygen production (60 min light
    period, c. 200 μmol m−2 s−1), and after a 30 min dark period, oxygen
    consumption (60 min dark period). Net community production and
    community respiration were measured during the light and dark
    periods, respectively, with gross primary production calculated as
    the net community production minus community respiration. All
    three measurements are presented as mmol O2 h

    −1 m−2. To assess
    whether any changes in respiration rates were driven by an altered
    sessile invertebrate community for the 12 month communities, the

    T A B L E 1  Summary of the carbonate chemistry for the 350 and 900 μatm locations. The pHT (350 μatm, n = 1964; 900 μatm, n = 10,818),
    salinity (350 μatm, n = 1964; 900 μatm, n = 10,818) and total alkalinity (AT; 350 μatm, n = 56; 900 μatm, n = 47) are measured values. All
    other values were calculated using the carbonate chemistry system analysis program CO2SYS: Seawater pCO2, dissolved inorganic carbon
    (DIC), bicarbonate (HCO3

    −), carbonate (CO3
    2−), carbon dioxide (CO2), saturation states for calcite (Ωcalcite) and aragonite (Ωaragonite).

    Values are presented as mean, with standard deviation below

    Location pHT
    Salinity
    (psu)

    AT
    (μmol kg−1)

    pCO2
    (μatm)

    DIC
    (μmol kg−1)

    HCO3

    (μmol kg−1)
    CO3

    2−
    (μmol kg−1) Ωcalcite Ωaragonite

    ‘350 μatm’ 8.137 34.504 2264.29 316.057 1962.694 1740.629 211.979 5.087 3.301

    0.056 0.427 15.34 47.466 34.376 55.084 22.221 0.534 0.348

    ‘900 μatm’ 7.788 34.351 2268.33 841.148 2125.785 1984.889 115.150 2.771 1.805

    0.106 0.484 19.45 291.762 39.381 52.510 21.308 0.512 0.336

    Note: ‘350 μatm’ carbonate chemistry data are sourced from Agostini et al. (2018), ‘900 μatm’ carbonate chemistry data are averaged across Agostini
    et al. (2018) and original data collected between 2017/05/31 and 2017/08/08.

        |  5HARVEY Et Al.

    community composition and total cover of the sessile fauna (lo-
    cated on the underside of the tile) was assessed, following the same
    procedure as the community analysis (based on the percent cover
    of the sessile faunal species).

    2.4  |  Statistical analysis

    Statistical analyses were conducted using R (version 3.6.0; R Core
    Team, 2019), with the ‘vegan’ (Oksanen et al., 2019), ‘mvabund’
    (Wang et al., 2012) and base ‘stats’ package used for statistical anal-
    ysis, and the ‘ggplot2’ (Wickham, 2016) and ‘ggpubr’ (Kassambara,
    2019) packages used for figure production. For each of the analyses
    performed, the package and specific function used in R are listed
    below as ‘package::function’.

    For the seasonal experiment, differences in community compo-
    sition (based on percentage cover) between pCO2 (two levels: 350
    and 900 μatm) and Season (two levels: ‘Cold Period’ and ‘Warm
    Period’) were assessed using non-metric multidimensional scaling
    (nMDS; vegan::metaMDS) and a permutational analysis of variance
    (PERMANOVA) based on Bray–Curtis dissimilarity (vegan::vegdist
    and vegan::adonis). To test for differences in the percentage cover
    of the specific algal functional groups, we employed a two-way
    generalized linear model (GLM; Family: Binomial (Link = Logit)),
    with pCO2 (two levels: 350 and 900 μatm) and Season (two levels:
    ‘Cold Period’ and ‘Warm Period’) as fixed factors (stats::glm).

    For the reciprocal transplant experiment, differences in commu-
    nity composition (based on percentage cover) between Community
    Origin (two levels: 350 and 900 μatm), Community Destination (two
    levels: 350 and 900 μatm) and Time (two levels: 6 and 12 months) as
    fixed factors were assessed using an nMDS (vegan::metaMDS) and
    a PERMANOVA based on Bray–Curtis dissimilarity (vegan::vegdist
    and vegan::adonis). As the starting communities at 6 months could
    influence the resulting communities at 12 months on a particular
    tile, we accounted for this repeated measure in the PERMANOVA by
    using the ‘strata’ argument within vegan::adonis. To test for differ-
    ences in the percentage cover of the specific algal functional groups
    at 12 months, we employed a two-way GLM (Family: Binomial
    (Link = Logit)), with Community Origin (two levels: 350 and 900 μatm)
    and Community Destination (two levels: 350 and 900 μatm) as fixed
    factors (stats::glm).

    Differences in community production and respiration of the
    12 month communities were tested using a two-way GLM (Family:
    Gaussian (Link = Identity)), with Community Origin (two levels: 350
    and 900 μatm) and Community Destination (two levels: 350 and
    900 μatm) as fixed factors (stats::glm). To assess whether changes
    in respiration were driven by an altered sessile invertebrate commu-
    nity for the 12 month communities, we assessed for differences in
    the community composition (based on percentage cover) and total
    percentage cover of the sessile fauna. Differences in the community
    composition of the sessile fauna, with Community Origin (two lev-
    els: 350 and 900 μatm), and Community Destination (two levels: 350
    and 900 μatm) as fixed factors, were assessed using a PERMANOVA

    based on Bray–Curtis dissimilarity (vegan::vegdist and vegan::adonis).
    Differences in the total percentage cover of the sessile fauna were
    tested using a two-way GLM (Family: Gaussian (Link = Identity)),
    with Community Origin (two levels: 350 and 900 μatm) and
    Community Destination (two levels: 350 and 900 μatm) as fixed fac-
    tors (stats::glm).

    Post hoc comparisons for all PERMANOVAs were achieved
    using a Bonferroni-corrected pairwise PERMANOVA. For the re-
    ciprocal transplant, only comparisons chosen a priori were tested;
    these were between Time (two levels: 6 and 12 months) for each
    of the four combinations of Community Origin (two levels: 350
    and 900 μatm) and Community Destination (two levels: 350 and
    900 μatm). The assumptions of the generalized linear models
    (GLMs) were met, with the response variable independent, the
    mean–variance relationship suitable (assessed by plotting the
    residuals vs. fits; mvabund::manyglm) and dispersion parameters
    not being under- or over-dispersed (assessed using quasibinomial
    error distributions in R; stats::glm). Post hoc testing of the GLM
    for the gross primary production was achieved using TukeyHSD
    (multcomp::glht).

    3   |   R E S U LT S

    3.1  |  Effects of pCO2 and season on community
    composition

    Overall community composition following 6 months settlement
    was highly separated by nMDS, being significantly affected by both
    ‘pCO2’ and ‘Season’ (PERMANOVA: pCO2*Season, F1,25 = 4.351,
    p = 0.012; Figure 2a; Table 2). Community composition was simi-
    lar between seasons at 350 μatm (post hoc Bonferroni-adjusted,
    p = 0.102), but greatly differed between seasons at 900 μatm (post
    hoc Bonferroni-adjusted, p = 0.018). The typical newly settled com-
    munity at 350 μatm during the ‘Warm Period’ largely comprised
    of a low coverage (<10%) of microalgae and turf algae (Figure 2b) as well as corticated foliose algae, and a high coverage (~75%) of both corticated macrophytes (Figure 2c, including Chondracanthus tenellus (Harvey) Hommersand, 1993) and crustose coralline algae (Figure 2d, including Lithophyllum sp. Philippi, 1837). The community at 350 μatm during the ‘Cold Period’ was similar, the only difference being a significantly lower cover (approximately half) of corticated macrophytes relative to the ‘Warm Period’ (Figure 2c).

    Relative to 350 CO2, at 900 μatm the newly settled community
    during the ‘Warm Period’ showed increased coverage by microalgae
    and turf (from almost zero to ~30% coverage; Figure 2b), as well as
    increased corticated foliose algae (tenfold increase; including the
    species Zonaria diesingiana J. Agardh, 1841), but a threefold de-
    crease in corticated macrophytes (Figure 2c) and a fourfold decrease
    in crustose coralline algae (Figure 2d). This resulted in a community
    with roughly similar coverage of the microalgae and turf, corticated
    macrophytes and crustose coralline algae. During the ‘Cold Period’,
    the 900 μatm communities showed a high coverage of microalgae

    6  |    HARVEY Et Al.

    and turf algae (~72%; Figure 2b), but an absence of corticated foliose
    algae. Both the corticated macrophytes (Figure 2c) and crustose cor-
    alline algae (Figure 2d) were similar in coverage between the ‘Warm
    Period’ and ‘Cold Period’ at 900 μatm CO2.

    3.2  |  Effects of early-stage composition and pCO2
    on subsequent community succession

    After performing a reciprocal transplant of the established
    6 month communities for a further 6 months, it was found that

    regardless of the conditions under which the communities were
    established (i.e. ‘Community Origin’), the communities converged
    to form similar communities based on the pCO2 conditions they
    were currently residing in (i.e. ‘Community Destination’; Figure 3;
    Table 3).

    At 12 months the 350 μatm CO2 recruitment tiles typically had
    a corticated macrophyte upperstorey (Figure 4b, including Codium
    coactum Okamura, 1930 and Chondracanthus tenellus), with an un-
    derstorey of microalgae and turf algae (Figure 4a), as well as crustose
    coralline algae (Figure 4c, including Lithophyllum sp.). Both the ‘350
    to 350 μatm’ and the ‘900 to 350 μatm’ communities had a higher
    total cover than at 6 months (significant increase by ~40% cover for
    both). This was due to an increased cover of understorey microalgae
    and turf algae in the ‘350 to 350 μatm’ community (Figure 4a), and
    increased corticated macrophytes (Figure 4b; both as understorey
    and canopy) for both communities. Overall the ‘350 to 350 μatm’
    community was most complex, followed by the ‘900 to 350 μatm’
    community (Figure 4d).

    At 900 μatm CO2, a typical community after 12 months lacked
    canopy algae and was predominantly comprised of turf algae
    (Figure 4a, including Biddulphia biddulphiana (J.E. Smith) Boyer
    1900) with some corticated foliose algae (including Zonaria diesing-
    iana), and minimal cover of corticated macrophytes (Figure 4b) and

    F I G U R E 2  Community composition and percentage cover of algal functional groups. (a) nMDS of community composition based on
    algal functional groups. Treatments are displayed by pCO2 (‘350 μatm’ – blue; ‘900 μatm’ – red) and season (‘Warm Period’ – triangles; ‘Cold
    Period’ – crosses). (b–d) Percentage cover (%) of (b) microalgae and turf algae, (c) corticated macrophytes, and (d) crustose coralline algae
    following 6 months settlement at either 350 μatm pCO2 (‘Warm Period’ – darker blue, ‘Cold Period’ – lighter blue) or 900 μatm pCO2 (‘Warm
    Period’ – darker red, ‘Cold Period’ – lighter red). Two-way GLM (pCO2*Season) results are presented in the top-right of (b)–(d). See Table S1
    for more detailed statistics

    (a)

    (c)

    T A B L E 2  PERMANOVA summary of pCO2 (350 μatm vs.
    900 μatm) and Season (‘Cold Period’ vs. ‘Warm Period’) for the algal
    communities. For p-values, * (p < 0.05), ** (p < 0.01), *** (p < 0.001)

    Term df
    Sum
    Sq.

    Mean
    Sq. F p

    pCO2 1 1.995 1.995 32.58 0.001***

    Season 1 0.378 0.378 6.18 0.003**

    pCO2*Season 1 0.266 0.266 4.35 0.012*

    Residuals 22 1.347 0.061

    Total 25 3.986

        |  7HARVEY Et Al.

    crustose coralline algae (Figure 4c). The community that was trans-
    planted into 900 μatm (‘350 to 900 μatm’) showed a decline in total
    cover (significant decrease by ~45% cover), and a large reduction in
    complexity (Figure 4d) due to the loss of corticated macrophytes
    and crustose coralline algae (Figure 4b,c). The community that was

    consistently maintained at 900 μatm (‘900 to 900 μatm’) showed an
    increased coverage by turf and microalgae (Figure 4a), but overall
    displayed similar levels of total cover (~100% cover at 6 months and
    ~120% cover at 12 months) and complexity (Figure 4d) relative to
    the initial 6 month community.

    F I G U R E 3  nMDS of community
    composition based on algal functional
    groups. Communities are grouped as
    those exposed to 350 μatm throughout
    (darker blue), transplanted from 900 to
    350 μatm (lighter blue), transplanted from
    350 to 900 μatm (lighter red) and exposed
    to 900 μatm throughout (darker red). The
    initial starting 6 month communities are
    displayed with open symbols and dashed
    lines, and the 12 month communities
    following the transplant are displayed with
    solid symbols and lines

    Term df Sum Sq. Mean Sq. F p

    Origin 1 0.910 0.910 13.05 0.001***

    Destination 1 0.669 0.669 9.59 0.001***

    Time 1 1.267 1.267 18.17 0.001***

    Origin*Destination 1 0.141 0.141 2.03 0.002**

    Origin*Time 1 0.490 0.490 7.03 0.023*

    Destination*Time 1 1.16 1.16 16.66 0.003**

    Origin*Destination*Time 1 0.074 0.074 1.06 0.541

    Residuals 23 1.604 0.070

    Total 30 6.136

    T A B L E 3  PERMANOVA summary
    of Community Origin (‘Origin’: 350 vs.
    900 μatm), Community Destination
    (‘Destination’: 350 μatm vs. 900 μatm),
    and Time (‘6 months’ vs.’12 months’) for
    algal communities grown on settlement
    tiles off Shikine Island, Japan. For
    p-values, * (p < 0.05), ** (p < 0.01), *** (p < 0.001)

    F I G U R E 4  Percentage cover (%) of
    functional groups following a further
    6 month settlement: (a) microalgae and
    turf algae, (b) corticated macrophytes,
    (c) and crustose coralline algae. (d)
    structural complexity of the communities.
    Communities are grouped as those
    exposed to 350 μatm throughout (darker
    blue), transplanted from 900 to 350 μatm
    (lighter blue) or 350 to 900 μatm
    (lighter red) and exposed to 900 μatm
    throughout (darker red). Two-way GLM
    results are presented in the top-right
    of each. See Table S2 for more detailed
    statistics

    (a)

    (b)

    (c) (d)

    p = 0.59
    p < 0.01

    8  |    HARVEY Et Al.

    3.3  |  Effects of early-stage composition and pCO2
    on community production

    Net community production was reduced for the 350 μatm com-
    munities compared to the 900 μatm communities (Figure 5a).
    This was due to greatly increased community respiration rates
    for those communities at 350 μatm (Figure 5b), which resulted
    in the ‘350–350 μatm’ communities showing no net community
    production, and the ‘900–350 μatm’ communities having a nega-
    tive net community production (Figure 5a). Overall, the resulting
    communities had similar levels of gross community production
    (Figure 5c). A significant interaction was observed for gross com-
    munity production (Figure 5c); however, all post hoc comparisons
    were non-significant.

    3.4  |  Effects of early-stage composition and pCO2
    on sessile invertebrate community

    To assess whether increased respiration was driven by an altered
    sessile invertebrate community for the 12 month communities, the
    community composition and total cover of the sessile fauna was
    assessed. The sessile invertebrate community composition of the
    12 month communities was altered by both ‘Community Origin’ and
    ‘Community Destination’ (PERMANOVA: F1,14 = 15.07, p < 0.001 and F1,14 = 3.68, p = 0.042, respectively; Figure S2), with a greater cov- erage of ascidians (Didemnidae) and hydrozoans (Leptomedusae) in the elevated pCO2 conditions (‘350–900 μatm’ and ‘900–900 μatm’) and a greater coverage of polychaetes (Serpulidae) in the reference pCO2 conditions (‘350–350 μatm’ and ‘900–350 μatm’; see Figure S3). However, the overall coverage of sessile fauna did not significantly differ between the treatments (GLM: ‘Community Origin’ t = 1.17, p = 0.27 and ‘Community Destination’ t = −0.18, p = 0.86), with no interaction (GLM: ‘Community Origin’*‘Community Destination’ t = −0.07, p = 0.95).

    4   |   D I S C U S S I O N

    Ocean acidification alters the competitive abilities of algae and so is
    expected to change the structure, composition and functioning of
    both coastal and open ocean marine habitats (Cornwall et al., 2017;
    Hall-Spencer & Harvey, 2019). Observations at volcanic CO2 seeps in
    the photic zone have shown profound ecosystem shifts towards sim-
    plified non-calcareous communities that are often algal dominated
    with lower biodiversity and reduced ecological complexity (Agostini
    et al., 2018; Connell et al., 2018; Foo et al., 2018; González-Delgado
    & Hernández, 2018; Kroeker, Gambi, et al., 2013). The underlying
    processes by which ocean acidification affects the structure of
    shallow-water marine communities are not clearly established and
    require additional investigation, although significant progress has
    been made in recent years (see Kroeker, Gambi, et al., 2013; Kroeker
    et al., 2011, 2012; Porzio et al., 2013; Teixidó et al., 2018; Vizzini
    et al., 2017). We found that an enriched CO2 environment had a
    positive effect on r-selected, fast-growing microalgae and turf algal
    species in the early stages of community development. This species-
    poor, low complexity, early-successional stage was then locked-in as
    it inhibited the settlement and growth of corticated macrophytes.
    We highlight the potential ecological processes responsible for this
    change in a temperate rocky reef community exposed to enriched
    CO2 conditions.

    Ocean acidification altered competitive dominance after
    6 months with substrata in the acidified conditions being dominated
    by microalgae and turf algae, rather than the corticated macrophytes
    and crustose coralline algae that dominated in reference pCO2 condi-
    tions. This pattern was temporally consistent, showing that regardless
    of the growing season ocean acidification truncates the normal suc-
    cessional trajectories of communities (Baggini et al., 2014; Kroeker
    et al., 2012). The benefits of seawater acidification to opportunistic

    F I G U R E 5  (a) Net community production (mmol O2 h
    −1 m−2), (b)

    Community respiration (mmol O2 h
    −1 m−2), and (c) Gross community

    production (mmol O2 h
    −1 m−2). Communities are grouped as those

    exposed to 350 μatm throughout (darker blue), transplanted
    between locations (900 to 350 μatm, lighter blue; and 350 to
    900 μatm, lighter red), and exposed to 900 μatm throughout
    (darker red). See Table S3 for more detailed statistics

    (a)
    (b)
    (c)

    p
    p
    p

    p
    p
    p
    p
    p
    p

        |  9HARVEY Et Al.

    species, such as turf algae, over others (including calcareous species)
    is well established (Connell et al., 2013, 2018; Cornwall et al., 2017).
    Lowered carbonate saturation is a stressor to calcified macroalgae
    (Brinkman & Smith, 2015; Enochs et al., 2015; Fabricius et al., 2011;
    Kamenos et al., 2016) whilst turf algae and other fast-growing oppor-
    tunistic species can use the additional carbon from ocean acidifica-
    tion to grow and compete for resources (Harvey et al., 2019; Kroeker
    et al., 2012; Porzio et al., 2013), thereby attaining dominance (Connell
    et al., 2018). The shift from typical coastal habitat-forming species
    (such as corals and kelp forests) to turf algal dominance causes a loss
    of structural complexity and associated ecosystem services (O’Brien
    & Scheibling, 2018; Rogers et al., 2014).

    Before turf algae overgrew the recruitment tiles in acidified con-
    ditions (between 6 and 12 months, see Figure 3b), both crustose
    coralline algae and corticated macrophytes recruited onto the sub-
    strata. This suggests that the divergence in community composition
    was not due to limited recruitment or a physiological intolerance to
    the acidified conditions, but was driven by altered competitive in-
    teractions (Crook et al., 2016; Kroeker et al., 2012) and/or loss of
    compensatory processes (Connell et al., 2018; Ghedini et al., 2015).
    Although bottom-up control helps stimulate algal growth on coral
    and rocky reefs, grazing pressure determines whether turf algae
    dominate (Mumby et al., 2006). The grazing pressure of large benthic
    invertebrates in acidified sites is thought to be lowered due to phys-
    iological impacts (Calosi et al., 2013; O’Donnell et al., 2010) which
    cause their size and abundance to be reduced compared to the ref-
    erence pCO2 areas (Connell et al., 2018, White Island, New Zealand;
    Harvey et al., unpublished, this site and Vulcano, Italy). Similarly, the
    observed number of sea urchin feeding halos has also previously
    been found to be reduced in the Ischia CO2 seeps (Kroeker, Gambi,
    et al., 2013). Fish communities also play a key role in top-down con-
    trol and, in our site, communities included more herbivorous fishes
    than the surrounding non-acidified areas (Cattano et al., 2020).
    Clearly, the increased turf algae supported a greater herbivorous
    fish population than in the reference conditions, but those same
    herbivores alone are not able to control the increased growth of the
    boosted turf algae (also see Baggini et al., 2015; Connell et al., 2018).
    Although outside the scope of this study, this may support the no-
    tion that the fish are preferentially consuming different algae other
    than the turfs and further reinforcing the ecological shift.

    After 12 months, assemblages in reference pCO2 conditions
    continued to gain species through time and had developed more
    structurally complex communities with clearly defined understorey
    and canopy species. The assemblages in the elevated pCO2 became
    arrested in terms of their successional development due to competi-
    tion for space by the turf algae. A similar overgrowth and dominance
    by turf algae was observed on recruitment tiles in the Ischia CO2
    seep (Kroeker et al., 2012; Porzio et al., 2013). This community de-
    velopment in our study resulted in a similar community composition
    at 6 and 12 months with only the abundance of the turf algae being
    increased. At 12 months, communities on the tiles were visually in-
    distinguishable from the surrounding rocky substrata (Figure S4).
    Similar declines in macroalgal diversity with increasing pCO2 have

    also been demonstrated in Methana, Greece (Baggini et al., 2014).
    The simplification of marine ecosystems has been observed across
    CO2 seeps (Agostini et al., 2018; Brown et al., 2018; Cigliano et al.,
    2010; Fabricius et al., 2011; Kroeker, Gambi, et al., 2013; Vizzini
    et al., 2017), with such changes leading to a functional biodiversity
    loss in the system (Teixidó et al., 2018). It is likely that such sim-
    ple systems are maintained by reinforcing feedback loops (sediment
    trapping, changes in physicochemical environment, and recruit-
    ment inhibition) that facilitate turf algal dominance. Turf algae can
    inhibit successional development by reducing primary substratum
    availability (Airoldi, 1998; Connell & Russell, 2010) and by trapping
    sediment which alters settlement surface chemistry and reduces the
    survival of other recruits (Airoldi, 2003; Gorman & Connell, 2009).
    Such dominance by short-lived species, which then locks the system
    in place, can lead to decreased stability in the system (Stachowicz
    et al., 2007), with implications for the functioning of the system
    under future ocean acidification (Teixidó et al., 2018).

    In terms of community dynamics, both the reference and ele-
    vated pCO2 conditions appeared to overwhelm any ecological re-
    sistance that would have otherwise resisted ecosystem change.
    This was demonstrated by the established algal communities that
    were transplanted from reference to elevated pCO2 conditions con-
    verging (in terms of community composition) to almost match the
    community formed under elevated pCO2 conditions (and vice versa).
    This suggests that acidification-driven ecological shifts to simplified
    turf algae communities will occur regardless of the state that the
    community is in, and means that the community successional trajec-
    tory is not fixed from the initial bare substratum during primary or
    secondary succession. The prevention of such shifts by ecosystem
    management will require a focused effort on resilience building in
    order to mitigate the future degradation of ecosystems (Billé et al.,
    2013; Falkenberg et al., 2013). In contrast, the convergence of the
    communities transplanted from elevated pCO2 conditions to the ref-
    erence pCO2 conditions could mean that recovery from a degraded
    state is possible. This would likely be due to sufficient compensatory
    processes at our reference pCO2 location, and/or the turf algae los-
    ing its competitive edge in the absence of elevated pCO2. Therefore,
    a combination of conservation strategy and meaningful reductions
    in atmospheric CO2 emissions could achieve substantial recovery
    of the abundance, structure and function of shallow coastal marine
    ecosystems (Duarte et al., 2020).

    Despite possessing highly divergent communities, gross oxygen
    production was similar between all of the transplanted tiles. Net ox-
    ygen production, however, was positive in the acidified conditions,
    but balanced between productivity and respiration for the reference
    pCO2 communities due to elevated respiration. Ecosystems that
    are more developed and stable will tend towards rates of oxygen
    production and respiration being equal, tending to not accumulate
    further biomass. Early-stage ecosystems will tend to have a higher
    productivity per biomass, but will be lacking in terms of biomass and
    species diversity (Cooke, 1967). This further supports the concept
    that the algal community developing under elevated pCO2 is ar-
    rested into a typical early-stage community dominated by r-selected

    10  |    HARVEY Et Al.

    species. Previous studies in CO2 seeps have generally focussed on
    the primary production or photophysiology of individual species
    of algae (e.g. Celis-Plá et al., 2015; Porzio et al., 2018, 2020) with
    the aim of assessing their physiological response to ocean acidifica-
    tion, rather than the effects on overall community net production
    (making comparisons difficult). The sessile invertebrate communi-
    ties differed in community composition between the reference and
    elevated pCO2 sites, but not in percentage cover, suggesting that
    they were not a sizeable contributor toward such large changes in
    net oxygen production. Instead, the decreased net oxygen was likely
    driven by the greater algal biomass (as well as low surface to vol-
    ume ratio) of the more highly structurally complex reference com-
    munity compared to the high surface to biomass ratio of the turf
    algae. Taken together, this suggests that the greater net production
    stimulated by ocean acidification does not translate into enhanced
    ecosystem benefits, such as increased community cover, biomass,
    biodiversity or structural complexity, as well as an altered sessile in-
    vertebrate community.

    Natural analogues provide a number of benefits for advancing
    our understanding about the responses of shallow-water marine
    communities to ocean acidification conditions, but they are not
    perfect analogues. Carbonate chemistry at some CO2 seeps can be
    highly variable (Rastrick et al., 2018), and areas in close proximity
    to CO2 vents can be enriched in some metals and toxins (Vizzini
    et al., 2013; Zitoun et al., 2020). It is possible to reduce such con-
    founding factors by avoiding toxic areas and only selecting sites a
    suitable distance away, since contamination from hydrothermal flu-
    ids can be quickly diluted by mixing with seawater (Agostini et al.,
    2015; Pichler et al., 2019). The gas being released at our study site
    is 98 ± 3 (SD) % CO2, and although concentrations of hydrogen
    sulfide are detected at the main vent, they are below detection
    limits ~50 m away from the main vents (Agostini et al., 2015) and
    the study site used in this study is more than 300 m away from
    the main vent. The possibility remains that other trace elements
    or heavy metals (as yet unmeasured) may be present either in the
    water or bioaccumulated in biota, as has been shown in other CO2
    seeps (Mirasole et al., 2020; Mishra et al., 2020; Vizzini et al., 2013;
    Zitoun et al., 2020), which could influence the response of marine
    organisms to ocean acidification. An additional consideration for
    CO2 seeps is that they demonstrate the consequences of future
    ocean acidification but in the absence of concurrent ocean warming
    (Hughes et al., 2017), and temperatures will mediate the response
    of organisms and communities to future ocean acidification. Such
    an issue can be addressed by comparing CO2 seep systems under
    different thermal regimes and assess the consistency of responses
    (Johnson et al., 2012), or by manipulating temperature along CO2
    gradients (Alessi et al., 2019). Despite these caveats, the use of CO2
    seeps is still invaluable for providing a window into the future state
    of organisms, communities and ecosystems to future ocean acidifi-
    cation (Rastrick et al., 2018).

    In conclusion, ocean acidification can set the course of suc-
    cessional development in algal communities, benefitting turf algae
    whilst causing reduced algal biomass, diversity and complexity.

    Altered carbonate chemistry can enable opportunistic r-selected
    species to competitively exclude other species and lock the com-
    munity in a species-poor early successional stage. The ecological
    process responsible for this shift in community composition was
    not simply altering community trajectory during primary succes-
    sion, as the same shift occurred in preestablished communities.
    This highlights that without reducing atmospheric CO2 emissions
    we may increasingly observe the loss of large algal habitats and the
    spread of fast-growing, small opportunistic species that can utilize
    additional inorganic carbon. By understanding the ecological pro-
    cesses responsible for driving shifts in community composition, we
    can begin to better assess how communities are likely to be altered
    by ocean acidification. Finally, our results show that the recovery of
    shallow-water marine communities is possible if meaningful reduc-
    tions in CO2 emissions are implemented, as encouraged by the Paris
    Agreement.

    A C K N O W L E D G E M E N T S
    We thank the technical staff at ‘Shimoda Marine Research Center,
    University of Tsukuba’ for their assistance aboard RV Tsukuba II and
    at the study site, and the Japan Fisheries agencies of Nijima/Shikine
    Island (Tokyo prefecture) for their support. This project was heavily
    supported by the ‘International Education and Research Laboratory
    Program’, University of Tsukuba. This work was also supported by
    JSPS KAKENHI Grant Number 17K17622, and we acknowledge
    funding support from the Ministry of Environment, Government of
    Japan (Suishinhi: 4RF-1701).

    Some of the images used within the graphical abstract are courtesy
    of the Integration and Application Network, University of Maryland
    Center for Environmental Science (ian.umces.edu/symbo ls/).

    C O N F L I C T O F I N T E R E S T
    The authors declare no conflicts of interest.

    A U T H O R C O N T R I B U T I O N S
    B.P.H. conceived the idea, designed the methodology, analysed
    the data and led the writing of the manuscript. B.P.H. and K.K.
    carried out the image analysis. B.P.H. and S.A. performed the
    oxygen production measurements. All authors assisted with field
    work, contributed critically to the drafts and gave final approval
    for publication.

    D ATA AVA I L A B I L I T Y S TAT E M E N T
    Biological data (Figures 2–5 and Tables 2 and 3) and associated car-
    bonate chemistry data (Figure 1; Figure S1; Table 1) are stored on
    Zenodo (https://doi.org/10.5281/zenodo.4280018).

    O R C I D
    Ben P. Harvey https://orcid.org/0000-0002-4971-1634
    Koetsu Kon https://orcid.org/0000-0003-1379-0702
    Sylvain Agostini https://orcid.org/0000-0001-9040-9296
    Shigeki Wada https://orcid.org/0000-0001-6893-7498
    Jason M. Hall-Spencer https://orcid.org/0000-0002-6915-2518

    http://ian.umces.edu/symbols/

    https://doi.org/10.5281/zenodo.4280018

    https://orcid.org/0000-0002-4971-1634

    https://orcid.org/0000-0002-4971-1634

    https://orcid.org/0000-0003-1379-0702

    https://orcid.org/0000-0003-1379-0702

    https://orcid.org/0000-0001-9040-9296

    https://orcid.org/0000-0001-9040-9296

    https://orcid.org/0000-0001-6893-7498

    https://orcid.org/0000-0001-6893-7498

    https://orcid.org/0000-0002-6915-2518

    https://orcid.org/0000-0002-6915-2518

        |  11HARVEY Et Al.

    R E F E R E N C E S
    Abràmoff, M. D., Magalhães, P. J., & Ram, S. J. (2004). Image processing

    with ImageJ. Biophotonics International, 11(7), 36–42.
    Agostini, S., Harvey, B. P., Wada, S., Kon, K., Milazzo, M., Inaba, K., &

    Hall-Spencer, J. M. (2018). Ocean acidification drives community
    shifts towards simplified non-calcified habitats in a subtropi-
    cal-temperate transition zone. Scientific Reports, 8, 11354. https://
    doi.org/10.1038/s4159 8-018-29251 -7

    Agostini, S., Wada, S., Kon, K., Omori, A., Kohtsuka, H., Fujimura, H.,
    Tsuchiya, Y., Sato, T., Shinagawa, H., Yamada, Y., & Inaba, K. (2015).
    Geochemistry of two shallow CO2 seeps in Shikine Island (Japan)
    and their potential for ocean acidification research. Regional Studies
    in Marine Science, 2(Suppl.), 45–53.

    Airoldi, L. (1998). Roles of disturbance, sediment stress, and substratum re-
    tention on spatial dominance in algal turf. Ecology, 79(8), 2759–2770.

    Airoldi, L. (2003). The effects of sedimentation on rocky coast assem-
    blages. Oceanography and Marine Biology: An Annual Review, 41,
    169–171.

    Albright, R., Caldeira, L., Hosfelt, J., Kwiatkowski, L., Maclaren, J. K.,
    Mason, B. M., Nebuchina, Y., Ninokawa, A., Pongratz, J., Ricke, K.
    L., Rivlin, T., Schneider, K., Sesboüé, M., Shamberger, K., Silverman,
    J., Wolfe, K., Zhu, K., & Caldeira, K. (2016). Reversal of ocean acid-
    ification enhances net coral reef calcification. Nature, 531(7594),
    362–365.

    Albright, R., Takeshita, Y., Koweek, D. A., Ninokawa, A., Wolfe, K.,
    Rivlin, T., Nebuchina, Y., Young, J., & Caldeira, K. (2018). Carbon
    dioxide addition to coral reef waters suppresses net community
    calcification. Nature, 555, 516–519. https://doi.org/10.1038/
    natur e25968

    Alessi, C., Giomi, F., Furnari, F., Sarà, G., Chemello, R., & Milazzo, M. (2019).
    Ocean acidification and elevated temperature negatively affect re-
    cruitment, oxygen consumption and calcification of the reef-building
    Dendropoma cristatum early life stages: Evidence from a manipula-
    tive field study. Science of The Total Environment, 693, 133476.

    Algueró-Muñiz, M., Alvarez-Fernandez, S., Thor, P., Bach, L. T., Esposito,
    M., Horn, H. G., Ecker, U., Langer, J. A. F., Taucher, J., Malzahn, A.
    M., Riebesell, U., & Boersma, M. (2017). Ocean acidification ef-
    fects on mesozooplankton community development: Results from
    a long-term mesocosm experiment. PLoS One, 12(4), e0175851.
    https://doi.org/10.1371/journ al.pone.0175851

    Baggini, C., Issaris, Y., Salomidi, M., & Hall-Spencer, J. (2015). Herbivore
    diversity improves benthic community resilience to ocean acidi-
    fication. Journal of Experimental Marine Biology and Ecology, 469,
    98–104. https://doi.org/10.1016/j.jembe.2015.04.019

    Baggini, C., Salomidi, M., Voutsinas, E., Bray, L., Krasakopoulou, E., &
    Hall-Spencer, J. M. (2014). Seasonality affects macroalgal commu-
    nity response to increases in pCO2. PLoS One, 9(9), e106520.

    Billé, R., Kelly, R., Biastoch, A., Harrould-Kolieb, E., Herr, D., Joos, F.,
    Kroeker, K., Laffoley, D., Oschlies, A., & Gattuso, J.-P. (2013). Taking
    action against ocean acidification: A review of management and
    policy options. Environmental Management, 52(4), 761–779. https://
    doi.org/10.1007/s0026 7-013-0132-7

    Brinkman, T. J., & Smith, A. M. (2015). Effect of climate change on crus-
    tose coralline algae at a temperate vent site, White Island, New
    Zealand. Marine and Freshwater Research, 66(4), 360–370.

    Brown, N. E. M., Milazzo, M., Rastrick, S. P. S., Hall-Spencer, J. M.,
    Therriault, T. W., & Harley, C. D. G. (2018). Natural acidification
    changes the timing and rate of succession, alters community struc-
    ture, and increases homogeneity in marine biofouling communities.
    Global Change Biology, 24(1), e112–e127. https://doi.org/10.1111/
    gcb.13856

    Brown, N. E. M., Therriault, T. W., & Harley, C. D. G. (2016). Field-based
    experimental acidification alters fouling community structure and
    reduces diversity. Journal of Animal Ecology, 85(5), 1328–1339.
    https://doi.org/10.1111/1365-2656.12557

    Calosi, P., Rastrick, S. P. S., Graziano, M., Thomas, S. C., Baggini, C.,
    Carter, H. A., Hall-Spencer, J. M., Milazzo, M., & Spicer, J. I. (2013).
    Distribution of sea urchins living near shallow water CO2 vents is
    dependent upon species acid-base and ion-regulatory abilities.
    Marine Pollution Bulletin, 73(2), 470–484.

    Cattano, C., Agostini, S., Harvey, B. P., Wada, S., Quattrocchi, F., Turco,
    G., Inaba, K., Hall-Spencer, J. M., & Milazzo, M. (2020). Changes in
    fish communities due to benthic habitat shifts under ocean acidi-
    fication conditions. Science of The Total Environment, 725, 138501.
    https://doi.org/10.1016/j.scito tenv.2020.138501

    Celis-Plá, P. S. M., Hall-Spencer, J. M., Horta, P. A., Milazzo, M.,
    Korbee, N., Cornwall, C. E., & Figueroa, F. L. (2015). Macroalgal
    responses to ocean acidification depend on nutrient and light
    levels. Frontiers in Marine Science, 2, 26. https://doi.org/10.3389/
    fmars.2015.00026

    Cigliano, M., Gambi, M. C., Rodolfo-Metalpa, R., Patti, F. P., & Hall-
    Spencer, J. M. (2010). Effects of ocean acidification on invertebrate
    settlement at volcanic CO2 vents. Marine Biology, 157(11), 2489–
    2502. https://doi.org/10.1007/s0022 7-010-1513-6

    Connell, J. H., & Slatyer, R. O. (1977). Mechanisms of succession in nat-
    ural communities and their role in community stability and organi-
    zation. The American Naturalist, 111(982), 1119–1144. https://doi.
    org/10.1086/283241

    Connell, S. D., Doubleday, Z. A., Foster, N. R., Hamlyn, S. B., Harley, C.
    D., Helmuth, B., Kelaher, B. P., Nagelkerken, I., Rodgers, K. L., Sarà,
    G., & Russell, B. D. (2018). The duality of ocean acidification as a
    resource and a stressor. Ecology, 99(5), 1005–1010. https://doi.
    org/10.1002/ecy.2209

    Connell, S. D., Kroeker, K. J., Fabricius, K. E., Kline, D. I., & Russell, B.
    D. (2013). The other ocean acidification problem: CO2 as a re-
    source among competitors for ecosystem dominance. Philosophical
    Transactions of the Royal Society B: Biological Sciences, 368(1627),
    20120442. https://doi.org/10.1098/rstb.2012.0442

    Connell, S. D., & Russell, B. D. (2010). The direct effects of increasing
    CO2 and temperature on non-calcifying organisms: Increasing the
    potential for phase shifts in kelp forests. Proceedings of the Royal
    Society B: Biological Sciences, 277(1686), 1409–1415.

    Cooke, G. D. (1967). The pattern of autotrophic succession in lab-
    oratory microcosms. BioScience, 17(10), 717–721. https://doi.
    org/10.2307/1294089

    Cornwall, C. E., Revill, A. T., Hall-Spencer, J. M., Milazzo, M., Raven, J.
    A., & Hurd, C. L. (2017). Inorganic carbon physiology underpins
    macroalgal responses to elevated CO2. Scientific Reports, 7, 46297.
    https://doi.org/10.1038/srep4 6297

    Crook, E. D., Kroeker, K. J., Potts, D. C., Rebolledo-Vieyra, M., Hernandez-
    Terrones, L. M., & Paytan, A. (2016). Recruitment and succession in
    a tropical benthic community in response to in-situ ocean acidifi-
    cation. PLoS One, 11(1), e0146707. https://doi.org/10.1371/journ
    al.pone.0146707

    Done, T. (1992). Effects of tropical cyclone waves on ecological and geo-
    morphological structures on the Great Barrier Reef. Continental
    Shelf Research, 12(7–8), 859–872.

    Duarte, C. M., Agusti, S., Barbier, E., Britten, G. L., Castilla, J. C., Gattuso,
    J.-P., Fulweiler, R. W., Hughes, T. P., Knowlton, N., Lovelock, C. E.,
    Lotze, H. K., Predragovic, M., Poloczanska, E., Roberts, C., & Worm,
    B. (2020). Rebuilding marine life. Nature, 580(7801), 39–51. https://
    doi.org/10.1038/s4158 6-020-2146-7

    Enochs, I. C., Manzello, D. P., Donham, E. M., Kolodziej, G., Okano,
    R., Johnston, L., Young, C., Iguel, J., Edwards, C. B., Fox, M. D.,
    Valentino, L., Johnson, S., Benavente, D., Clark, S. J., Carlton, R.,
    Burton, T., Eynaud, Y., & Price, N. N. (2015). Shift from coral to
    macroalgae dominance on a volcanically acidified reef. Nature
    Climate Change, 5(12), 1083–1088. https://doi.org/10.1038/nclim
    ate2758

    Fabricius, K. E., Langdon, C., Uthicke, S., Humphrey, C., Noonan, S.,
    De’ath, G., Okazaki, R., Muehllehner, N., Glas, M. S., & Lough, J.

    https://doi.org/10.1038/s41598-018-29251-7

    https://doi.org/10.1038/s41598-018-29251-7

    https://doi.org/10.1038/nature25968

    https://doi.org/10.1038/nature25968

    https://doi.org/10.1371/journal.pone.0175851

    https://doi.org/10.1016/j.jembe.2015.04.019

    https://doi.org/10.1007/s00267-013-0132-7

    https://doi.org/10.1007/s00267-013-0132-7

    https://doi.org/10.1111/gcb.13856

    https://doi.org/10.1111/gcb.13856

    https://doi.org/10.1111/1365-2656.12557

    https://doi.org/10.1016/j.scitotenv.2020.138501

    https://doi.org/10.3389/fmars.2015.00026

    https://doi.org/10.3389/fmars.2015.00026

    https://doi.org/10.1007/s00227-010-1513-6

    https://doi.org/10.1086/283241

    https://doi.org/10.1086/283241

    https://doi.org/10.1002/ecy.2209

    https://doi.org/10.1002/ecy.2209

    https://doi.org/10.1098/rstb.2012.0442

    https://doi.org/10.2307/1294089

    https://doi.org/10.2307/1294089

    https://doi.org/10.1038/srep46297

    https://doi.org/10.1371/journal.pone.0146707

    https://doi.org/10.1371/journal.pone.0146707

    https://doi.org/10.1038/s41586-020-2146-7

    https://doi.org/10.1038/s41586-020-2146-7

    https://doi.org/10.1038/nclimate2758

    https://doi.org/10.1038/nclimate2758

    12  |    HARVEY Et Al.

    M. (2011). Losers and winners in coral reefs acclimatized to ele-
    vated carbon dioxide concentrations. Nature Climate Change, 1(3),
    165–169. https://doi.org/10.1038/nclim ate1122

    Falkenberg, L. J., Connell, S. D., & Russell, B. D. (2013). Disrupting the
    effects of synergies between stressors: Improved water quality
    dampens the effects of future CO2 on a marine habitat. Journal of
    Applied Ecology, 50(1), 51–58. https://doi.org/10.1111/1365-2664.
    12019

    Foo, S. A., Byrne, M., Ricevuto, E., & Gambi, M. C. (2018). The carbon
    dioxide vents of Ischia, Italy, a natural system to assess impacts
    of ocean acidification on marine ecosystems: An overview of re-
    search and comparisons with other vent systems. Oceanography
    and Marine Biology: An Annual Review, 56, 237–310. https://doi.
    org/10.1201/97804 29454 455-9

    Garilli, V., Rodolfo-Metalpa, R., Scuderi, D., Brusca, L., Parrinello, D.,
    Rastrick, S. P. S., Foggo, A., Twitchett, R. J., Hall-Spencer, J. M., &
    Milazzo, M. (2015). Physiological advantages of dwarfing in surviv-
    ing extinctions in high-CO2 oceans. Nature Climate Change, 5(7),
    678–682. https://doi.org/10.1038/nclim ate2616

    Gattuso, J.-P., Magnan, A., Billé, R., Cheung, W. W. L., Howes, E. L.,
    Joos, F., Allemand, D., Bopp, L., Cooley, S. R., Eakin, C. M., Hoegh-
    Guldberg, O., Kelly, R. P., Pörtner, H.-O., Rogers, A. D., Baxter, J.
    M., Laffoley, D., Osborn, D., Rankovic, A., Rochette, J., … Turley,
    C. (2015). Contrasting futures for ocean and society from differ-
    ent anthropogenic CO2 emissions scenarios. Science, 349(6243),
    aac4722. https://doi.org/10.1126/scien ce.aac4722

    Gaylord, B., Kroeker, K. J., Sunday, J. M., Anderson, K. M., Barry, J.
    P., Brown, N. E., Connell, S. D., Dupont, S., Fabricius, K. E., Hall-
    Spencer, J. M., Klinger, T., Milazzo, M., Munday, P. L., Russell, B.
    D., Sanford, E., Schreiber, S. J., Thiyagarajan, V., Vaughan, M. L.
    H., Widdicombe, S., & Harley, C. D. G. (2015). Ocean acidification
    through the lens of ecological theory. Ecology, 96(1), 3–15.

    Ghedini, G., & Connell, S. D. (2016). Organismal homeostasis buffers the
    effects of abiotic change on community dynamics. Ecology, 97(10),
    2671–2679. https://doi.org/10.1002/ecy.1488

    Ghedini, G., Russell, B. D., & Connell, S. D. (2015). Trophic compensation
    reinforces resistance: Herbivory absorbs the increasing effects of
    multiple disturbances. Ecology Letters, 18(2), 182–187. https://doi.
    org/10.1111/ele.12405

    Godbold, J. A., & Solan, M. (2013). Long-term effects of warming and
    ocean acidification are modified by seasonal variation in species re-
    sponses and environmental conditions. Philosophical Transactions of
    the Royal Society B: Biological Sciences, 368(1627), 20130186.

    González-Delgado, S., & Hernández, J. C. (2018). The importance of
    natural acidified systems in the study of ocean acidification: What
    have we learned? Advances in Marine Biology, 80, 57–99. https://doi.
    org/10.1016/bs.amb.2018.08.001

    Gordillo, F. J., Figueroa, F. L., & Niell, F. X. (2003). Photon-and carbon-use
    efficiency in Ulva rigida at different CO2 and N levels. Planta, 218(2),
    315–322.

    Gorman, D., & Connell, S. D. (2009). Recovering subtidal forests in hu-
    man-dominated landscapes. Journal of Applied Ecology, 46(6), 1258–
    1265. https://doi.org/10.1111/j.1365-2664.2009.01711.x

    Gruner, D. S., Smith, J. E., Seabloom, E. W., Sandin, S. A., Ngai, J. T.,
    Hillebrand, H., Harpole, W. S., Elser, J. J., Cleland, E. E., Bracken, M.
    E. S., Borer, E. T., & Bolker, B. M. (2008). A cross-system synthesis
    of consumer and nutrient resource control on producer biomass.
    Ecology Letters, 11(7), 740–755.

    Hall-Spencer, J. M., & Harvey, B. P. (2019). Ocean acidification impacts
    on coastal ecosystem services due to habitat degradation. Emerging
    Topics in Life Sciences, 3(2), 197–206. https://doi.org/10.1042/
    ETLS2 0180117

    Hall-Spencer, J. M., Rodolfo-Metalpa, R., Martin, S., Ransome, E., Fine, M.,
    Turner, S. M., Rowley, S. J., Tedesco, D., & Buia, M.-C. (2008). Volcanic
    carbon dioxide vents show ecosystem effects of ocean acidification.
    Nature, 454(7200), 96–99. https://doi.org/10.1038/natur e07051

    Harvey, B. P., Agostini, S., Kon, K., Wada, S., & Hall-Spencer, J. M. (2019).
    Diatoms dominate and alter marine food-webs when CO2 rises.
    Diversity, 11(12). https://doi.org/10.3390/d1112 0242

    Harvey, B. P., Agostini, S., Wada, S., Inaba, K., & Hall-Spencer, J. M. (2018).
    Dissolution: The Achilles’ heel of the triton shell in an acidifying
    ocean. Frontiers in Marine Science, 5, 371. https://doi.org/10.3389/
    fmars.2018.00371

    Harvey, B. P., Gwynn-Jones, D., & Moore, P. J. (2013). Meta-analysis
    reveals complex marine biological responses to the interactive
    effects of ocean acidification and warming. Ecology and Evolution,
    3(4), 1016–1030. https://doi.org/10.1002/ece3.516

    Harvey, B. P., McKeown, N. J., Rastrick, S. P. S., Bertolini, C., Foggo, A.,
    Graham, H., Hall-Spencer, J. M., Milazzo, M., Shaw, P. W., Small, D.
    P., & Moore, P. J. (2016). Individual and population-level responses
    to ocean acidification. Scientific Reports, 6, 20194. https://doi.
    org/10.1038/srep2 0194

    Hillebrand, H., Gruner, D. S., Borer, E. T., Bracken, M. E. S., Cleland, E.
    E., Elser, J. J., Harpole, W. S., Ngai, J. T., Seabloom, E. W., Shurin,
    J. B., & Smith, J. E. (2007). Consumer versus resource control of
    producer diversity depends on ecosystem type and producer com-
    munity structure. Proceedings of the National Academy of Sciences
    of the United States of America, 104(26), 10904–10909. https://doi.
    org/10.1073/pnas.07019 18104

    Hughes, T. P., Barnes, M. L., Bellwood, D. R., Cinner, J. E., Cumming, G.
    S., Jackson, J. B. C., Kleypas, J., van de Leemput, I. A., Lough, J. M.,
    Morrison, T. H., Palumbi, S. R., van Nes, E. H., & Scheffer, M. (2017).
    Coral reefs in the Anthropocene. Nature, 546(7656), 82–90. https://
    doi.org/10.1038/natur e22901

    Hughes, T. P., Rodrigues, M. J., Bellwood, D. R., Ceccarelli, D., Hoegh-
    Guldberg, O., McCook, L., Moltschaniwskyj, N., Pratchett, M. S.,
    Steneck, R. S., & Willis, B. (2007). Phase shifts, herbivory, and the
    resilience of coral reefs to climate change. Current Biology, 17(4),
    360–365. https://doi.org/10.1016/j.cub.2006.12.049

    IPCC. (2013). Climate change 2013 – The physical science basis: Working
    group I contribution to the fifth assessment report of the IPCC (No.
    0521880092; p. 1535). Cambridge University Press.

    Jenkins, S. R., Hawkins, S. J., & Norton, T. A. (1999). Interaction between
    a fucoid canopy and limpet grazing in structuring a low shore inter-
    tidal community. Journal of Experimental Marine Biology and Ecology,
    233(1), 41–63.

    Johnson, C. R., & Mann, K. H. (1988). Diversity, patterns of adaptation,
    and stability of Nova Scotian kelp beds. Ecological Monographs,
    58(2), 129–154. https://doi.org/10.2307/1942464

    Johnson, V. R., Russell, B. D., Fabricius, K. E., Brownlee, C., & Hall-
    Spencer, J. M. (2012). Temperate and tropical brown macroalgae
    thrive, despite decalcification, along natural CO2 gradients. Global
    Change Biology, 18(9), 2792–2803.

    Kamenos, N. A., Perna, G., Gambi, M. C., Micheli, F., & Kroeker, K. J.
    (2016). Coralline algae in a naturally acidified ecosystem persist
    by maintaining control of skeletal mineralogy and size. Proceedings
    of the Royal Society B: Biological Sciences, 283(1840), 20161159.
    https://doi.org/10.1098/rspb.2016.1159

    Kassambara, A. (2019). ggpubr: “ggplot2” based publication ready plots. R
    Package Version 0.2.4. https://CRAN.R-proje ct.org/packa ge=ggpubr

    Kelly, E. L. A., Eynaud, Y., Clements, S. M., Gleason, M., Sparks, R. T., Williams,
    I. D., & Smith, J. E. (2016). Investigating functional redundancy versus
    complementarity in Hawaiian herbivorous coral reef fishes. Oecologia,
    182(4), 1151–1163. https://doi.org/10.1007/s0044 2-016-3724-0

    Kerfahi, D., Harvey, B. P., Agostini, S., Kon, K., Huang, R., Adams, J. M., &
    Hall-Spencer, J. M. (2020). Responses of intertidal bacterial biofilm
    communities to increasing pCO2. Marine Biotechnology. https://doi.
    org/10.1007/s1012 6-020-09958 -3

    Koch, M., Bowes, G., Ross, C., & Zhang, X.-H. (2013). Climate change and
    ocean acidification effects on seagrasses and marine macroalgae.
    Global Change Biology, 19(1), 103–132. https://doi.org/10.1111/
    j.1365-2486.2012.02791.x

    https://doi.org/10.1038/nclimate1122

    https://doi.org/10.1111/1365-2664.12019

    https://doi.org/10.1111/1365-2664.12019

    https://doi.org/10.1201/9780429454455-9

    https://doi.org/10.1201/9780429454455-9

    https://doi.org/10.1038/nclimate2616

    https://doi.org/10.1126/science.aac4722

    https://doi.org/10.1002/ecy.1488

    https://doi.org/10.1111/ele.12405

    https://doi.org/10.1111/ele.12405

    https://doi.org/10.1016/bs.amb.2018.08.001

    https://doi.org/10.1016/bs.amb.2018.08.001

    https://doi.org/10.1111/j.1365-2664.2009.01711.x

    https://doi.org/10.1042/ETLS20180117

    https://doi.org/10.1042/ETLS20180117

    https://doi.org/10.1038/nature07051

    https://doi.org/10.3390/d11120242

    https://doi.org/10.3389/fmars.2018.00371

    https://doi.org/10.3389/fmars.2018.00371

    https://doi.org/10.1002/ece3.516

    https://doi.org/10.1038/srep20194

    https://doi.org/10.1038/srep20194

    https://doi.org/10.1073/pnas.0701918104

    https://doi.org/10.1073/pnas.0701918104

    https://doi.org/10.1038/nature22901

    https://doi.org/10.1038/nature22901

    https://doi.org/10.1016/j.cub.2006.12.049

    https://doi.org/10.2307/1942464

    https://doi.org/10.1098/rspb.2016.1159

    https://CRAN.R-project.org/package=ggpubr

    https://doi.org/10.1007/s00442-016-3724-0

    https://doi.org/10.1007/s10126-020-09958-3

    https://doi.org/10.1007/s10126-020-09958-3

    https://doi.org/10.1111/j.1365-2486.2012.02791.x

    https://doi.org/10.1111/j.1365-2486.2012.02791.x

        |  13HARVEY Et Al.

    Kroeker, K. J., Gambi, M. C., & Micheli, F. (2013). Community dynam-
    ics and ecosystem simplification in a high-CO2 ocean. Proceedings
    of the National Academy of Sciences of the United States of America,
    110(31), 12721–12726. https://doi.org/10.1073/pnas.12164 64110

    Kroeker, K. J., Kordas, R. L., Crim, R., Hendriks, I. E., Ramajo, L., Singh, G.
    S., Duarte, C. M., & Gattuso, J.-P. (2013). Impacts of ocean acidifica-
    tion on marine organisms: Quantifying sensitivities and interaction
    with warming. Global Change Biology, 19(6), 1884–1896. https://doi.
    org/10.1111/gcb.12179

    Kroeker, K. J., Micheli, F., & Gambi, M. C. (2012). Ocean acidification
    causes ecosystem shifts via altered competitive interactions.
    Nature Climate Change, 3(2), 156–159.

    Kroeker, K. J., Micheli, F., Gambi, M. C., & Martz, T. R. (2011). Divergent
    ecosystem responses within a benthic marine community to ocean
    acidification. Proceedings of the National Academy of Sciences of
    the United States of America, 108(35), 14515–14520. https://doi.
    org/10.1073/pnas.11077 89108

    Li, W., Gao, K., & Beardall, J. (2012). Interactive effects of ocean acid-
    ification and nitrogen-limitation on the diatom Phaeodactylum tri-
    cornutum. PLoS One, 7(12), e51590. https://doi.org/10.1371/journ
    al.pone.0051590

    Ling, S. D., Scheibling, R. E., Rassweiler, A., Johnson, C. R., Shears, N.,
    Connell, S. D., Salomon, A. K., Norderhaug, K. M., Pérez-Matus, A.,
    Hernández, J. C., & Clemente, S. (2015). Global regime shift dynam-
    ics of catastrophic sea urchin overgrazing. Philosophical Transactions
    of the Royal Society B: Biological Sciences, 370(1659), 20130269.

    Milazzo, M., Alessi, C., Quattrocchi, F., Chemello, R., D’Agostaro, R., Gil,
    J., Vaccaro, A. M., Mirto, S., Gristina, M., & Badalamenti, F. (2019).
    Biogenic habitat shifts under long-term ocean acidification show
    nonlinear community responses and unbalanced functions of asso-
    ciated invertebrates. Science of The Total Environment, 667, 41–48.
    https://doi.org/10.1016/j.scito tenv.2019.02.391

    Milazzo, M., Rodolfo-Metalpa, R., Chan, V. B. S., Fine, M., Alessi, C.,
    Thiyagarajan, V., Hall-Spencer, J. M., & Chemello, R. (2014). Ocean
    acidification impairs vermetid reef recruitment. Scientific Reports, 4,
    4189. https://doi.org/10.1038/srep0 4189

    Mirasole, A., Scopelliti, G., Tramati, C., Signa, G., Mazzola, A., & Vizzini,
    S. (2020). Evidences on alterations in skeleton composition and
    mineralization in a site-attached fish under naturally acidified
    conditions in a shallow CO2 vent. Science of The Total Environment,
    143309. https://doi.org/10.1016/j.scito tenv.2020.143309

    Mishra, A. K., Santos, R., & Hall -Spencer, J. M. (2020). Elevated trace
    elements in sediments and seagrasses at CO2 seeps. Marine
    Environmental Research, 153, 104810. https://doi.org/10.1016/j.
    maren vres.2019.104810

    Moulin, L., Grosjean, P., Leblud, J., Batigny, A., Collard, M., & Dubois, P.
    (2015). Long-term mesocosms study of the effects of ocean acid-
    ification on growth and physiology of the sea urchin Echinometra
    mathaei. Marine Environmental Research, 103, 103–114. https://doi.
    org/10.1016/j.maren vres.2014.11.009

    Mumby, P. J., Dahlgren, C. P., Harborne, A. R., Kappel, C. V., Micheli,
    F., Brumbaugh, D. R., Holmes, K. E., Mendes, J. M., Broad, K., &
    Sanchirico, J. N. (2006). Fishing, trophic cascades, and the process
    of grazing on coral reefs. Science, 311(5757), 98–101.

    O’Brien, J., & Scheibling, R. (2018). Turf wars: Competition between foun-
    dation and turf-forming species on temperate and tropical reefs and
    its role in regime shifts. Marine Ecology Progress Series, 590, 1–17.

    O’Donnell, M. J., Todgham, A. E., Sewell, M. A., Hammond, L. M.,
    Ruggiero, K., Fangue, N. A., Zippay, M. L., & Hofmann, G. E. (2010).
    Ocean acidification alters skeletogenesis and gene expression in
    larval sea urchins. Marine Ecology Progress Series, 398, 157–171.

    Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn,
    D., Minchin, P. R., O’Hara, R. B., Simpson, G. L., Solymos, P., Stevens,
    M. H. H., Szoecs, E., & Wagner, H. (2019). The vegan package.
    Vegan: Community ecology package. R Package Version 2.5-5. https://
    Cran.r-Proje ct.Org/Packa ge=vegan

    Pichler, T., Biscéré, T., Kinch, J., Zampighi, M., Houlbrèque, F., & Rodolfo-
    Metalpa, R. (2019). Suitability of the shallow water hydrothermal
    system at Ambitle Island (Papua New Guinea) to study the effect
    of high pCO2 on coral reefs. Marine Pollution Bulletin, 138, 148–158.
    https://doi.org/10.1016/j.marpo lbul.2018.11.003

    Porzio, L., Arena, C., Lorenti, M., De Maio, A., & Buia, M. C. (2020).
    Long-term response of Dictyota dichotoma var. intricata (C. Agardh)
    Greville (Phaeophyceae) to ocean acidification: Insights from high
    pCO2 vents. Science of The Total Environment, 731, 138896. https://
    doi.org/10.1016/j.scito tenv.2020.138896

    Porzio, L., Buia, M. C., Ferretti, V., Lorenti, M., Rossi, M., Trifuoggi,
    M., Vergara, A., & Arena, C. (2018). Photosynthesis and mineral-
    ogy of Jania rubens at low pH/high pCO2: A future perspective.
    Science of The Total Environment, 628–629, 375–383. https://doi.
    org/10.1016/j.scito tenv.2018.02.065

    Porzio, L., Garrard, S. L., & Buia, M. C. (2013). The effect of ocean acid-
    ification on early algal colonization stages at natural CO2 vents.
    Marine Biology, 160(8), 2247–2259. https://doi.org/10.1007/s0022
    7-013-2251-3

    R Core Team. (2019). R: A language and environment for statistical comput-
    ing. R Foundation for Statistical Computing. Retrieved from https://
    www.R-proje ct.org/

    Rastrick, S. S. P., Graham, H., Azetsu-Scott, K., Calosi, P., Chierici, M.,
    Fransson, A., Hop, H., Hall-Spencer, J., Milazzo, M., Thor, P., &
    Kutti, T. (2018). Using natural analogues to investigate the effects
    of climate change and ocean acidification on Northern ecosys-
    tems. ICES Journal of Marine Science, 75(7), 2299–2311. https://doi.
    org/10.1093/icesj ms/fsy128

    Rogers, A., Blanchard, J. L., & Mumby, P. J. (2014). Vulnerability of coral
    reef fisheries to a loss of structural complexity. Current Biology,
    24(9), 1000–1005. https://doi.org/10.1016/j.cub.2014.03.026

    Stachowicz, J. J., Bruno, J. F., & Duffy, J. E. (2007). Understanding
    the effects of marine biodiversity on communities and ecosys-
    tems. Annual Review of Ecology Evolution and Systematics, 38,
    739–766.

    Steneck, R. S., & Dethier, M. N. (1994). A functional group approach
    to the structure of algal-dominated communities. Oikos, 69(3),
    476–498.

    Sunday, J. M., Fabricius, K. E., Kroeker, K. J., Anderson, K. M., Brown, N.
    E., Barry, J. P., Connell, S. D., Dupont, S., Gaylord, B., Hall-Spencer,
    J. M., Klinger, T., Milazzo, M., Munday, P. L., Russell, B. D., Sanford,
    E., Thiyagarajan, V., Vaughan, M. L. H., Widdicombe, S., & Harley, C.
    D. G. (2017). Ocean acidification can mediate biodiversity shifts by
    changing biogenic habitat. Nature Climate Change, 7, 81–85. https://
    doi.org/10.1038/nclim ate3161

    Teixidó, N., Gambi, M. C., Parravacini, V., Kroeker, K., Micheli, F., Villéger,
    S., & Ballesteros, E. (2018). Functional biodiversity loss along nat-
    ural CO2 gradients. Nature Communications, 9(1), 5149. https://doi.
    org/10.1038/s4146 7-018-07592 -1

    Vizzini, S., Di Leonardo, R., Costa, V., Tramati, C. D., Luzzu, F., &
    Mazzola, A. (2013). Trace element bias in the use of CO2 vents as
    analogues for low pH environments: Implications for contamina-
    tion levels in acidified oceans. Estuarine, Coastal and Shelf Science,
    134, 19–30.

    Vizzini, S., Martínez-Crego, B., Andolina, C., Massa-Gallucci, A., Connell,
    S. D., & Gambi, M. C. (2017). Ocean acidification as a driver of com-
    munity simplification via the collapse of higher-order and rise of
    lower-order consumers. Scientific Reports, 7(1), 4018. https://doi.
    org/10.1038/s4159 8-017-03802 -w

    Wang, Y., Naumann, U., Wright, S. T., & Warton, D. I. (2012). Mvabund–
    An R package for model-based analysis of multivariate abundance
    data. Methods in Ecology and Evolution, 3(3), 471–474.

    Wickham, H. (2016). ggplot2: Elegant graphics for data analysis.
    Springer-Verlag.

    Wilson, S. S., Furman, B. T., Hall, M. O., & Fourqurean, J. W. (2020). Assessment
    of hurricane Irma impacts on south Florida seagrass communities using

    https://doi.org/10.1073/pnas.1216464110

    https://doi.org/10.1111/gcb.12179

    https://doi.org/10.1111/gcb.12179

    https://doi.org/10.1073/pnas.1107789108

    https://doi.org/10.1073/pnas.1107789108

    https://doi.org/10.1371/journal.pone.0051590

    https://doi.org/10.1371/journal.pone.0051590

    https://doi.org/10.1016/j.scitotenv.2019.02.391

    https://doi.org/10.1038/srep04189

    https://doi.org/10.1016/j.scitotenv.2020.143309

    https://doi.org/10.1016/j.marenvres.2019.104810

    https://doi.org/10.1016/j.marenvres.2019.104810

    https://doi.org/10.1016/j.marenvres.2014.11.009

    https://doi.org/10.1016/j.marenvres.2014.11.009

    https://Cran.r-Project.Org/Package=vegan

    https://Cran.r-Project.Org/Package=vegan

    https://doi.org/10.1016/j.marpolbul.2018.11.003

    https://doi.org/10.1016/j.scitotenv.2020.138896

    https://doi.org/10.1016/j.scitotenv.2020.138896

    https://doi.org/10.1016/j.scitotenv.2018.02.065

    https://doi.org/10.1016/j.scitotenv.2018.02.065

    https://doi.org/10.1007/s00227-013-2251-3

    https://doi.org/10.1007/s00227-013-2251-3

    https://www.R-project.org/

    https://www.R-project.org/

    https://doi.org/10.1093/icesjms/fsy128

    https://doi.org/10.1093/icesjms/fsy128

    https://doi.org/10.1016/j.cub.2014.03.026

    https://doi.org/10.1038/nclimate3161

    https://doi.org/10.1038/nclimate3161

    https://doi.org/10.1038/s41467-018-07592-1

    https://doi.org/10.1038/s41467-018-07592-1

    https://doi.org/10.1038/s41598-017-03802-w

    https://doi.org/10.1038/s41598-017-03802-w

    14  |    HARVEY Et Al.

    long-term monitoring programs. Estuaries and Coasts, 43(5), 1119–
    1132. https://doi.org/10.1007/s1223 7-019-00623 -0

    Witkowski, C. R., Agostini, S., Harvey, B. P., van der Meer, M. T. J.,
    Sinninghe Damsté, J. S., & Schouten, S. (2019). Validation of car-
    bon isotope fractionation in algal lipids as a $p$CO2 proxy using
    a natural CO2 seep (Shikine Island, Japan). Biogeosciences, 16(22),
    4451–4461. https://doi.org/10.5194/bg-16-4451-2019

    Zitoun, R., Connell, S. D., Cornwall, C. E., Currie, K. I., Fabricius, K.,
    Hoffmann, L. J., Lamare, M. D., Murdoch, J., Noonan, S., Sander,
    S. G., Sewell, M. A., Shears, N. T., van den Berg, C. M. G., & Smith,
    A. M. (2020). A unique temperate rocky coastal hydrothermal vent
    system (Whakaari–White Island, Bay of Plenty, New Zealand):
    Constraints for ocean acidification studies. Marine and Freshwater
    Research, 71(3), 321–344.

    S U P P O R T I N G I N F O R M AT I O N
    Additional supporting information may be found online in the
    Supporting Information section.

    How to cite this article: Harvey BP, Kon K, Agostini S, Wada
    S, Hall-Spencer JM. Ocean acidification locks algal
    communities in a species-poor early successional stage. Glob
    Change Biol. 2020;00:1–14. https://doi.org/10.1111/
    gcb.15455

    https://doi.org/10.1007/s12237-019-00623-0

    https://doi.org/10.5194/bg-16-4451-2019

    https://doi.org/10.1111/gcb.15455

    https://doi.org/10.1111/gcb.15455

    L E T T E R
    Ocean acidification reduces coral recruitment by disrupting

    intimate larval-algal settlement interactions

    Christopher Doropoulos,1,2*

    Selina Ward,1 Guillermo

    Diaz-Pulido,2,3 Ove

    Hoegh-Guldberg2,4 and

    Peter J. Mumby1,2

    Abstract
    Successful recruitment in shallow reef ecosystems often involves specific cues that connect planktonic

    invertebrate larvae with particular crustose coralline algae (CCA) during settlement. While ocean acidification

    (OA) can reduce larval settlement and the abundance of CCA, the impact of OA on the interactions between

    planktonic larvae and their preferred settlement substrate are unknown. Here, we demonstrate that CO2
    concentrations (800 and 1300 latm) predicted to occur by the end of this century significantly reduce coral
    (Acropora millepora) settlement and CCA cover by ‡ 45%. The CCA important for inducing coral settlement
    (Titanoderma spp., Hydrolithon spp.) were the most deleteriously affected by OA. Surprisingly, the only preferred

    settlement substrate (Titanoderma) in the experimental controls was avoided by coral larvae as pCO2 increased,

    and other substrata selected. Our results suggest OA may reduce coral population recovery by reducing coral

    settlement rates, disrupting larval settlement behaviour, and reducing the availability of the most desirable

    coralline algal species for successful coral recruitment.

    Keywords
    Acropora, coral, crustose coralline algae, electivity, Hydrolithon, ocean acidification, recruitment, settlement,

    Titanoderma.

    Ecology Letters (2012) 15: 338–346

    INTRODUCTION

    The effects of ocean acidification (OA) have raised concerns about

    coral reef ecosystem function by reducing the calcification rates of

    benthic organisms important to maintaining habitat structure and

    biodiversity (Hoegh-Guldberg et al. 2007; Kroeker et al. 2010).

    Anthropogenic emissions of carbon dioxide (CO2) have increased

    atmospheric CO2 from approximately 280 ppm prior to the year 1750

    to > 380 ppm in 2005 (Jansen et al. 2007), and these are continuing to

    rise (Le Quere et al. 2009). The absorption of this atmospheric CO2 by

    the oceans has reduced global pH by 0.1 units and carbonate

    saturation state by 20% since 1800 (Orr et al. 2005). Numerous

    laboratory studies have demonstrated that corals (Schneider & Erez

    2006; Anthony et al. 2008), calcifying algae (Anthony et al. 2008;

    Kuffner et al. 2008), and coral reef communities (Langdon et al. 2000;

    Andersson et al. 2009) have reduced calcification in seawater with

    lower pH due to depleted carbonate saturation.

    Ecological processes pivotal to coral reef resilience, including coral

    recruitment, herbivory, trophic integrity, and connectivity (Knowlton

    2001; Mumby et al. 2007), under high CO2 levels have hardly been

    investigated (Doney et al. 2009). Yet, growing evidence suggests that

    interactions between species are altered as CO2 increases. Under

    conditions of OA, corals in contact with fleshy macroalgae had higher

    mortality (Diaz-Pulido et al. 2011), and fish mortality increased as OA

    reduced the ability of juvenile fish to detect their predators (Munday

    et al. 2010). Furthermore, it has been suggested that turf algae can

    decrease the recruitment of crustose coralline algae (CCA) (Kuffner

    et al. 2008; Russell et al. 2009) and kelp (Connell & Russell 2010)

    because of greater space occupation at elevated pCO2. While these

    examples illustrate that ecological interactions can be altered as CO2
    increases, potential interactions of OA on coral recruitment have not

    been addressed.

    Recruitment is critical to community recovery as it represents a

    crucial process in the development of populations in the post-

    disturbance period. A key ecological process in the formation of

    coral reefs is the settlement of coral larvae from the plankton to

    the reef substrata. Many larvae test benthic substrates for

    microhabitat suitability prior to settlement (i.e. attachment and

    metamorphosis), with the selection of optimal microhabitats critical

    in the post-settlement survival of benthic invertebrates (Raimondi

    & Keough 1990; Harrington et al. 2004). Different benthic algae

    offer both inductive and inhibitive settlement cues for planktonic

    invertebrate larvae (Rodriguez et al. 1993; Kuffner et al. 2006; Diaz-

    Pulido et al. 2010), and larvae often search for appropriate substrata

    associated with specific CCA and microbial communities for

    successful settlement (Morse et al. 1988; Johnson & Sutton 1994;

    Heyward & Negri 1999; Negri et al. 2001; Webster et al. 2004).

    While recent evidence demonstrates the settlement of coral larvae is

    reduced as pCO2 increases (Albright et al. 2010; Albright &

    Langdon 2011; Nakamura et al. 2011), the interactions between

    planktonic larvae and the CCA community under elevated CO2
    levels are unknown.

    1School of Biological Sciences, University of Queensland, St Lucia, Qld 4072,

    Australia
    2Australian Research Council Centre of Excellence for Coral Reef Studies,

    University of Queensland, St Lucia, Qld 4072, Australia

    3Griffith School of Environment and Australian Rivers Institute, Nathan Campus,

    Griffith University, Nathan, QLD 4111, Australia
    4Global Change Institute, University of Queensland, St Lucia, Qld 4072, Australia

    *Correspondence: E-mail: c.doropoulos@uq.edu.au

    Ecology Letters, (2012) 15: 338–346 doi: 10.1111/j.1461-0248.2012.01743.x

    � 2012 Blackwell Publishing Ltd/CNRS

    Here, we test the hypothesis that elevated pCO2 (400 control, 800

    and 1300 latm) alters the recruitment of a spawning coral (Acropora
    millepora) by affecting the benthic algal community structure, and the

    interactions between the substrata and larvae during settlement. We

    used a mechanistic approach with three complementary experiments

    to investigate how OA reduces larval settlement. First, to investigate

    whether OA caused a shift in the community structure of the

    settlement substrata to alter coral settlement, we preconditioned

    settlement tiles in treatment seawater for 60 days prior to conducting

    6 day settlement assays on those tiles in ambient seawater (expt. 1).

    Second, we conducted the reciprocal experiment by isolating the

    exposure of elevated pCO2 seawater to the coral larvae and settlement

    substrata during the 6 days settlement phase only (expt. 2). Finally, we

    explored whether there was a combined effect on coral settlement

    when the settlement substrata and coral larvae were both exposed to

    elevated pCO2 for 60 and 6 days respectively (expt. 3). From this

    series of experiments, we show that OA decreases coral settlement

    rates by reducing the availability of specific CCA preferred for larval

    settlement, as well as interfering with the interaction between larvae

    and CCA by altering the settlement behaviour of the coral larvae, such

    that previously avoided substrata are preferentially selected as pCO2
    increases in all three conditions.

    MATERIAL AND METHODS

    CO2 treatments and general protocol

    Coral settlement experiments were conducted from October to

    December 2009, at Heron Island Research Station, southern Great

    Barrier Reef (GBR). Settlement substrata and coral larvae were

    exposed to three treatments, which represented control (pH 8.04,

    401 latm), and two elevated (pH 7.79, 807 latm; pH 7.60,
    1299 latm) levels of future CO2 concentrations (Table 1). Treatments
    were based on the worst-case stabilisation levels V (pCO2 700–

    850 latm) and VI (pCO2 > 900 latm) predicted by the Intergovern-
    mental Panel on Climate Change (IPCC) (Meehl et al. 2007). These

    were chosen for the experiment as current CO2 emissions are tracking

    the most carbon intensive levels (A1FI) predicted by the IPCC (Le

    Quere et al. 2009).

    As pH is reduced in a predictable manner by elevated pCO2, the

    CO2 levels of the experimental seawater were controlled by adjusting

    the pH of the seawater in 200 L sumps (Table 1) (see Diaz-Pulido

    et al. 2011 for system details). Briefly, the total pH of the seawater was

    continuously measured with temperature compensated pH electrodes

    (InPro4501VP; Mettler-Toledo, Melbourne, Victoria, Australia),

    which maintained the targeted pH levels with a control unit

    (Aquatronica, AEB technologies, Italy) that opened solenoid valves

    that injected CO2 into the seawater when pH exceeded the desired

    threshold. The calibration of the pH probes was checked daily, and

    recalibrated with Mettler-Toledo calibration buffers to 0.01 pH units

    when necessary. Alkalinity was measured on seawater samples taken

    every 6 h over a spring tidal cycle (2.4 m range) at the end of the study

    period to capture the largest variation in the seawater alkalinity and

    consolidate the pH treatments to the CO2 levels. Alkalinity replicates

    within a sample were analysed until a maximum 2% error was met,

    using a Metrohm auto-titrator at Edith Cowan University, WA. The

    carbonate chemistry of the control and experimental seawater was

    calculated with CO2SYS (Lewis & Wallace 2006) using pH, total

    alkalinity, salinity (35.4 ppt ± 0.2 SEM; n = 8), and temperature as

    the inputs, with the constants from Mehrbach et al. (1973) refitted by

    Dickson & Millero (1987).

    The settlement tile CO2 conditioning and settlement assays were

    conducted on tiles in replicate tanks in the outdoor flow-through

    aquarium system (details of the CO2 exposure times, tank and tile

    replication are described below in the protocols under each

    experiment and in supplementary Fig. S1). The three treatments were

    fed from the 200 L sumps into replicate 12 L tanks at a mean flow

    rate of 2.4 (± 0.2 SEM) L min
    )1

    , and each tank had a small

    powerhead for extra seawater circulation. This flow rate and water

    movement maintained the target pH levels, which were verified

    regularly with a portable SG2 SevenGo
    TM

    pH meter. Replicate tanks

    were randomised on the aquarium table under shade-cloth to account

    for the heterogeneity in light, which averaged 406 (± 18 SEM)

    lmol m)2 s)1 between 6 AM to 6 PM.

    Settlement tile preparation

    Unglazed terracotta settlement tiles (� 5 · 5 · 0.5 cm) were initially
    preconditioned on the Heron Island reef flat (23� 26¢ 42.2¢¢ S, 151�
    54¢ 47.0¢¢ E) for 5 months to develop a microbial and encrusting
    community important to coral settlement (Heyward & Negri 1999).

    Tiles were collected and carefully cleaned of fouling organisms using a

    toothbrush, tweezers, and a plastic scraper. The tiles were then

    randomly placed in replicate 12 L aquaria and conditioned in the

    control and elevated CO2 treatments for 60 days prior to the

    settlement assays. During this time the walls of the aquaria were

    cleaned regularly to minimize any algal growth. Settlement tiles were

    orientated horizontally at the bottom of the tanks and were stacked in

    tile pairs with a 0.5 cm spacer, maximising the amount of cryptic

    surfaces available for settlement, as coral larvae generally settle in

    cryptic areas in shallow habitats (Wallace 1985).

    Coral larvae collection

    Gravid adult colonies of Acropora millepora were located on the Heron

    Island reef flat around the time of the predicted spawning (2nd Dec

    2009). Acropora millepora was chosen as a model organism as it is

    Table 1 Summary of the physical and chemical seawater values for CO2 treatment levels

    Treatment

    Temp*

    pH*

    TA* pCO2� HCO3
    )� CO3

    2)�

    XAragonite��C lmol kg)1 latm lmol kg)1 lmol kg)1

    Control 26.0 (± 0.6) 8.04 (± 0.01) 2355 (± 14) 401 (± 11) 1800 (± 20) 227 (± 4) 3.6 (± 0.07)

    Stabilisation level V 26.0 (± 0.6) 7.79 (± 0.01) 2365 (± 20) 807 (± 14) 2019 (± 23) 142 (± 3) 2.3 (± 0.06)

    Stabilisation level VI 26.0 (± 0.6) 7.60 (± 0.01) 2363 (± 19) 1299 (± 21) 2125 (± 19) 97 (± 3) 1.6 (± 0.06)

    *Temperature, pH, and total alkalinity are means (± SEM) of five replicates.

    �pCO2, bicarbonate, carbonate and aragonite saturation state (X) were calculated using CO2 SYS (Lewis & Wallace 2006).

    Letter Elevated CO2 alters CCA-larval interactions 339

    � 2012 Blackwell Publishing Ltd/CNRS

    commonly found on GBR and Indo-Pacific shallow reef flats. Five

    colonies were collected and transported to outdoor aquarium facilities

    where they were housed in 60 L flow-through aquaria until they

    released their egg-sperm bundles. The bundles were broken apart by

    gently stirring and agitating the water, and gametes from the different

    colonies were collected and cross-fertilized. Fertilization took place

    for 2 h, after which the embryos were collected and reared in a

    laboratory at 25 �C in ambient seawater, using 200 L sumps with
    aeration. At least half the seawater was changed every few hours for

    the first 24 h and every 6–12 h thereafter. This removed dead larvae

    and unfertilized gametes, to minimise contamination of the rearing

    sumps. The larvae developed cilia and began swimming 3 days after

    spawning, after which they were used for the settlement assays.

    Swimming A. millepora larvae were randomly removed from the

    rearing sumps, added to the experimental aquaria, and allowed 6 days

    to settle (i.e. attach and metamorphose). The number of larvae added

    to each tank during the settlement assays was standardised to 150

    (± 10) per tile. After this time, the tiles were removed from the tanks

    at random and inspected for settlement with a dissecting microscope.

    Experiment 1

    To isolate whether changes to the benthic community altered coral

    settlement, settlement tiles were conditioned at 400, 800, and

    1300 latm pCO2 for 60 days. Following the conditioning period,
    coral settlement assays were conducted for 6 days on those tiles with

    control seawater only. Three replicate tanks per treatment, with 8 tiles

    and 1200 (± 80) larvae per tank, were used for

    the experiment (Fig. S1).

    Experiment 2

    A reciprocal experiment was conducted to determine whether

    settlement was altered by elevated pCO2 exposure of the coral larvae

    and the benthic community during the settlement assays only.

    Settlement assays were conducted for 6 days using the three CO2
    seawater treatments described above with settlement tiles that were

    conditioned with control seawater only. Two replicate tanks per

    treatment, with 6 tiles and 900 (± 60) larvae per tank, were used for

    the experiment (Fig. S1).

    Experiment 3

    Finally, to investigate the combined effect of prolonged exposure of

    elevated pCO2 on the benthic community and the settling larvae,

    settlement tiles were conditioned in the three CO2 treatments for

    60 days prior to conducting 6 day settlement assays on those tiles in

    the treatment seawater described above. Three replicate tanks per

    treatment, with 10 tiles and 1500 (± 150) larvae per tank, were used

    for the experiment (Fig. S1).

    Response variables and data analyses

    We analysed total larval settlement, benthic community and CCA

    community cover of the settlement tiles, and coral settlement

    substrate preferences for each of the three experiments. The number

    of settled (i.e. attached and metamorphosed) coral larvae was initially

    quantified for all orientations of each tile. However, we only analysed

    the undersides of each tile (for this and all other variables) as the

    number of corals settled in this orientation accounted for ‡ 95% of
    the total settlement.

    The benthic community of the settlement tiles was quantified by

    placing a grid on a tile, and evaluating the dominant substrate in a

    square (7.5 mm
    2
    ) using a dissecting microscope, with 224–377 squares

    per tile. The substrata were characterised into eight major benthic

    groups which were: bare tile, CCA, dead crustose coralline algae

    (DCCA), endolithic algae found in dead crustose coralline algae

    (EDCCA), turf algae found on dead crustose coralline algae

    (TDCCA), turf algae (Turf), encrusting fleshy algae (EFA), and other

    organisms which included biofilm, bryozoans, foraminifera, and other

    encrusting organisms (Other). CCA specimens were identified to the

    finest taxonomic resolution where possible and included nine CCA

    taxa (see Appendix S1 in Supporting Information for details on CCA

    identification). When CCA specimens could not be identified to genus

    or species, they were placed in to an Unknown CCA group, which

    represented � 6% of the total CCA community. See supplementary
    Fig. S2 for images of the dominant benthic groups and CCA taxa.

    The substrate settled on by each individual was quantified to

    investigate larval settlement behaviour using Vanderploeg and Scavia�s
    electivity index (E*). This index is analogous to Ivlev�s E, but
    incorporates a selectivity coefficient and the number of substrata

    available for settlement (Lechowicz 1982). Therefore: E* = [Wa )
    (1 ⁄ n)] ⁄ [Wa + (1 ⁄ n)], where n is the total number of substrate types
    available on each tile and W is the selectivity coefficient for substrate

    �a� determined by: Wa = [ra ⁄ pa] ⁄
    P

    (ra ⁄ pa),(rb ⁄ pb)…(rz ⁄ pz), r is the
    proportion of coral larvae settled on substrata a to z on each tile, and

    p is the proportion of substrata a to z available for settlement on each

    tile. A substrate was selected at random for larval settlement when E*

    was � 0, preferably settled on when E* was > 0, and avoided for
    settlement when E* was < 0.

    The number of coral larvae settled per tile was analysed with a

    generalised linear mixed effects model using Poisson distribution. We

    tested the effects of elevated pCO2 on counts of coral settlement

    amongst CO2 treatment (3 levels, fixed) with replicate tanks as a

    random effect and nested in CO2 treatment. The effect of elevated

    pCO2 on the percent cover of the broad benthic community and CCA

    community composition were tested using a mixed effects permuta-

    tional MANOVA (PERMANOVA), with the same fixed and random

    factors described for the previous model. When significant differences

    were detected (P < 0.05), pair-wise comparisons were performed to

    investigate treatment effects. In multivariate analyses, SIMPER

    analysis was used to determine the variables that characterised the

    dissimilarity amongst treatments. Univariate ANOVA was conducted

    within CCA cover to determine any significant treatment effects. All

    percentage cover data were sin
    )1 �x transformed to meet require-

    ments of homogeneity (permDISP) prior to analysis. Finally, we tested

    the effect of CO2 treatment (3 levels, fixed) on coral settlement

    behaviour with tanks as replicates using PERMANOVA. In all

    PERMANOVA main effect and pair-wise tests, we used the P-values

    generated by 99 999 permutations when the number of unique

    permutations were large, and the Monte Carlo asymptotic P-value

    otherwise (Anderson 2005).

    RESULTS

    We report the results of each of the three settlement experiments in

    turn, describing the impacts of OA on overall coral settlement density,

    340 C. Doropoulos et al. Letter

    � 2012 Blackwell Publishing Ltd/CNRS

    the structure of benthic substrata, and settlement behaviour of the

    larvae. Results are summarised in Table 2.

    Experiment 1

    A reduction in the cover of CCA and shift in the CCA community

    from elevated CO2 decreased coral settlement in the OA treatments.

    The reduction in settlement decreased significantly from an average of

    11.0 individuals per 25 cm
    2

    in the control, to 1.6 and 5.5 individuals at

    800 and 1300 latm, respectively (Table 2; Fig. 1a). The cover of
    CCAs changed dramatically in the elevated CO2 treatments with a

    significant decline of � 50% (ANOVA: F2,6 = 13.283; P = 0.014;
    Table 2; supplementary Fig. S3a). The CCA community structure also

    changed as pCO2 increased, with three out of ten coralline algal taxa

    declining with increasing CO2 concentrations (MANOVA: F2,6 = 3.286;

    P = 0.017; Table 2). Titanoderma spp., Hydrolithon boreale, and

    H. farinosum were the species that characterised the loss of CCA

    cover in both the elevated CO2 treatments (supplementary Fig. S3b).

    The settlement behaviour of the larvae, as measured by their

    substrate selectivity, was significantly affected by the exposure of

    settlement tiles to elevated pCO2 prior to the settlement assays

    (MANOVA: F2,6 = 4.291; P = 0.004; Table 2). Titanoderma spp. was the

    only preferred settlement substrate in the control treatment

    (E* = 0.8) and there were lower rates of settlement on all other

    substrata than would be expected by chance (supplementary Fig. S4a).

    At 800 latm, larvae did not show any clear settlement preferences and
    most substrata were avoided (supplementary Fig. S4b), while the

    larvae showed a weak preference for H. farinosum (E* = 0.2) at

    1300 latm (supplementary Fig. S4c).

    Experiment 2

    As expected, there were no differences between the broad community

    composition, the CCA percent cover, or the CCA community

    amongst the settlement tiles that were allocated for use in these

    settlement assays (Table 2). Yet, exposure of coral larvae and the

    settlement tiles to elevated pCO2 during the 6 day settlement assays

    significantly reduced coral settlement, as it declined from an average

    of 11.9 individuals per 25 cm
    2

    in the control, to 4.7 and 2.8 individuals

    at 800 and 1300 latm, respectively (Table 2; Fig. 1b). A similar
    disruption to larval settlement behaviour occurred to that found when

    only the tiles were pre-exposed to elevated pCO2 for a prolonged

    period of time (exp. 1). Again, coral larvae preferred to settle on

    Titanoderma spp. (E* = 0.75) in controls (supplementary Fig. S5a),

    most substrata were avoided at 800 latm (supplementary Fig. S5b),
    and a weak preference for H. farinosum (E* = 0.2) was found at

    1300 latm (supplementary Fig. S5c).

    Experiment 3

    Again, settlement was reduced when the tiles were conditioned in the

    CO2 treatments for 60 days, and 6 day settlement assays were

    conducted on those tiles under elevated pCO2. The magnitude of the

    effect was similar to whether the tiles were conditioned in the CO2
    treatments for 60 days prior to the 6 day settlement assays with

    control seawater only (exp.1), or whether the larvae and tiles were

    exposed to the CO2 treatments during the 6 day settlement assays

    only (exp. 2) (Table 2). Increased CO2 reduced the settlement of A.

    millepora from an average of 9.7 individuals per 25 cm
    2

    in the control,

    to 5.2 and 4.2 individuals at 800 and 1300 latm, respectively (Fig. 1c).
    The reduction in settlement was significant between the control and

    highest CO2 treatment (P = 0.046) and marginally significant between

    the control and intermediate treatment (P = 0.060).

    The changes in tile community structure were similar to those in

    experiment 1, but the effects of OA appeared to be less variable in this

    experiment. As a result, the wider benthic community structure on the

    tile undersides was found to differ significantly amongst the CO2
    treatments (MANOVA: F2,6 = 2.612; P = 0.003; Table 2; Fig. 2a).

    The loss of coralline algae was partly replaced by an increase of 8% in

    the cover of �bare tile� (Fig. 2a). As in experiment 1, OA led to a
    significant reduction in the cover of CCAs on the tiles (ANOVA:

    F2,6 = 40.538; P = 0.002; Table 2), characterised by declines in

    Titanoderma spp., H. boreale, and H. farinosum (Fig. 2b).

    Coral settlement behaviour was again altered significantly by

    elevated pCO2 (MANOVA: F2,6 = 4.224; P = 0.004; Table 2;

    Fig. 3). Of the 19 substrata available, Titanoderma spp. was again the

    only preferred settlement substrate in the control (E* = 0.6), while all

    other substrata were avoided (Fig. 3a). At 800 latm, Hydrolithon
    reinboldii was the only preferred coral settlement substrate (E* = 0.3),

    and all other settlement substrata were either randomly settled on or

    avoided (Fig 3b). No substrate was preferred for settlement at

    1300 latm, with random settlement on bare tile (E* = )0.05), and all
    other substrata were avoided (Fig 3c).

    DISCUSSION

    In our study, the settlement density of coral larvae decreased by

    ‡ 45% as pCO2 increased from 400 to 800 and 1300 latm in all three

    Table 2 Changes to the response variables in Experiments 1, 2, and 3, comparing elevated CO2 treatments (800 and 1300 latm) to the controls (400 latm)

    Response variable

    Experiment 1 Experiment 2 Experiment 3

    800 latm 1300 latm 800 latm 1300 latm 800 latm 1300 latm

    1. Total settlement fl 82%*** fl 45% fl 58%*** fl 75%*** fl 50% fl 60%*
    2. Benthic community structure NS NS NS NS CCA CCA**

    3. CCA cover fl 47%* fl 52%* NS NS fl 42%* fl 63%***
    4. CCA community structure Titanoderma** Titanoderma* NS NS NS Titanoderma**

    5a). Overall settlement behaviour Titanoderma* Titanoderma* NS NS Sporolithon* Titanoderma**

    5b). Selectivity from Titanoderma fl 72% fl 74% fl 69% fl 65% fl 35% fl60%

    SIMPER analysis determined the variable that characterised the difference between the control and elevated CO2 treatments in multivariate analyses. Coral behaviour (5) is

    divided into the change in settlement preferences of the larvae (5a) among the substrate community, and (5b) from Titanoderma spp., the only preferred settlement substrate in

    the controls. Significance values are indicated by: NS = non-significant, * = < 0.05, ** = < 0.01, *** = < 0.001.

    Letter Elevated CO2 alters CCA-larval interactions 341

    � 2012 Blackwell Publishing Ltd/CNRS

    experiments. The reduction in settlement was accompanied by a

    profound decline in the cover of CCA when the settlement substrata

    were conditioned in elevated CO2 treatments for 60 days prior to the

    settlement assays (expt. 1 & 3). While recent studies have also found

    inverse relationships between elevated pCO2 and rates of coral

    settlement (Albright et al. 2010; Albright & Langdon 2011; Nakamura

    et al. 2011), and overall CCA cover (Hall-Spencer et al. 2008; Kuffner

    et al. 2008; Russell et al. 2009; Fabricius et al. 2011), our study is the

    first to directly link benthic community cover with coral settlement

    and it provides three important novel insights. First, we identified the

    most susceptible CCA to OA and found that they are the most

    important taxa for coral settlement, particularly Titanoderma. Secondly,

    we discovered that OA reduced the affinity between the settling larvae

    and Titanoderma, their preferred settlement substrate. Third, we found

    that similar changes in settlement behaviour occurred under all three

    (a)

    (b)

    (c)

    Figure 1 Coral (Acropora millepora) settlement rates on experimental tiles (25 cm
    2
    ) in

    response to increasing pCO2. Assays occurred on (a) settlement tiles conditioned in

    treatment seawater for 60 days prior to 6 day larval settlement assays on those tiles

    with control seawater (n = 3); (b) settlement tiles and larvae exposed to treatment

    seawater for 6 days during the settlement assays on tiles conditioned in control

    seawater only (n = 2); and (c) settlement tiles and larvae exposed to treatment

    seawater for 60 and 6 days, respectively (n = 3). Data are means ± SEM.

    Significance values comparing elevated CO2 treatments to the control are indicated

    by: * = < 0.05, ** = < 0.01, *** = < 0.001.

    (a)
    (b)

    Figure 2 Percent cover of (a) the broad benthic community and (b) the crustose

    coralline algae community in response to increasing pCO2. Settlement tiles were

    exposed to the treatments for 66 days, which involved a 60 day pre-exposure

    period prior to the 6 day settlement assays (expt. 3). CCA = crustose coralline

    algae. DCCA = dead crustose coralline algae. EDCCA = endolithic algae in dead

    crustose coralline algae. TDCCA = turf on dead crustose coralline algae.

    Turf = filamentous algal turf. EFA = encrusting fleshy algae. Other = biofilm,

    carbonate, bryozoans, encrusting foraminifera, and unidentified. H. boreale = Hy-

    drolithon boreale. H. farinosum = Hydrolithon farinosum. H. reinboldii = Hydrolithon

    reinboldii. Data are

    means ± SEM; n = 3.

    342 C. Doropoulos et al. Letter

    � 2012 Blackwell Publishing Ltd/CNRS

    experimental conditions. As we explain below, this surprising result

    implies that coral settlement behaviour is mediated by cues associated

    with coralline algae that appear to be highly sensitive to elevated

    pCO2.

    We designed our experiments to distinguish the effects of OA on

    the settling organisms (corals) from the settlement surfaces (the

    benthic community on the tiles). In experiment 1, we subjected tiles to

    a 60 day exposure to elevated pCO2 that resulted in profound changes

    to the coralline algal assemblage. When these tiles were then placed in

    control (ambient) conditions with coral planulae, the settlement

    behaviour of the larvae was disrupted. Because the larvae never

    experienced OA conditions in this experiment, the result implies that

    prolonged exposure of substrates to OA may alter the cues associated

    with CCA that are used by larvae to settle preferentially on Titanoderma.

    To examine the influence of OA on the settling larvae themselves

    (expt. 2), we exposed them to OA treatments during the 6 day

    settlement assays. In this case, all benthic substrata were precondi-

    tioned in control seawater prior to the experiment and were exposed

    to the treatment seawater for the 6 day period during the assays.

    Again, we found the same qualitative disruption to larval settlement

    behaviour, suggesting that the 6 day exposure of the benthic

    community to OA disrupted the signalling from the CCA, and

    potentially that the larvae may also be directly affected by elevated

    pCO2. In the third experiment, we found a similar qualitative result

    when larvae were subjected to OA and offered settlement substrates

    that had also been exposed to the treatments for a 60 day period.

    There are two possible explanations of our results. The most

    parsimonious explanation is that even a daily exposure of the benthos

    to OA disrupts the signalling mechanisms used by coral planulae to

    preferentially settle upon Titanoderma. That is, the outcome for

    settlement was the same whether the tiles were pre-conditioned for

    60 days, causing profound changes in coralline cover, or 6 days during

    the experiment. This explanation is consistent with all three

    experiments and accounts for the lack of an additive impact of OA

    on settlement when both larvae and substrates were exposed to OA.

    The most likely mechanism is that larval settlement behaviour is

    mediated by bacteria and ⁄ or chemical cues associated with CCA
    because settlement was disrupted even when the cover of coralline

    algae was unchanged (exp. 2). These results imply that the cues

    associated with algal morphogens and ⁄ or the bacterial communities
    associated with the CCA thalli are highly sensitive to changes in water

    chemistry. It was recently demonstrated that microbial communities

    associated with biofilms grown on glass slides were altered after

    11 days in elevated pCO2 (Witt et al. 2011). There is a precedent for

    the role of bacteria in facilitating settlement (Johnson & Sutton 1994;

    Negri et al. 2001; Webster et al. 2004), but neither the taxon specificity

    (to Titanoderma) nor the sensitivity to OA have been shown, and future

    work should isolate whether it is changes in bacterial communities

    and ⁄ or the morphogens associated with CCA that alters the
    preference of larval settlement under elevated CO2.

    An alternative, albeit not mutually exclusive, explanation is that

    coral settlement on CCA is disrupted by exposure of either partner to

    (a)
    (b)
    (c)

    Figure 3 Coral (Acropora millepora) settlement behaviour of the

    substrata that the coral larvae preferred (> 0), avoided (< 0), or

    randomly (� 0) settled on in response to (a) 400, (b) 800, and (c)
    1300 latm pCO2 using Vanderploeg and Scavia�s electivity index
    (E*). Settlement assays occurred with settlement tiles and larvae

    exposed to the treatments for 60 and 6 days, respectively (expt. 3).

    See Fig. 2 for abbreviated substrate definitions. Data are

    means ± SEM; n = 3.

    Letter Elevated CO2 alters CCA-larval interactions 343

    � 2012 Blackwell Publishing Ltd/CNRS

    OA conditions (i.e. exposure of either the larvae or the algae). This

    explanation is consistent with recent reports of the impacts of elevated

    pCO2 on coral larvae metabolic rate (Albright & Langdon 2011;

    Nakamura et al. 2011) and metamorphosis (Albright et al. 2010;

    Albright & Langdon 2011; Nakamura et al. 2011), and on fish larvae

    olfactory ability (Dixson et al. 2010; Munday et al. 2010). It has also

    been shown that invertebrate larvae become less discriminating in the

    selection of their preferred substrate for settlement when they are

    under stress (Marshall & Keough 2003). Thus, the coral larvae may

    have lost their selectivity for Titanoderma due to the stress related to

    OA. Yet, this explanation is not entirely satisfactory for the following

    reasons. Firstly, we have to accept that the similarity in outcome from

    manipulating the settlement substrata versus the planulae is coinci-

    dental (i.e. the disruption to either partner has the same overall

    outcome). Secondly, we cannot easily account for the absence of a

    clear additive effect when both partners were perturbed simulta-

    neously.

    Previous studies of coral recruitment in both spawning and

    brooding corals, including those from the families Acroporidae,

    Agariciidae, Pocilloporidae, and Poritidae, and stemming from both

    the Atlantic and Indo-Pacific, have found that coral larvae have an

    innate ability to settle preferentially on a single CCA genus,

    Titanoderma, and that ensuing survival is greatest on this substrate

    compared to any other (Harrington et al. 2004; Arnold et al. 2010;

    Price 2010; Ritson-Williams et al. 2010). Here, we found that OA

    presents two problems for settling corals. Not only is Titanoderma

    exceptionally sensitive to OA, such that its availability is compro-

    mised, but larval behaviour switches from a high preference to settle

    on Titanoderma to avoidance. Corals have previously been shown to

    settle on other substrata, including Hydrolithon spp. and bare tile that

    the larvae preferentially settled upon at the elevated CO2 treatments in

    this study, but this occurs at lower rates of settlement and survival

    (Harrington et al. 2004; Arnold et al. 2010; Price 2010; Ritson-Williams

    et al. 2010). Titanoderma has been proposed to be a good facilitator of

    coral settlement because it does not slough off tissue and therefore

    provides a persistent substratum for recruits, while some species of

    Hydrolithon and other CCA slough their tissue to remove fouling

    organisms (Harrington et al. 2004; Ritson-Williams et al. 2010). Thus,

    while the settlement of corals onto previously avoided substrata at

    elevated pCO2 in our study implies that their post-settlement survival

    may be reduced, empirical investigations of the long-term survival of

    recruits on different substrates at elevated pCO2 are needed to directly

    test this hypothesis as it may be an adaptive trait.

    Titanoderma is a cryptic, early successional species with relatively

    rapid growth, creeping morphology and delicate, thin thalli

    (< 500 lm) (Steneck 1986; Ringeltaube & Harvey 2000; Littler & Littler 2003). Its morphology and cryptic, opportunistic nature make it

    indicative of fresh substratum with relatively benign levels of stress

    such as parrotfish grazing or sediment scour. Such environments are

    likely to be ideal for coral settlement because new substratum is likely

    to possess fewer competitors (Vermeij & Sandin 2008) and parrotfish

    predation can be problematic for coral recruits (Penin et al. 2010).

    However, we hypothesize that some of the traits that make Titanoderma

    such an important settlement inducer might predispose a particular

    sensitivity to OA. It has previously been demonstrated that elevated

    pCO2 decreases the abundance and recruitment of coralline algae, in

    both field (Hall-Spencer et al. 2008; Fabricius et al. 2011) and

    laboratory (Kuffner et al. 2008; Russell et al. 2009) settings. While

    these reports (Hall-Spencer et al. 2008; Kuffner et al. 2008) found an

    inverse competitive relationship between CCA and turf cover as pCO2
    increased, we found that reduced CCA was accompanied by an

    increase in the amount of bare tile rather than turf. This suggests that

    the reduction of CCA cover was not a consequence of space

    competition with turfs, but a direct effect of OA on CCA. In our

    study, the three most sensitive taxa of CCA to elevated pCO2
    (H. boreale, H. farinosum, and Titanoderma spp.), are all early successional

    species with rapid growth and thin thalli (< 500 lm) (Steneck 1986; Ringeltaube & Harvey 2000; Littler & Littler 2003). In contrast, later

    successional CCA taxa that have thicker crusts (> 500 lm) (e.g.
    Sporolithon, Neogoniolithon, Porolithon) (Steneck 1986; Ringeltaube &

    Harvey 2000; Littler & Littler 2003) may be more resistant to OA.

    While further studies are needed to test this hypothesis, our results

    show that increasing levels of dissolved CO2 are likely to have

    profound consequences for the functional diversity of coralline algal

    communities.

    Our research has demonstrated the ecological mechanics of how

    ocean acidification may interfere with a critical process important to

    the resilience of a diverse marine ecosystem. This occurred by a

    reduction to the abundance of the preferred substrate for larval

    settlement, and a disruption to an intimate ecological interaction

    between the coral larvae and its preferred substrate during settlement.

    The altered interaction between coral settlement and the CCA

    community suggests that future recruitment of individuals may be

    impaired by CO2 concentrations predicted to be reached this century.

    These type of impacts of increased CO2 on non-trophic ecological

    interactions between species are just starting to be experimentally

    demonstrated (e.g. Connell & Russell 2010; Dixson et al. 2010; Diaz-

    Pulido et al. 2011), but suggest profound consequences on the

    recovery potential of shallow marine ecosystems (e.g. coral reefs)

    following local and global disturbances.

    ACKNOWLEDGEMENTS

    We thank M. Cowlin, A. Noel, M. Nitschke, M. Smith, O. McIntosh,

    and the staff at Heron Island Research Station for their technical

    assistance in the field; K. Anthony for designing and providing the

    experimental aquarium system; J. Pandolfi for providing laboratory

    facilities to process the tiles; and, to the four anonymous referees who

    provided constructive criticism of the original manuscript. This

    research was financially supported by an ARC Discovery Grant

    awarded to O. Hoegh-Guldberg, S. Ward and G. Diaz-Pulido, a QLD

    Smart Futures PhD Scholarship to C. Doropoulos, and an ARC

    Laureate Fellowship to PJ. Mumby. All work was conducted under

    GBRMPA permit number 31597.1.

    AUTHOR CONTRIBUTIONS

    CD, SW, GDP and PJM designed the study, CD and SW conducted

    the study, CD and GDP collected the data, and CD and PJM analysed

    the data. CD wrote the first draft of the manuscript, and all the

    authors contributed substantially to the interpretation and final

    version of the paper.

    REFERENCES

    Albright, R. & Langdon, C. (2011). Ocean acidification impacts multiple early life

    history processes of the Caribbean coral Porites astreoides. Glob. Change Biol., 17,

    2478–2487.

    344 C. Doropoulos et al. Letter

    � 2012 Blackwell Publishing Ltd/CNRS

    Albright, R., Mason, B., Miller, M. & Langdon, C. (2010). Ocean acidification

    compromises recruitment success of the threatened Caribbean coral Acropora

    palmata. Proc. Natl Acad. Sci. USA, 107, 20400–20404.

    Anderson, M.J. (2005). PERMANOVA: A FORTRAN Computer Program for

    Permutational Analysis of Variance. Department of Statistics, University of

    Auckland, New Zealand.

    Andersson, A.J., Kuffner, I.B., Mackenzie, F.T., Jokiel, P.L., Rodgers, K.S. &

    Tan, A. (2009). Net loss of CaCO3 from a subtropical calcifying community due

    to seawater acidification: mesocosm-scale experimental evidence. Biogeosciences, 6,

    1811–1823.

    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. Proc. Natl Acad. Sci. USA, 105, 17442–17446.

    Arnold, S.N., Steneck, R.S. & Mumby, P.J. (2010). Running the gauntlet: inhibitory

    effects of algal turfs on the processes of coral recruitment. Mar. Ecol. Prog. Ser.,

    414, 91–105.

    Connell, S.D. & Russell, B.D. (2010). The direct effects of increasing CO2 and

    temperature on non-calcifying organisms: increasing the potential for phase shifts

    in kelp forests. P. R. Soc. B-Biol. Sci., 277, 1409–1415.

    Diaz-Pulido, G., Harii, S., McCook, L.J. & Hoegh-Guldberg, O. (2010). The impact

    of benthic algae on the settlement of a reef-building coral. Coral Reefs, 29, 203–

    208.

    Diaz-Pulido, G., Gouezo, M., Tilbrook, B., Dove, S. & Anthony, K.R.N. (2011).

    High CO2 enhances the competitive strength of seaweeds over corals. Ecol. Lett.,

    14, 156–162.

    Dickson, A.G. & Millero, F.J. (1987). A comparison of the equilibrium constants

    for the dissociation of carbonic acid in seawater media. Deep-Sea Res., 34, 1733–

    1743.

    Dixson, D.L., Munday, P.L. & Jones, G.P. (2010). Ocean acidification disrupts the

    innate ability of fish to detect predator olfactory cues. Ecol. Lett., 13, 68–75.

    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.

    Fabricius, K.E., Langdon, C., Uthicke, S., Humphrey, C., Noonan, S., De�ath, G.
    et al. (2011). Losers and winners in coral reefs acclimatized to elevated carbon

    dioxide concentrations. Nature Clim. Change, 1, 165–169.

    Hall-Spencer, J.M., Rodolfo-Metalpa, R., Martin, S., Ransome, E., Fine, M.,

    Turner, S.M. et al. (2008). Volcanic carbon dioxide vents show ecosystem effects

    of ocean acidification. Nature, 454, 96–99.

    Harrington, L., Fabricius, K., De�Ath, G. & Negri, A. (2004). Recognition and
    selection of settlement substrata determine post-settlement survival in corals.

    Ecology, 85, 3428–3437.

    Heyward, A.J. & Negri, A.P. (1999). Natural inducers for coral larval metamor-

    phosis. Coral Reefs, 18, 273–279.

    Hoegh-Guldberg, O., Mumby, P.J., Hooten, A.J., Steneck, R.S., Greenfield, P.,

    Gomez, E. et al. (2007). Coral reefs under rapid climate change and ocean

    acidification. Science, 318, 1737–1742.

    Jansen, E., Overpeck, J., Briffa, K.R., Duplessy, J.-C., Joos, F., Masson-Delmotte, V.

    et al. (2007). Palaeoclimate. In: Climate Change 2007: The Physical Science Basis.

    Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental

    Panel on Climate Change (eds Solomon, S., Qin, D., Manning, M., Chen, Z.,

    Marquis, M. & Averyt, K.B., et al.). Cambridge University Press, Cambridge, pp.

    433–498.

    Johnson, C.R. & Sutton, D.C. (1994). Bacteria on the surface of crustose coralline

    algae induce metamorphosis of the crown-of-thorns starfish Acanthaster planci.

    Mar. Biol., 120, 305–310.

    Knowlton, N. (2001). The future of coral reefs. Proc. Natl Acad. Sci. USA, 98, 5419–

    5425.

    Kroeker, K.J., Kordas, R.L., Crim, R.N. & Singh, G.G. (2010). Meta-analysis reveals

    negative yet variable effects of ocean acidification on marine organisms. Ecol.

    Lett., 13, 1419–1434.

    Kuffner, I.B., Walters, L.J., Becerro, M.A., Paul, V.J., Ritson-Williams, R. &

    Beach, K.S. (2006). Inhibition of coral recruitment by macroalgae and cyano-

    bacteria. Mar. Ecol. Prog. Ser., 323, 107–117.

    Kuffner, I.B., Andersson, A.J., Jokiel, P.L., Rodgers, K.S. & Mackenzie, F.T. (2008).

    Decreased abundance of crustose coralline algae due to ocean acidification.

    Nature Geosci., 1, 114–117.

    Langdon, C., Takahashi, T., Sweeney, C., Chipman, D., Goddard, J., Marubini, F.

    et al. (2000). Effect of calcium carbonate saturation state on the calcification rate

    of an experimental coral reef. Global Biogeochem. Cy., 14, 639–654.

    Le Quere, C., Raupach, M.R., Canadell, J.G., Marland, G., Bopp, L., Ciais, P. et al.

    (2009). Trends in the sources and sinks of carbon dioxide. Nature Geosci., 2, 831–836.

    Lechowicz, M.J. (1982). The sampling characteristics of electivity indexes. Oecologia,

    52, 22–30.

    Lewis, P.D.E. & Wallace, D.W.R. (2006). MS excel program developed for CO2
    system calculations. Carbon Dioxide Information Analysis Center, Oak Ridge

    National Laboratory, U.S. Department of Energy, Oak Ridge.

    Littler, D.S. & Littler, M.M. (2003). South Pacific Reef Plants: A Diver�s Guide to the
    Plant Life of South Pacific Coral Reefs. Offshore Graphics, Washington, D.C.

    Marshall, D.J. & Keough, M.J. (2003). Variation in the dispersal potential of non-

    feeding invertebrate larvae: the desperate larva hypothesis and larval size.

    Mar. Ecol. Prog. Ser., 255, 145–153.

    Meehl, G.A., Stocker, T.F., Collins, W.D., Friedlingstein, P., Gaye, A.T.,

    Gregory, J.M. et al. (2007). Global climate projections. In: Climate Change 2007:

    The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment

    Report of the Intergovernmental Panel on Climate Change (eds Solomon, S., Qin, D.,

    Manning, M., Chen, Z., Marquis, M. & Averyt, K.B., et al.). Cambridge University

    Press, Cambridge, pp. 747–845.

    Mehrbach, C., Culberso, C.H., Hawley, J.E. & Pytkowic, R.M. (1973). Measurement

    of apparent dissociation-constants of carbonic-acid in seawater at atmospheric

    pressure. Limnol. Oceanogr., 18, 897–907.

    Morse, D.E., Hooker, N., Morse, A.N.C. & Jensen, R.A. (1988). Control of larval

    metamorphosis and recruitment in sympatric Agaricid corals. J. Exp. Mar. Biol.

    Ecol., 116, 193–217.

    Mumby, P.J., Harborne, A.R., Williams, J., Kappel, C.V., Brumbaugh, D.R., Micheli,

    F. et al. (2007). Trophic cascade facilitates coral recruitment in a marine reserve.

    Proc. Natl Acad. Sci. USA, 104, 8362–8367.

    Munday, P.L., Dixson, D.L., McCormick, M.I., Meekan, M., Ferrari, M.C.O. &

    Chivers, D.P. (2010). Replenishment of fish populations is threatened by ocean

    acidification. Proc. Natl Acad. Sci. USA, 107, 12930–12934.

    Nakamura, M., Ohki, S., Suzuki, A. & Sakai, K. (2011). Coral larvae under ocean

    acidification: survival, metabolism, and metamorphosis. PLoS ONE, 6, e14521.

    Negri, A.P., Webster, N.S., Hill, R.T. & Heyward, A.J. (2001). Metamorphosis of

    broadcast spawning corals in response to bacteria isolated from crustose algae.

    Mar. Ecol. Prog. Ser., 223, 121–131.

    Orr, J.C., Fabry, V.J., Aumont, O., Bopp, L., Doney, S.C., Feely, R.A. et al. (2005).

    Anthropogenic ocean acidification over the twenty-first century and its impact on

    calcifying organisms. Nature, 437, 681–686.

    Penin, L., Michonneau, F., Baird, A.H., Connolly, S.R., Pratchett, M.S., Kayal, M.

    et al. (2010). Early post-settlement mortality and the structure of coral assem-

    blages. Mar. Ecol. Prog. Ser., 408, 55–64.

    Price, N. (2010). Habitat selection, facilitation, and biotic settlement cues affect

    distribution and performance of coral recruits in French Polynesia. Oecologia, 163,

    747–758.

    Raimondi, P.T. & Keough, M.J. (1990). Behavioural variability in marine larvae.

    Aust. J. Ecol., 15, 427–437.

    Ringeltaube, P. & Harvey, A. (2000). Non-geniculate coralline algae (Corallinales,

    Rhodophyta) on Heron Reef, Great Barrier Reef (Australia). Bot. Mar., 43, 431–454.

    Ritson-Williams, R., Paul, V.J., Arnold, S.N. & Steneck, R.S. (2010). Larval settle-

    ment preferences and post-settlement survival of the threatened Caribbean corals

    Acropora palmata and A. cervicornis. Coral Reefs, 29, 71–81.

    Rodriguez, S.R., Ojeda, F.P. & Inestrosa, N.C. (1993). Settlement of benthic marine

    invertebrates. Mar. Ecol. Prog. Ser., 97, 193–207.

    Russell, B.D., Thompson, J.A.I., Falkenberg, L.J. & Connell, S.D. (2009). Syner-

    gistic effects of climate change and local stressors: CO2 and nutrient-driven

    change in subtidal rocky habitats. Glob. Change Biol., 15, 2153–2162.

    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.

    Steneck, R.S. (1986). The ecology of coralline algal crusts: convergent patterns and

    adaptive strategies. Annu. Rev. Ecol. Syst., 17, 273–303.

    Vermeij, M.J.A. & Sandin, S.A. (2008). Density-dependent settlement and mortality

    structure the earliest life phases of a coral population. Ecology, 89, 1994–2004.

    Letter Elevated CO2 alters CCA-larval interactions 345

    � 2012 Blackwell Publishing Ltd/CNRS

    Wallace, C.C. (1985). Seasonal peaks and annual fluctuations in recruitment of

    juvenile scleractinian corals. Mar. Ecol. Prog. Ser., 21, 289–298.

    Webster, N.S., Smith, L.D., Heyward, A.J., Watts, J.E.M., Webb, R.I., Blackall, L.L.

    et al. (2004). Metamorphosis of a scleractinian coral in response to microbial

    biofilms. Appl. Environ. Microbiol., 70, 1213–1221.

    Witt, V., Wild, C., Anthony, K.R.N., Diaz-Pulido, G. & Uthicke, S. (2011). Effects

    of ocean acidification on microbial community composition of, and oxygen

    fluxes through, biofilms from the Great Barrier Reef. Environ. Microbiol., 13,

    2976–2989.

    SUPPORTING INFORMATION

    Additional Supporting Information may be downloaded via the online

    version of this article at Wiley Online Library (www.ecologyletters.com).

    As a service to our authors and readers, this journal provides

    supporting information supplied by the authors. Such materials are

    peer-reviewed and may be re-organized for online delivery, but are not

    copy-edited or typeset. Technical support issues arising from

    supporting information (other than missing files) should be addressed

    to the authors.

    Editor, J. Emmett Duffy

    Manuscript received 15 November 2011

    First decision made 24 December 2011

    Manuscript accepted 8 January 2012

    346 C. Doropoulos et al. Letter

    � 2012 Blackwell Publishing Ltd/CNRS

    Your Name

    Interactionsbetween Plant Semiochemicals and Insects

    There are many methods of communication prevalent in species interactions. However,
    some methods allow species from even different kingdoms to communicate with each other.
    Plants, in order to communicate with insects, release signals known as semiochemicals, which
    are packets of chemicals used to deliver some sort of message. These semiochemicals vary in
    their effects, and different plants have evolved different kinds of semiochemicals for certain
    situations. Examining these chemicals allows humans to understand the varying kinds of insect-
    plant interactions, as well as give humans a means by which insects can be communicated to
    through artificial semiochemicals. These examples of the powerful and efficient effects of
    semiochemicals can show us the importance of cross-species communication.

    I. There are several introductory elements to semiochemicals which must be known.

    Plants interact with insects by the way of ‘odor plumes’ carrying plant volatiles that affect the
    insect’s olfactory senses (Beyaert and Hilker, 2014).

    Semiochemials, in a general sense, operate with a type of ‘push-pull strategy’ when used on
    insects (Cook et al. 2006).

    One basic function of plant semiochemicals is to repel insects. Sometimes this has the added
    effect of ‘inhibiting’ the insect’s ability to sense pheromones (Reddy and Guerrero, 2004).

    Normally, plants release defensive compounds only during the day. Some plants, such as the
    tobacco plant, have evolved to release compounds at night to deal nocturnal herbivores (De
    Moraes et al. 2001).

    II. While some plants use semiochemicals as a basic ‘push’ protection, others are able to
    use them to ‘pull’ insects towards them.

    Some plants, such as orchids, trick insects into thinking they are potential mates not only through
    visual mimicry, but also through semiochemicals (Dettner and Liepert, 1994).

    Some kinds of insects take in plant compounds to use as pheromones of their own (Landolt and
    Phillips, 1997)

    Some plants are able to attract other organisms that prey on the herbivores they are being
    attacked by using volatiles (Bernasconi et al. 1998).

    There is a specific method by which hunter and parasitoid organisms find their host with plant
    semiochemicals (Stowe et al. 1995).

    III. Humans have been able to synthesize semiochemicals for their own uses.

    Your Name

    Humans are able to use semiochemicals as an alternative to normal means of pest control
    (Agelopoulos et al. 1999).

    There are multiple benefits for using semiochemicals for pest management compared to
    insecticides (Witzgall et al. 2010).

    Conclusion:

    As a basis of many examples of insect-plant interactions, semiochemicals are incredibly
    important signals to research and utilize. As studies have shown, the complex effects of
    semiochemicals are invaluable to ensure a plant’s survival against insect herbivory and to bolster
    other relationships with them. However, they are also able to become a boon not only to plants,
    but also to humans, who are able to manufacture semiochemicals in order to manipulate insects
    much more safely than conventional methods. Due to their significant utility as tools,
    semiochemicals can no longer be ignored as solutions to the multitude of problematic insect-
    plant situations

    References

    Agelopoulos N, Birkett MA, Hick AJ, Hooper AM, Pickett JA, Pow EM, Smart LE,
    Smiley DWM, Wadhams LJ, Woodcook CM (1999) Exploiting semiochemicals in insect
    control. Pesticide Science 55:225-235.

    Bernasconi ML, Turlings TCJ, Ambrosetti L, Bassetti P, Dorn S (1998) Herbivore-
    induced emissions of maize volatiles repel the corn leaf aphid, Rhopalosiphum maidis.
    Entomologia Experimentalis et Applicata 87:133-142.

    Beyaert I, Hilker M (2014) Plant odour plumes as mediators of plant-insect interactions.
    Biol. Rev. 89:68-81.

    Cook SM, Khan ZR, Pickett JA (2006) The use of push-pull strategies in integrated pest
    management. Annu. Rev. Entomol. 52:375-400.

    Dettner K, Liepert C (1994) Chemical mimicry and camouflage. Annu. Rev. Entomol.
    39:129-154.

    Landolt PJ, Phillips TW (1997) Host plant influences on sex pheromone behavior of
    phytophagous insects. Annu. Rev. Entomol. 42:371-391.

    Moraes CM, Mescher MC, Tumlinson JH (2001) Caterpillar-induced nocturnal plant
    volatiles repel conspecific females. Nature 410:577-580.

    Your Name

    Reddy GVP, Guerrero A (2004) Interactions of insect pheromones and plant
    semiochemicals. Trends in Plant Science 9:253-261.

    Stowe MK, Turlings TCJ, Loughrin JH, Lewis WJ, Tumlinson JH (1995) The chemistry
    of eavesdropping, alarm, and deceit. Proc. Natl. Acad. Sci. USA 92:23-28.

    Witzgall P, Kirsch P, Cork A (2010) Sex pheromones and their impact on pest
    management. J Chem Ecol 36:80-100.

    Steps in the Preparation of an Annotated Bibliography

    Outline:

    This outline is an overview of the paper that outlines the scope of the paper, describing the topics

    to be covered and the order. This outline will contain all the detail you need to write a complete

    paper. If you prepare your outline correctly, it should be almost as long as the actual paper

    (don’t freak out, that’s a good thing). Below is a general example of what your outline should

    look like, except that yours will be much longer.

    Make sure that your outline includes the following:

    1. Topic sentences for each section (marked by Roman numerals)

    2. Subtopics:

    • Major key points

    3. References for each point that you intend to cover.

    Example

    Title

    • Should inform the reader of what the paper is about.

    • When constructing a title, choose informative over cute.

    I. Introduction: Topic sentence that states the basic idea/premise of your paper. For
    example, if you are examining the role of neuropeptides in parental behavior, your first

    sentence might introduce parental behavior and it’s significance in species survival

    (Reference- see format for in-text citations).

    a. Introduce your system

    i. Define the system, its function, types present, etc. (Reference- see format

    for in-text citations)

    ii. If you are comparing organisms, briefly introduce the organisms

    (Reference- see format for in-text citations)

    b. The main focus of the paper

    c. Provide the scientific reasoning as to why part b is interesting

    d. Be brief and concise

    II. Main Body. Topic sentence focusing on your major points (Reference- see format for in-
    text citations).

    a. You may choose to devote one section to describe the behavior/ ecology /

    scientific relevance or problem that you are focusing on.

    i. Background information

    1. Generalities of the taxon the species belongs to

    2. Behavior and ecology of the species

    3. Scientific relevance of the species (e.g. research breakthroughs that

    have been possible thanks to this species)

    4. Etc. (Reference- see format for in-text citations)

    5. Include as many sections as you deem necessary to cover your

    main ideas.

    III. Conclusion
    a. Summary

    b. Significance

    IV. References (see reference instructions)

    JosephMartinez

    10/16/2020

    Topic in Ecology

    Title: Review of The Impact of Climate Change an Wildfires and It’s Ecological Ramifications

    I. Introduction: This section will focus on introducing and providing background for wildfires

    (and its significance ecologically). The introduction will also introduce the concept of

    climate changes as an amplifying force for intense wildfires in order to set up the

    structure for the rest of the review paper.

    A. Wildfires are naturally occurring phenomena that may temporarily change an

    ecosystem’s composition, however most modern wildfires have had devastating

    effects on ecosystems (Akaike et al., 1974).

    B. Anthropogenic climate change makes intense wildfire more common (Abatzoglou

    & Williams,

    2016).

    C. The main contributing factors are high temperatures, more severe droughts,

    stronger winds, and more frequent lighting strikes- all side effects of climate

    change.

    II. Main Body: This section seeks to explore how wildfires naturally start and what makes

    an intense wildfire, using evidence from the American West as well as the Australian

    outback for a more global perspective. The main body will also break down how each

    factor that contributes to intense wildfires is being amplified by climate change.

    A. The Conditions Necessary for an “Intense” Wildfire- ​The main point of this section

    is that wildfires needs certain conditions to thrive:

    1. Wildfires need hot weather in order to take hold (Nature, 2019).

    2. Wildfires need dry vegetation for “fuel” (Nature, 2019).

    3. Wildfires need strong winds for oxygenation and to spread over long

    distances (Nature, 2019).

    4. Wildfires need a “spark” (lightning, campfire, arson, cigarette) in order to

    ignite the initial flame (Nature, 2019).

    B. How Climate Change is Amplifying these Conditions-​ As a follow up the the

    previous section, it will be explained here how each of these conditions have

    been amplified due to climate change:

    1. Climate change contributes to ever hotter air and surface temperatures,

    leading to “hot weather” (Hansen et al., 2006).

    2. Climate change contributes to prolonged and intense droughts leading to

    vast quantities of dry vegetation (Littell, Peterson, Riley, Liu, & Luce,

    2016).

    3. Climate change has been linked to contributing to stronger and faster

    winds, which are an essential source of oxygenation and spreading

    mechanisms for wildfires (Zeng et al., 2019).

    4. Climate change has been linked to an increase in lightning frequency, one

    of the most common “sparks” that ignite wildfires (Romps, Seeley,

    Vollaro, & Molinari, 2014).

    C. The Ecological Impacts Of Intense Wildfires Globally-​ In this section, the

    ecological effects of wildfires will be explored in order to understand how

    damaging more frequent wildfires will be in the future as climate change

    progresses.

    1. Wildfires release large amounts of previously trapped carbon into the

    atmosphere, creating a feedback loop where wildfire emissions worsen

    climate change which make wildfires more common and devastating

    which leads to more emissions (van der Werf et al., 2017).

    2. Wildfires produce a hydrophobic soil after its been charred and cleared of

    vegetation, leading to more pronounced runoffs after rain. The result is a

    large amount of contaminated water flowing into nearby bodies of water

    dramatically changing their concentrations of nitrogen and phosphorus

    and introducing heavy metals (Hallema et al., 2018).

    3. Intense wildfires can change and reduce the regional biodiversity through

    making certain areas unsuitable for some plants and animals after the

    dramatic change to the environment (Stevens-Rumann et al., 2017).

    III. Conclusion: In this section, the paper will attempt to tie in all the evidence and reiterate

    its results in a concise way. In addition to that, it will point out the relevance of the paper

    holds given current events. This section will also explain the significance and importance

    of intense wildfires being caused by climate change and what we can expect in the

    future as climate change progresses and wildfires become more frequent and

    devastating.

    IV. Citations

    A. Williams, A., Abatzoglou, J., Gershunov, A., Guzman-Morales, J., Bishop, D.,

    Balch, J., & Lettenmaier, D. (2019, August 04). ​Observed Impacts of

    Anthropogenic Climate Change on Wildfire in California​. Retrieved October 15,

    2020, from

    https://agupubs.onlinelibrary.wiley.com/doi/full/10.1029/2019EF001210

    B. Abatzoglou, J., & Williams, A. (2016, October 18).​ Impact of anthropogenic

    climate change on wildfire across western US forests​. Retrieved October 15,

    2020, from ​https://www.pnas.org/content/113/42/11770

    C. LeRoy, W., Anthony LeRoy Westerling Anthony LeRoy Westerling

    http://orcid.org/0000-0003-4573-0595\, Westerling, A., Anthony LeRoy Westerling

    https://agupubs.onlinelibrary.wiley.com/doi/full/10.1029/2019EF001210

    https://www.pnas.org/content/113/42/11770

    http://orcid.org/0000-0003-4573-0595 Google Scholar Find this author on

    PubMed Search for more papers by this author, One contribution of 24 to a

    discussion meeting issue ‘The interaction of fire and mankind’., & Al., E.

    (2016, June 05).​ Increasing western US forest wildfire activity: Sensitivity to

    changes in the timing of spring​.

    Retrieved October 15, 2020, from

    https://royalsocietypublishing.org/doi/10.1098/rstb.2015.0178

    D. Romps, D., Seeley, J., Vollaro, D., & Molinari, J. (2014, November 14).

    Projected increase in lightning strikes in the United States due to global warming​.

    Retrieved October 15, 2020, from

    https://science.sciencemag.org/content/346/6211/851.abstract

    E. Liu, Y., Stanturf, J., & Goodrick, S. (2009, October 03). ​Trends in global

    wildfire potential in a changing climate​. Retrieved October 15, 2020, from

    https://www.sciencedirect.com/science/article/abs/pii/S0378112709006148

    F. Stevens-Rumann, C., Kemp, K., Higuera, P., Harvey, B., Rother, M., Donato, D.,

    Veblen, T. (2017, December 12).​ Evidence for declining forest resilience to

    wildfires under climate change​. Retrieved October 15, 2020, from

    https://onlinelibrary.wiley.com/doi/abs/10.1111/ele.12889

    G. Higuera, P., Abatzoglou, J., Littell, J., & Morgan, P. (n.d.). ​The Changing

    Strength and Nature of Fire-Climate Relationships in the Northern Rocky

    Mountains​, U.S.A., 1902-2008. Retrieved October 15, 2020, from

    https://journals.plos.org/plosone/article?id=10.1371%2Fjournal.pone.0127563

    H. Running, S. (2006, August 18).​ Is Global Warming Causing More, Larger

    Wildfires?​ Retrieved October 15, 2020, from

    https://science.sciencemag.org/content/313/5789/927

    I. Westerling, A., Hidalgo, H., Cayan, D., & Swetnam, T. (2006, August 18).

    Warming and Earlier Spring Increase Western U.S. Forest Wildfire Activity​.

    https://royalsocietypublishing.org/doi/10.1098/rstb.2015.0178

    https://science.sciencemag.org/content/346/6211/851.abstract

    https://www.sciencedirect.com/science/article/abs/pii/S0378112709006148

    https://onlinelibrary.wiley.com/doi/abs/10.1111/ele.12889

    https://journals.plos.org/plosone/article?id=10.1371%2Fjournal.pone.0127563

    https://science.sciencemag.org/content/313/5789/927

    Retrieved October 15, 2020, from

    https://science.sciencemag.org/content/313/5789/940

    J. Dennison, P., Brewer, S., Arnold, J., & Moritz, M. (2014, April 25).​ Large

    wildfire trends in the western United States, 1984–2011​. Retrieved October 15,

    2020, from ​https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2014GL059576

    K. Littell, J., McKenzie, D., Peterson, D., & Westerling, A. (2009, June 01).

    Climate and wildfire area burned in western U.S. ecoprovinces, 1916–2003​.

    Retrieved October 15, 2020, from

    https://esajournals.onlinelibrary.wiley.com/doi/10.1890/07-1183.1

    L. Littell, J., Peterson, D., Riley, K., Liu, Y., & Luce, C. (2016, April 19).​ A review of

    the relationships between drought and forest fire in the United States​. Retrieved

    October 15, 2020, from ​https://onlinelibrary.wiley.com/doi/abs/10.1111/gcb.13275

    M. Harvey, B. (2016, October 15).​ Human-caused climate change is now a key

    driver of forest fire activity in the western United States​. Retrieved October 15,

    2020, from ​https://www.pnas.org/content/113/42/11649

    N. Akaike, H., CD. Allen, D., BJ. Bentz, J., C. Boisvenue, S., Breiman, L., DD.

    Breshears, N., RH. Waring, S. (1974, January 01).​ Forest ecosystems,

    disturbance, and climatic change in Washington State, USA​. Retrieved October

    15, 2020, from ​https://link.springer.com/article/10.1007/s10584-010-9858-x

    O. Zeng, Z., Ziegler, A., Searchinger, T., Yang, L., Chen, A., Ju, K., . . . Wood, E.

    (2019, November 18).​ A reversal in global terrestrial stilling and its implications

    for wind energy production​. Retrieved October 15, 2020, from

    https://www.nature.com/articles/s41558-019-0622-6

    P. Hallema, D., Sun, G., Caldwell, P., Norman, S., Cohen, E., Liu, Y., . . . McNulty,

    S. (2018, April 10).​ Burned forests impact water supplies​. Retrieved October 15,

    2020, from ​https://www.nature.com/articles/s41467-018-03735-6

    https://science.sciencemag.org/content/313/5789/940

    https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2014GL059576

    https://esajournals.onlinelibrary.wiley.com/doi/10.1890/07-1183.1

    https://onlinelibrary.wiley.com/doi/abs/10.1111/gcb.13275

    https://www.pnas.org/content/113/42/11649

    https://link.springer.com/article/10.1007/s10584-010-9858-x

    https://www.nature.com/articles/s41558-019-0622-6

    https://www.nature.com/articles/s41467-018-03735-6

    Q. Van der Werf, G., Randerson, J., Giglio, L., Van Leeuwen, T., Chen, Y., Rogers,

    B., . . . Kasibhatla, P. (2017, September 12). ​Global fire emissions estimates

    during 1997–2016​. Retrieved October 15, 2020, from

    https://essd.copernicus.org/articles/9/697/2017/

    R. Hansen, J., Sato, M., Ruedy, R., Lo, K., Lea, D., & Medina-Elizade, M.

    (2006, September 26).​ Global temperature change​. Retrieved October 15, 2020,

    from ​https://www.pnas.org/content/103/39/14288

    S. The complexities of wildfires​. (2019, January 30). Retrieved October 15, 2020,

    from https://www.nature.com/articles/s41561-019-0311-0

    https://essd.copernicus.org/articles/9/697/2017/

    https://www.pnas.org/content/103/39/14288

    Invasive reptile species of Florida.

    I. Introduction:

    The problem of invasive species is relevant all over the world. In the state of Florida, this

    issue is especially acute due to the hospitable climate that due to warm temperatures and

    increases air moisture makes it incredibly easy for newly introduced species to thrive. The

    invasive species may potentially damage the environment in many ways, human economy,

    health, safety, and negatively impact native species. Among other animals, the reptiles comprise

    quite a long list of the invasive species in Florida: Argentine black and white tegu (Tupinambis

    merianae), black spiny-tailed iguana (Ctenosaura similis), brown anole (Anolis sagrei), the

    Burmese python (Python bivittatus), common house gecko (Hemidactylus frenatus), green

    iguana (Iguana iguana), Mediterranean gecko (Hemidactylus turcicus), Nile Monitor (Varanus

    niloticus). In this review paper, some of the most invasive reptile species in the state of Florida

    will be discussed. Their history, origin, and impact on native species and the environment will be

    addressed. In final part will focus on the investigation of steps and measures taken or planned to

    be taken in the future to reduce or eliminate these species from the state.

    II. History and origin of reptile invasive species in Florida.

    In this section, the history of invasive reptiles will be discussed, the ways of the invasion

    as well as some aspects of the dispersal. There are multiple ways the species may be introduced.

    It may be a natural or human-facilitated event, accidental or deliberate, such as pet trade or zoo

    escape. The key aspects of their success will be presented (Engeman et al., 2011).

    i. Green iguana (Krysko et al., 2007)

    ii. Nile monitor (Enge et al., 2004), (Wood et al., 2016).

    iii. Burmese python (Wilson et al., 2011).

    iv. Black and White Tegu (Pernas et al., 2012).

    v. Human mediated dispersal on the example of common house gecko (Short &

    Petren, 2011), (Muller et al., 2020).

    III. Impact on native species.

    This section will cover particular examples of some invasive reptiles on native species. The

    introduction of the genetic variation as one of the effects will be introduced. Besides, the

    interesting effect of the attempt of their extraction will be considered.

    i. Brown anole effect on native lizard species in Florida (Campbell, 2000).

    ii. Birds predation by Burmese python (Dove et al., 2011).

    iii. The decline of the tree snail due to predation by green iguana in Key Biscayne, Florida

    (Townsend et al., 2005).

    iv. Discussion of the genetic paradox of the increase in genetic variation during the invasion

    of brown anole (Kolbe et al., 2004).

    v. Removal of some invasive species such as green iguanas may negatively affect their

    predators, such as gray fox and raccoons (Meshaka et al., 2007).

    IV. Influence on the environment.

    i. It is important to understand how certain species such as the Burmese python, select the

    locations of their habitat. It would allow identifying the consequences, areas most

    affected, and the scale and direction of the possible spread. (Walters et al., 2016)

    ii. Burmese python as hosts for native mosquito communities may introduce negative

    implications such as an increase in mosquito population and transmission of mosquito-

    vectored pathogens (Reeves et al., 2018).

    V. Economic effects.

    This section will touch upon some economic implications caused by the invasion of reptiles

    using the green iguana as an example.

    i. Associated costs and policy implications due to green iguana invasion and consequent

    infrastructural damages may be counted in billions of dollars. (Sementelli et al., 2008).

    VI. Invasive reptile species management.

    Steps currently that are being taken to reduce the populations of invasive reptile species will

    be discussed in this section as well as some proposed solutions.

    ii. Annual event National Invasive Species Awareness Week that takes place since 2010

    raises the awareness and seeks for solutions (NISAW).

    iii. Overview of Burmese python management steps taken in 2013 (Mazzotti, 2016).

    VII. Conclusion

    The invasion of the species into the non-native environment may not always have

    negative consequences; in some rare cases, it might be even beneficial. However, in the case of

    Florida, the invasive reptile species bring a lot of harm to not only the local environment and

    ecology but also are quite costly to be managed to reduce the destructive effects of their

    presence. It is a task of high importance to preserve the abundant natural habitat of the southern

    state and limit the new invasions on the legislative level. This paper reviewed some of the most

    abundant and damaging species, the ways and history of their introduction, the effects they

    caused, and the possible ways of the control for their population numbers and spreading.

    References

    Campbell, T. S. (2000). Analysis of the effects of an exotic lizard (Anolis sagrei) on a

    native lizard (Anolis carolinensis) in Florida, using islands as experimental units. PhD thesis,

    Univ. Tennessee

    Dove, C. J., Snow, R. W., Rochford, M. R., Mazzotti, F. J. (2011). Birds Consumed by

    the Invasive Burmese Python (Python Molurus Bivittatus) in Everglades National Park, Florida,

    USA. The Wilson Journal of Ornithology, 123(1), 126.

    Enge, K. M., Krysko, K. L., Hankins, K. R., Campbell, T. S., & King, F. W. (2004).

    Status of the Nile monitor (Varanus niloticus) in southwestern Florida. Southeastern Naturalist,

    3(4), 571-582.

    Wood, J. P., Dowell, S. A., Campbell, T. S., Page, R. B. (2016) Insights into the

    Introduction History and Population Genetic Dynamics of the Nile Monitor (Varanus niloticus)

    in Florida, Journal of Heredity, Volume 107, Issue 4, July 2016, Pages 349–362, https://doi-

    org.ezproxy.fiu.edu/10.1093/jhered/esw014

    https://doi-org.ezproxy.fiu.edu/10.1093/jhered/esw014

    https://doi-org.ezproxy.fiu.edu/10.1093/jhered/esw014

    Townsend, J. H., Slapcinsky, J., Krysko, K. L., Donlan, E. M., Golden, E. A. (2005).

    Predation of a Tree Snail Drymaeus multilineatus (Gastropoda: Bulimulidae) by Iguana iguana

    (Reptilia: Iguanidae) on Key Biscayne, Florida. Southeastern Naturalist, 4(2), 361.

    Kolbe, J., Glor, R., Rodríguez Schettino, L. (2004). Genetic variation increases during

    biological invasion by a Cuban lizard. Nature 431, 177–181 https://doi.org/10.1038/nature02807

    Krysko, K. L., Enge, K. M., Donlan, E. M., Seitz, J. C. (2007) Distribution, Natural

    History, and Impacts of the Introduced Green Iguana in Florida, Iguana: Conservation, Natural

    History, and Husbandry of Reptiles, International Reptile Conservation Foundation, 14 (3): 142–

    151

    Mazzotti F. J., Rochford M., Vinci J., Jeffery B. M., Eckles J. K., Dove C., Sommers K.

    P. (2016). Implications of the 2013 Python Challenge® for Ecology and Management of Python

    Molorus bivittatus (Burmese Python) in Florida. Southeastern Naturalist, 15, 63.

    Meshaka, W., Bartlett, R.,Smith, H. (2004) Colonization success by Green Iguanas in

    Florida. Iguana: Journal of the International Iguana Society. State Museum of Pennsylvania,

    Zoology and Botany. 11 (3): 154–161.

    Muller, B. J., Andrews, R. M., Schwarzkopf, L., Pike, D. A. (2020). Social context alters

    retreat- and nest-site selection in a globally invasive gecko, Hemidactylus frenatus. Biological

    Journal of the Linnean Society, 129(2), 388–397. https://doi-

    org.ezproxy.fiu.edu/10.1093/biolinnean/blz188

    Reeves, L. E., Krysko, K. L., Avery, M. L., Gillett-Kaufman, J. L., Kawahara, A. Y.,

    Connelly, C. R., & Kaufman, P. E. (2018). Interactions between the invasive Burmese python,

    Python bivittatus, and the local mosquito community in Florida, USA. PloS one, 13(1),

    e0190633. https://doi.org/10.1371/journal.pone.0190633

    Sementelli, A., Smith, H. T., Meshaka, W., Engeman, R. M. (2008). Just Green Iguanas?

    The Associated Costs and Policy Implications of Exotic Invasive Wildlife in South Florida.

    Public Works Management Policy 2008; 12; 599.DOI: 10.1177/1087724X08316157

    https://doi.org/10.1038/nature02807

    https://doi-org.ezproxy.fiu.edu/10.1093/biolinnean/blz188

    https://doi-org.ezproxy.fiu.edu/10.1093/biolinnean/blz188

    https://doi.org/10.1371/journal.pone.0190633

    Short, K. H., Petren, K. (2011). Multimodal dispersal during the range expansion of the

    tropical house gecko Hemidactylus mabouia. Ecology and evolution, 1(2), 181–190.

    https://doi.org/10.1002/ece3.18

    Walters, T. M., Mazzotti, F. J., Fitz, C. (2016). Habitat Selection by the Invasive Species

    Burmese Python in Southern Florida. Journal of Herpetology, 50(1), 50.

    Pernas, T., Giardina, D. J., McKinley, A., Parns, A., Mazzotti, F. J. (2012). First

    Observations of Nesting by the Argentine Black and White Tegu, Tupinambis merianae, in

    South Florida. Southeastern Naturalist, 11(4), 765.

    Willson, J. D., Dorcas, M. E., Snow, R. W. (2011). Identifying plausible scenarios for the

    establishment of invasive Burmese pythons (Python molurus) in Southern Florida. Biological

    Invasions, 7, 1493.

    https://doi.org/10.1002/ece3.18

    What Will You Get?

    We provide professional writing services to help you score straight A’s by submitting custom written assignments that mirror your guidelines.

    Premium Quality

    Get result-oriented writing and never worry about grades anymore. We follow the highest quality standards to make sure that you get perfect assignments.

    Experienced Writers

    Our writers have experience in dealing with papers of every educational level. You can surely rely on the expertise of our qualified professionals.

    On-Time Delivery

    Your deadline is our threshold for success and we take it very seriously. We make sure you receive your papers before your predefined time.

    24/7 Customer Support

    Someone from our customer support team is always here to respond to your questions. So, hit us up if you have got any ambiguity or concern.

    Complete Confidentiality

    Sit back and relax while we help you out with writing your papers. We have an ultimate policy for keeping your personal and order-related details a secret.

    Authentic Sources

    We assure you that your document will be thoroughly checked for plagiarism and grammatical errors as we use highly authentic and licit sources.

    Moneyback Guarantee

    Still reluctant about placing an order? Our 100% Moneyback Guarantee backs you up on rare occasions where you aren’t satisfied with the writing.

    Order Tracking

    You don’t have to wait for an update for hours; you can track the progress of your order any time you want. We share the status after each step.

    image

    Areas of Expertise

    Although you can leverage our expertise for any writing task, we have a knack for creating flawless papers for the following document types.

    Areas of Expertise

    Although you can leverage our expertise for any writing task, we have a knack for creating flawless papers for the following document types.

    image

    Trusted Partner of 9650+ Students for Writing

    From brainstorming your paper's outline to perfecting its grammar, we perform every step carefully to make your paper worthy of A grade.

    Preferred Writer

    Hire your preferred writer anytime. Simply specify if you want your preferred expert to write your paper and we’ll make that happen.

    Grammar Check Report

    Get an elaborate and authentic grammar check report with your work to have the grammar goodness sealed in your document.

    One Page Summary

    You can purchase this feature if you want our writers to sum up your paper in the form of a concise and well-articulated summary.

    Plagiarism Report

    You don’t have to worry about plagiarism anymore. Get a plagiarism report to certify the uniqueness of your work.

    Free Features $66FREE

    • Most Qualified Writer $10FREE
    • Plagiarism Scan Report $10FREE
    • Unlimited Revisions $08FREE
    • Paper Formatting $05FREE
    • Cover Page $05FREE
    • Referencing & Bibliography $10FREE
    • Dedicated User Area $08FREE
    • 24/7 Order Tracking $05FREE
    • Periodic Email Alerts $05FREE
    image

    Our Services

    Join us for the best experience while seeking writing assistance in your college life. A good grade is all you need to boost up your academic excellence and we are all about it.

    • On-time Delivery
    • 24/7 Order Tracking
    • Access to Authentic Sources
    Academic Writing

    We create perfect papers according to the guidelines.

    Professional Editing

    We seamlessly edit out errors from your papers.

    Thorough Proofreading

    We thoroughly read your final draft to identify errors.

    image

    Delegate Your Challenging Writing Tasks to Experienced Professionals

    Work with ultimate peace of mind because we ensure that your academic work is our responsibility and your grades are a top concern for us!

    Check Out Our Sample Work

    Dedication. Quality. Commitment. Punctuality

    Categories
    All samples
    Essay (any type)
    Essay (any type)
    The Value of a Nursing Degree
    Undergrad. (yrs 3-4)
    Nursing
    2
    View this sample

    It May Not Be Much, but It’s Honest Work!

    Here is what we have achieved so far. These numbers are evidence that we go the extra mile to make your college journey successful.

    0+

    Happy Clients

    0+

    Words Written This Week

    0+

    Ongoing Orders

    0%

    Customer Satisfaction Rate
    image

    Process as Fine as Brewed Coffee

    We have the most intuitive and minimalistic process so that you can easily place an order. Just follow a few steps to unlock success.

    See How We Helped 9000+ Students Achieve Success

    image

    We Analyze Your Problem and Offer Customized Writing

    We understand your guidelines first before delivering any writing service. You can discuss your writing needs and we will have them evaluated by our dedicated team.

    • Clear elicitation of your requirements.
    • Customized writing as per your needs.

    We Mirror Your Guidelines to Deliver Quality Services

    We write your papers in a standardized way. We complete your work in such a way that it turns out to be a perfect description of your guidelines.

    • Proactive analysis of your writing.
    • Active communication to understand requirements.
    image
    image

    We Handle Your Writing Tasks to Ensure Excellent Grades

    We promise you excellent grades and academic excellence that you always longed for. Our writers stay in touch with you via email.

    • Thorough research and analysis for every order.
    • Deliverance of reliable writing service to improve your grades.
    Place an Order Start Chat Now
    image

    Order your essay today and save 30% with the discount code Happy