(Advances in Marine Biology 46) Alan J Southward - P A Tyler - C M Young - Lee A Fuiman-Academic Press (2003)
(Advances in Marine Biology 46) Alan J Southward - P A Tyler - C M Young - Lee A Fuiman-Academic Press (2003)
(Advances in Marine Biology 46) Alan J Southward - P A Tyler - C M Young - Lee A Fuiman-Academic Press (2003)
*The full list of contents for volumes l-37 can be found in volume 38
ix
X Series Contents for Last Ten Years
V
Benthic Foraminifera (Protista) as Tools
in Deep-water Palaeoceanography:
Environmental Influences on
Fauna1 Characteristics
1. Introduction ................................................................... 3
2. Deep-sea Environments ....................................................... 5
3. Methodology: Sieve Sizes, Sampling Devices and Replication ............... 6
4. Aspects of Deep-sea Foraminiferal Ecology ................................... 7
4.1. Introduction .............................................................. 7
4.2. Small-scale patterns ..................................................... 8
4.3. Regional patterns ........................................................ 14
5. Fauna1 Approaches to Reconstructing Palaeoceanography ................... 15
6. Organic Matter Fluxes ......................................................... 18
6.1. General considerations .................................................. 18
6.2. Reconstructing annual flux rates ........................................ 19
6.3. Responses to seasonally varying fluxes ................................. 29
6.4. Are calcareous species more responsive than other foraminifera? ..... 31
7. Oxygen Concentrations ....................................................... 33
7.1. General considerations .................................................. 33
7.2. Qualitative approaches .................................................. 35
7.3. Quantitative approaches ................................................. 37
8. Bottom-water Hydrography ................................................... 39
8.1. General considerations .................................................. 39
8.2. Carbonate undersaturation .............................................. 40
8.3. Current flow ............................................................. 41
9. Water Depth ................................................................... 43
Foramintferal research lies at the border between geology and biology. Benthic
,foramimfera are a major component of marine communities, highly sensitive to
environmental injhrences, and the most abundant benthic organisms preserved
in the deep-sea fossil record. These characteristics make them important tools
for reconstructing ancient oceans. Much of the recent work concerns the
search for palaeoceanographic proxies, particularly ,for the key, parameters of
surface primary productivity and bottom-water oxygenation. At small spatial
scales, organic3ux and pore-water oxygen profiles are believed to control the
depths at which species live within the sediment (their ‘microhabitats’).
Ep[faunal/shallow infaunal species require oxygen and labile food and prefer
relatively oligotrophic settings. Some deep infaunal species can tolerate anoxia
and are closely linked to redox,fronts MYthin the sediment,. they consume more
refractory organic matter, and flourish in relatively eutrophic environments.
Food and oxygen availability are also key j&actors at large (i.e. regional)
spatial scales. Organic flux to the sea floor, and its seasonality, strongly
influences fauna1 densities, species compositions and diversity parameters.
Species tend to be associated with higher or lowerjux rates and the anmral
flux range of 2-3 g C,, m-’ appears to mark an important fauna1 boundary.
The oxygen requirements of benthic foramin~fera are not well understood. It
has been proposed that species distributions reJect oxygen concentrations up
to,fairly high values (3mll-’ or more). Other evidence suggests that oxygen
only begins to afleet community parameters at concentrations < 0.5 ml I-‘.
D@erent species clearly have dtyerent thresholds, however, creating species
successions along oxygen gradients. Other factors such as sediment type,
llydrostatic pressure and attributes of bottom-water masses (particularly
carbonate undersaturation and current ,$ow) influence ,foramintferal distribu-
tions, particularly on continental margins where strong seafloor environmental
gradients exist. Epifaunal species living on elevated substrata are directly
exposed to bottom-water masses and,flourish where suspended,food particles
are advected by strong currents. Biological interactions, e.g. predation and
competition, must also play a role, although this is poorly understood
and d$&ult to quanttfy. Despite often clear qualitative links between
BENTHIC FORAMINIFERA
3
1. INTRODUCTION
and structure of the test wall (Loeblich and Tappan, 1987, 1989; Sen Gupta,
1999). In modern oceans, the most important orders are the following:
AIlogromiida: organic wall, usually monothalamous
Astrorhizidu: agglutinated wall, organic cement, monothalamous
Testuluriidu: agglutinated wall, calcitic cement, polythalamous
Lituolidu: agglutinated wall, organic or calcitic cement, polythalamous
Trochamminidu: agglutinated wall, organic cement, trochospiral
arrangement of chambers
Mifiolidu: wall often with a white, “porcellaneous” appearance in reflected
light, composed of high-Mg calcite, imperforate, usually polythalamous
Lugenidu: wall glassy (‘hyaline’) when fresh, composed of low-Mg calcite,
monolamellar, perforate, monothalamous or polythalamous
Rohertinidu: wall glassy when fresh, composed of aragonite, perforate,
multilocular
Buliminidu: wall glassy when fresh, composed of low-Mg calcite, bilamellar,
perforate, multilocular; chamber arrangement high trochospiral, triserial,
biserial or uniserial; aperture often with toothplate
Rot&da: wall glassy when fresh, composed of low-Mg calcite, bilamellar,
perforate, multilocular; chamber arrangement low trochospiral,
planispiral, or irregular.
It is important to note that, according to recent molecular studies, there is
no phylogenetic distinction between the organic-walled and agglutinated
monothalamous taxa, traditionally referred to the orders Allogromiidu and
Astrorhiziidu respectively (Pawlowski et al., 2001; papers in Cedhagen et al.,
2002). These foraminifera are represented by a series of evolutionary
lineages, many of which include both wall types.
2. DEEP-SEA ENVIRONMENTS
The deep sea lies beyond the shelf break (usually located at around 200m
water depth) and, in a general sense, is a more uniform environment than the
continental shelf. It is characterised by a lack of light, high pressures,
generally low temperatures and constant salinities (Tyler, 1995). Primary
production is confined to chemosynthetic communities located around vents
and cold seeps. The vast majority of organisms are sustained by organic
matter derived from phytoplankton primary production settling through
the water column or by laterally advected material. Although sometimes
considered as a single habitat, substantial environmental differences exist
within the deep sea, particularly between continental margins and
abyssal plains (Berger and Wefer, 1992; Gooday and Rathburn, 1999;
6 ANDREW J. GOODAY
Etter and Mullineaux, 2000). Bathyal continental slopes and rises are
physically much more heterogeneous than abyssal plains. They are often
topographically complex (Mellor and Paull, 1994) and subject to vigorous
current activity and catastrophic mass movements (Masson et al., 1996).
Compared with abyssal plains, sedimentation rates are usually higher on the
continental slope and the sediments are more heterogeneous, less well
oxidised and richer in animal life (Etter and Grassle, 1992; Bett, 2001).
Continental slopes have experienced dramatic oceanographic changes linked
to global climatic fluctuations in the geologic past, particularly during the
Pliocene and Pleistocene glacial cycles. By comparison, the abyssal sea floor
is relatively uniform and quiescent with gently undulating topography.
Because the amount of organic matter reaching the ocean floor decreases
with increasing depth, abyssal environments are typically more food limited
than continental margins that also receive laterally advected organic matter
from the continental shelf. High productivity areas associated with upwelling
or major river discharges are more common on continental margins (Diaz
and Rosenburg, 1995; Rogers, 2000). These environmental contrasts have
ecological consequences for benthic communities. In a general sense, one
would expect sediment types, near-bottom currents and oxygen depletion
coupled with organic enrichment to exert a greater influence on continental
margins (e.g. Schaff et al., 1992; Schmiedl et aE., 1997; Levin et al., 2000) and
patterns of food inputs derived from surface production to be more
important on abyssal plains (e.g. Loubere, 1991; Smith et al., 1997).
(e.g. Gooday, 1988, 1996; Gooday and Lambshead, 1989; Mackensen et al.,
2000; Rathburn et al., 2001; Gooday and Hughes, 2002) (see Figure 7 on
page 57). To ensure maximum comparability, studies ideally should be
based on several different size fractions (> 150, 125-150, 63-125 pm).
Because fine fractions are very time consuming to analyse, it may be
necessary to split samples. Wet samples can be split using the Askii splitter
of Elmgren (1973). The more elaborate device designed by Jensen (1982) is
also very effective.
Many of the earlier ecological studies on deep-sea benthic foraminifera
were based on box core or even Van Veen grab samples. More recently, the
use of hydraulically dampened multiple corers (‘multicorers’) of different
design (e.g. Barnett et ql., 1984) has become widespread. This is an
important technical advance since multicorers retain light, flocculent surface
material such as phytodetritus that is rarely present in box cores (Thiel et al.,
1989). Bett et al. (1994) showed that multicorers sample metazoan
meiofauna much more efficiently than box corers. Recent work on the
UK continental margin suggests that box corers even underestimate
macrofaunal densities by a factor of > 2 compared with multicorers 10 cm
in diameter (Bett, in press). These differences presumably arise because
lighter-bodied, surface-dwelling organisms are blown away by the bow wave
generated by the box corer. Further fauna1 losses from box corers may occur
as the overlying water is drained on deck. Nevertheless, box corers retain
sandy sediments more reliably than multicorers and their greater surface
area permits the recognition of sedimentary features, biogenic and other
habitat structures that may be important for interpreting foraminiferal
assemblages (Schtinfeld, 2002a, 2000~).
Because populations often exhibit considerable small-scale patchiness
(e.g. Gooday and Lambshead, 1989), samples for living foraminifera should
ideally be replicated, for example, by taking one multicore from each
of several deployments. One solution to the additional sorting load imposed
by replication is to take several subcores from a standard multicore
using a cut-off syringe. A 20ml syringe has a cross-sectional area of
3.45 cm* compared to 25.5cm2 in the case of a 57mm internal diameter
multicorer tube.
4.1. Introduction
Foraminifera are one of the principle eukaryotic life forms in the deep sea
and often constitute a substantial proportion of benthic biomass (Snider
8 ANDREW J. GOODAY
et al., 1984; Altenbach and Sarnthein, 1989; Gooday et al., 1992; Kroncke
et al., 2000). Where bottom waters are well oxygenated, live assemblages are
highly diverse, often with well over 100 morphospecies occurring in
relatively small volumes of surface sediment (Gooday et al., 1998). These
assemblages include taxa with organic, agglutinated and calcareous test
walls. The proportion of calcareous foraminifera tends to decline with
increasing water depth (Douglas, 1981; Jorissen et al., 1998; Hughes et al.,
2000), probably reflecting a decrease in the organic carbon flux to the sea
floor. At great depths, carbonate dissolution becomes important (Berger,
1979) and below the Carbonate Compensation Depth (CCD: generally
> 4000-5500m, but considerably shallower in some areas around
Antarctica), faunas consist almost entirely of taxa with agglutinated or
organic tests (Saidova, 1967). Many of them are undescribed soft-walled
forms belonging to groups such as the Komokiacea (Tendal and Hessler,
1977; Schriider et al., 1989; Gooday, 1990) which disintegrate rapidly after
death. Foraminifera play an important role in deep-sea ecology, for
example, by processing of fresh organic material deposited on the sea floor
(Moodley et al., 2002) as prey for other organisms (Gooday et al., 1992)
and by providing habitat structure (Levin, 1991).
The use of benthic foraminifera in palaeoceanography is based on
ecological observations made at spatial scales ranging from centimetres (e.g.
sediment microhabitats) to 100-1000 km2 (regional distributions). One
overriding factor, the organic matter flux to the sea floor, pervades much of
the recent literature on deep-sea foraminiferal ecology (Jorissen, 1999). The
organic flux delivers food to the benthos. It is also inversely related to
bottom-water oxygenation and controls oxygen profiles and other geo-
chemical gradients within the sediment. These, in turn, influence foramini-
fera and other sediment-dwelling organisms. In some areas, regional fauna1
patterns also clearly reflect other factors, notably the imprint of bottom-
water hydrography.
During the 1980s it was recognised that species tend to occupy distinct
horizontal levels within the sediment profile rather than being confined to
the surface layer (Basov and Khusid, 1983; Corliss, 1985; Gooday, 1986).
Various terms have been used to categorise these microhabitats; for
example, epifaunal (O-l cm), shallow (O-2 cm), intermediate infaunal
(14 cm), transitional (04 cm), deep infaunal ( > 4 cm) (Corliss, 1991;
Rathburn and Corliss, 1994; Rathburn et al., 1996; Mackensen, 1997).
Jorissen (1999) considers these schemes too rigid and recognises instead four
basic patterns: (1) type A - population maximum near sediment surface, (2)
BENTHIC FORAMINIFERA
9
Well-oxygenated
continental margins and
Food increases - - -
- - - Oxygen increases
the food is located. Eutrophic systems are oxygen limited and species are
concentrated near the surface into order to avoid anoxic conditions deeper
in the sediment profile. Maximum penetration is found in intermediate
(‘mesotrophic’) settings where both food and oxygen are available well
below the sediment/water interface. This basic scheme has been refined by
Jorissen et al. (1998), Jorissen (1999), van der Zwaan et al. (1999) and
Fontanier et al. (2002) who make the following suggestions: (1) the organic
flux is the pre-eminent parameter controlling foraminiferal microhabitats;
(2) Oxygen is not a limiting factor for deep infaunal (Type C) species that
occur below the subsurface oxic/anoxic interface. These species may be more
closely linked to subsurface accumulations of organic matter (Rathburn and
Corliss, 1994) or to populations of anaerobic bacteria associated with redox
boundaries (Jorissen et al., 1998; Fontamer et al., 2002), (3) Biological
interactions, particularly competition for labile food material, play a role in
determining where foraminifera live within the sediment profile. The TROX
model and its successors provide a useful framework for understanding
how various factors may interact to control foraminiferal microhabitats,
although they are qualitative and cannot be used to reconstruct values for
parameters such as organic fluxes directly.
Corliss and colleagues (Corliss, 1985, 1991; Corliss and Chen, 1988;
Roscoff and Corliss, 1991; Rathburn and Corliss, 1994) related microhabitat
preferences to calcareous test morphotypes. (1) “Epifaunal” species (those
living in the top 1 cm of sediment, i.e. shallow infaunal of some authors)
tend to have either milioline coiling, trochospiral tests with rounded, plano-
convex or biconvex shapes and pores either absent or confined to one side of
the test (Figures 2A-F, 3H-I). (2) Infaunal species (those living at > 1 cm
depth) tend to have tests that are rounded and planispiral or flattened ovoid,
flattened tapered, tapered and cylindrical or spherical in shape with pores
present all over the test (Figure 3A-G). There are many exceptions to these
generalisations, and microhabitats cannot always be predicted from
morphotypes (Jorissen 1999), but assignments seem to be accurate in most
(-75%) cases (Buzas et al., 1993). Thus the linkage between test
morphotypes and microhabitats, although imperfect, provides a basis for
analysing relationships between foraminiferal faunas, depth in the sediment,
and hence food and oxygen availability.
version of TROX model reproduced from “Modem Foraminifera” (editor B.K. Sen
Gupta), 1999, p. 175, Benthic forammiferal microhabitats below the sediment-water
interface, F. Jorissen, Figure 10.9, with kind permission of Kluwer Academic
Publishers. The original version of the TROX model was published in Marine
Micropaleontology Vol. 26, F.J. Jorissen, H.C. de Stigter, J.G. Widmark, A
conceptual model explaining benthic foraminiferal microhabitats, pp. 3-l 5, 1995,
with permission from Elsevier Science.
12 ANDREW J. GOODAY
Surface primary productivity/ Abundance of foraminiferal Transfer function links Herguera and Berger (1992)
organic matter flux to tests > 150 urn “benthic foraminiferal
sea floor accumulation rate” (BFAR)
to productivity
Organic matter flux to sea floor Assemblages of “high produc- Assemblages indicate high Sarnthein and
tivity” taxa (e.g. organic matter flux to Altenbach (1995);
Globobulimina, Melonis) sea floor, with or without Altenbach et al. (1999)
corresponding decrease in
oxygen concentrations; high
percentages of some species
characteristic of particular
flux ranges
Organic matter flux to sea floor Ratio between infaunal and Infaunal morphotypes tend to Corliss and Chen (1988)
epifaunal morphotypes dominate in high produ-
ctivity areas
Surface ocean productivity and Principle components analysis Requires large dataset for Loubere (1991, 1994, 1996);
organic carbon flux to of species abundance data calibration Loubere and Fariduddin
sea floor (1999a)
Seasonality in organic matter Relative abundance of “phyto- Reflects seasonally pulsed Thomas et al. (1995)
flux detritus species” inputs of labile organic b
matter to sea floor =L
Seasonality in surface ocean Discriminant function Loubere (1998); Loubere and :
productivity and organic analysis of assemblage data Fariduddin (1999a) 3
c
carbon flux to sea floor from E. Pacific Ocean (low
seasonality) and Indian
Ocean (highly seasonal)
Oxygen-deficient bottom- and (i) Characteristic species asso- (i) Species not consistently (i) Sen Gupta and Machain-
pore-water ciations associated with particular Castillo (1993), Bernhard z
range of oxygen concen- f? al. (1997) 5
trations and also found in 0
high productivity areas 6
(ii) Transfer function based on (ii) Proportion of different (ii) Kaiho (1991, 1994, 1999); $
relative frequency of infau- morphotypes also related Van der Zwaan et ul. z
nal and epifaunal morpho- to organic flux (in Kouwenhoven, 2000) $
types
(iii) Patterns of species diver- (iii) Oxygen-deficient environ- (iii) Den Dulk et ~1. (1998); P
sity and dominance ments characterised by Gooday et ul. (2000)
low diversity/high domi-
nance assemblages
CaCOj corrosive bottom Abundance of Nuttallides Distribution of N. umhon@r (i) Mackensen et al. (1995)
waterjoligotrophic umhon@r linked to (i) corrosive bot- (ii) Loubere (1991)
conditions tom water (broadly corre-
sponds to Antarctic Bottom
Water); (ii) highly oligo-
trophic conditions.
Current flow Characteristic associations of Species are suspension feeders Mackensen et al. (1995);
sessile epifaunal species that capture food particles Schanfeld (1997, 2002a,c)
living on raised substrates advected by currents
Water depth (i) Bathymetric ranges of abun- (i) Ranges depend on organic (i) Phleger (1960); Phlumm and
dant species in modern matter fluxes to sea floor Frerichs (1976); Culver
oceans and therefore largely local, (1988)
although broad distinction
between shelf, slope and
abyssal depth zones is pos-
sible.
(ii) Ratio between planktonic (ii) Ratio is independent of flux (ii) Van der Zwaan et ~11.(1990,
and benthic tests intensity; estimates become 1999)
less accurate with increasing
water depth
18 ANDREW J. GOODAY
The search for proxies of particulate organic matter (POM) fluxes to the
sea floor is a major goal in palaeoceanography. Much of the recent
geologically orientated research on deep-sea benthic foraminifera has
addressed this issue (e.g. Jorissen et al., 1998; papers in Jorissen and
Rohling, 2000; Morigi et al., 2001). On continental margins, refractory
organic material is transported down the continental slope by various
mechanisms, including nepheloid layers, turbidity currents and down-
canyon currents. A large proportion of the labile POM arriving at the ocean
floor, however, originates from phytoplankton primary production in the
overlying water column. This is particularly true in central oceanic areas
where the POM flux largely reflects the intensity of surface primary
production and lateral advection from slope and shelf areas is not a
significant factor. The material that settles out below the zone of winter
mixing constitutes the long-term export production to the ocean interior
(Berger and Wefer, 1990). In open-ocean settings, only a small fraction
(O.Ol-1.0%) of this exported material reaches the bottom and this fraction
decreases with increasing water depth (Suess, 1980; Berger et al., 1988, 1989;
Berger and Wefer, 1992). The flux at 2000m shows a linear relation with
levels of primary production below production levels of 200 g Cm-* y-‘, but
at higher levels the flux remains constant, for reasons that are not well
understood (Lampitt and Antia, 1997).
Although the complex processes by which organic matter derived from
surface production is delivered to the ocean floor (‘bentho-pelagic coupling’)
are understood in general terms, actual rates of supply are more difficult to
determine accurately (Berger and Wefer, 1992; Murray, 2001). Estimates are
often derived from empirical equations that incorporate primary produc-
tion, export production, and flux rate data obtained from sediment traps
(Suess, 1980; Pace et al., 1987; Berger et al., 1988, 1989; Berger and Wefer,
1990, 1992). These parameters are not necessarily well constrained. In
particular, primary productivity estimates may vary by a factor of 2-3 and
exhibit considerable variability, both spatially and temporally (Berger et al.,
1988; Herguera, 2000). Oxygen fluxes across the sediment-water interface,
obtained by measuring either sediment pore water oxygen profiles or
sediment community oxygen consumption (SCOC), provide a more direct
BENTHIC FORAMINIFERA 19
and time-averaged measure of POM fluxes (e.g. Loubere et al., 1993; Graf
et al., 1995; Jahnke, 1996, 2002; Rowe et al., 1997; Sauter et al., 2001). Even
this approach is not without problems since oxygen fluxes reflect inputs of
refractory carbon (e.g. redeposited material) of limited nutritional value to
foraminifera, as well as labile material. These data are still relatively scarce,
although they can be extrapolated using other measures as proxies (Jahnke,
1996). Thus, despite considerable improvements in our knowledge of ocean-
wide and global patterns of POC fluxes, values at particular localities will
often be subject to substantial uncertainties, a fact that complicates the task
of calibrating flux proxies (Berger et al., 1994).
Another complicating factor is that primary production and export flux
usually have a more or less distinct seasonal component (Berger and Wefer,
1990; Lampitt and Antia, 1997) that is transmitted down through the
oceanic water column (Asper et al., 1992; Turley et al., 1995), leading to the
seasonally pulsed deposition of phytodetritus on the sea floor (Billett et al.,
1983). In the temperate abyssal NE Atlantic Ocean, these deposits deliver an
estimated 224% of spring-bloom production to the benthos (Turley et al.,
1995). The strength of seasonality in the vertical flux is related to the nature
of the pelagic ecosystem (Lampitt and Antia, 1997) i.e. the “plankton
climate” provinces of Longhurst (1996, 1998). It is most intense at high
latitudes and least intense in tropical regions (Fischer et al., 1988; Berger
and Wefer, 1990; Wefer and Fischer, 1991; Ramseier et al., 1999). Berger
and Wefer (1990) suggest that export production is higher in strongly
seasonal systems compared with more constant ones, although this is not
confirmed by sediment trap data (Lampitt and Antia, 1996).
Area (water depth) Size fraction Oxygen (ml I--‘) Characteristic species Reference
NW African margin off Cap > 250 urn > 1.0 Bulimina marginutu, Lutze (1980)
Barbas & Cap Blanc Chilostomellu oolina, Lutze and Coulbourne
Globobulimina spp., (1984)
Uvigerina peregrina
NW Africa off Cap Blanc > 150um 4.5 Globobulimina pyrula, Melonis Jorissen et al. (1998)
barleeanum, Uvigerina
peregrina
Tropical Atlantic >63um 5.0 Alabaminella weddeliensis Fariduddin and
Loubere (1997)
Eastern South Atlantic: > 125 urn 2.7-5.1 Bulimina spp., Uvigerina Schmiedl and
lower slope off Cunene auberiana, Fursenkoina Mackensen (1997)
River (800-2000 m) mexicana, Vulvulineria
laevigatu
Lower slope off Cunene River > 125 urn 5.2 Melonis spp., U. peregrinu, Schmiedl and
(3000-4000 m) Globobulimina turgida, Mackensen (1997)
Chilostomella oolina,
Nonionella opima,
Cassidulinu renijorme
East Pacific Rise >63pm 3.3-3.1 Hispid Uvigerina; Melonis Loubere ( 199 1)
burleeanum
Eastern equatorial and North > 63 pm -1.8-3.5 A. wrcldellensis, Bulimina alazi- Loubere (1996)
Pacific Ocean nensis, Chilostomellu oolina,
Globobulimina sp.,
Sphueroidinu bulloides,
Stainforthia sp.
NE US slope (350-500 m) > 250 pm -3.0 Globobulimina spp.. Bulimina Miller and Lohmann (1982)
aculeota
North Carolina slope off Cape > 300 pm -4.0 Globobulimina auriculata Gooday et al. (2001)
Hatterras (X50 m) dominant; Ammobaculites
agglutinans, Hormosina
dentalinijormis also
important
22 ANDREW J. GOODAY
Higher-flux species
Trifarina fornasinii lo-30 100-300 Inner and outer shelf
Uvigerina mediterranea 2-9 150-250 Slope (20&1000m)
Uvigerina peregrina 2-20 100-300 Lower slope
(700-2000 m)
Hoeglundina elegans 2.5515 20&280 Lower slope
(40&2000 m)
Sphaeroidina bulloides 3-12 9&300 Slope (70&1000m)
Bolivina albatrossi 5-15 100-300 Slope (300-1000m)
Cibicidoides 2.5-20 90-300 Slope (25&l 500 m)
pseudoungerianus
Globobulimina spp. >3 Slope
Intermediate-flux species
Cibicidoides kullenbergi 14 80-250 Lower slope/rise
(2000-4000 m)
Lower-flux species
Cibicidoides wuellerstorfL 0.2-3.0 15-100 Lower slope - abyssal
’ Pyrgo rotalaria 0.2-2.5 15-200 Lower slope - abyssal
Eponides tumidulus < 0.4 < lo-25 Abyssal
‘Epistominella arctica 0.03-2.0 Upper slope,
Arctic Ocean
‘Stetsonia hovarthi < 0.4 < l&25 Abyssal
Oridorsalis umbonatus < 1.5 Abyssal
Nuttallides umbonifera < 1.5 Abyssal
Species spanning wide
flux range
Epistominella exigua 0.9-100 Slope-abyssal
Melonis zaandami 2-l Slope-abyssal
Cribrostomoides 0.410 Slope-abyssal
subglobosum
Adercotryma glomeratum 0.03312 Slope-abyssal,
Arctic Ocean
‘Often reported as Pj~go murrhinu or Pyrgo murrhyna; Pyrgo rotalaria is the senior synonym
(Thies, 1991).
“Epl~tominella arctica and Stetsonia hovarthi are considered synonyms by Scott and Vilks
(1991).
24 ANDREW J. GOODAY
estimate flux rates < 2 g Corg me2 with reasonable confidence. In addition,
some infaunal species (e.g. Uvigerina mediterranea Hofker 1932) tend to be
associated with a specific range of flux values irrespective of their abundance.
Sarnthein and Altenbach (1995) and Altenbach et al. (1999) suggest that the
annual flux range of 2-3 g Corg m-* marks an upper threshold of dominance
for many species characteristic of oligotrophic abyssal settings and a lower
threshold of dominance for species adapted to more eutrophic shelf and
bathyal environments. Other authors have also reported evidence for an
ecological boundary around the same flux levels (De Rijk et al., 2000;
Jian et al., 1999; Morigi et al., 2001; Weinelt et al., 2001).
These approaches rely on the relationship between organic fluxes and the
relative abundance of infaunal and epifaunal morphotypes. The use of
morphotypes as flux indicators is complicated by the fact that they are also
related to oxygen availability, at least at low oxygen concentrations. This is
discussed in Section 7.
Corliss and Chen (1988) reanalysed the data of Mackensen et al. (1985)
from the Norwegian margin (dead assemblage, O-1 cm layer, > 125 urn
fraction), assigning the species to either infaunal or epifaunal morphotypes.
Both categories occurred between 200m and 500m water depth, infaunal
morphotypes predominated between 500 m and 1500 m, epifaunal morpho-
types were most abundant below this depth (Corliss and Chen, 1988).
A similar pattern was observed by Roscoff and Corliss (1991) in the
Greenland-Norwegian Sea. Below 800 m depth, these patterns correlated
well with the organic carbon content of surface sediments; high organic
carbon values were associated with dominance by infaunal morphotypes,
low carbon values with epifaunal morphotypes. Corliss and Chen (1988)
suggest that the switch from infaunal to epifaunal dominance occurs within
the yearly organic carbon flux range 3-6 g Corg m-*. Despite the imperfect
relationship between microhabitats and morphotypes referred to above, the
Corliss and Chen (1988) approach can provide a general indication of
organic fluxes levels. The quality of the available food is also important.
Deep infaunal species living below the level at which oxygen disappears
from the sediment pore water apparently consume more degraded organic
matter than epifaunal and shallow-infaunal species (Goldstein and Corliss,
1994; Fontanier et al., 2002). The abundance of the former and scarcity of
the latter at a site off Cap Blanc (NW African margin) overlain by well-
oxygenated bottom water (4.5 ml/l) was attributed by Jorissen et al. (1998)
to the lack of freshly deposited labile detritus compared to the relatively
large amounts of more degraded material available deeper in the sediment.
BENTHIC FORAMINIFERA 25
Thus, as van der Zwaan et al. (1999) conclude, infaunal morphotypes reflect
the abundance of organic matter stored within the sediment, rather than the
flux of fresh material.
As well as the overall scale of the flux (i.e. the annual flux rate), the degree
of seasonality in its delivery to the sea floor is an important parameter
(Gooday, 2002; Hayward et al., 2002; Loubere and Fariddudin, 1999b).
Seasonal delivery of food leads to temporal fluctuation in the population
densities of some benthic foraminiferal species that are not reflected in dead
assemblages (Douglas et al., 1980). Indeed, seasonality generally is a difficult
parameter to detect in the palaeoceanographic record (Smart et al., 1994;
Thomas and Gooday, 1996).
Pulsed fluxes of phytodetritus, usually reflecting seasonal surface
production, occur in the temperate North Atlantic Ocean (Hecker, 1990b;
Rice et al., 1994), the Greenland-Norwegian Sea (Graf et al., 1995), the
Southern Ocean (Mackensen et al., 1993), the monsoon-influenced Arabian
Sea (Pfannkuche et al., 2000) and the NE Pacific Ocean (Smith and Druffel,
1998). Beaulieu (2002) provides a detailed review of the occurrence,
composition and origin of phytodetritus, its geochemical significance and
fate on the seafloor. A limited number of small foraminiferal species are
physically associated with phytodetrital deposits. In open-ocean areas,
where the overall organic matter flux is not too high, the best-known are
Epistominella exigua (Brady) (Figure 2E-F) and Alabaminella weddellensis
(Earland), both species with cosmopolitan distributions at abyssal depths.
In the temperate NE Atlantic Ocean, individuals are commonly found living
within phytodetrital aggregates (Gooday, 1988, 1993, 1996). Epistominella
exigua is also associated with strong seasonality in primary production in
the Indian Ocean (Loubere, 1998). These “phytodetritus species” are
probably enrichment opportunists that undergo rapid population increase
when presented with a good food supply (Gooday and Rathburn, 1999).
Experiments suggest that shallow-water foraminiferal species grow con-
tinuously and rapidly when adequate food is available but slowly when food
is scarce (Bradshaw, 1961). Deep-water species also undergo rapid
population growth when presented with a pulse of algal food under
experimental conditions (Heinz et al., 2001, 2002).
On bathyal continental margins, where conditions are more eutrophic,
other small benthic species respond to phytodetritus. In the NE Atlantic
Ocean, the two most common are Nonionellu iridea Heron-Allen & Earland
and Eponides pusillus Parr (= Eilohedra nipponica (Kuwano) of Wollenburg
and Mackensen, 1998), both of which are abundant at BENBO Site C (1900 m
water depth) and at a 1340-m site in the Porcupine Seabight (Gooday and
Lambshead, 1989; Gooday and Hughes, 2002). In the high Arctic.
Epistominellu arctica Green 1960 is an opportunist that reproduces rapidly
during short-pulsed, local phytoplankton blooms (Wollenburg and Kuhnt,
2000). It apparently prefers more oligotrophic conditions than either
30 ANDREW J. GOODAY
modest annual flux intensity. Other “phytodetritus species” may have rather
different ecological requirements. Statistical analyses of samples from the
Atlantic (Fariduddin and Loubere, 1997) Pacific (Loubere, 1996; Hayward
et al., 2002), and Indian (Loubere, 1998) Oceans tend to group Alabamineflu
weddellensis with high productivity species, rather than with E. exigua. Fossil
occurrences support the view that A. weddellensis and E. exigua have different
ecological characteristics (Nees and Struck, 1999; Okhushi et al., 2000).
Species do not always exhibit the same response to pulsed food inputs
across their entire range. Epistominella arctica is an opportunist in the High
Arctic (Wollenburg and Kuhnt, 2000) but at the temperate BENBO Site C it
shows only a modest numerical increase in post-bloom (July 1998) samples
with phytodetritus compared to pre-bloom (May 1998) samples devoid of
phytodetritus, and is never associated directly with these deposits (Gooday
and Hughes, 2002). If this tiny species is adapted to very low productivity
combined with extremely short pulses of phytodetritus, as suggested by
Thomas et al. (1995) then conditions at Site C are probably not ideal for it.
Similarly, Stainforthiu jkiformis (Williamson) occurs in relatively low
number in “live” assemblages at this seasonal bathyal site but is reported to
be highly opportunistic in shallower, continental shelf settings (Alve, 1994).
These observations suggest that foraminiferal species may exhibit different
life-history characteristics in different areas. Where conditions are optimal,
they may react opportunistically to a fluctuating food supply. Near the
edges of their range, however, factors close to the tolerance limit exert
strong controls that dampen opportunistic responses.
300
250
200
150
E9) 100
z
E 50
ii 0
a
2 3000-
s 2500.
.g
= 2cm.
Major taxa
7. OXYGEN CONCENTRATIONS
content umol l-l]= 7.9602 + 5.95 [% oxyphilic species]. This proxy, which
was calibrated using modern foraminiferal data from the Mediterranean Sea
and Atlantic and Indian Oceans, depends on the fact that oxyphilic species
require free oxygen and therefore are confined to fully oxic microenviron-
ments close to the sediment surface. Kouwenhoven (2000) used this method
to provide a generally convincing reconstruction of oxygenation history of
the late Miocene (6.3-8.1 million years ago) Mediterranean (Monte de1
Casino section in Northern Italy). However, in places, it yielded oxygen
estimates that were either unrealistically low or peaked to unrealistically high
values, depending on whether or not particular species were included in
the calculations. Jannink et al. (in Jannink, 2001) proposed a very similar
transfer function [Oxygen content umol l-l] = 7.23 + 5.62 [% oxyphilic
species] with an R2 value of 0.66, also based on Mediterranean and
Atlantic and Indian Ocean material representing a wide variety of product-
ivity regimes. They argued that the abundance of oxyphilic taxa is regulated
by the volume of aerated sediment (i.e. living space) rather than bottom-
water oxygen values as such. When applied to a core from the North
Adriatic, this proxy yielded oxygenation estimates for the past 160 years that
corresponded well with historical data. This method seems to produce
plausible results when applied cautiously, although it may require further
refinement.
8. BOTTOM-WATER HYDROGRAPHY
9. WATER DEPTH
characteristics, all of which are also to some extent related to water depth
(Hayward et al., 2002). It seems likely that organic fluxes to the sea floor are
particularly important in determining the bathymetric distributions of
foraminiferal species (Haake et al., 1982; van der Zwaan et al., 1999). Fluxes
decrease with increasing water depth at any particular locality, and also vary
from region to region. Analysing the relation between the percentage
abundance of species and flux intensity, Altenbach et al. (1999) concluded
that, down to a water depth of about 1000 m, species “patterns are depth
dependent with reduced influence from organic matter flux”. Below this
limit, patterns of species abundance follow flux values down to depths of
2000 m, below which species tend to be very widely distributed. In the
Mediterranean Sea, there is a clear relation between flux and bathymetric
distribution of species. De Rijk et al. (2000) described a shallowing of the
lower water depth limit for many species from west to east, corresponding
to a change from eutrophic to oligotrophic conditions. They conclude that
“...the bathymetric distribution of the dominant foraminiferal taxa seems
indeed to be controlled by the level of the organic flux to the sea floor.”
De Stigter et al. (in De Stigter, 1996, Chapter 6 therein) found that
foraminiferal species lived at greater water depths in the South Adriatic
Basin during the late Pleistocene compared with their present-day
distributions. They attributed this upward bathymetric shift to a decrease
in primary productivity and POC flux to the sea floor since the late
Pleistocene.
These results imply that benthic foraminifera can provide only a broad
indication of bathymetry (Murray, 1991). Although sucessions of species
invariably occur with increasing water depth along continental margin
transects (Pujos-Lamy, 1972; Pflumm and Frerichs, 1976; Haake, 1980;
Lutze, 1980; Douglas and Woodruff, 1981) species depth ranges are not
consistent between regions. The distributions of modern foraminiferal
species and species assemblages can therefore be used to reconstruct
palaeobathymetry only when applied to fossil faunas in the same study area
(e.g. Culver, 1988; Hayward, 1990; Kamp et al., 1998; Akimoto et al., 2002).
A more generalised approach to estimating palaeodepths exploits the
relation between the abundance of planktonic and benthic foraminifera and
organic fluxes. Building on the ideas of earlier authors (e.g. Berger and
Diester Haas, 1988), van der Zwaan et al. (1990, 1999) suggested that the
planktonic/benthic ratio (P/B ratio = the percentage of planktonic forami-
niferal tests in the total assemblage) reflects water depth. A better relation
between the P/B ratio and water depth was obtained when infaunal species,
which are less closely linked to the freshly settled flux, were excluded. This
relationship can be used to determine approximate palaeodepths and works
fairly well as long as the ratios are not distorted by selective dissolution or
by high inputs of organic matter and oxygen depletion. It has been used, for
BENTHIC FORAMINIFERA 45
(continued)
5
Table 4. Continued.
BENBO Site A
O-1 cm, > 125 urn
3569m (Aug. 1997) 19 3830 11 62 - 27.3 15.8 38.5 230 4398 50 77 31.2 32.6 36.1 30.9 H,L
3576m (May 1998) 59 20 23.0 20.3 391 62 31.4 24.3
3567m (July 1998) 103 17 17.4 75.0 298 43 25.1 26.2
BENBO Site B
O-1 cm, > 125 urn 136 326 17 39 17.8 21.5 16.2 21.3 274 360 49 41 32.3 25.1 19.7 24.7 H,L
I 1OOm (Aug. 1997)
BENBO Site C
O-l cm, > 125 urn -6.0 68 247 19 44 19.0 30.4 3X.2 18.2 430 315 56 61 28.2 37.0 27.4 14.2 H,L
1926m (Aug. 1997)
L
BENBO Site C zi
O-1 cm; > 63 urn -6.0 3
1913 m (May 1998) 397 1247 32 42 19.9 18.9 31.6 27.8 2265 1306 158 78 48 22 13.0 26.2 L
1913m (May 1998) 237 2013 23 41 16.4 20.0 38.4 19.1 2115 73 24 19.3 I n
1963 m (July 1998) 981 1668 33 39 15.1 15.8 46.1 24.5 6547 1855 176 62 34 22 20.4 24.5 8
2818 1265 38 44 12.8 17.7 44.0 29.5 1350 68 23 21.6 T
1980m (July 1998)
‘,*Porcupine Abyssal
Plain: > 63 pm Oxic .z
z
4845 m: O-l cm 210 15 11.7 38.0 1
&lOcm 181 28 21.7 40.9 E 0
11908#70: O-1 cm 1386 64 16.6 35.0 1763 102 25.1 27.5 6
I 1908#5: O-1 cm 813 53 16.8 36.2 1010 87 25.6 29.1 B
11908#32: O-l cm 1034 58 17.6 34.5 1234 87 26.0 2X.9 3
5
Madeira Abyssal Plain T,
>63pm Oxic zID
4940 m: O-1 cm 54 13 38.8 217 80
O-10 cm 61 16 36.1 240 83
12174#24: O-1 cm 182 24 18.2 30.2 311 58 33.3 17.7
12174#15: O-1 cm 167 29 21.8 2X.7 298 59 33.5 16.1 E
12174#88: O-l cm 65 18 18.0 21.5 152 45 37.2 21.5
Bay of Biscay:
> 150pm
140m 4.9 Xmm
O-l cm 925 22 13.0 37.9 1221 31 15.1 26.5
0-IOcm 1584 25 14.3 34.9 1989 36 16.8 30.2
553 m: 4.8 27 mm
O-l cm 914 34 13.4 37.7 1040 45 18.1 38.8
O-1Ocm 1205 37 13.6 44.2 1345 49 18.6 33.8
1012m: 4.4 20 mm F
Cl cm 150 18 15.4 35.6 208 28 22.2 19.7
0-1Ocm 385 25 15.8 27.3 476 37 22.9 28.8
1264m: 4.7 64 mm
Glcm 12 14 14 39.0 16 16 16 39.5
O-IOcm 105 18 17.7 41.7 122 25 23.5 33.6
1993 m: 5.8 63mm
O-1 cm 107 13 12.8 36.5 125 22 20.6 40.8
0-IOcm 159 18 15.4 47.7 179 27 22.2 32.4
%
(continued)
Table 4. Continued.
Off NW Africa:
> 150pm
1195m 3.67 1.5mm
c-1 cm 558 39 19.8 32.8 848 63 27.3 21.6
&lOcm 891 44 19.4 24.8 1297 69 28.2 17.0
1525m 4.29 l.Omm
O-lcm 132 26 24 26.5 223 40 30.7 15.7
&lOcm 253 32 23 16.6 449 48 31.7 11.4
2005 m 4.44 2.2 mm G
O-lcm 104 23 22.6 24.0 187 33 27.0 13.6
O-1Ocm 179 28 22.2 22.3 272 39 27.8 14.7 4
2530m 4.48 3.8mm :
0-lcm 65 16 16 27.7 214 33 24.3 20.6 F
O-10 cm 249 22 13.8 39.4 449 40 22.4 21.8 c
3010m 4.88 2.5mm
O-l cm 76 17 17 42.1 139 27 25.6 23.0 E
0-IOcm 166 22 19.7 11.2 359 40 26.0 20.9 5
“SE Atlantic:
O-lcm; >125pn
250-700m 2-3 287 572 41 43
400-1000m >3 225 602 35 40
1000-2000m >3 235 426 52 53
2000-3000m >3 230 343 50 64
30004000m >3 264 334 53 53
400&5000m >3 300 543 52 59
>5000m >3 330 360 44 38
“Arctic Ocean:
17-38gCm-'y-' flux:
37-lOOm(?n=4) 27-33mm 594 676 49 37
6-SgCm--y-' flux:
200-250m(n=4) 25mm 616 756 56 51
O-2gCme2yp' flux: K
500-1000m (n=9) 3%30mm 657 691 66 63
100&2000m (n=7) 4-35mm 464 640 57 51
200s3000m (n=S) 2-13mm 612 648 41 47
300&4000m (n=4) 3-10mm 503 604 39 40
(Levin et al., 2001). Foraminifera exhibit similar trends (Table 4). The
numbers of “live” (rose Bengal stained) calcareous and total hard-shelled
species were very low (maximum 8 and 12, respectively) in the Santa
Barbara Basin (02 <0.5mll-‘) (Bernhard et al., 1997). More species (28
calcareous, 41 total hard shelled) occurred in the core of the Oman margin
oxygen minimum zone (02 = 0.13 ml 1-l) but this reflects the large number
of specimens examined (4414 and 5505, respectively) (Gooday et al., 2000).
In general, 20 or more “live” calcareous species and 30 or more total live
species, are present at oxic sites, except where the number of specimens
examined is < 200. One exception is Rathburn and Corliss’ (1994) 4000 m
site in the Sulu Sea (02 = l.l9mll-‘) where only 14 calcareous species were
represented among 252 “live” specimens. Where comparative data are
available, the number of dead species is often higher than the number of
‘live’ species.
Relative levels of species diversity can be related to the occupancy of
sediment microhabitats, as summarised in the TROX model (Figure 1).
Diversity is believed to exhibit a parabolic relation to productivity and food
supply in deep-sea and other ecosystems (Levin et al., 2001). One would
therefore expect diversity to be lowest at the oligotrophic and eutrophic
extremes of the model, and highest at intermediate levels of food input, i.e.
in mesotrophic systems. This expectation is generally supported by field
evidence. In the Arctic Ocean, where food supply is severely limited by
permanent ice cover, foraminiferal diversity is positively correlated with the
organic carbon flux; i.e. higher productivity = higher diversity (Wollenburg
and Mackensen, 1998; Wollenburg and Kuhnt, 2000; Wollenburg et ul.,
2001). This presumably represents the ascending, left-hand side of the
parabolic curve shown in Figure 1. In “normal” oxic deep-sea settings,
foraminiferal assemblages are typically highly diverse (Gooday et al., 1998;
Gooday, 1999). As discussed in Section 7, where a high organic matter flux
is combined with low oxygen concentrations, species richness and diversity
decrease and dominance increases; i.e. higher productivity = lower diversity.
Such situations correspond to the right-hand, descending side of the curve.
Diversity parameters can be helpful in palaeoceanographic reconstruc-
tions (Thomas and Gooday, 1996; van der Zwaan et al., 1999) provided that
significant dissolution has not occurred. In the northern Arabian Sea,
changes in the quantity and quality of organic matter arriving at the sea floor,
and corresponding variations in the thickness and intensity of the OMZ, have
been inferred from shifts in foraminiferal diversity (den Dulk, 2000). These
changes may have been linked to fluctuations in the intensity of monsoon-
induced upwelling in response to orbital and sub-orbital (precessional)
forcing mechanisms. Den Dulk et al. (1998) studied Quaternary cores
spanning 120,000 years from the Pakistan margin in the Northern Arabian
Sea. Two foraminiferal assemblages were recognised, one characterised by
BENTHIC FORAMINIFERA 53
low diversity and high dominance, the other by high diversity and low
dominance. The low diversity assemblages recurred every 23,000 years and
possibly reflected enhanced summer surface productivity (and therefore
intensified OMZ development) linked to the precessional component of
orbital forcing. A more sustained period of low diversity occurred under
glacial conditions, perhaps related to a strengthening of the NE monsoon
which led to higher winter productivity and hence lower bottom-water
oxygen concentrations (den Dulk, 1998). In a detailed multiproxy study of
shorter cores (spanning the last 30,000 years) from the same margin, von Rad
et al. (1999) reported a switch from low to high foraminiferal diversity on the
Pakistan margin during brief, late Quaternary to early Holocene climatic
oscillations (Younger Dryas, Heinrich Events 1 and 2) when surface
productivity was believed to be unusually low. Jian et al. (1999) studied
fluctuations in benthic foraminiferal assemblages in the South China Sea over
the last 40,000 years. They observed that species diversity (Shannon-Wiener
index) decreased as surface productivity (inferred from the proportion of
infaunal species) increased. In the Ionian Sea (E. Mediterranean Sea), on the
other hand, Schmiedl et al. (1998) reported very low Shannon-Wiener values
associated with very sparse faunas during interglacial periods. Glacial
deposits were characterised by high abundance, high diversity faunas. The
low diversity faunas persist in the modern E. Mediterranean, and probably
reflect the extremely food-poor nature of this basin. These low-diversity
faunas correspond, respectively, to the descending (eutrophic) and ascending
(oligotrophic) sides of the parabolic curve (Figure 1).
Diversity parameters are particularly useful as indicators of changing
environmental conditions in absence of extant species. Examples include
the increasing eutrophication and dysoxia that apparently affected
Mediterranean basins at the end of the Palaeogene (Kouwenhoven et al.,
1997) and continental margin basins in the North and South Atlantic Oceans
in the mid- and Late Cretaceous (Koutsoukos et al., 1990; Holbourn et al.,
1999, 2001a, b). In contrast to modern dysoxic environments, however,
foraminiferal abundance in some Cretaceous deposits was unusually low,
probably because oxygen depletion was very intense (Holbourn et al., 2001a).
There is little evidence to link deep-sea foraminiferal diversity to current
activity. However, enhanced bottom flow strongly affects the structure and
composition of benthic faunas (Hall, 1994) and leads to changes in
metazoan macrofaunal diversity (Levin et a/., 1994; Gage et ul., 1995; Gage,
1996, 1997) and taxonomic composition (Levin and DiBacco, 1995).
Differences have been observed in the taxonomic/functional composition
of predominantly agglutinated foraminiferal assemblages between quiescent
areas and those subject to strong currents (Kaminski, 1985; Kaminski and
Schriider, 1987; Murray and Alve, 1994; Kuhnt and Collins, 1995). Hydro-
dynamically active regions are widespread in the deep sea (Hollister and
54 ANDREW J. GOODAY
related to hydrostatic pressure and this parameter must set upper and lower
limits to bathymetric ranges through its effect on cell biochemistry.
9) Assemblage parameters (abundance, diversity, species richness,
dominance) provide palaeoceanographic indicators that are independent
of the identity of species and morphotypes. High productivity/low oxygen
regimes are characterised by high abundance, high dominance and relatively
low species richness and diversity. More oligotrophic, well-oxygenated
settings are characterised by lower abundance, low dominance and high
levels of species richness and diversity. Highly oligotrophic environments
are characterised by low abundance and low species richness and diversity.
Site BAIOO m
80
70
80
s 50
40
30
20
10
0
90
80 Site C:1900 m
70
60
s 50
40
30
20
10
0
ioo-
go-
80-
70-
60-
s so-
40-
30-
20
10
0
with calcitic cement. In order to bridge the gap between dead and fossil
assemblages, Mackensen et al. (1990) introduced the concept of “potential
fossil assemblages”, derived from modern dead assemblages by subtracting
the non-resistant agglutinated species. Potential fossil assemblages do not
reflect the effects of dissolution and other destructive processes and
must therefore be considered as “ideal” fossil assemblages. Calcareous
tests may also disappear as a result of dissoluton by corrosive pore waters
(Mackensen and Douglas, 1989) leaving a predominately or entirely
agglutinated assemblage (Gradstein and Berggren, 1981; Murray and Alve,
1994).
Loubere (1989) and Loubere and Gary (1990) presented computer models
and evidence from box cores (Gulf of Mexico, 102&l 170 m water depth)
suggesting that microhabitat preferences can also influence the susceptibility
of a species to taphonomic destruction. They argue that much of the
destruction of foraminiferal tests occurs in the surface l-2cm of sediments
where disturbance by benthic animals is most intense (see also de Stigter,
1996). It therefore has a particular impact on epifaunal and shallow infaunal
species. Deep infaunal species live below this surface zone and therefore
largely escape these destructive processes. Loubere et al. (1993) developed
these arguments further, emphasising the ways in which the organic flux and
bottom-water oxygenation interact to influence the formation of fossil
assemblages. They suggested that assemblage generation depends on: (1) the
distribution of living foraminifera within the sediment profile, (2) the rates
of test production in different sediment layers, (3) the rates of destructive
taphonomic processes (the ‘taphonomic filter’) and the variation of these
rates within the sediment profile, (4) the style of bioturbation and the depth
to which it extends (among other things, this will influence the extent to
which deep-infaunal tests are exposed to taphonomic processes in surface
sediments). The final fossil assemblage reflects a combination of these four
processes (Figure 9), the importance of which change along gradients of
organic carbon flux, oxygen availability, and depth in the sediment profile.
On the basis of these ideas, Loubere et al. (1993) and Loubere (1997)
suggested that in well-oxygenated, low-flux settings (e.g. central oceanic
regions), most test production will occur close to the sediment surface and
be subject to intense and uniform taphonomic processes leading to a
substantial loss of tests in the fossil assemblage. In these environments, low
sedimentation rates will also promote the formation of a homogeneous dead
assemblage (de Stigter, 1996). Where a moderate organic flux is combined
with well-oxygenated bottom water (e.g. bathyal continental margins),
the fossil assemblage will form over a thicker sediment layer, species will
be distinctly stratified within this habitable zone, and deep infaunal
foraminifera will be subject to less taphonomic loss than those originating
in near-surface layers. In high flux, low-oxygen settings (e.g. oxygen
60
Standing
Stock
Production
Rate
..*
.,..*.
*...
._..
*.*
ANDREW
Sediment
Mixing Rate
J. GOODAY
7
3Jlumpy
Uiffd
of the ODA. ATAs from slope, rise and abyssal plain settings were
interpreted in terms of organic inputs, current flow, and water mass
properties (Murray and Alve, 1994). Nevetheless, some modification of the
signal must occur in deep-sea settings. Harloff and Mackensen (1997)
reported that live and dead assemblages in the Scotian Sea and Argentine
Basin correspond fairly closely. However, the potential fossil assemblages
(sensu Mackensen et al., 1990) defined on the basis of Principal Component
Analysis and consisting of species likely to fossilise, were generally more
extensive than the corresponding dead assemblages and therefore embraced
a wider range of environmental conditions.
The quantification of proxies, particularly for organic flux to the sea floor
and bottom-water oxygen concentrations, remains a central challenge for
palaeoceanographers. Reliable, globally applicable, quantitative, proxies for
these parameters based on benthic foraminifera may always remain elusive
because foraminiferal biology is so complex. Nevertheless, considerable
progress has been made recently in a number of areas. Studies based on
samples collected over a wide geographical area (e.g. Mackensen et al., 1995;
Wollenburg and Mackensen, 1998) reveal qualitative relationships between
fauna1 patterns and environmental parameters. Large data sets relating
species abundances to a single parameter along an environmental gradient
provide the basis for a more quantitative approach. For example, the per-
centage abundance of species in the NE Atlantic Ocean has been related to
the organic carbon flux to the sea floor (Altenbach et al., 1999) and multi-
variate approaches have been used to develop transfer functions linking
species assemblages to surface primary productivity (e.g. Loubere, 1994;
BENTHIC FORAMINIFERA 65
ORGANIC
FLUXES
I I I I
t t r, v t t
l Stress due 0 Fresh food l Direct effects
to CaCOs leads to faster on cell bio-
dissolution reproduction chemistry impose
leads to and iarge bathymetric
increased populations of limits
mortality some epifaunal
species
t
l Elimination of oxyphilic
species, abundant food,
Local effects
reduced macrofaunal
on benthic
competition/predation leads
foraminifera
to large populations of
infaunalspecies
Figure II Environmental gradients that act over regional spatial scales (outer
rectangle) and their effects on foraminiferal faunas at local scales (inner rectangle).
Some of the regional gradients (carbonate undersaturation, current flow, oxygena-
tion) are water mass attributes. Organic fluxes and bottom-water oxygen act to
modify local geochemical gradients within the sediment and these, in turn, influence
fauna1 characteristics (species, morphotype composition, relative abundance of
epifaunal/shallow infaunal vs. deep infaunal species) by accelerating or decelerating
rates of reproduction. Current flow, organic fluxes, hydrostatic pressure and
carbonate undersaturation have a more direct effect on fauna1 characteristics. Biotic
interactions involving metazoan meio- and macro-fauna (not shown) will also
influence foraminiferal faunas.
result, they usually look for relations between species and groups of
species (rather than assemblage parameters) and particular physical and
chemical variables.
Despite these contrasting approaches, there is considerable potential for
synergy between palaeoceanography and biology (Gooday, 1994; Nees and
Struck, 1999). Biologists and geologists share a common interest in many
basic issues in deep-sea ecology and have often addressed them in the
same geographical settings. Biology underpins the accurate reading of
palaeoenvironmental signals, both fauna1 and geochemical, carried by fossil
foraminifera. Palaeoceanographic studies, in turn, provide a record of
fauna1 responses to changes in the environment over time scales that are
much longer than those available to biologists (Cronin and Raymo, 1997;
Den Dulk et al., 1998). It has long been known that deep-sea foraminiferal
assemblages have responded over geological time to environmental
fluctuations and recent studies reveal just how sensitive they are to rapid
climatic oscillations (Cannariato et al., 1999). The long temporal perspective
( lo3 to lo6 or more years) provided by the palaeoceanographic record offers
unique insights into the historical and macroecological processes that have
helped to shape modern communities (Lawton, 1999). These are only now
beginning to be exploited by marine biologists, for example, in the
interpretation of large-scale patterns of genetic differentiation and species
diversity in the deep sea (Rex et al., 1997; Quattro et al., 2001; Stuart et al.,
2002). A broad perspective that combines biological and geological
approaches to the study of benthic foraminifera (e.g. Loubere and
Fariduddin, 1999b; Levin et al., 2001) may ultimately lead to a more
complete understanding of the biology of these remarkable and immensely
successful organisms.
ACKNOWLEDGEMENTS
REFERENCES
Akimoto, K., Hattori, M. and Oda, M. (2002). Late Cenozoic paleobathymetry and
paleogeography in the South Fossa-Magna and Enshunada regions, Japan, based
on planktic and benthic foraminifera. Marine Geology 187, 89-118.
Alley, R. B., Meese, D. A., Shuman, C. A., Gow, A. J., Taylor, K. C., Grootes, P.
M., White, J. W. C., Ram, M., Waddington, E. D., Mayewski, P. A. and Zielinski,
A. G. A. (1993). Abrupt increase in Greenland snow accumulation at the end of
the Younger Dryas event. Nature 362, 527-529.
Altenbach, A. V. (1988). Deep-sea foraminifera and flux rates of organic carbon.
Revue Palkobiologie, Vol. Spec. 2, 119-120.
Altenbach, A. V. and Sarnthein, M. (1989). Productivity record in benthic
foraminifera. In “Productivity of the Oceans: Past and Present” (W. H. Berger,
V. S. Smetacek and G. Wefer, eds.), pp. 255-269, Wiley-Interscience, Chichester.
Altenbach, A. V., Pflaumann, U., Schiebel, R., Thies, A., Timm, S. and Trauth, M.
(1999). Scaling percentages and distributional patterns of benthic foraminifera
with flux rates of organic carbon. Journal qf Foraminiferal Research 29.
173-185.
Alve, E. (1990). Variations in estuarine foraminiferal biofacies with diminishing
oxygen conditions in Drammensfjord, SE Norway. In “Paleoecology,
Biostratigraphy, Paleoceanography and Taxonomy of Agglutinated
Foraminifera.” (C. Hemleben, M. A. Kaminski, W. Kuhnt and D. B. Scott,
eds.), pp. 661694, Kluwer Academic Publishers, Dordrecht, Boston, London.
Alve, E. (1994). Opportunistic features of the foraminifer Stainforthia fusiformis
(Williamson): evidence from Frierfjord, Norway. Journal of Micropalaeontolog~
13, 24.
Alve, E. (1999). Colonization of new habitats by benthic foraminifera: a review.
Earth Sciences Reviews 46, 167-185.
Alve, E. and Bernhard, J. M. (1995). Vertical migratory response of benthic
foraminifera to controlled oxygen concentrations in an experimental mesocosm.
Marine Ecology Progress Series 116, 137-151.
Alve, E. and Murray, J. W. (1995). Experiments to determine the origin and
palaeoenvironmental significance of agglutinated foraminiferal assemblages. In
“Proceedings of the Fourth International Workshop on Agglutinated
Foraminifera, Krakow, Poland, September 12-19, 1993” (M. A. Kaminski, S.
Geroch, M. A. Gasinski, eds.), pp. 1-l 1, Grzybowski Foundation Special
Publication no. 3, Krakow, Poland.
Asper, V. L., Deuser, W. G., Knauer G. A. and Lohrenz, S. E. (1992). Rapid
coupling of sinking particles between surface and deep ocean waters. Nature 357,
670-672.
Baas, J. H., Schiinfeld, J. and Zahn, R. (1998). Mid-depth oxygen drawdown during
Heinrich events: evidence from benthic foraminiferal community structure, trace-
fossil tiering, and benthic S13C at the Portuguese Margin. Marine Geology 152,
25-55.
Barmawidjaja, D. M., Jorissen, F. J., Puskaric, S. and van der Zwaan, G. J. (1992).
Microhabitat selection by benthic foraminifera in the Northern Adriatic Sea.
Journal of Foraminiferal Research 22, 297-3 17.
Barnett, P. R. O., Watson, J. and Connelly, D. (1984). A multiple corer for taking
virtually undisturbed samples from shelf, bathyal and abyssal sediments.
Oceanologica Acta 7, 401408.
BENTHIC FORAMINIFERA 71
Cenozoic Glacial Ages” (Turekian, K., ed.), pp. 71-181. Yale University Press.
New Haven, Connecticut.
Jahnke, R. A. (1996). The global ocean flux of particulate organic carbon: Area1
distribution and magnitude. Global Biogeochemical Cycles 10, 71-88.
Jahnke, R. A. (2002). Global distribution and magnitude of deep particulate organic
carbon fluxes estimated by benthic flux measurements (abstract). In “International
Workshop on Global Ocean Productivity and the Fluxes of Carbon and Nutrients:
Combining Observations and Models” (R. Schlitzer, P. Monfray and N.
Hoepfner, eds), pp. 1617. JRC-ISPRA, Italy 24-27 June 2002.
Jannink, N. T. (2001). Seasonality, biodiversity and microhabitats in benthic
foraminiferal communities. Geologica Ultraiectina, no. 203, l-l 92.
Jannink, N. T., Zachariasse, W. J. and van der Zwaan, G. J. (1998). Living (Rose
Bengal stained) benthic Foraminifera from the Pakistan continental margin
(northern Arabian Sea). Deep-Sea Research Z 45, 1483-1513.
Jensen, P. (1982). A new meiofaunal splitter. Annales Zoologici Fennici 19,
233-236.
Jian, Z., Wang, L., Kienast, M., Sarnthein, M., Kuhnt, W., Lin, H. and Wang, P.
(1999). Benthic foraminiferal paleoceanography of the South China Sea over the
last 40,000. Marine GeologJl 156, 159-186.
Jian, Z., Huang, B., Kuhnt, W. and Lin, H.-L. (2001). Late Quaternary upwelling
intensity and East Asian monsoon forcing in the South China Sea. Quaternar?
Research, 55, 363-370.
Jorissen. F. (1999). Benthic foraminiferal microhabitats below the sediment-water
interface. In “Modern Foraminifera” (B. K. Sen Gupta, ed.), pp. 161-179. Kluwer
Academic Publishers, Dordrecht, Boston, London.
Jorissen. F. and Rohling, E. T. (eds.), (2000). Foraminiferal proxies of paleopro-
ductivity. Marine Micropalaeontology 40, 131-344.
Jorissen, F. J. and Wittling, I. (1999). Ecological evidence from live-dead
comparisons of benthic foraminiferal faunas off Cape Blanc (Northwest Africa).
Palaeogeography, Palaeoclimatology, Palaeoecology 149, 151-170.
Jorissen, F. J., de Stigter, H. C. and Widmark, J. G. V. (1995). A conceptual model
explaining benthic foraminiferal microhabitats. Marine Micropaleontology 26,
3-15.
Jorissen, F., Wittling, I., Peypouquet, J. P., Rabouille, C. and Relexans, J. C. (1998).
Live benthic foraminiferal faunas off Cap Blanc, NW Africa: community structure
and microhabitats. Deep-Sea Research Z 45, 2157-2188.
Josefson, A. B. and Widbom, B. (1988). Differential response of benthic macrofauna
and meiofauna to hypoxia in the Gullmar fjord basin. Marine Biology 100,
3 l-40.
Kaiho, K. (1991). Global changes of Paleogene aerobic/anaerobic benthic
foraminifera and deep-sea circulation. Palaeogeography, Palaeoclimatolog~~,
Palaeoecology 83, 65-85.
Kaiho, K. (1994). Benthic foraminiferal dissolved-oxygen levels in the modern ocean.
Geology 22, 7 19-722.
Kaiho, K. (1999). Effect of organic carbon flux and dissolved oxygen on the benthic
foraminiferal oxygen index (BFOI). Marine Micropaleontology 37, 67-76.
Kaminski, M. A. (1985). Evidence for control of abyssal agglutinated foraminiferal
community structure by substrate disturbance: results from the HEBBLE area.
Marine Geo1og.v 66, 113-131.
Kaminski, M. A. and Schriider, C. J. (1987). Environmental analysis of deep-
sea agglutinated foraminifera: can we distinguish tranquil from disturbed
80 ANDREW J. GOODAY
1. Introduction ..................................................................... 92
2. Distribution and Natural History ............................................... 93
3. Sex Ratio and Size at Sexual Maturity ......................................... 95
4. Neoteny ......................................................................... 96
5. Protandric Hermaphroditism in E. asiatica ..................................... 99
6. Mating Habits.. ................................................................. 104
7. Spermatophores and Sperm Transfer .......................................... 106
7.1. Morphology of spermatophores ........................................... 106
7.2. Histochemistry of spermatophoric components .......................... 107
7.3. Origin of spermatophores ................................................. 111
7.4. Spermatophore dehiscence ............................................... 111
7.5. Adaptive role of spermatophores in sperm transfer ...................... 111
8. Moulting Pattern of E. asiatica-A Case Study ................................. 112
8.1. Moult cycle stages ......................................................... 113
8.2. Size-related moulting frequency in E. asiatica ............................. 117
8.3. Endocrine regulation of moulting ......................................... 117
8.4. Nutritional control of moulting in Emerita ................................ 121
9. Reproductive Cycle ............................................................. 122
9.1. Method of estimating reproductive cycle ................................. 123
9.2. Reproductive cycle in E. asiatica ........................................... 124
9.3. Reproductive cycle of E. asiatica in relation to size ....................... 128
9.4. Egg production ............................................................ 129
9.5. Effect of temperature on egg development on the pleopods ............. 131
10. Interrelationship Between Moulting and Reproduction ....................... 135
10.1. Role of haemolymph lipoproteins in moulting and reproduction ....... 136
10.2. Endocrine regulation of moulting and reproduction ..................... 138
1. INTRODUCTION
Exposed sandy beaches look superficially barren, but can have an abundant
invertebrate infaunal community. The mole or sand crabs, including various
species of Emerita, are often dominant inhabitants. Being a suspension
feeder, Emerita is well represented in beaches characterised by large waves,
wide surf zones, fine sands and gentle slopes (Dugan et al., 1995). The crabs
play an important role in the economy of a sandy coast, contributing in a
major way to secondary benthic production. The distribution pattern of
each species is characteristic in that it is generally limited to long coastlines,
though occasionally extending to offshore islands. In North America,
E. analoga has a long distribution on the west coast; whereas E. talpoida
inhabits predominantly the east coast. Two New World species, E.
portoricensis and Hippa pac$ca, have island distributions (Efford, 1976).
In peninsular South India there are two species, Emerita asiatica (= E.
emeritus) and E. holthuisi, the former inhabiting the east coast and the latter,
the west coast. Figure 1 depicts the geographical distribution of the nine
species of the genus Emerita.
94 T. SUBRAMONIAM AND V. GUNAMALAI
Ovigerous Females
4 6 8 IO 12 20 22 24 26 28 30 32
Carapace length f mm )
4. NEOTENY
One recurrent feature of the life history of the sand crab genus Emerita is
that with few exceptions the males are always smaller than the females
(Table 1). In at least five species, the males are known to become sexually
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 97
mature soon after their arrival on the beach as megalopas. The smallest
mature males vary among species from 2.5 to 6 mm CL, whereas the females
are not usually mature until they exceed 12mmCL, except in E.
portoricensis which matures at 8mm CL. Although female maturity is
attained as juvenile adults, mature males retain several larval characters.
Subramoniam (1977b) described the secondary sexual characters, along with
other morphological characters of the neotenic males of E. asiatica. The
males lack the pleopods on the abdominal segments that are characteristic
of mature females. However, they possess short stumps of the natatory
pleopods found in the megalopa (Menon, 1933). Small,males of E. tafpoidu
also retain the stumps of the pleopods (Efford, 1967). They also show a
general simplicity of the appendages associated with their small size. For
example, the antennae are simple and do not have the regularly arranged,
closely packed setal net of the larger animals.
On the fifth thoracic leg of E. tulpoidu, situated at the inner side of the base
of the coxa, there is a triangular sac, called a sperm sac (Wharton, 1942)
T. SUBRAMONIAM AND V. GUNAMALAI
The problem of sex reversal in the sand crab Emerita is a story in itself. The
occurrence of neoteny in several species of Emerita and the fact that the
males die before reaching the size at which the females become sexually
mature, result in considerable deviation from the 1 : 1 sex ratio. By applying
a size-related sex ratio method, Barnes and Wenner (1968) found a sigmoid
curve, characteristic of protandry in other crustaceans such as the deep sea
prawns (Yaldwyn, 1966) and proposed for the first time a sex reversal
hypothesis for E. analoga. In support, Eickstaedt, (1969) and Knapp and
Wenner (as reported in Barnes and Wenner, 1968) postulated that some
males, kept under laboratory conditions, changed sex. However, a series of
laboratory culture experiments as well as natural environment observations
on E. analoga (Auyong, 1981; Wenner and Haley, 1981; Conan et al., 1975)
E. asiatica (Subramoniam, 1977b), E. talpoida (Diaz, 1981), E. portoricensis
(Sastre, 1991) and an Island species, Hippa paczjka (Haley, 1979) suggest
that the apparent anomaly in the sex ratio results only from the differential
growh of males and females. Wenner and Haley (1981) summarised the
arguments in favour of and against the sex reversal hypothesis for the hippid
mole crabs, basing them mainly on population and sex ratio studies and
laboratory experiments on differential growth and moult increments of
males and females in different size groups. In all the above studies, there was
no direct observation of the male gonads during the period when they might
change to females, and hence the possibility of sex-reversal in Emerita
species was at that time not resolved.
Unequivocal existence of functional protandric hermaphroditism was
demonstrated in E. asiatica by Subramoniam (1979~2, 1981). The main
reproductive events enumerated in the Figure 4 indicate that neotenous
males continue to grow after serving an active normal male life, deviate
from normal sexual behaviour, gradually lose their secondary and primary
sexual characters, and undergo sex reversal by acquiring female characters
around 19 mm CL. The disappearance of genital papillae around 15 mm CL,
is the first visible sign of sex reversal. Concurrently, spermatogonial
activity in testes ceases but hyperactivity of the mesodermic cells ensues
100 T. SUBRAMONIAM AND V. GUNAMALAI
piizz- -2 0
4 Carapace
length(mm)
above and
22
Secondary Females
I
Inter sexuals with
22 functionals ovary and
j Vitello~~i;~ytes I’. 3 oviduct. secondary
19 females
I8
Critical stage for
sex reversal
15
Genital pappillae
l5 disappear. Teshcular
activity ceases.
A.G. disappears.
1;
8,5 Hetero-sexual raping
degenerating A.G.
a
Genital pappillae persist.
Potential males. Not
G found in wmal mating
E
Non males destined to 5 Mature males involved
become primary innormal mating. A.G.
females -2 3,5 gland-veryactive
FEMALE MALE
(Figure 5A). Such hyperactivity of the mesodermal cells in the testis during
periods of sexual inactivity has also been shown in the crayfish Pontastacus
leptodactylus leptodactylus (Amato and Payen, 1978). In the size range of
19-22mmCL, these males possess a gonad comprising inactive testicular
and active ovarian portions. The ovarian anlagen spread along the mid-
dorsal line of the paired testis. During sex reversal, following the formation
of separate ovarian anlagen, the ovarian structure contains typical follicle
cells that have either migrated from the testicular region or differentiated
from a prefollicular mesodermic tissue of the gonad (Figure 5B). These cells
attach themselves to the vitellogenic oocytes, probably mediating uptake of
yolk proteins from the haemolymph (Charniaux-Cotton, 1975b). The newly
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 101
formed median limb beyond the fused posterior extremity of the testis,
however, lacks testicular elements (Figure SC). It is composed of a central
germarium surrounded by previtellogenic and vitellogenic oocytes. The vas
deferens in these intersexuals is intact but its opening is occluded. They also
begin to possess three pairs of pleopods and a pair of functional oviducts
formed at the base of the coxae of the third walking legs. The external
morphology of the reproductive system in the intersexual animals as
compared with the testis and ovary of the normal crabs is depicted in Figure
6. These intersexuals with a functional ovary could be easily identified by the
possession of only a few eggs on the pleopods.
The malacostracan crustaceans are generally gonochoristic with geneti-
cally determined sex. The genes for male morphogenesis act in the presence
of an androgenic hormone and for females in its absence (Charniaux-
Cotton, 1960). Sex reversal via protandric hermaphroditism has also been
reported in other crustaceans (Ghiselin, 1969; Policansky, 1982). Inversion
of sexual phenotype is influenced by epigamous factors exerted during
growth (Gallien, 1959). In malacostracan crustaceans, the hermaphroditic
potentialities are governed by the androgenic gland hormone (Charniaux-
Cotton, 1965a). The sequential disappearance of primary and secondary
male characters during the changeover phase of E. asiutica reported above,
and the concomitant assumption of female characters strongly suggest
6. MATING HABITS
burrowing and reburrowing when the tides are in and out, as well as filter
feeding. Again, mating between equally sized partners would be a
cumbersome process in the intertidal region, as the larger males clinging
onto females could be brushed off by waves, or by sand when the female
burrows. Coupled with aggregation behaviour (Efford, 1965; Cubit, 1969)
the neotenous male of Emerita species has probably evolved as a means to
increase the chance of fertilisation in the unstable habitat.
The occurrence of large non-functional but sexually mature males in the
population of E. asiatica merits further comments on their sexual behaviour.
The size-increase of non-functional large males suggests that the specialisa-
tion towards neoteny in this species is still incomplete (Subramoniam,
1977b). A peculiar mating behaviour of these larger males has also been
reported by Subramoniam (1979b). A male of 8.5 mm CL was observed to
deposit a spermatophore ribbon on the ventral side of a freshly moulted
immature female. Deposition of spermatophores by larger males on
immature, helpless freshly moulted females indicates indiscriminate copula-
tion, amounting to raping, in E. asiatica. Incidentally, Kittredge et al. (197 1)
provided some experimental evidence for sex pheromonal activity of the
moulting hormone, crustecdysone, in a number of decapods. Whether such
pheromonal attraction to the moulting females could trigger spermato-
phoral deposition by males on the freshly moulted female E. asiatica
is conjectural and needs experimental support,
SP
(continued)
Table 2 Continued.
e
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 111
The moult cycle stages have been determined in E. asiatica using the criteria
of changes in the cuticular morphology, epidermal retraction and setagenic
events occurring in the pleopod. Furthermore, an aggregation of hemocytes
characteristic of moulting stages was also evaluated throughout the moult
cycle stages in E. asiatica, using microscopic observation on the pleopodal
lumen (Gunamalai and Subramoniam, 2002). Four major stages, namely
postmoult, intermoult, premoult and ecdysis have been distinguished. The
defining features of different moult cycle stages are given in Table 3.
Postmoult stage refers to the crab immediately after ecdysis. During this
period, the soft and pliable new cuticle undergoes hardening. The moulted
animal is inactive during this phase, which lasts for 30min; thereafter, it
regains activity and burrows in the sand. The pleopods are soft and
transparent. The setae are thin-walled, and their lumen is wide and
prominent with a granular matrix filling up the space (Figure 9). This stage
is further divided into A,, A2 and B.
Post moult
Al 5-6h Freshly moulted crabs; Pleopods soft and transparent;
cuticle soft and pliable; setal shaft thin walled; setal
crab not active; after lumen wide and filled with
15-30 min becomes granular matrix; setal base
active and burrows into evenly arranged on pleopods
sand, if moulting is
outside the burrow
A2 24h Exoskeleton pliable and No change in pleopods
soft but begins to
harden
B 4d Carapace continues to Pleopods hard and rigid
harden
Intermoult
Cl Sd Exoskeleton remains Setal lumen becomes narrow;
hard; lateral side of the setal wall thickened; setal
carapace depressed by cone visible
finger pressure
G 45d Exoskeleton evenly hard Setal cone prominent, a
throughout body surface tube-like structure observed
under setal articulation;
epidermis condensed with
setal articulation (node)
G 334d Carapace attains rigidity No changes in pleopods
on dorso-lateral sides
Premoult
Do 34d No changes in Appearance of setal groove
exoskeleton at base of pleopod; no
epidermal retraction
Do, Same as stage De Apolysis starts; narrow gap
between old cuticle and
epidermis evident; setal
groove extends up to tip of
pleopods
DI 2-3 d Exoskeleton becomes Retracted zone between old
brittle cuticle and epidermis widens;
tip of new setae still within
setal groove; new cuticle
appears wavy
Dl, No further changes in New setae protrude into
exoskeleton. retracted zone.
D,,, Same as above. New cuticle clearly seen as
a layer
(continued)
BREEDING BIOLOGY OF THE SAND CRAB, EMEf?/TA 115
Tahk 3 Continued.
This stage represents the emergence of the crab through the ecdysial sutures
of the old cuticle. As a result of endocuticular resorption, the old cuticle is
thin and friable. The first ecdysial suture appears in the intersegmental
membrane connecting the cephalothorax and the abdomen. When flexed
116 T. SUBRAMONIAM AND V. GUNAMALAI
D
Figure IO Moulting sequence in Emerita asiatica. (A) first phase of ecdysis
in which the ecdysial suture is visible; (B and C) exposed part of its
cephalothorax and abdominal region seen through the ecdysial suture (dorsal
view); (D) fully moulted animal with soft exoskeleton. From Gunamalai and
Subramoniam (2002).
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 119
L
50-
;
t
$
a
30-
- IO - 17 mm
- 18-22mm
- 23-33mm
IO
t
OL I t I I I I 1 I I 1
J F M & M J J a s 0 N D
400 - . Immature
I Maturing first time
2 300- A Repetitive Spawner
\
P
g 200-
N
0
A8 C Do 01 02 03-4
Moulting stoges
c3 DO Dl 0; 0;’ D2 D~A E
Despite the fact that moulting is under hormonal control, environment may
also play a significant role in determining the seasonality of moulting
frequency at population level. For marine invertebrates in general,
temperature, photoperiod, salinity and availability of food are known to
exert influence on the vital physiological processes relating to growth and
reproduction (Giese and Pearse, 1974). In Emerita species, the evidence
indicates that abundance of food materials and the accumulation of
nutrients have an influence on the seasonality and intensity of moulting
(Siegel, 1984). As a filter feeder, Emerita might thus depend on the seasonal
abundance of plankton to control the moulting process. While studying the
122 T. SUBRAMONIAM AND V. GUNAMALAI
9. REPRODUCTIVE CYCLE
Like many other decapod crustaceans, Emerita carries the eggs on the
pleopods of the abdominal segments where they are hatched and released as
zoea larvae. The breeding season of these crabs can therefore be determined
by plotting the percentage of ovigerous females against time (Boolootian
et al., 1959; Knudsen, 1960). Although several workers on Emerita have
used incidence of ovigerous forms to determine the reproductive cycle, this
method has inherent difficulties in the estimation of gonad changes inside
the animal. For example, in the species inhabiting the temperate seas, egg
124 T. SUBRAMONIAM AND V. GUNAMALAI
masses remain on the pleopods long after cessation of gonadal activities for
the particular reproductive season. Reproductive activities such as the
formation and maturation of gametes may start well ahead of the spawning
season, thereby obscuring the correct commencement point of the
reproductive cycle. Therefore, more accurate quantitative methods such as
gonad index and histological examination are needed to assess the cyclic
seasonal reproduction. The gonad index can be calculated in several ways,
but usually it is the ratio of the gonad wet weight to the wet weight of the
whole animal expressed as a percentage (Giese and Pearse, 1974). It rests
upon the assumption that the ratio of body parts varies little with changes in
size of the animal. While several studies have employed the gonad index
method to delineate the reproductive cycle of Emerita species, a study on
the egg development in the pleopod may also be taken into consideration,
especially when the crab breeds continuously throughout the year
(Boolootian et al., 1959). Egg mass index is calculated as a percentage of
the weight of the whole animal (Eickstaedt, 1969). As a corollary to egg
mass index, seasonality in the pleopodal egg development can also be
assessed by studying the mean developmental stages of eggs on the berried
females in various months of the year. This method obviously necessitates a
classification of the stages in egg development leading to the hatching of
zoea larvae (Subramoniam, 1979a).
Figure 13 Annual fluctuations in the gonad, egg mass and hepatic indices of
Emerita asiatica; range of carapace length given above bottom axis. From
Subramoniam (1979a).
40-
20-
I I I I I I I I I I I I I
JFMAMJJASOND
16 18 20 22 24 26 26 30 32
Carapace length I mm )
Emerita produces a large number of yolky eggs and attaches them to the
setae of the endopodite of the pleopods. The abdomen, with the egg-
carrying pleopods, is tightly flexed beneath the thorax. This gives protection
to the developing embryos on the pleopods while the crab is inside the
burrow. In Emerita, age of maturity, breeding frequency, and clutch size
have a bearing on fecundity. As pointed out by Wenner et al. (1974), the
maturation age of the female crab could vary in different populations of the
same species, owing to differences in food availability. Thus, the E. analoga
population from Santa Cruz Island attained sexual maturity at a lower
carapace length due to poor food availability and slower growth rate.
Conversely, with abundant food availability on the Santa Barbara coast, the
130 T. SUBRAMONIAM AND V. GUNAMALAI
growth rate was not only high but there was also an increase in size of the
females at sexual maturity. A season-dependent size at sexual maturity is
also suggested for E. analoga by Eickstaedt (1969).
Wenner et al. (1987) studied egg production in E. analoga in reference to
the size and the year class at three California locations. They found that the
overall pattern of egg number as a function of size was similar for the first
two year classes, but egg production by the few third year crabs was highly
variable at the San Clemente site. Interestingly, the slope of the regressions
of size and egg number for each year class showed significant variation. The
slope was quite steep for first year and less so for the second year, but in the
third year the slope decreased considerably. This may suggest that the egg
laying intensity is inversely proportional to the size of the crab. However,
the number of eggs per spawning by the individual crabs always increased
linearly with the size of the laying female. Several authors who worked on E.
analoga at different beaches in north and south America also provided data
on the number of eggs produced as a function of size, although the number
per brood varied with season and locality (Osorio et al., 1967; Efford, 1969;
Eickstaedt, 1969). The size-related fecundity has also been determined
for the tropical species E. asiatica (Figure 16) indicating again that the egg-
laying capacity increases in direct relation to its body mass (Subramoniam,
1977a). In another population of E. asiutica from the east coast of India,
0 I 1 I I I I I
22 24 26 28 30 32
Carapace length (mm1
Figure 16 Relationship between carapace length and the number of eggs carried
of Emerita asiatica. From Subramoniam
in the pleopods (1977a).
BREEDING BIOLOGY OF THE SAND CRAB, EMERKA 131
Using several sets of published and unpublished data on the sand crab
E. analoga, which is widely distributed along the west coast of the Americas,
Wenner et al. (1991) plotted the egg development time as a function of
temperature. They found that the egg development time varied from 40 days
at 25°C to 100 days at 12°C. Furthermore, the duration of embryonic
development on the pleopod may also have a direct relation to the frequency
of spawning in these crabs. Fusaro (1980) provided experimental evidence
that decreased egg development time resulting from increased seawater
temperature has a positive effect on the number of egg batches produced per
female. Understandably, egg production by populations of E. analoga living
in cooler waters may be depressed relative to those populations experiencing
warmer water conditions.
Eickstaedt (1969) calculated the monthly mean egg development of the
berried females in the natural population to estimate the variation in the
time of egg development, as influenced by environmental factors such as
Table 7 Fecundity profiles (number of eggs produced per female) of Emerita species in relation to body size and geographical
occurrence.
temperature and salinity. His results agreed well with those of Fusaro in that
a reduction in the mean egg development time coincides with high breeding
activity in the summer month of August. While the monthly mean egg
development time shows wide variation in temperate species such as
E. analoga, in the tropical species, E. asiatica, egg development time is
almost the same in all months of the year (Subramoniam, 1979a). The
proportion of various stages in the egg development of berried females in
different months of the year 1975-76 is given in Figure 17. It is clear from
the figure that almost all stages, except stage X (hatching stage) are
obtainable at any time in different individuals of the population, suggesting
that egg development leading to the release of zoea larvae may be a
p = 28
MEo=4.60
Table 8 Haemolymph protein levels during the moult cycle stages versus different
size classes (10-l 7 mm CL immature females; 18-22 mm CL, first maturing females;
23-33 mm CL, continuously reproducing females). Numbers within parentheses
indicate the total number of crabs analysed within each stage. From Gunamalai and
Subramoniam (2002).
MSP{
Lp III - - Lp III
-Lp II
Lpl-
?LPl
l-l&-
HCY
SP
1
P,
Mole Immature Mature
Female Female
Figure 18 Comparison of haemolymph lipoproteins from male, immature and
mature (Vitellogenic) females of Emerita asiatica. Msp = male specific protein;
Lpi = lipoprotein I; LpII = lipoprotein II; LPI11 = lipoprotein III;
Hey = hemocyanin; Sp = simple protein. From Gunamalai (2001).
activities occur and, from the onset of premoult stage the protein registers
yet another peak followed by a sharp decline in the late premoult stages
(Table 8).
;
BREEDING BIOLOGY OF THE SAND CRAB, EMfR/TA 141
the major yolk protein resistant to proteolytic cleavage (Berman and Lasky,
1985) during yolk degradation.
Glycolipids of the major yolk protein have been reported for the first
time in E. asiatica (Tirumalai and Subramoniam, 1992). Glycolipid formed
the minor lipid species and constituted 2% of the total lipid fraction of the Lv
II. Furthermore, Tirumalai and Subramoniam (2001) have demonstrated
the presence of both glucose (monoglycosylceramide) and galactose
(diglycosylceramide) containing glycolipids in the lipovitellin of E. asiatica.
The galactolipids of yolk/yolk precursor protein may be involved in
the recognition of its receptors on the oocyte membrane (van Berkel et (II.,
1985).
Amino acid composition of the major yolk protein Lv II is given in
Table 11. A characteristic feature of E. usiatica yolk protein is the high
content of acidic amino acids such as aspartic acid and glutamic acid, the
latter alone constituting 18.9 mole percent. The Lv II contains three amino
acids with potential glycosylation sites such as serine, threonine and
asparagine for the glycosylation of 0- and iv-linked oligosaccharides. Lv II,
however, contained less basic amino acids such as lysine and the sulphur
containing amino acid, methionine (Tirumalai, 1996).
High levels of lipids are a defining character of eggs of marine
invertebrates, constituting the main source of metabolic energy during egg
maturation and embryonic development (Holland, 1978). The percentage
distribution of different lipid species, including phospolipids, neutral lipids,
and glycolipids in the eggs and embryos of E. asiatica is presented in
142 T. SUBRAMONIAM AND V. GUNAMALAI
ND = Not done
Table 12. Phospholipids formed by far the greatest fraction of the total
lipids in both freshly laid eggs and Lv II, as has been reported for the ovary
of many crustaceans (Teshima and Kanazawa, 1983; Lautier and
Lagarrigue, 1988; Teshima et al., 1989). As many as seven phospholipid
species have been separated from the lipovitellin and eggs of E. nsiatica,
using thin layer chromatography (Tirumalai and Subramoniam, 1992). They
are: (1) lysophosphatidylecholine; (2) sphingomylin; (3) phosphatidyl-
choline; (4) phosphatidylenositol; (5) phosphatidylserine; (6) phosphatidy-
lethanolamine and (6) cardiolipin. However, phosphatidyl choline and
phosphatidyl serine were the predominant phospholipid species. These
phospholipid species, accumulated within the eggs, have an important role
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 143
Cl2 17.214
Cl4 5.164
Cl6 6.886
Cl8:O 3.442
Cl&l 3.452
cc+2 8.610
Cl%3 17.206
C 20.4 13.827
C 22 10.329
Unidentified fatty acids 13.916
Emerita asiatica yolk proteins also contain several metal ions such as
copper, iron, sodium, and calcium, also phosphorus (Table 14; Tirumalai
and Subramoniam, 1992). These ions constituted as much as 3.5% of the
purified major yolk protein. The calcium and copper are bound to lipid in
Lv II, whereas the iron, phosphorus and sodium are both lipid and protein
bound. The metalloprotein nature of Emerita lipovitellin assumes develop-
mental significance inasmuch as lipovitellins serve important functions
during embryogenesis of oviparous eggs. A characteristic feature of
vertebrate yolk protein is its high phosphate content, helping in skeletal
formation during embryogenesis (Wahli, 1988). Crustaceans lack an internal
skeleton, but secrete a calcareous exoskeleton as armour. Whereas in
BREEDING BIOLOGY OF THE SAND CRAB, EMER/?X 145
ND = Not detected
Emerita species fasten the eggs to the pleopodal hairs where the eggs
develop and hatch out as larvae. The biochemical composition
of E. asiatica eggs shows them to be a rich source of nutrition. The
yolk comprises a glycolipocarotenoprotein complex, free lipids and
glycogen granules. During maturation in the ovary the eggs also acquire
various other organic and inorganic components needed for embryogen-
esis and also early larval development. The embryos also absorb water
and salts from the environment during the course of their development.
A special feature of Emerita eggs is the large proportion of lipid in the
yolk, forming as much as 30% of the lipovitellin (Tirumalai and
Subramoniam, 1992), in addition to a significant quantity of free lipids.
Lipid accumulation is a strategy to decrease density and to reduce
energy cost of egg carriage in pelagic crustaceans, which are characterised
by abbreviated development coupled with an extended period of
incubation (Herring, 1973), but this has little bearing on benthic species
such as Emerita. Studies on yolk utilisation in Emerita are limited to only
two species, namely E. holthuisi (Vijayaraghavan et al., 1976) and E.
asiatica (Subramoniam, 1991).
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 147
Protein 3.219 2.224 2.433 2.121 2.484 2.357 1.692 1.272 0.889
Free carbohydrates 0.293 0.489 0.718 0.969 1.161 1.714 2.000 3.090 4.000 +
Protein-bound
Glycogen 0.012
0.178 0.012
0.173 0.119
0.03 1 0.139
0.051 0.066
0.036 0.121
0.093 0.100
0.089 0.114
0.089 0.117
0.115 i
Lipid sugars 4.609 3.837 4.795 5.606 5.806 2.286 2.000 2.318 2.722 g
ii
-c
Table 18 Fluctuation of enzymatic activity during embryonic development in the crab Emerita asiatica. 0
n
Embryonic “Esterase activity #Protease activity sGlycosidase activity (PM p-nitrophenol released/lOmg embryo) 2m
stage (nmol napthol/mg in enzyme units
protein per min) (1 ug leucine a-Giucosidase p-Glucosidase cr-Galactosidase /LGalactosidase g
equivalent/30 min) E
- z
I ND* 5.5 - PI
II ND ND - -
III 0.1198 8.69 - - - 3
IV 0.1983 ND 0.058 0.018 0.054 0.086 2
V 0.3086 12.6 0.079 0.069 0.096 0.153
VI 0.1585 ND 0.085 0.036 0.157 0.172
VII 0.115 10.1 0.116 0.03 1 0.287 0.197
VIII 0.0523 9.35 0.075 0.028 0.195 0.112
IX - 0.037 0.025 0.165 0.062
H-lnnnnnn~
-El
I
- E2
+
--E3
-E4
-E5
0 1 2 3 4 5 6 7 0 9
Stage
Figure 19 Zymogram of esterases from the ovary and different egg develop-
mental stages in Emeritu asiatica. (El-E5 represent different isozymes). Redrawn
from Subramoniam (1991).
converted the biochemical value given above (Table 16, 17) into energy
value by considering the energy equivalents for total carbohydrates as
17.3 kJ g-’ dry weight, protein as 23.5 kJ gg’ and lipid as 39.5 kJ g-’ (Brody,
1968). The energy equivalent of total carbohydrates was calculated by
pooling free carbohydrates, glycogen and protein-bound sugars. It is evident
from Table 19 that the energy derived from the proteins is continuously
utilised from stage I of egg development (75.6 J per 10 mg dry tissue) to the
last stage (20.9 J per 10 mg dry tissue).
On the other hand, the carbohydrate-based energy is continuously built-
up from 8.4 J per 10 mg (stage I) to 73.2 J per 10 mg (stage IX). There is also
an apparent increase in the lipid energy from stage II to stage V. There is
then considerable utilisation of lipid energy in stage VI and VII. These
stages correspond to the maximum yolk clearance coupled with faster
organogenesis (eye and appendages development). However, there is a
considerable retention of lipid energy in the last two stages of embryonic
development, which may facilitate utilisation of stored energy in the absence
of adequate food for the free-swimming zoea larva.
These data suggest that the mobilisation of energy sources especially in
the first phase of embryogenesis has changed the energy profile during egg
development in E. asiatica. In the freshly laid eggs, the major energy source
is lipid (68.4%) followed by protein (28.4%). Carbohydrate is a very poor
source of energy (3.3%) in the beginning of the embryonic development
(Table 20). However, prior to hatching, the energy profile changes
dramatically. Protein contributes only 10.2% and lipid 53.3% at the end
of embryonic development. Interestingly, the carbohydrate-based energy
source has substantially increased to 36.3%. Furthermore, from the above
BREEDING BIOLOGY OF THE SAND CRAB, EMEFUTA 155
ISOPODA
Ligia oceanica 24.93 30.0 Pandian (1972)
Probopyrus 32.90 5.0 Anderson (1977)
pandalicola
DECAPODA
Macrobrachium 29.39 18.0 Balasundaram (1980)
nobilli
M. lamarrei 26.48 4.0 Katre (1977)
M. idella 26.08 29.0 Vijayaraghavan and
Easterson (1974)
Crangon crangon 24.76 23.0 Pandian (1967)
Homarus americanus 27.78 35.0 Pandian (1970b)
H. gammarus 25.84 26.0 Pandian (1970a)
Pagurus bernhardus 25.34 21.0 Pandian and Schumann
(1967)
Caridina nilotica 62.67 60.0 Ponnuchamy et al.
(1979)
Emerita holthuisi 17.95 77.0 Vijayaraghavan et al.
(1976)
Emerita asiatica 26.61 24.2 Subramoniam
(1991)
p-carotene, with its concentration varying between 15.4 ug g-’ wet weight
and 16.1 ugg-’ wet weight in the early stages of embryonic development.
After maintaining almost the same level up to stage V, B-carotene started
declining gradually to reach a low level of 3.7 ug g-’ wet weight in the newly
hatched out larvae. Alpha carotene also showed a declining trend during
embryogenesis of E. asiatica. Obviously, these two parent carotenoids of
dietary origin undergo bioconversion into more oxidised forms such as
hydroxy and ketocarotenoids. The involvement of B-carotene in the
production of ketocarotenoids such as echinenone, canthoxanthin and
astaxanthin is also evidenced in other crustacean species (Herring, 1968;
Hsu et al., 1970). That oxidation of B-carotene takes place via isocrypto-
xanthin is revealed by the occurrence of this intermediate compound in all
stages analysed, with the level declining as development proceeds.
Kour and Subramoniam (1992) suggested a possible biosynthetic pathway
of carotenoids during embryogenesis in E. asiutica (Figure 20). It can be
seen from the figure that astaxanthin is the final product of ,&carotenoid
metabolism. Free astaxanthin is found in all stages of embryonic
development. However, esterified astaxanthin is found only in the last
bz
R
E
zi
E
0
6
2
Table 22 Carotenoid content in different egg developmental stages of Emerita asiatica &g/g wet weight). Data from Kour and o
Subramoniam (1992). 71
2m
Carotenoids Stage
?
I III V VII VIII IX X G
s
u-carotene 0.853*0.056 0.921 ho.189 1.490 f 0.026 0.833 i 0.013 0.960 f 0.012 0.03 1 f 0.002 - .m
B-carotene 15.560*0.122 16.072*0.141 15.445 *0.087 14.320&0.097 12.220~0.034 7.220&I-0.034 3.7OOiO.069
Lutein 2.080 f 0.067 - - - - - - B
- - 3
Echinenone 0.846 f 0.031 1.96OiO.036 3.540f0.036
Isozeaxanthin 4.373 f 0.068 1.5OOiO.019 3.540*0.039 3.38OiO.048 0.031 &O.OlO - - 3
Zeaxanthin 4.034 + 0.045 - 4.5 10 •t 0.058 4.093 f 0.248 5.971 * 0.372 ~
Canthaxanthin - - - 2.972hO.323 5.806*0.528 4.613+0.264 2.606+0.264
cY-doradexanthin - - - - - - 0.666 zk 0.117
Isocryptoxanthin 6.712*0.198 5.10010.197 3.91O~kO.153 3.630*0.161 2.540i0.236 2.136k0.142 2.104+0.173
Free astaxanthin 0.600*0.022 0.216kO.016 0.686iO.034 0.608&0.016 1.192+0.055 2.440%0.100 0.848&0.044
Esterilied astaxanthin - - - - - 4.280 f 0.018
158 T. SUBRAMONIAM AND V. GIJNAMALAI
the eye and appendages are well developed (Subramoniam et al., 1999). The
third major peak at prehatching stage (VIII) is correlated with the
deposition of the embryonic cuticle of the zoea larva of E. asiatica,
as suggested by Goudeau et al. (1990) for the control of secretory activities
related to the synthesis of embryonic envelopes in European lobsters.
Apart from acting as a morphogenetic hormone that controls several
developmental events, including the secretion of embryonic cuticle and
moulting, the accumulation of significant quantities of polar and
apolar conjugates and their possible catabolism to other products such as
20, 26-dihydroxyecdysone (McCarthy and Skinner, 1979) and ecdysonic
acids (Lachaise and Lafont, 1984) would suggest their elimination through
storage excretion.
- Estrcdiol 17~3
600-
Embryonic stages
Like many other littoral benthic invertebrates, Emerita has a pelagic larval
phase. Some Emerita species can have as many as seven zoeal stages that are
spent in the open oceanic waters before metamorphosis to the megalopa
stage, which then migrates back to the sandy seashore for settlement.
The description of Emerita larvae dates back to 1877 when Smith
described three zoeal stages namely second, third and last zoea and a
megalopa collected from the plankton for a species described under the
generic name Hippa (= Emerita talpoida). Subsequently, Faxon (1879)
described the first zoeal stage hatched from the eggs in the laboratory. Much
later, Menon (1933), Johnson and Lewis (1942) and Sankolli (1967)
described larval development in three other species, E. asiatica, E. analoga
and E. holthuisi respectively. Menon described five zoeal stages from
the plankton. Similarly, Johnson and Lewis also described five zoeal
stages from the plankton, and the first stage from the laboratory-hatched
larvae.
Figure 22 Zoeal stages of Emerita talpoida: (A) First zoea; (B) Second zoea; (C)
Third zoea; (D) Fourth zoea. Redrawn from Rees (1959).
first achieved by Rees (1959) with E. talpoida. This study described up to six
zoeal stages before the megalopa stage. The number of zoeal stages could
also extend to a seventh stage in certain individuals in laboratory culture.
The details of different zoeal stages as well as the megalopa larva are given
in Figures 22-24. In general, there is a uniformity of morphological
structures in the first zoea of E. talpoida as compared with other Emerita
species such as E. analoga and E. asiatica.
The stage I zoea is characterised by a smoothly rounded carapace that is
translucent and colourless. The shape of the carapace changes somewhat in
stage III into a more or less pear-shaped structure. The lateral spines which
are not present in stage I zoea are characteristic of the subsequent stages.
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 163
Figure 23 Zoeal stages of Emeritu ralpoida: (E) Fifth zoea; (F) Sixth zoea.
Redrawn from Rees (1959).
The rostrum is short and broad in stage I zoea and continues to elongate
and reaches about one and a half times the length of the carapace. The
eyestalks are short and thick and lie close against. the carapace, directed
somewhat posteriorly. In the subsequent zoeal stages, the eyestalks increase
in length and the eyes are carried somewhat farther forward than in the first
stage. In the megalopa stage, which resembles the adult, the eyes are still
relatively large as compared to the adult.
The antennules in stage I zoea are short unjointed appendages which are
thick at the base and taper to a blunt point where three setae of about equal
164 T. SUBRAMONIAM AND V. GUNAMALAI
length are borne. These setae increase to four in number in stage IV, six in
stage V and eleven in stage VI.
The antennae in stage I zoea are rather stubby appendages, produced on
the outer side into a spine-like process. From the base of the outer spine,
there arises a somewhat slender dentiform process of about the same length.
At the base of this inner process, there is a much smaller spine. The form of
the antennae remains relatively unchanged through the first four zoeal
stages, the first indication of a flagellum not appearing until the fifth
zoeal stage. At this stage, the rudiment of the flagellum is visible as
a conspicuous knob, which lengthens enormously in the VI stage zoea. In
the megalopa stage, the antenna possesses all the important features of the
adult form.
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA
165
The mandible in the stage I zoea consists of an armed crown on its ventral
edge followed by sharp triangular teeth. These appendages change very
little, except for a general increase in size, throughout the zoeal stages. In the
megalopa stage, the mandible has undergone a complete change in structure
and function. It is no longer an organ of mastication but is adapted, as in
the adult, for the purpose of scraping the antennae and passing food to
the mouth.
In the case of the maxillae and the maxillipeds, the structures remain more
or less unchanged in the zoeal stages except for the increase in the number of
setae in the second maxilla and the maxillipeds. In the megalopa stage, these
structures possess all the parts of the adult appendage.
The abdomen in the stage I zoea is composed of five segments projecting
almost straight downward from the carapace, and is flexed so that the telson
is carried beneath and parallel to the carapace. At this stage, no rudiments
of the abdominal appendages are visible. The sixth segment is consolidated
with the telson; this becomes apparent when the uropods appear in the III
stage zoea. The uropod consists of a short basal segment with a long,
flattened lobe extending from it. In stage IV zoea, the four free segments
of the abdomen bears two small round thickenings on its inner side, the
evidence of future pleopods, which eventually appear in the stage VI zoea.
The pleopods are uniramous, unsegmented and appear on the second
through fifth abdominal segments. The abdomen in the megalopa stage is
composed of six segments, which are similar in form and proportion to
those of the adult. In contrast to the uniramous pleopods of the zoeal stages,
the megalopa stage bears four pairs of biramous pleopods.
In the temperate species, such as E. analoga, eggs are laid in the summer
months and after incubation on the pleopods for about a month, give rise to
zoeae, which are released into the plankton. Johnson (1940) estimated the
time spent as zoea in the plankton to be about four and a half months, after
analysing planktonic materials collected from tows off the coast of
California. Following this, Johnson and Lewis (1942) described the zoeal
stages of E. analoga from the plankton and sugge.sted that they passed
through at least five stages before moulting to the megalopa. The duration
of larval development is also variable; for E. talpoidu the laboratory rearing
took 30 days (Rees, 1959) whereas, E. rathbunae took about 90 days
(Knight, 1967). From laboratory rearing, Efford (1970) observed that the
zoea larvae of E. analoga passed through as many as 9 moults in a total time
duration of 130 days, before metamorphosing into megalopa. He also
166 T. SUBRAMONIAM AND V. GUNAMALAI
observed that moulting into megalopa could occur at 8th or 9th zoeal moult.
Although Hanson (1969) contended that the larval development in the
Hippidae was temperature dependent, such variations in the larval duration,
especially from the estimates of laboratory rearing and plankton analysis
could not be explained in terms of temperature difference. The duration
of the later stages is also so variable that there is a possibility that one or
two stages from metamorphosis to megalopa are skipped if conditions are
ideal for settlement on the beach. Similarly, the larvae may have the
ability to delay metamorphosis if settlement conditions are unfavourable.
This feature increases the chances for selection of a suitable substratum,
thus contributing to the success of the population (Thorson, 1950; Wilson,
1952).
Along the ocean coasts where the shelf is rather narrow and the deep sea
is not far off, strong currents may carry the larvae away from their littoral
and shallow water habitat. Johnson (1939), correlated water movements and
the dispersal of pelagic zoea larvae of E. anafoga along the southern
Californian coast. He observed that the fourth zoeal stage of this sand crab
is taken in plankton hauls at a distance of 125-130 miles from the mainland
shores. However, the first zoeal stage, with a 4 week larval life in laboratory
rearing, was found in plankton taken less than 20miles from the shore.
Evidently, such a long journey offshore for development and metamor-
phosis into megalopa would cause a large wastage of larvae as suggested for
oyster larvae by Korringa (1947).
After spending a varying period of time in the plankton, zoea larvae
of E. analoga metamorphose into megalopae and start arriving in large
numbers in early April with a peak influx in early June at the Scripps beach,
La Jolla (Efford, 1965). However, Wenner (personal communication quoted
by Efford, 1970) observed the arrival of megalopae in the winter months of
1965-1966 on the beaches at Goleta, just south of Point Conception. Such
differences in the recruitment period on different beaches along the west
coast of North America may suggest that the timing of maximum
recruitment perhaps depends largely on the distribution of the later zoeal
stages in relation to local hydrographic conditions (Johnson, 1940; Efford,
1965; Barnes and Wenner, 1968; Cox and Dudley, 1968).
Seasonality of the megalopa settlement in temperate species can be
related to the seasonal reproductive cycle. With tropical species, such as
the E. asiatica, that breed all through the year, we would expect to have a
continuous or near continuous settlement pattern of megalopae. Year-
round egg laying, coupled with continuous embryonic development of
pleopodal eggs results in uninterrupted release of zoea larvae into the
plankton. Hence, larval availability for metamorphosis to the megalopa
stage and settlement occurs throughout the year. Yet, even on tropical
beaches, seasonality in the megalopa settlement has been reported,
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 167
in the beach could not only affect filter feeding, but also the channelling of
metabolic energy to egg production, in addition to causing direct damage to
the exposed egg masses.
However, thermal effluents from an atomic power station at Kalpakkam,
located south of Madras on the east coast of India, directly affect the
distribution of E. asiatica in the vicinity. Figure 2.5 shows the distribution
pattern of Emerita in the MAPS (Madras Atomic Power Station)
region. Significantly, in the impact zone having an elevated seawater
temperature of 35”C, the crab is completely absent. However, as we move
away from the impact zone, with normalisation of seawater temperature,
the Emerita population gradually increases. No difference in reproductive
activity was found between these populations and a population in the control
region (Station 1). Another significant observation is that there was no
megalopa settlement in the impact zone. This may suggest that both young
and adult Emerita are sensitive to elevated temperatures caused by thermal
effluents and they move to safer areas on either side of the impact zone.
1 2 3 4 5 6 7 8 9
Stations
15. CONCLUSIONS
ACKNOWLEDGEMENTS
REFERENCES
Pandian, T. J. (1970b). Yolk utilisation and hatching time in the Canadian lobster
Homarus americanus. Marine Biology 7, 249-254.
Pandian, T. J. (1972). Egg incubation and yolk utilisation in the
isopod Ligia oceanica. Proceedings of the Indian National Science Academy
38,430441.
Pandian, T. J. (1994). Arthropoda-Crustacea. In “Reproductive Biology of
Invertebrates” (K. G. Adiyodi and R. G. Adiyodi, eds.), Vol. VI, Part B.
pp. 39-166.
Pandian, T. J. and Schumann, K. H. (1967). Chemical composition and caloric
content of egg and zoea of the hermit crab, Eupagurus bernhardus. Helgoltinder
Wissenschaftliche Meeresuntersuchungen 16, 225-230.
Panikkar, N. K. and Jeyaraman, R. (1966). Biological and oceanographic differences
between the Arabian Sea and Bay of Bengal as observed from the Indian region.
Proceedings of the Indian Academy qf Sciences 64B, 231-240.
Parvathy, K. (1970). Blood sugars in relation to chitin synthesis during cuticle
formation in Emerita asiatica. Marine Biology 5, 108-l 12.
Perry, D. M. (1980). Factors influencing aggregation patterns in the sand crab
Emerita analoga (Crustacea: Hippidae). Oecologia 45, 379-384.
Pillai, C. K. and Subramoniam, T. (1985). Yolk utilization as an adaptive strategy of
terrestrialization in the freshwater crab, Paratelphusa hydrodromus (Herbst).
Physiological Zoology 58, 445-457.
Pochon-Masson, J. (1983). Arthropoda-Crustacea. In “Reproductive Biology of
Invertebrates, Vol. II. Spermatogenesis and Sperm Function” (K. G. Adiyodi and
R. G. Adiyodi, eds.), pp. 407449. John Wiley, Chichester.
Policansky, D. (1982). Sex change in plants and animals. Annual Review of Ecology
and Systematics 13, 471495.
Ponnuchamy, R., Ayyappan, A., Reddy, S. R. and Katre, S. (1979). Yolk and copper
utilisation during embryogenesis of the freshwater prawn, Caridina nilotica.
Proceedings of the Indian Academy of Science 88, 353-362.
Pravalli, A. S. (1990). Protease activity in the developing eggs of Emerita asiatica
(Milne Edwards). M. Phil. Dissertation, University of Madras, India.
Quesnel, V. C. (1975). Breeding season and breeding size of female Emerita
portoricensis Schmitt (Crustacea: Anomura) in Trinidad, West Indies. Journal of
the Trinidad and Tobago Field Naturalists Club 1, 54-59.
Quinitio, E. T., Yamauchi, K., Harra, A. and Fugi, A. (1991). Profiles of
progesterone and estradiol-like substances in the haemolymph of female
Pandalus kessleri during an annual reproductive cycle. General and Comparative
Endocrinology 81, 343-348.
Radha, T. and Subramoniam, T. (1985). Origin and nature of spermatophoric mass
of the spiny lobster, Panulirus homarus. Marine Biology 86, 13-19.
Ramachandran, N. (1992). Phospholipase activity in the developing eggs of Emerita
asiatica (Milne Edwards). M. Phil. Dissertation, University of Madras, India.
57 PP.
Rao, R. K., Fingerman, S. W. and Fingerman, M. (1973). Effects of exogenous
ecdysones on the moult cycles of fourth and fifth stage American
lobsters, Homarus americanus. Comparative Biochemistry and Physiology 44A.
1105-l 120.
Rees, G. H. (1959). Larval development of the sand crab Emerita talpoidu (Say) in
the laboratory. Biological Bulletin 117, 356-370.
Sankolli, N. K. (1965). On a new species of Emerita (Decapoda: Anomura) from
India, with a note on Emerita emeritus (L). Crustaceana 8, 48-54.
BREEDING BIOLOGY OF THE SAND CRAB, EMERlTA 179
1. INTRODUCTION
Table 1 Estimates of ranges and median dates when coral bleaching events may
occur annually based on threshold temperatures proposed to induce coral bleaching
locally and on projections of increasing ocean temperatures by four global climate
models (derived from Hoegh-Guldberg, 1999).
Locality Threshold Projected dates for 10 bleaching
temperature (“C) events/decade
Range Median
Jamaica 29.2 2010-2030 2020
Phuket 30.2 200&2040 2020
Tahiti 28.3 2035-2045 2040
Raratonga 28.3 2020-2040 2030
Southern GBR 28.3 2020-2060 2040
Central GBR 29.2 2025-2050 2037
Northern GBR 30.0 202&2040 2030
Averaging the media values for the ranges derived from seven regions
suggests that this could occur worldwide by about 2030. Various scenarios
have been proposed to describe reef conditions resulting from continuing
and repetitive bleaching events (Done, 1999). However, we should recognize
that reef corals have been a subject of research for only a little over a
hundred years, and that the last 30 years have produced the vast majority of
observations and measurements on reef corals and their association with
symbiotic zooxanthellae. Little is known regarding the capacity of corals or
zooxanthellae to adapt or acclimatize to elevated temperatures, or the rates
at which any such adjustment to stressful temperatures may occur. The
purpose of this review is to summarize the information that is available on
coral bleaching, focus on processes that may act as adaptive mechanisms
and suggest needed research in this area.
worldwide, from 27°C in Rapa Nui (Easter Island) during 2000 where
summer ambient maximum is normally about 25°C (Wellington et al., 2001)
to 35-36°C during 1998 in the Arabian Gulf (George and John, 1998;
Wilkinson et al., 1998; Riegl, 1999, 2002), where normal ambient summer
open water maxium usually ranges up to 34°C (Coles, 1988). Clearly,
maximum water temperatures normally occurring in particular geographic
areas have principally determined the upper temperature tolerances
of corals, indicating that the corals are adjusted to ambient conditions
(Figure 1). This implies a capacity for reef corals and/or their algal
symbionts to adapt to higher temperatures over as yet unknown periods of
time. What is not clear is whether adjustment can occur through phenotypic
acclimatization to acute stress conditions and the mechanisms involved,
or require longer-term adaptation involving selection and breeding of
eurythermal genotypes.
Tropical Pocillopora
y=, pmi + 22.278
f=o.90
Hawaiian Pocillopora
Y=104~6m+21 636
l=o.99
photochemical systems of the light reaction, with the release of oxygen, and
fixation of organic carbon in the dark reaction. At higher light intensities the
rates of processes can become saturated, with photosaturation occurring as
early as 09:OOh in shallow water corals (Brown, 1997b). With elevated
temperatures, the rates of these processes increase to a level where more
protons are produced in the light reaction than can be utilized to form
organic carbon in the dark reaction. In the first studies of bleaching-related
190 S. L. COLES AND BARBARA E. BROWN
process the excitation energy is dissipated as heat via the xanthophyll cycle
in a process termed nonphotochemical quenching (NPQ), which is well
documented in both higher plants and algae (Demmig-Adams and Adams,
1993; Olaizola and Yamamoto, 1994; Olaizola et al., 1994; Owens, 1994;
Wilkinson, 2000). In coral symbiotic zooxanthellae, heat dissipation is
achieved by the reversible interconversion of the xanthophylls, diadino-
xanthin, and diatoxanthin. These xanthophylls were first identified in coral
zooxanthellae by Jeffrey and Haxo (1968), and an active xanthophyll cycle
in corals was described by Ambarsari et al. (1997) and Brown et al. (1999b).
Figure 2 shows a pronounced cycling of photoprotective xanthophylls in
response to diurnal irradiance changes which induce photoinhibition in the
shallow water coral Goniastrea aspera. When sea temperatures rise above
the normal ambient maxima, corals become more susceptible to the effects
of damaging solar radiation (Brown, 1997b; Hoegh-Guldberg, 1999); thus
the xanthophyll cycle becomes a key photoprotective defense. Indeed it
has been claimed that those corals more capable of dissipating excess
excitation energy through NPQ are less prone to temperature bleaching
(Warner er al., 1996).
Another possible protective mechanism against stressful light levels may
be fluorescent coral pigments, which have been indicated to reduce coral
bleaching by reflecting and/or fluorescing absorbed light (Salih et al., 1998b,
2000; Dove et al., 2001). A total of 124 species of corals were found to have
morphs containing fluorescent pigments on the Great Barrier Reef, often
growing alongside of morphs of the same species without such pigments
(Salih et al., 2000). Corals containing such fluorescent capacity were found
to bleach significantly less than nonfluorescent colonies of the same species
growing in the same area. Nonfluorescent corals were significantly more
photoinhibited during peak irradiance periods, and bleaching resistance
measured as tissue dinoflagellate biomass correlated significantly with
fluorescent pigment concentrations in coral tissue. The protective capacity
of these pigments may have important implications for long-term survival of
corals exposed to thermal stress. At Phuket Thailand, Brown et al. (2002~)
found abundant fluorescent pigment in cores from bleaching-resistant west-
facing surfaces of G. aspera compared with low concentrations in bleaching-
prone east-facing surfaces. Fluorescent pigments were most abundant in the
endoderm surrounding the symbiotic algae, suggesting a photoprotective
function. Such potential protective capacity has far-ranging implications for
long-term survival of corals when additionally stressed by high temperature,
although recent preliminary work by Dove (pers. comm. to BEB) suggests
that some of these pigments are easily denatured by elevated sea
temperature.
An internal defense mechanism that may substantially influence coral
tolerance to bleaching and mortality is change in heat shock proteins (Hsps)
S. L. COLES AND BARBARA E. BROWN
I I 1
0.5 -
0.4 T, ; T
0.3
0.2
T : -I- :
0.1
0.0 ’
I I I I
20th 21st 22nd 23rd 24th
January 1998
The capacity of corals and reefs to adapt to elevated temperatures has been
the subject of a number of reviews (Gates, 1990; Buddemeier and Fautin,
1993; Glynn, 1993; Brown, 1997a,b; Buddemeir and Smith, 1999; Done,
1999), and the mechanisms of phenotypic adaptation for corals have been
discussed in Brown (1997a) and Gates and Edmunds (1999). As reviewed in
Brown (1997a), adaptations by corals to elevated temperature or light
regimes can occur under a range of time scales and conditions. Terms
referring to these adjustments have been variously used, and we herein
follow the terminology of Brown (1997a) and Gates and Edmunds (1999).
Although acclimation has been used ambiguously to refer to adaptation
over the long term, e.g., Ware (1997), acclimation more properly means
changes in tolerances under laboratory or other experimental conditions,
generally over the short term. Acclimatization refers to phenotypic
changes by an organism to stresses in the natural environment that result
in the readjustment of the organism’s tolerance levels. These phenotypic
responses are usually reversible and are limited by the organism’s genotype,
which determines the boundaries beyond which acclimatization
cannot occur. Finally, selective adaptation occurs when the more stenotopic
members of a population are eliminated by the environmental stress,
leaving the more tolerant organisms to reproduce and recruit to available
habitat.
The primary evidence of long-term selection for temperature tolerant
corals is based upon the linkage of thermal thresholds to maximum ambient
temperature environments previously described, and reports of corals
surviving temperatures well in excess of normally accepted limits. Gardiner
(1903) observed abundant corals in a tidepool in the Laccadives at water
temperatures up to 56°C and Kinsman (1964) noted massive Porizes at over
40°C near Abu Dhabi, Arabian Gulf. Motoda (1940), Orr and Moorhouse
(1933) and Vaughan (1914) reported corals surviving temperatures up to
38-39°C in Palau, Australia, and Florida, respectively. More recently
Tomascik et al. (1997) reported a variety of corals living at 34-37°C near a
thermal vent in Indonesia, with one species growing in the vent at 42°C. On
Ofu Island, American Samoa, Craig et al. (2001) found 52 coral species,
including nine Acropora taxa, to survive daily temperatures as high as
34.5% for up to 3 h exposures daily for 35 days during the summer of
1998-99 with virtually no bleaching. Meesters and Bak (1993) found
recovery of experimentally damaged bleached Porites ‘asteroides in the
thermal effluent of a power station in Curacao to be just as high as that of
196 S. L. COLES AND BARBARA E. BROWN
normal colonies, and that the corals regained normal pigmentation at higher
temperatures, suggesting acclimatization to have occurred at temperatures
averaging 1.3”C above ambient conditions.
However, there have been few controlled experiments on reef corals that
have attempted to determine the capacity of reef corals for even short-term
acclimation to elevated temperatures. Experiments by Coles (1973) and
Coles and Jokiel(l978) described in Brown (1997a) indicated that Hawaiian
Montipora verrucosa acclimated for 56 days at l-2% above summer
maxima had higher survival for 5 days at stress temperatures of 30-32.5”C
than did ambient controls. Clausen and Roth (1975) showed shifts in coral
calcification rates of Hawaiian P. damicornis corresponding to incubation
temperature, suggesting a capacity for short-term acclimation. Glynn and
D’Croz (1990) found corals from an upwelling area in the Gulf of Panama
to undergo greater bleaching at 30°C in controlled experiments than the
same species from the nonupwelling Gulf of Chiriqui, where ambient
temperatures were higher and more stable. Al-Sofyani and Davies (1992)
found that respiration rates of Echinopora gemmacea in the Red Sea did not
change with a 6°C seasonal change in seawater temperature, suggesting
acclimatization for this species, while respiration rates of Stylophora
pistillata indicated no such acclimatization. Berkelmans and Willis (1999)
found that the winter bleaching threshold of P. damicornis on the Great
Barrier reef was 1°C lower than the summer threshold for this species, and
proposed that the winter temperature bleaching threshold of 31-32°C was a
reliable predictor of subsequent mortality observed when the stressed corals
were returned to the field and observed for 84 days. This possibility of
seasonal acclimatization, while intriguing, was not fully supported by these
experiments, since postexposure observations were not made on corals
during the summer trials, and lack of postexposure information on the fate
of controls during the winter trial make the results subject to question. Also,
these experiments did not find differences in thermal thresholds between
corals from the reef flat compared to the reef slope, or from different reefs
that had shown contrasting bleaching susceptibility. Such differences would
be expected from Berkelmans’ (2002) conclusion that cross-shelf and
latitudinal differences in coral bleaching thresholds correspond to tempera-
ture regimes on the Great Barrier Reef, suggesting thermal adaptation at
spatial scales of ca. 10-100 km.
Observations comparing bleaching under field conditions during the
1997-98 periods of anomalous high temperatures at Ko Phuket Thailand
(Dunne and Brown, 2001; Brown et al., 2002b) with previous episodes
in 1991 and 1995 have indicated a complex interaction of light with
temperature that may act to induce bleaching protection. Despite similar
temperature elevations and durations in 1997 and even higher temperatures
in 1998 than the two previous periods, bleaching was considerably less
CORAL BLEACHING 197
in 1997 and 1998 than during previous episodes. High temperatures in 1997
and 1998 were preceded by periods of higher than normal light intensity that
was indicated to stimulate photoprotective defenses in both coral host and
algae when the sea temperature was lower than stress levels, and this
tolerance then persisted through the periods of maximum temperature-light
stress (Brown et al., 2002b). Anomalous low tides in 1997 and 1998 also
accentuated the high light environment experienced by the corals in the area
(Dunne and Brown, 2001).
Complex interaction between sea temperature and light was also evident
at the colony level at this Thailand site, where the west sides of colonies of
G. usperu showed superior thermal tolerance to the east sides both in the
field during major bleaching events as well as in laboratory experiments
(Brown et al., 2OOOb; 2002a,c). In this example (Plate la) west sides
of colonies are exposed to high h-radiance in the dry season (November
to May) and, as a result, may show solar bleaching. However, when
anomalously high sea temperatures cause extensive bleaching on the reef in
May, such bleaching is mainly restricted to the east sides of G. aspera
colonies (Plate 1b). It appears that exposure of western surfaces of the coral
to a high irradiance environment in the field subsequently conferred
tolerance to high sea temperatures due to improved photoprotective
defences on the west sides without alteration of the zooxanthellae genotype
(Brown et al., 2002a).
Recent experiments revealed increases of 10 to 50 fold for molecular
biomarkers of stress and host stress proteins of G. usperu during elevated
temperature (33°C) exposures (Figure 3). Higher levels of oxidative stress
occurred on east sides than west sides, concomitant with higher concentra-
tions of defenses, such as Hsps and oxidative enzymes (Brown et al., 2002~).
Interestingly, in this experiment the differences lie in the host defenses rather
than those of the algae. This model is useful in showing that, in this shallow
water coral, limited acclimatization to high temperature does occur in the
field, that the timescale for acclimatization is relatively short (days-weeks-
months) and that photoprotection in the host can be an important defense
against elevated sea temperatures.
Observations comparing the responses of corals in the eastern Pacific to
elevated temperatures that occurred during the ENS0 events of 1983-84
(Glynn, 1983, 1984; Glynn and D’Croz, 1990) and 1997-98 (Glynn et al.,
2001; Jimenez et al., 2001) suggest that corals or coral assemblages may
become more thermally resistant or tolerant of bleaching with repeated
bleaching events. Elevations of sea surface temperatures (SSTs) and
durations of elevations in the Gulfs of Panama and Chiriqui, the
Galapagos Islands, and the coast of Ecuador were of similar magnitude
during the 1987-88 and 1982-83 ENS0 events (Glynn et al., 2001; Podesta
and Glynn, 2001). However, coral bleaching and mortality from 1997-98
198 S. L. COLES AND BARBARA E. BROWN
F@ure 3 Concentrations (pg ug-’ except Ubiquitin and Cu: Zn SOD in ng ug-‘)
of 12 molecular markers in soluble protein in Goniastrea uspera held at an elevated
temperature of 33°C for three days at Phuket, Thailand. Open bars represent west
sides of colonies, shaded bars east sides. Markers included three indicators of
oxidative stress: (Chydroxynoneal [HNE], alondialdahyde [MDA] and ubiquitin)
four coral host-specific biomarkers: (oxidative enzymes copper/zinc superoxide
dismutase [Cu : Zn SOD] and manganese superoxide dismutase [MnSOD] and heat
shock proteins Hsp60 and Hsp70, and five symbiotic algae host-specific biomarkers:
Cu : Zn SOD, MnSOD, Hsp60, Hsp70, and chloroplast small heat shock protein
(ChlsHsp). Bar represent means f one standard error. Significant differences:
* < 0.05, ** < 0.01, *** < 0.001 (from Figure 2 in Brown et al., 2002~).
200 S. L. COLES AND BARBARA E. BROWN
was substantially less than in 1982-83 (Glynn et al., 2001; Podesta and
Glynn, 2001). Coral mortality from the 1982-83 event was 97-99% in the
Galapagos Islands, 85% in the Gulf of Panama, and 75% in the Gulf of
Chiriqui. By contrast, mortality in 1987-88 was 26% in the Galapagos, 13%
in the Gulf of Chiriqui, and undetectable in Gulf of Panama (Glynn et al.,
2001). Although these comparisons are not unequivocal due to differences
in seasonal timing of anomalies, duration of exposures (Podesta and Glynn,
2001) or upwelling (Glynn et al., 2001), the lower bleaching and mortality
that occurred in 1997-98 suggest that selection for resistant species or
genotypes of corals and zooxanthellae may have occurred during prior
ENSO-related temperature events (Podesta and Glynn, 2001). Jimenez
et al. (2001) also report higher bleaching and mortality to corals on
Costa Rican reefs in 1982-83 than in 1997-98, despite the temperature stress
from the later event having been as strong or stronger than in 1982-83.
Unfortunately, no information is provided concerning the prevailing light
climates in this region for the two major El Nino events that would clarify
whether differences in solar radiation might have influenced the generally
lower bleaching that occurred in 1997-98.
Coral bleaching was also minimal in the Society Islands during the 1998
event, but the cause there was attributed to reduced light during the event.
Mumby et al. (2001) found no coral bleaching in the Society Islands in 1998
despite high temperature anomalies, but attributed lack of bleaching to high
cloud cover and reduced light levels during the. period of elevated
temperatures. Statistical analyses of bleaching occurrence based on
cumulative temperature elevations, wind speed, and cloud cover predicted
the correct scenario for the 1998 event only when high cloud cover was
included in the analysis, indicating that the interactive effect of cloud cover
can reverse bleaching predictions based solely on temperature elevation.
Other findings suggest that coral populations can adapt to localized
temperature conditions. Cook et al. (1990) found that Bermuda corals at
lagoon and inshore sites, where they were subject to higher and more
variable temperatures, were more resistant to bleaching in 1987 than the
same species at offshore sites. Similar patterns have been observed on the
Great Barrier Reef (Marshall and Baird, 2000) and the East Pacific
(Guzman and Cortes, 2001). Berkehnans (2002) proposed that thermal
adaptation had taken place over both local (10s of km) and regional (100s to
1000s of km) scales in the Great Barrier Reef, although Berkelmans and
Oliver (1999) concluded that inshore reefs were more prone to bleaching
than offshore reefs because of higher inshore temperatures and probably
reduced circulation. An indication of localized thermal adaptation was
found in the Colombian Pacific (Vargas-Angel et al., 2001), where coral
responses to the 1997-98 elevated temperatures showed less bleaching and
lower mortality in an area where long-term temperatures were consistently
CORAL BLEACHING 201
higher by 0.5-1.0%. Although Bruno et al. (2001) did not find significant
differences in bleaching between sites at 3-5m compared with lO-12m
depths from the severe 1997-98 ENS0 event in Palau, Coles (unpublished
report) found high coral survival in nearshore compared with offshore areas
in August 1999, one year after the event. Corals at various nearshore sites
around the island of Babeldoab in 1999 were abundant, well pigmented, and
in apparently healthy condition at temperatures up to 31.7”C, equivalent to
the temperatures that occurred during the bleaching event (Bruno et al.,
2001). Coral coverage and species composition in these nearshore areas
was indistinguishable from observations made on surveys in 1997. By con-
trast, virtually all Acroporu and many other species on offshore reefs were
dead in 1999.
These examples indicate a capacity for selective adaptation by various
coral species to elevated temperatures. However, nothing is known about
the conditions or time frame under which this capacity was acquired. The
critical question pertaining to large-scale survival of corals and continued
viability of coral reefs over the next century is whether the temperature
tolerances of corals and their symbionts can adjust rapidly enough to
a changing ocean temperature environment, and whether the maximum
temperatures that ultimately occur will exceed adaptation capacity.
Attempts to predictively model reef conditions that may result from rising
sea temperatures have usually used fixed coral thermal tolerances (Hoegh-
Guldberg, 1999) predicting coral declines and phase shifts to algal-
dominated reefs over the next century. However, models comparing
projected global seawater change with various estimates of acclimation
(i.e., adaptation) times (Ware et al., 1996) suggest that, although probable
bleaching events are likely to increase over the next century, development
of higher temperature thresholds in 25-50 years may dramatically reduce
bleaching probabilities and frequencies. This suggests that models projecting
future conditions for reef corals and coral reefs could utilize specific
information relative to thermal acclimatization and adaptation of corals and
their symbionts. Especially needed are data on the timeframe required for
selective adaptation to both gradually increasing temperature and to
infrequent temperature increases in order to project the eventual impacts of
both global warming and El Nino events.
Plate 2 (a) Bleached corals near the entrance to Suva Harbor in March 2000.
Extensive mortality and wave breakage of branching and arborescent colonies
followed the bleaching event. (b) Coral recolonization near this reef in March 2002,
showing competition between colonies for available habitat space was already
underway. Settlement of new colonies was observed as early as three months
following the end of the bleaching event in 2000. (Pictures and information provided
by Ed Lovell.)
204 S. L. COLES AND BARBARA E. BROWN
Pacific reefs of Costa Rica following the 1982-83 ENSO. They attribute this
recovery to corals more tolerant of thermal stress, and note that mortality of
such corals was very limited during the 1997-98 ENS0 warming. Kayanne
et al. (2002) noted that recovery of Montipora to prebleaching conditions
two years after the 1998 bleaching event had resulted in high mortality in the
southern Ryukus, although Muntipora patches with coverage of less than
10% did not recover in that time period. Mortality and recovery varied
among the other genera surveyed, with low mortality and little overall
change shown for Heliopora and massive Porites, high mortality and
moderate recovery for branching Porites and Acropora, and high mortality
with no recovery shown for Pavona.
A recent review (Fitt et al., 2001) has emphasized that reliable conclusions
about coral bleaching and mortality should be based on measurements of a
variety of environmental factors, such as duration of thermal stress, light
intensity, and quality (Warner et al., 2002). It was considered that
substantial reductions in algal symbiont concentrations, i.e., subliminal
bleaching, can be normal annual events. Fitt et al. (2001) also question
whether bleaching is a meaningful indicator for coral mortality, given the
lack of information linking zooxanthellae loss to coral death. Going further,
the available information is, in our view, insufficient to provide definitive
conclusions about the long-term fate of corals and reefs impacted by coral
bleaching. Uncertainties remain concerning the tropical seawater tempera-
ture environment and frequency of thermal events in the next century. We
are only beginning to acquire basic information on bleaching thresholds,
and the capacity of corals and their symbionts to acclimatize or adapt to
increasing temperatures or thermal events. Limited information is available
concerning linkages between bleaching and mortality, reproduction,
recruitment, and the capability of coral assemblages to recover and
reestablish after a bleaching event. Even less information is available as to
whether coral acclimatization and adaptation can occur sufficiently fast to
adjust to temperature anomalies that may occur.
Uncertainties also remain concerning the interaction of the stresses which
induce coral bleaching with other sources of coral stress and reef alteration
(Buddemeir and Smith, 1999), such as nitrification and eutrophication,
increased macroalgal growth that may result from overfishing of herbivores
and reduced coral growth rates that result from ocean pH changes related to
increased atmospheric CO2 (Pittock, 1999). The combined effects of these
and other important factors with temperature and light effects on coral
survival and propagation may be additive, synergistic, or neutral, but not
necessarily negative in all cases. Turbid environments, generally considered
to inhibit coral growth and survival, may shield corals from high light
intensities and act as refugia for corals during times of thermal stress, and
contribute to acclimatization and adaptation (Meesters et al., 2002). This
attests to the potential importance of nonreef communities containing
resistant corals, both locally and globally, in providing recruits during
periods of large-scale disturbance (Buddemeir and Smith, 1999).
Various scenarios resulting from mass coral bleaching have been
presented by Done (1999), which include coral tolerance and adaptation,
shifting of coral populations to smaller size classes, changing of species,
compositions toward more tolerant coral species with decreasing diversity,
208 S. L. COLES AND BARBARA E. BROWN
Stress
Low, Intermittent 4 -, Hiih, Frequent
Symptom
Mechanism Short Term Acclimatization Long Term Adaptation, Reef Coral Redtiion
of Present Coral ape&s Selaotlon for Tolerant and/or Elimination, Coral
and/or Phenotypee Spscles and/or Gsnotypsa Spa&s Extinctions
Coral Dominance d---.-.-b Corals and Coralline Algae 4-w Macroalgae Dominance
Reef Composition
9. CONCLUSIONS
For the last 20 years corals and coral reefs have globally undergone repeated
stress from periodic elevation of seawater temperatures that is unprece-
dented in approximately one hundred years during which scientists have
been studying corals and their environmental responses. If these stresses
continue and seawater baseline temperature increases in the next century,
the tolerances of corals and their symbiotic zooxanthellae will be severely
210 S. L. COLES AND BARBARA E. BROWN
tested in many parts of the world where corals and coral reefs are the
dominant biotope.
There is ample evidence that global temperature, including SST, has risen
substantially and that the rise is continuing (Wigley et al., 1997; National
Research Council, 2002; Hansen, 2003). Responses to this warming have
been shown by both terrestrial and aquatic ecosystems (Parmesan and
Yohe, 2003). However, the rise has been most pronounced in the Atlantic
and at higher latitudes in the northern hemisphere (Hansen, 2003), and
changes have been less obvious in some tropical seas. Recent analyses of
satellite SST and in situ seawater temperatures (Liu et al., 2002; Strong et al.,
pers. comm.) suggest that, with ENS0 events excluded, the overall trend in
SSTs in certain tropical waters, notably the western tropical Pacific, has
been stable for the last two decades and in some regions temperature has
fallen. There has also been some controversy about tropical temperatures
during past “greenhouse” periods in the Eocene and Cretaceous (Zachos
et al., 2002).
Thus, projections of a steadily increasing baseline of SSTs underlying
periodic ENS0 events (Hoegh-Guldberg, 1999) may not apply to all tropical
regions. Even if SST warming occurs generally in the tropics and
temperature anomalies associated with ENS0 periods continue, there is
evidence that a degree of adaptability, not yet rigorously defined, exists for
corals and their zooxanthellae, suggesting that these organisms could
continue to dominate coral reefs. We base this conclusion on demonstrated
differences in coral thermal thresholds linked to ambient temperatures, both
locally and regionally, on experimentally demonstrated protective mecha-
nisms such as HSPs, coral fluorescent pigments, and zooxanthellae
adaptability, on limited experimental evidence for acclimatization and/or
adaptation, and on the rapid recovery of corals and reefs that has been
observed following bleaching events.
Repeated bleaching events followed by various levels of coral mortality
during the last two decades has led to the perception among many reef
scientists and the general public that coral bleaching is likely to result in
degradation and demise of coral reefs as a major tropical biotope within the
next 50 years. Although most of the available information and projections are
not encouraging in terms of the environmental stresses that are likely to
occur, there are also indications that reef corals have “potential for greater
physiological tolerance than might have been previously expected” (Done,
1999), and “possess effective mechanisms of adaptation and acclimation that
have ensured their survival and recurrence over geologic time” (Buddemeir
and Smith, 1999). Additional research is needed to clarify the potential for
corals and zooxanthellae to adapt to increasing temperatures occurring in
both brief events and over the long term. Since recruitment plays a major
function in reef recovery after bleaching events, it will be critically important
CORAL BLEACHING 211
to clarify the tolerance of coral larvae and newly settled juvenile corals versus
adult stages, and determine the importance of habitat diversity in providing
refuges for juveniles, both during and after bleaching events. Carefully
managed, long-term monitoring programs with high statistical power need to
be established or continued on reefs worldwide to clarify initial and long-term
impacts of coral bleaching events, and to test whether certain environmental
factors may provide resistance and resilience to coral bleaching (Done, 1999;
West, 2001; West and Salm, in press). This information could then be used to
establish criteria for protected areas to provide refugia as sources of
recruitment for coral reef recovery after bleaching events (Salm and Coles,
2001; Salm et al., 2001). Only after considerably more basic research has been
completed will we be able to make meaningful projections of the long-term
impacts of coral bleaching.
The biologist’s scope for understanding the complex interactions of
environmental stresses on coral bleaching and the equally complex
responses of the coral/algal symbiosis to these stresses may be significantly
expanded in the future by the application of environmental genomics.
Recent developments in DNA and protein-based technologies offer an
enormous increase in the efficiency of gene discovery and characterization,
placing focus specifically upon those genes that are upregulated as a result of
stress. Attempts to understand just how well corals may adjust to rising
seawater temperatures will need to focus increasingly on genetic variation,
both in terms of selection and phenotypic plasticity for ecophysiological
traits. Regarding phenotypic plasticity, Pigliucci (1996) comments “the old
metaphor of genes as blueprints for the organism has to be abandoned in
favor of a more complex view that sees organismal properties emerging from
local and limited genetic effects.” Work on noncoral organisms has shown
that there is considerable genetic variation for phenotypic plasticity in
natural populations and that this variation is both character and
environment specific (Via et al., 1993; Ackerly et al., 2000). Targeting
those ecophysiological processes that appear to confer thermal tolerance
in corals (e.g., xanthophyll cycling capability, HSPs, and oxidative enzyme
production to name but a few) and identifying the genes responsible for
plasticity in these traits in coral/algal symbioses from different environments
would be major advances in our understanding of the scope of corals to
survive an era of global warming.
ACKNOWLEDGEMENTS
These concepts expressed in this review have been influenced through many
years of observations and discussions with researchers in the field of coral
212 S. L. COLES AND BARBARA E. BROWN
biology, including those who may not totally agree with all of the
conclusions. These include stimulating conversations on coral bleaching
that occurred among participants in the workshop on Coral Bleaching and
Marine Protected Areas. Mitigating Coral Bleaching Impact Through MPA
Design, held at Bishop Museum in May 2001, namely R. Salm, B. Causey,
T. Done, P. Glenn, W. Heyman, P. Jokiel, G. Llewellyn, D. Obura, J. Oliver,
and J. West. Important input has also come from A. Salih, T. Nahaky,
T. McCleod, and E. Lovell, who kindly provided the photos for Plate 2.
Two anonymous reviewers and A.J. Southward provided very helpful
comments and suggested changes that resulted in major improvements to
the article. Figure 1 is reprinted by permission of University of Hawaii
Press, and Figures 2 and 3 by permission of Inter-Research. Thanks to
The Natural Environment Research Council, The Royal Society, and The
Leverhulme Trust in the United Kingdom for supporting research
conducted by BEB in Thailand over the last 23 years that has provided
insight to some of the issues raised by this review. Contribution No. 2003-
001 to the Pacific Biological Survey.
REFERENCES
Ackerly, D. D., Dudley, S. A., Sultan, S. E., Schimitt, J., Coleman, J. S., Linder,
C. R., Sandquist, D. R., Geber, M. A., Evans, A. S., Dawson, T. E. and
Lechowicz, M. J. (2000). The evolution of plant ecophysiological traits: recent
advances and future directions. Bioscience 50, 979-995.
Al-Sofyani, A. and Davies, P. S. (1992). Seasonal variation in production and
respiration of Red Sea corals. “Proceedings of the Seventh International Coral
Reef Symposium, Guam”, Vol. 1, pp. 351-357.
Ambarsari, I., Brown, B. E., Barlow, R. G., Britton, G. and Cummings, D. G.
(1997). Fluctuations in algal chlorophylls and carotenoid pigments during solar
bleaching in the coral Goniustrea aspera at Phuket, Thailand. Marine Ecology
Progress Series 159, 303-307.
Aronson, R. B., Precht, W. F., Macintyre, I. G. and Murdoch, T. 5. T. (2000). Coral
bleachout in Belize. Nature 405, 36.
Aronson, R. B., Precht, W. F., Toscano, M. A. and Koltes, K. H. (2002). The 1998
bleaching event and its aftermath on a coral reef in Belize. Marine Biology 141,
435447.
Ashbumer, M. and Bonner, J. J. (1979). The induction of gene activity in Drosophila
by heat shock. Cell 17, 241-254.
Baird, A. H. and Marshall, P. A. (1998). Mass bleaching of corals on the Great
Barrier Reef. Coral Reefs 17, 376.
Baird, A. H. and Marshall, P. A. (2002). Mortality, growth and reproduction in
scleractinian corals following bleaching on the Great Barrier Reef. Marine Ecology
Progress Series 237, 133-141.
Baker, A. C. (2001). Reef corals bleach to survive change. Nature 411, 765-766.
CORAL BLEACHING
213
Banin, E., Israely, T., Kushmaro, A., Loya, Y., Orr, E. and Rosenberg, E. (2000).
Penetration of the coral-bleaching bacterium Vibrio shiloi into Oculina patagonica.
Applied Environmental Microbiology 66, 303 l-3036.
Banin, E., Israely, T., Fine, M., Loya, Y. and Rosenberg, E. (2001). Role of endo-
symbiotic zooxanthellae and coral mucus in the adhesion of the coral-bleaching
pathogen Vibrio shiloi to its host. FEMS Microbiology Letters 199, 33-37.
Beckman, K. B. and Ames, B. N. (1988). The free radical theory of aging matures.
Physiological Reviews 78, 547-58 1.
Ben-Haim, Y. and Rosenberg, E. (2002). A novel Vibrio sp. pathogen of the coral
Pocillopora damicornis. Marine Biology 141, 41-55.
Berkelmans, R. (2001). Bleaching, upper thermal limits and temperature adaptation
in reef coral. Ph. D. thesis, James Cook University, Townsville, 178 pp.
Berkelmans, R. (2002). Time-integrated thermal bleaching thresholds of reefs and
their variation on the Great Barrier Reef. Marine Ecology Progress Series 229,
73382.
Berkelmans, R. and Oliver, J. K. (1999). Large-scale bleaching of corals on the Great
Barrier Reef. Cowl Reefs 18, 55-60.
Berkelmans, R. and Willis, B. L. (1999). Seasonal and local spatial patterns in the
upper thermal limits of corals on the inshore Central Great Barrier Reef. Coral
Reefs 18, 219-228.
Black. N. A., Voellmy, R. and Szmant, A. M. (1995). Heat shock protein induction
in Montastrca filveoiata and Aiptasia pallida exposed to elevated temperatures.
Biological Bulletin 188, 234-240.
Boesch, D. F., Field. J. C. and Scavia’D (2000). “The potential consequences of
climate variability and change on coastal areas and marine resources”. NOAA
Coastal Ocean Program. Decision Analysis Series No. 21.
Boschma, H. (1925). The nature of the association between Anthozoa
and zooxanthellae. Proceedings c~f’ the National Acudemy qf’ Science, USA 11.
65-67.
Boschma, H. (1926). On the food of reef corals. Proceedings of the Academy of
Science, Amsterdam 29, 993-997.
Brown, B. E. (1987). Worldwide death of corals - natural cyclical events or man-
made pollution? Marine Pollution Bulletin 18, 9-13.
Brown, B. E. (1997a). Adaptations of reef corals to physical environmental stress.
Advances in Marine Biology 31~ 222-299.
Brown, B. E. (1997b). Coral bleaching: causes and consequences. Coral Re<fk 16,
S1299S138.
Brown, B. E. and Suharsono (1990). Damage and recovery of coral reefs affected by
El Nino related seawater warming in the Thousand Islands, Indonesia. Coral Reefi
8, 1633170.
Brown, B. E., Ambarsari, I., Warner, M. E., Fitt, W. K., Dunne, R. P., Gibb, S. W.
and Cummings, D. G. (1999b). Diurnal changes in photochemical efficiency and
xanthophyll concentrations in shallow water reef corals: evidence for photoinhibi-
tion and photoprotection. Coral Re<fy 18, 99-105.
Brown, B. E., Clarke, K. R. and Warwick, R. M. (2002b). Serial patterns of
biodiversity change in corals across shallow reef flats in Ko Phuket, Thailand, due
to the effects of local (sedimentation) and regional (climatic) perturbations.
Marine Biology 141, 21.-29.
Brown, B. E., Downs, C. A., Dunne, R. P. and Gibb, S. W. (2002~). Exploring the
basis of thermotolerance in the reef coral Goniastrea aspera. Marine Ecology
Progress Series 242. I I9- 129.
214 S L. COLES AND BARBARA E. BROWN
Fabricius, K. E. (1999). Tissue loss and mortality in soft corals following mass-
bleaching. Coral Reefs 18, 54.
Fagerstrom, J. A. and Rougerie, A. (1994). 1994 Coral Bleaching Event, Society
Islands, French Polynesia. Marine Pollution Bulletin 29, 34-35.
Fang, L., Huang, S. and Lin, K. (1997). High temperature induces the synthesis of
heat-shock proteins and the elevation of intercellular calcium in the coral Acropora
grandis. Coral Reefs 16, 1277131.
Feingold, J. S. (2001). Responses of three coral communities to the 1997798 El Nino-
Southern oscillation: Galapagos Islands, Equador. Bulletin of Marine Science 69,
61-77.
Fine, M., Banin, E., Israely, T., Rosenberg, E. and Loya, Y. (2002a). Ultraviolet
radiation prevents bleaching in the Mediterranean coral Oculina patagonica.
Marine Ecology Progress Series 226, 249-254.
Fine, M., Oren, U. and Loya, Y. (2002b). Bleaching effect on regeneration and
resource translocation in the coral Oculina patagonica. Marine Ecology Progress
Series 234, 119-125.
Fitt, W. K. and Warner, M. E. (1995). Bleaching patterns of four species of
Caribbean reef corals. Biological Bulletin 189, 298-307.
Fitt, K., Brown, B. E., Warner, M. E. and Dunne, R. P. (2001). Coral bleaching:
interpretation of thermal tolerance limits and thermal thresholds in tropical corals.
Coral Reefs 20, 5 l-65.
Gardiner. J. S. (1903). “The Fauna and Geography of the Maldive and Laccadive
Archipelagoes”. University Press, Cambridge.
Gates, R. D. (1990). Seawater temperature and sublethal coral bleaching in Jamaica.
Coral Recfk 8, 1933197.
Gates, R. D. and Edmunds, P. J. (1999). The physiological mechanisms of
acclimatization in tropical reef corals. American Zoologist 39, 3043.
Gates, R. D., Baghdasarian, G. and Muscatine, L. (1992). Temperature stress causes
host cell detachment in symbiotic cnidarians: implications for coral bleaching.
Biological Bulletin 182, 324-332.
George, D. and John, D. (1998). High sea temperatures along the coast of Abu
Dhabi (UAE), Arabian Gulf ~ their impact upon corals and macroalgae. Reef
Encounter 23, 21-23.
Gleason, M. G. (1993). Effects of disturbance on coral communities: bleaching in
Moorea, French Polynesia. Coral Re<f:y 12, 193-201.
Glynn, P. W. (1983). Extensive “bleachmg” and death of reef corals on the Pacific
coast of Panama. Environmental Conservation 10, 1499153.
Glynn, P. W. (1984). Widespread coral mortality and the 1982283 El Nino warming
event. Environmental Conservation 11, 133-146.
Glynn, P. W. (1991). Coral reef bleaching in the 1980s and possible connections with
global warming. Trends in Ecology and Evolution 6, 1755178.
Glynn, P. W. (1993). Coral reef bleaching: ecological perspectives. Coral Recfk 12.
l-17.
Glynn, P. W. and D’Croz, L. (1990). Experimental evidence for high temperature
stress as the cause of El Nino-coincident coral mortality. Coral Recfh 8, 18 l-1 9 1.
Glynn, P. W., Mate, J. L., Baker, A. C. and Calderon, M. 0. (2001). Coral bleaching
and mortality in Panama and Equador during the 1997798 El Nino Southern
Oscillation event: spatial/temporal patterns and comparisons with the 1982-1983
event. Bulletin of‘ Marine Science 69, 79-109.
Goreau. T. J. (1992). Bleaching and reef community change in Jamaica: 195 l-1 99 1.
American Zoologist 32, 683-695.
CORAL BLEACHING 217
Jimenez, C., Cortes, J., Leon, A. and Ruiz, E. (2001). Coral bleaching and mortality
associated with the 1997-98 El Nino in an upwelling environment in the eastern
Pacific (Gulf of Papagyo, Costa Rica). Bulletin of Marine Science 69, 15 1-169.
Jokiel, P. L. and Coles, S. L. (1974). Effects of heated effluent on hermatypic corals
at Kahe Point, Oahu. Pacific Science 28, 18.
Jokiel, P. L. and Coles, S. L. (1977). Effects of temperature on the mortality and
growth of Hawaiian reef corals. Marine Biology 43, 201-208.
Jokiel, P. L. and Coles, S. L. (1990). Response of Hawaiian and other Indo-Pacific
reef corals to elevated temperature. Coral Reefs 8, 155-162.
Jokiel, P. L. and Guinther, E. B. (1978). Effects of temperature on reproduction in
the hermatypic coral Pocillopora damicornis. Bulletin of Marine Science 28,
786789.
Jones, R. S. and Randall, R. H. (1973). A study of biological impact caused by
natural and man-induced changes on a tropical reef. Marine Laboratory,
University of Guam, Technical Report 7.
Jones, R. J., Berkelmans, R. and Oliver, J. K. (1997). Recurrent bleaching of corals
at Magnetic Island (Australia) relative to air and seawater temperature. Marine
Ecology Progress Series 158, 289-292.
Jones, R. J., Hoegh-Guldberg, O., Larkum, W. D. and Schreiber, U. (1998).
Temperature-induced bleaching of corals begins with impairment of the carbon
dioxide fixation mechanism in zooxanthellae. Plant, Cell and Environment 21,
1219-1230.
Jones, R. J., Ward, S., Amri, A. Y. and Hoegh-Guldberg, 0. (2000). Changes in
quantum efficiency of Photosystem II of symbiotic dinoflagellates after heat stress,
and of bleached corals sampled after the 1998 Great Barrier Reef mass bleaching
event. Marine and Freshwater Research 51, 63-71.
Kayanne, H., Harii, S., Yoichi, I. and Akimoto, F. (2002). Recovery of coral
populations after the 1998 bleaching on Shiraho Reef, in the southern Ryukyus,
NW Pacific. Marine Ecology Progress Series 239, 93-103.
Kinsman, D. J. J. (1964). Reef coral tolerance of high temperatures and salinities.
Nature 202, 128&1282.
Kinzie, R. A., III, Takayama, M., Santos, S. R. and Coffroth, M. A. (2001). The
adaptive bleaching hypothesis: experimental tests of critical assumptions,
Biological Bulletin 200, 5 l-58.
Kushmaro, A., Banin, E., Loya, Y., Stackebrandt, E. and Rosenberg, E. (2001).
Vibrio shiloi sp. nov., the causative agent of bleaching of the coral Oculina
patagonica. International Journal of Systematic and Evolutionary Microbiology 51,
1383-1388.
Kushmaro, A., Loya, Y., Fine, M. and Rosenberg, E. (1996). Bacterial infection and
coral bleaching. Nature 380, 396.
Kushmaro, A., Rosenberg, E., Fine, M., Haim, Y. and Loya, Y. (1998). Effect of
temperature on bleaching of the coral Oculina patagonica by Vibrio AK- 1. Marine
EcoIogy Progress Series 171, 13 l-l 37.
Lasker, H. R., Peters, E. C. and Coffroth, M. A. (1984). Bleaching of reef
coelenterates in the San Blas Islands, Panama. Coral Reefs 3, 1833190.
Lesser, M. P. (1997). Oxidative stress causes coral bleaching during exposure to
elevated temperatures. Coral Reefs 16, 187-192.
Lesser, M. P., Stochaj, W. R., Tapley, W. R. and Schick, J. M. (1990). Bleaching in
coral reef anthozoans: effects of irradiance, ultraviolet radiation, and temperature
on the activities of protective enzymes against active oxygen. Coral Re<fi 8,
225-232.
CORAL BLEACHING 219
Lindahl, U., Ghman, M. C., and Schelten, C. K. (2001). The 1997/1998 mass
mortality of corals: effects on fish communities on a Tanzanian coral reef. Marine
Pollution Bulletin 422, 127-13 1.
Lindquist, S., and Craig , E. A. (1988). The heat shock proteins. Annual Review of
Genetics 22, 631-677.
Liu, G., Strong, A. E., and Casey, K. (2002). Investigation of sea surface temperature
tendencies during 1985-1999 using NOAA/NASA satellite AVHRR
oceans pathfinder data set. In “Abstracts of the 2002 Ocean Sciences Meeting,
Honolulu”.
Loh, W. K. W., Loi, T., Carter, D., and Hoegh-Guldberg, 0. (2001). Genetic
variability of the symbiotic dinoflagellates from the wide ranging coral species
Seriatopora hystrix and Acropora longicyathus in the Indo-West Pacific. Marine
Ecology Progress Series 222, 97-107.
Loya, Y., Sakai, K., Yamzato, K., Nakano, Y., Gambali, H., and van Woesik, R.
(2001). Coral bleaching, the winners and the losers. Ecology Letters 4, 122-131.
Marshall, P. A., and Baird, A. H. (2000). Bleaching of corals on the Great Barrier
Reef: differential susceptibilities among taxa. Coral Reefs 19, 155-163.
Mayer, A. G. (1914). The effects of temperature on tropical marine animals.
Publications qf the Carnegie Institution, Washington 183, 3-24.
Mayer, A. G. (1917). Is death from high temperature due to accumulation of
acid in the tissues? Proceedings qf the National Academy of’ Science, USA
3, 626627.
Mayer, A. G. (1918a). Toxic effects due to high temperature. Papers jiorn the
Carnegie Institution, Washington 12, 1733178.
Mayer, A. G. (1918b). Ecology of the Murray Island coral reef. Publications of’ the
Carnegie Institution, Washington 213, l-48.
Mayer, A. G. (1924). Structure and ecology of Samoan reefs. Publications qf the
Carnegie Institution, Washington 340, l-25.
McClanahan, T. R. (2000). Bleaching damage and recovery potential of Maldivian
coral reefs. Marine Pollution Bulletin 40, 587-597.
McClanahan, T. R., Muthiga, N. A., and Mangi, S. (2001). Coral and algal changes
after the 1998 coral bleaching: interaction with reef management and herbivores
on Kenyan reefs. Coral Reejk 19, 38&391.
Meesters, E. H., and Bak, R. P. M. (1993). Effects of coral bleaching on tissue
regeneration potential and colony survival. Marine Ecology Progress Series
96, 189-198.
Meesters, E. H., Nieuwland, G., Duineveld, G. C. A., Kok, A., and Bak, R. P. M.
(2002). RNA/DNA ratios of scleractinian corals suggest acclimatisation/adapta-
tion in relation to light gradients and turbidity regimes. Marine Ecology Progress
Series 227, 233-239.
Mendes, J. M., and Woodley, J. D. (2002). Effect of the 199551996 bleaching event
on polyp tissue depth, growth, reproduction and skeletal band formation in
Montastraea annularis. Marine Ecology Progress Series 235, 933102.
Michalek-Wagner, K., and Willis, B. L. (2001a). Impacts of bleaching on the soft
coral Lobophytum compac~tum. I. Fecundity, fertilization and offspring viability.
Coral Ret$ 19, 231-239.
Michalek-Wagner, K., and Willis, B. L. (2001b). Impacts of bleaching on the soft
coral Lobophytum compacturn. II. Biochemical changes in adults and their eggs.
Coral Reeji 19, 240-246.
Morimoto, R. I., Jurivich. D. A., Kroeger, P. E., Mathur, S. K., Murphy, S. P.,
A. N., Sarge, K., Abravaya, K., and Sistonen, L. T. (1994). Regulation of heat
220 S. L. COLES AND BARBARA E. BROWN
Quinn, N. J., and Kojis, B. L. (1999). Subsurface seawater temperature variation and
the recovery of corals from the 1993 coral bleaching event in waters off St.
Thomas, U. S. Virgin Islands. Bulletin of Marine Science 65, 201-214.
Reyes Bonilla, H. (2001). Effects of the 1997-98 El Nino-Southern Oscillation on
coral communities of the Gulf of California, Mexico. Bulletin of Marine Science
69, 251-266.
Richardson, L. L., Goldberg, W. M., Kuta, K. G., Aronson, R. B., Smith, G. W.,
Halas, J. C., and Miller, S. L. (1998). Florida’s mystery coral-killer identified.
Nature 392, 557-558.
Riegl, B. (1999). Corals in a non-reef setting in the southern Arabian Gulf (Dubai,
UAE): fauna and community structure in response to recurring mass mortality.
Coral Reejk 18, 63-73.
Riegl, B. (2002). Effects of the 1996 and 1998 positive sea-surface temperature
anomalies on corals, coral diseases and fish in the Arabian Gulf (Dubai, UAE).
Marine Biology 140, 29-40.
Roberts, D. A., Hofmann, G. E., and Somero, G. N. (1997). Heat-shock expression
in Mytilus californianus: acclimatization (seasonal and tidal-height comparison)
and acclimation effects. Biological Bulletin 192, 309-320.
Rodriguez-Lanetty, M., Loh, W., Carter, D., Hoegh-Guldberg, 0. (2001).
Latitudinal variability in symbiont specificity within the widespread scleractinian
coral Plesiastrea versipora. Marine Biology 138, 1175-l 181.
Salih, A. (2001). Mechanisms of light and temperature-induced bleaching of corals
and the sunscreening role of their fluorescent pigments. Ph. D. thesis, University of
Sydney, Sydney, 236 pp.
Salih, A., Hoegh-Guldberg, O., and Cox, G. (1998a). Bleaching responses of
symbiotic dinoflagellates in corals: the effects of light and elevated temperature on
their Morphology and Physiology. In “Proceedings of the Australian Coral Reef
Society 75th Anniversary Conference, Heron Island October 1997” (J. G.
Greenwood and N. J. Hall, eds.), pp. 199-216. School of Marine Science,
University of Queensland, Brisbane, Australia.
Salih, A., Hoegh-Guldberg, O., and Cox, G. (1998b). Photoprotection of symbiotic
dinoflagellates by fluorescent pigments in reef corals. In “Proceedings of the
Australian Coral Reef Society 75th Anniversary Conference, Heron Island
October” (J. G. Greenwood and N. J. Hall, eds.), pp. 217-230. School of
Marine Science, The University of Queensland, Brisbane, Australia.
Salih, A, Larkum, A., Cox, G., Kuhl, M., and Hoegh-Guldberg, 0. (2000).
Fluorescent pigments in corals are photoprotective. Nature 408, 850-853.
Salm, R. V., and Coles, S. L., eds. (2001). “Coral Bleaching and Marine Protected
Areas. Mitigating Coral Bleaching Impact Through MPA Design”. Bishop
Museum, Honolulu, 29-31 May. The Nature Conservancy Asia Pacific Coastal
Marine Program Report No. 0102. Honolulu.
Salm, R. V., Smith, S. E., and Llewellyn, G. (2001). Mitigating the impact of coral
bleaching through marine protected area design. In “Coral Bleaching: Causes,
Consequences and Response. Selected papers presented .at the 9th International
Coral Reef Symposium on Coral Bleaching: Assessing and Linking Ecological and
Socioeconomic Impacts, Future Trends and Mitigation Planning” (H. Z.
Schuttenberg, ed.), Vol. Coastal Management Report 2230, pp. 81-88. Coastal
Resources Center, University of Rhode Island, Narragansett.
Salvat, B. (1992). The 1991 bleaching event in the Society islands, French Polynesia.
Proceedings qj’the 7th International Coral Reef Symposium, (R. H. Richmond, ed.),
University of Guam Press, VOG Station, Guam 1, 73.
222 S. L COLES AND BARBARA E. BROWN
1. INTRODUCTION
At present, one of the key issues for both marine and terrestrial ecologists as
well as resource managers is to resolve and predict the impacts of global
change on ecosystem dynamics. The objective of these activities is the
development of ecosystem-based management strategies with the ultimate
goal of preserving the structure and functioning of ecosystems and
contributing to the sustainable management of natural resources.
Contingent upon developing such strategies is a clear understanding of
the bottleneck processes (both biotic and abiotic) that impact the population
dynamics of key trophic level species, and which are influenced by global
change. Resolution of these bottleneck processes has to date been
determined primarily via an approach whereby the growth and overall
condition as well as trophic links of individuals are related to in situ
conditions and assumptions about the survival potential of the individual
are scaled up to the population level. Such classical approaches to the
resolution of key processes are limited, as they are reliant upon temporal
snapshots of complex and highly variable (both spatially and temporally)
interactions obtained from individuals that might never survive to become
part of the reproductive population. In order to expand the temporal
window of resolution of key processes, an approach termed “characteristics
228 JOHANNE DALSGAARD ETAL.
of survivors” (Fritz et al., 1990; Taggart and Frank, 1990; St. John et al.,
2000), has recently been used to identify processes leading to enhanced
survival success. This approach is based upon the examination of
phenotypic and genotypic characteristics of individuals before and after
experiencing an event. if, after exposure to a specific process, a random
subset of survivors exists from the initial population, no phenotypic or
genotypic selective advantage exists with respect to that process. However,
if a particular characteristic results in an increased survival success, this
characteristic can be described as increasing the individual’s fitness. To date,
the characteristics of survivors approach has primarily been used to identify
survivors in terms of growth rates (Miller et al., 1988; Meekan and Fortier,
1996) food webs (St. John and Lund, 1996; Storr-Paulsen et al., 2003) and
transport processes (St. John et al., 2000). All of these studies employ a
biomarker approach to identify in situ processes contributing to enhanced
growth, condition and survival success. Specific biomarkers included in
these studies comprise otolith microstructure (e.g., Meekan and Fortier,
1996) and fatty acid trophic markers (e.g., St. John and Lund, 1996). The
application of otolith microstructure for the study of larval, juvenile and
adult fish is now a common tool in fish ecology (e.g., Campana, 1996)
however, the application of fatty acid trophic markers (FATM) to address
issues in marine science is so far relatively limited. Hence, the major
objective of this review is to summarize applications of fatty acids (FA) as
trophic markers in marine ecosystems and furthermore, to assess the future
prospects for their application in resolving ecosystem dynamic processes.
For three decades, detailed information on the FA composition of marine
organisms has been generated under the assumption that, among other
things, such data may provide valuable insight into predator-prey
relationships. Studies employing FATM have taken place in both marine
and freshwater pelagic systems as well as in demersal and deep-sea
applications. In order to constrain this review and avoid duplication we
will focus on applications in the pelagic marine system, and will not consider
other lipid biomarkers such as sterols and hydrocarbons (but see, e.g.,
Sargent and Whittle, 1981; Volkman et al., 1998). Furthermore, for an
introduction to FATM in freshwater ecosystems we refer readers to
Desvilettes et al. (1997) and Napolitano (1999).
In the first part of this review, we introduce the FATM concept and give
a chronological synopsis of the development and application of FATM in
marine food web research. General FA biochemistry is briefly presented and
the distribution of lipids and FA in marine organisms is discussed.
Subsequently, the dynamics of lipids and FA at the various trophic levels
(i.e., primary producers, zooplankton and fish) are described in more detail.
At the Grst trophic level microalgae are given most emphasis, as they are the
principal primary producers in the marine environment, supporting both
FATTY ACID TROPHIC MARKERS 229
pelagic and offshore benthic food webs (Parsons, 1963; Kayama et al.,
1989). In order to summarize and compare the information on algal groups,
the characteristic FA patterns of various marine microalgal classes are
visualized through a partial least squares (PLS) regression analysis based
on published laboratory culture studies, and the conclusions compared to
natural plankton communities. We comment briefly on macroalgae, which
are largely confined to shallow coastal regions. Here they may support local
benthic food webs (Ackman et al., 1968 and references therein), while they
generally have little importance in the pelagic marine environment. Finally,
FATM of heterotrophic bacteria and terrestrially derived organic matter are
summarized. In general, bacteria make important contributions in the
marine environment, particularly in microbial loop food webs, which
develop primarily in stratified and nutrient depleted areas (e.g., Cushing,
1989 and references therein). Terrestrial matter can be important in coastal
and estuarine ecosystems, and differences in the FA pattern between the
terrestrial and marine environment have been used to detect the entrainment
of terrestrial organic matter into coastal food webs. We look therefore
briefly at characteristic terrigenous FATM.
At the next trophic level, zooplankton form an essential link between
primary producers and higher order consumers (Sargent, 1976; Sargent and
Henderson, 1986). We focus on herbivorous calanoid copepods, as they are
the best studied group of zooplankton with respect to FATM. We describe
the uptake, incorporation and modification of dietary FA during different
life history stages, and give examples of studies that have verified the
conservative incorporation of specific phytoplankton-derived FATM by
copepods. Moreover, apart from incorporating and transferring dietary FA
from primary producers to higher trophic levels, calanoid copepods are
themselves important producers of specific FA and fatty alcohols (the
moieties of wax esters). Hence, we discuss their capacity to biosynthesize
such compounds de ytovo, focusing on those FA and fatty alcohols that can
be used to elucidate predator-prey relationships at higher trophic levels. The
FA characteristics of omnivorous and carnivorous copepods are subse-
quently discussed, and FA that have been used as markers of carnivory are
summarized. Lastly, the information on herbivorous, omnivorous and
carnivorous copepods is summarized and compared in a PLS regression
analysis based on FA and fatty alcohol compositional data of key Arctic
and Antarctic copepod species.
Next, we review the dynamics of FA in fish, primarily teleosts, which
principally catabolize and transform dietary FA (Sargent and Henderson,
1986). We describe the processes of uptake, incorporation and modification
of dietary FA, de ~OVObiosynthesis, mobilization of FA during starvation
and reproduction, and summarize studies that have validated the FATM
approach in this group of organisms.
230 JOHANNE DALSGAARD ETAL
Finally, in the last section of the paper, we review major marine food webs
in which FA have been used to trace or confirm predator-prey relationships
and key processes impacting on ecosystem dynamics, i.e. the Arctic, the
Antarctic, northwest Atlantic Ocean, the North Sea, the Gulf of Alaska,
Mediterranean Sea, upwelling and subtropical-tropical systems. The section
is introduced by a comparison between these different systems based on the
influence of stratification processes on phytoplankton group dominance.
The perfect trophic marker is a compound whose origin can be uniquely and
easily identified, that is inert and nonharmful to the organisms, that is not
selectively processed during food uptake and incorporation, and that is
metabolically stable and hence transferred from one trophic level to the next
in both a qualitative and quantitative manner. Such a marker would provide
essential insight into the dynamics of ecosystems by presenting unique
information on pathways of energy flows, i.e., crucial information on which
all ecosystem models are eventually built. However, such markers are
unfortunately rare if nonexistent and instead we have to be content with less
ideal components, a category to which FA belong.
In the case of FATM, these lipid components are in many circumstances
incorporated into consumers in a conservative manner, thereby providing
information on predator-prey relations. Moreover, contrary to the more
traditional gut content analyses, which provide information only on recent
feeding, FA provide information on the dietary intake and the food
constituents leading to the sequestering of lipid reserves over a longer period
of time (e.g., Hakanson, 1984; St. John and Lund, 1996; Kirsch et al., 1998;
Auel et al., 2002). This integrating effect helps to resolve the importance
of specific prey items and can validate prey utilization strategies based on
traditional stomach content analyses (Graeve et al., 1994b). Furthermore,
traditional stomach analyses suffer from the fact that food items in the gut
are frequently difficult to identify and are quantitatively biased due to
differential digestion rates of soft and hard parts. For example, exoskeletons
and otoliths may be retained in the stomachs whereas softer tissue parts are
rapidly digested, and hence, seldom observed (e.g., Iverson et al., 1997a
and references therein; Budge et al., 2002). These problems are partly
circumvented by FA but unfortunately replaced by other constraints. For
example, no single FA can be assigned uniquely to any one species and
depending on the condition and metabolic strategy of the consumer, FA are
not necessarily metabolically stable (e.g., Section 3.2 and 4.3). In addition,
the temporal dynamics, i.e., turnover rate of individual FA, can be species-
specific and are often linked to the metabolic condition or reproductive
FATTY ACID TROPHIC MARKERS
231
status of the organism (Section 3.3 and 4.4) and have seldom been
quantified (St. John and Lund, 1996; Kirsch et al., 1998). Consequently, FA
have so far only been used as qualitative and “semi-quantitative” food web
markers, the latter in concert with other tracers such as stable isotopes
(Kiyashko et al., 1998; Kharlamenko et al., 2001). It still remains to be
established whether they can be used for more than that. This is a serious
challenge given the fact that whereas the FA composition may be used to
elucidate the dietary source of lipid reserves, it is not possible to discern
whether an individual is incorporating or depleting reserves in its current
situation, using a marker which gives no indication of the temporal
dynamics of growth or conditional status.
linear, experimental food web consisting of(i) algae - brine shrimp, and (ii)
brine shrimp nauplii - Hydra at 10°C and 20°C. These authors found that
the FA composition of the neutral lipid (NL) fraction resembled the diet
more closely than did that of the polar lipid fraction, and that
polyunsaturated fatty acids (PUFA) were predominantly concentrated
within the polar lipid fraction.
In a study involving the culturing of twelve species of unicellular marine
algae from the phytoplankton classes Chrysophyceae, Cryptophyceae,
Bacillariophyceae (diatoms), Dinophyceae (dinoflagellates), Chlorophyceae
(green algae), Prasinophyceae, Rhodophyceae (red algae) and
Xanthophyceae, Ackman et al. (1968) discovered that despite large
variations of individual FA within the different taxonomic classes,
common features could still be recognized. Subsequently, consistent with
these findings, Jeffries (1970) performed the seminal work on the changes in
the FA composition accompanying a succession of species within a
natural plankton community (Figure 1). In this study, a succession from
diatoms to flagellates in Narragansett Bay, Rhode Island, was found to
be associated with a decrease in the 16:1/16:0 ratio from > 2 to ~0.3.
This trend was, furthermore, partly mirrored in locally abundant Acartia
sp., assumed to feed on the algae. Complementary but much less
pronounced trends were also evident in the 18:4/18:1 ratio, highest when
the 16:1/16:0 ratio was lowest and commensurate with the peak in flagellate
dominance.
Shortly thereafter, Lee et al. (1971b) demonstrated that dietary FA were
incorporated conservatively into the wax ester (WE) fraction of marine
copepods. In this study, Calanus helgolandicus fed on either diatoms
(Lauderia borealis, Chaetoceros curvisetus, Skeletonema cost&urn) or
dinoflagellates (Gymnodinium splendens) showed a WE fatty acid composi-
tion very similar to its prey. Such similarities were, however, not observed
between the diet and the WE fatty alcohol composition, which consisted
primarily of saturated and monounsaturated fatty alcohols, purported to
be biosynthesized de novo by the copepods. Subsequently, Sargent (1976)
concluded that calanoid copepods differ from phytoplankton in containing
high proportions of CZO and CZ2 monounsaturated FA and fatty alcohols
biosynthesized de novo, and that these moieties can be recognized (as FA) in
copepod predators.
To verify the potential of phytoplankton FA as trophic markers, Graeve
et al. (1994a) performed a feeding experiment with herbivorous, Arctic
calanoid copepods (see also Section 3.4). After 42 days on a diet of
Thalassiosira antarctica (diatom), the level of 16: l(n-7) in Calanus
jinmarchicus had strongly increased, whereas 18:4(n-3) was almost depleted.
Opposite trends were observed in C. hyperboreus fed on Amphidinium
FATTY ACID TROPHIC MARKERS 233
ONDJFMAMJJASONDJ-ONDJFMAMJJAS
1963 1964 1965 1966
carterae (dinoflagellate), i.e., the level of 16:l(n-7) had decreased, while the
level of 18:4(n-3) had increased.
Kharlamenko et al. (1995) demonstrated how a .combination of FA
identified in samples of pelagic diatoms, seston, microbial mats, sediments
and macroalgae collected in an isolated shallow-water hydrothermal
ecosystem in Kurile Islands, east Pacific, could be used to identify potential,
major food sources of locally abundant macrozoobenthic species. Their
ratio of 16:1(n-7)/16:0 and the 20:5(n-3) content indicated that diatoms were
a major food source of all species. Furthermore, some of the deposit and
234 JOHANNE DALSGAARD ETAL.
0.45
i Skeletonema ,
Heterocapsa /
0.10 I -I
0 2 4 6 8 10 12 14 16 18
Time (Days)
Figure 2 Validation of the 16:l(n- 7)/16:0 specific food web tracer in larval
North Sea cod (Gadus morhua) raised on food webs based on either Skeletonema
costatum, Heterocapsa triquetra, a 50% mix of the two or starved. The algae were fed
to Acartia tonsa and the resultant nauplii fed to the cod larvae. Each point represents
an average of five cod larvae. Redrawn with permission after St. John and
Lund (1996).
FATTY ACID TROPHIC MARKERS 235
Fatty acids are ubiquitous components of all living organisms where they
form an essential and integral part of neutral and polar lipids and constitute
important precursors of “local” hormones (eicosanoids). One major role of
polar lipids is to provide the basic matrix of the cellular membranes into
which cholesterol, proteins and other membrane constituents are embedded
(Spector and Yorek, 1985; Stubbs and Smith, 1990; Cook, 1996; Vance,
1996). The dual structural and functional role of polar lipids limits the type
of FA that are incorporated, consisting principally of PUFA of the (n-3)
and (n-6) series (reviewed by Vaskovsky, 1989). These particular FA provide
special conformational properties to the biomembranes, and assist tissue
specific cells in reacting to external stimuli such as, e.g., changing
environmental temperatures and light regimes (Sargent et al., 1993; Cook,
1996).
The principal role of neutral lipids (NL), which in marine systems consist
predominantly of triacylglycerols (TAG) and WE, is as an energetic reserve
of FA that are destined either for oxidation to provide energy (ATP) or for
incorporation into phospholipids (PL) (Sargent and Whittle, 1981; Hslmer,
1989; Lee and Patton, 1989). The NL content and the constituent FA is
linked to the physiological status of the organism, and is determined by the
rate of turnover of the lipid depots, i.e., the coupled processes of anabolism
and catabolism. An organism experiencing a dietary surplus of energy may
accumulate lipids either directly, in which case the FA composition is similar
to the diet (Ackman and McLachlan, 1977; Sargent and Whittle, 1981) or
after modifying the FA to suit particular physiological needs, e.g., for the
formation of reproductive tissue. The former situation underpins the belief
that FA in many cases can be used to explore predator-prey relationships,
i.e., that they can be used as trophic markers.
In reviewing the literature and the use of FATM, it has become apparent
that different approaches have been used for detecting dietary relationships.
Either the total lipid (TL) composition has been analysed, or individual lipid
classes have been evaluated separately. Neutral lipids are preferred for
resolving dietary contributions in “end” predators, since the FA composi-
tion of this lipid class usually reflects trophic influences much better than PL
(e.g., Bell and Dick, 1990; Stubbs and Smith, 1990; Parrish et al., 1995).
However, if the objective of the study is to characterize the FA signature of
potential prey organisms. and as most predators consume their prey whole,
236 JOHANNE DALSGAARD ETAL.
A
Acetyl-CoA
B
Acetyl-CoA
r
16.0 b 16:l(n-7)
+ + Desaturation
18 0 - 18:l(n-9) 18: l(n-7)
+ + +
20:0-b ZO:l(n-1 I) 20: l(n-9) 20:l(n-7) Z-carbon
+ + chain
22:l(n-II) 22: l(n-9) el&gation
+ f
24:&1-l 1) 24: l(n-9)
(see below) are much less abundant except in bloom conditions (e.g., Parrish
et al., 1995; Cripps et al., 1999).
Phytoplankton are the major providers of metabolic energy in pelagic
food webs (Parsons, 1963) which is transferred to higher trophic levels via
grazing by herbivorous and omnivorous planktivorous species including the
larvae of fish and larger invertebrates. Similarly, phytoplankton support
benthic food webs that, in addition, may receive considerable inputs from
macroalgae (seaweeds), belonging to one of the three classes: Chlorophyceae
(green algae), Rhodophyceae (red algae) or Phaeophyceae (brown algae)
(Raven et al., 1992). While a few species of the brown algae Sargassum are
free-floating (Raven et al., 1992) the dominant life phase in most macro-
algal species is benthic, and because of limited light availability they are
restricted to the shallower coastal areas (Levring, 1979; Kristiansen et al.,
1981). Here, they constitute an important refuge for fish and invertebrates
and are either grazed directly or, as is mostly the case, enter the detrital food
webs via microheterotrophs (Dunstan et al., 1988; Sherr and Sheer, 2000
and references therein; Graeve et al., 2002 and references therein).
temperature, light and nutrient availability being the three key factors
affecting the FA pattern of the local community.
The impact of these environmental factors has been studied primarily in
laboratory cultures, and has been reviewed, e.g., by Pohl and Zurheide
(1979), Cobelas (1989), Kayama et al. (1989), and Roessler (1990). Typically,
lower water temperatures result in an increase in the level of unsaturation
(e.g., Ackman et al., 1968; Pohl and Zurheide, 1982), whereas the impact of
light is ambiguous and more species-specific. In general, however, the level of
glycolipids, and hence (n-3) PUFA, increases under nonlimiting light
conditions, whereas photo-inhibition and reduced light intensities reportedly
lead to the accumulation of TAG (the major lipid storage product in algae),
which is richer in saturated fatty acids (SFA) and MUFA (Cohen et al., 1988;
Harrison et al., 1990; Mayzaud et al., 1990 and references therein; Thompson
et al., 1990; Sukenik and Wahnon, 1991; Smith et al., 1993; Parrish et al.,
1994). Algal growth, as previously mentioned, is influenced by the
availability of limiting nutrients (principally nitrogen, phosphorus or
silicate), which influence the transition from the exponential phase (non-
nutrient limited) to the stationary growth phase (nutrient limited), the latter
being characterized by the accumulation of TAG (see above for consequences
on FA patterns; Kattner et al., 1983; Morris et al., 1983; Ben-Amotz et al.,
1985; Harrison et al., 1990; Kattner and Brockmann, 1990; Mayzaud et al.,
1990; Fahl and Kattner, 1993; Reitan et al., 1994; Falk-Petersen et al., 1998;
Henderson et al., 1998). During the exponential growth phase of
phytoplankton blooms, carbon fixed through photosynthesis is allocated to
growth and cell division rather than lipid storage (e.g., Morris, 1981; Kattner
et cl/., 1983; Parrish and Wangersky, 1990). As a consequence, the relative
proportion of glycolipids is particularly high in this phase (Sargent and
Henderson, 1986; Roessler, 1990), and the concentration of (n-3) PUFA may
approach 50% of the TL content (e.g., Napolitano et al., 1997; Claustre et al.,
1989; Sargent et d., 1989; Falk-Petersen et al., 1998; Henderson et al., 1998).
This exponential algal growth phase occurs during spring bloom conditions
and the FA pattern of the exponentially growing algae is particularly evident
in field examinations of phytoplankton lipid dynamics (e.g., Kattner et al..
1983; Hama, 1991).
2.4.1. Microalgae
(e.g., Ackman et al., 1968; Chuecas and Riley, 1969; Pohl and Zurheide, 1979;
Kattner et al., 1983; Sargent et al., 1987; Cobelas and Lechado, 1989;
Mayzaud et al., 1990; Mourente et al., 1990; Fahl and Kattner, 1993; Viso and
Marty, 1993; Napolitano, 1999; Volkman et al., 1998).
Since the early 1960s a large number of laboratory studies have examined
the FA composition of marine microalgae (reviewed by Ackman et al., 1968;
Pohl and Zurheide, 1979; Cobelas and Lechado, 1989; Kayama et al., 1989).
In these studies, the algae have been cultured under a wide range of
treatment conditions, and have been analyzed using standard, organic-
solvent extraction and methylation procedures combined with thin layer
chromatography (TLC) and gas chromatography (GC) later combined with
mass spectrometry (GC-MS) (Ackman, 2002; Traitler, 1987). Many of the
earliest studies were characterized by incomplete compound separation and
loss of PUFA due to improper sample handling and storage protocols.
Hence, the results from these studies should be interpreted with caution
(discussed by A&man et al., 1968; Chuecas and Riley, 1969; Conte et al.,
1994). Subsequently, techniques have improved (especially column technol-
ogy), resulting in a higher degree of sensitivity. As a consequence, more
precise estimates of total FA contents may be obtained, and in addition,
more FA have been identified. For example, trace amounts of the very-long-
chain, highly-unsaturated-fatty-acids (VLC-HUFA) 28:7(n-6) and 28:8(n-3)
have been identified in several species of dinoflagellates (Mansour et al.,
1999a, b). Intriguingly, octacosaheptaenoic acid (28:7(n-6)) and other VLC-
HUFA had previously been detected in Baltic herring where they were
suspected to originate from the diet (Link0 and Karinkanta, 1970).
However, except for a few examples like this, these more unusual FA
usually occur only in trace amounts in phytoplankton (e.g., Nichols et al.,
1986) and are even more difficult to recognize in the consumers due to the
low levels of occurrence, limiting their potential as trophic markers (see also
Section 5.7.2; Ackman and Mclachlan, 1977; Mayzaud et al., 1999).
Aside from sample treatment and identification procedures, another
obstacle associated with the application of FATM has been the interpreta-
tion of the large data sets routinely produced in these types of analyses
(typically arrays of more than 30 FA determined simultaneously from one
or more samples). With the development of computer power, easily
accessible, multivariate statistical methods have advanced to become
particularly useful for interpreting such large data sets (e.g., Wold et al.,
1988; Frank, 1989; Kaufmann, 1992; Smith et al., 1997, 1999; Legendre
and Legendre, 1998). Here, we present the results of such an analysis,
indicating the patterns of FA similarities within and among eight classes
of microalgae (Bacillariophyeae, Chlorophyceae, Cryptophyceae,
Dinophyceae, Eustigmatophyceae, Prymnesiophyceae, Prasinophyceae and
Raphidophyceae). The outcome of the analysis is visualized in Figure 5,
ATTY ACID TROPHIC MARKERS 243
PC2,18% . Chlorophyceae
A Cryptophyceae
. Bacillariophyceae
n Dinophyceae
q Eustigmatophyceae
P Prasinophyceae
o Prymnesiophyceae
A Raphidophyceae
, 1 PC1,37%
C22P FA
B
Prymnesjophg%
i
Cl6 4(“-3) Din phyceae
Cl&l(W9) Cl&b
Cl6 3(fl-3)
-0.5 I ClGPUFA
/ , PC1,25%
-0.5 0.0 0.5
Data from Dustan et 01. (1994), Mansour et al. (1999b), Mourente rt ul. (1990), Napolitano et a/. (1990), Nichols et ~1. (1987, 1991), Parrish rt ~1.
(1990, 1994). Servel er (Il. (1994), Visa and Marty (1993), Volkman et o/. (1981, 1989). Values are mean i one standard deviation.
246 JOHANNE DALSGAARD ETAL.
other algal classes (e.g., Volkman et al., 1989; Napolitano et al., 1990;
Parrish et al., 1990; Thompson et al., 1990; Dunstan et al., 1994). Trace
amounts have, however, been detected in a few species of chlorophytes
(Chuecas and Riley, 1969) dinophytes (Mansour et al., 1999b), prasino-
phytes (Chuecas and Riley, 1969) prymnesiophytes (Ackman et al., 1968;
Chuecas and Riley, 1969) and rhodophytes (macroalgae; Graeve et a/.,
2002).
Prymnesiophytes (except for two species of Hymenomonas) and dino-
phytes separate from the other classes by positive anomalies of 18:0, 18: 1
(n-9), 18:4(n-3) CZ2 PUFA and DHA. Another important FA of these
two classes, though not included in the analysis, is 18:5(n-3). This FA
was identified for the first time by Joseph (1975) in several species of
Dinophyceae. Later, it has been identified in species of prymnesiophytes
(Volkman et al., 1981, 1989; Sargent et al., 1985; Claustre et al., 1990;
Napolitano et al., 1990; Viso and Marty, 1993) raphidophytes (Nichols
et al., 1987; Viso and Marty, 1993) prasinophytes (Viso and Marty, 1993)
and bacillariophytes (Reitan et al., 1994).
Chlorophytes (except for one species of Nannochloris) are discriminated
by 18:3(n-3) 18:2(n-6) and other (n-6) PUFA. The close association of the
chlorophytes with prasinophytes in Figure 5A is consistent with both classes
belonging to the same division of Chlorophyta (Viso and Marty, 1993). A
characteristic FA of both these classes is 16:4(n-3) (not included in the
analysis; Ackman et al., 1968; Viso and Marty, 1993) whereas the presence
of > CZO FA in prasinophytes distinguishes them from the chlorophytes
(Viso and Marty, 1993).
Cryptophytes, raphidophytes and eustigmatophytes can be distinguished
as more or less separate groups. However, together with prasinophytes their
variations are poorly explained by the model, clustering around the center
on the loading weight plot (Figure 5B). It should be emphasized that limited
FA compositional data were available for these four classes of microalgae.
Hence, they were only represented by four observations each, which are
really too few for ensuring stability of the model (Albano et gl., 1981; Wold
et al., 1988).
This was taken into account in a second analysis, considering only
Bacillariophyceae, Dinophyceae, Prymnesiophyceae and Chlorophyceae,
which all contributed sufficient sample sizes and, as mentioned in Section
2.1, dominate the phytoplankton biomass in most marine ecosystems
(except for Chlorophyceae). In this analysis, the first four PLS components
now explain 83% of the variance of the FA compositional data and 76% of
the variance contributed by microalgal “class-affiliation”. A plot of the
second vs. third principal component (Figure 6A) reveals that apart from
two species of Hymenomonas, two large clusters of prymnesiophytes can be
recognized consisting predominantly of Isoch~~sis spp. (upper ellipse). and
ATTY ACID TROPHIC MARKERS 24
. Chlorophyceae \r
. Bacillariophyceae
0 Dinophyceae
o Prymnesiophyceae
I I
’ PCZ, 21%
-1 0 1 2
&MPUFA
(n-3)PUFA Dinophyceae
.
CI~PGRFA
‘UFA
1
c,8 3(n-3j Bacillarioph ceae Cm4(n3)
c14.0
I PCZ, 22%
-0.5 0.0 0.5
Phaeocystis spp. and Chrysotilu spp. (lower ellipse). Consistent with their
position in Figure 6A close to the Dinophyceae, species of Zsochrysis are
characterized by a low 16:1/16:0 ratio, and high concentrations of Crs
PUFA and EPA (Conte et al., 1994). The loading weight plot (Figure 6B)
shows that the prymnesiophytes in general separate from the dinophytes by
higher, positive anomalies of 14:0, 16: l(n-7) lS:l(n-9) and 18:4(n-3) while
the dinophytes link positively with DHA, CzlPUFA and (n-3) PUFA.
These analyses re-emphasize that individual FA cannot be used as
taxonomic indicators of particular algal species or classes, whereas
combinations of FA reveal certain patterns when microalgae are compared
class-wise. This conclusion confirms the statement of Viso and Marty (1993)
who identified the need to combine several FA criteria to distinguish natural
assemblages of microalgae belonging to different taxonomic classes. To
date, most of the criteria (ratios of FA) that have been developed have
focused on bacillariophytes and dinophytes, reflecting the relative impor-
tance of these two classes in the marine environment. Here, in particular,
high values of 16:l(n-7)/16:0 (typically > 1) and CCr6/CC~s have been
associated with a dominance of bacillariophytes (e.g., Miyazaki, 1983:
Claustre et al., 1988, 1989; Mayzaud et al., 1990; Viso and Marty, 1993;
Budge and Parrish, 1998; Budge et al., 2001; Reuss and Poulsen, 2002)
whereas high values of 18:5(n-3)/18:3(n-3) and C (CrsPUFA, &PUFA)
have been associated with a dominance of dinophytes (Nichols et al.,
1984; Viso and Marty, 1993). Combining these criteria, i.e., high values of
CCr6/CC1s together with low values of 18:5(n-3)/18:3(n-3) has been
proposed as a means whereby bacillariophytes can be distinguished from
dinophytes (Viso and Marty, 1993). This could be further strengthened
by examining the ratio of 22:6(n-3)/20:5(n-3) as suggested by Budge
and Parrish (1998). Here, a value >l signals a dominance in the
contribution of dinophytes while conversely, a value < 1 is suggestive of a
greater contribution of bacillariophytes.
A quantitative summary of the PLS analyses and the criteria discussed
above is presented in Table 2 with mean values for the eight classes of
microalgae included in the analyses. It must be recognised that the figures in
the table should be perceived only as a very rough guideline of potentially
useful FATM.
2.4.2. Macroalgae
Boxed entries indicate that the particular FA criteria may be a useful tracer of the algal class.Based on data compiled from the literature: see Table 1
for references. “16:l(n-7) + 16:4(11-l) + EPA: b18:5(n-3) + DHA: ‘18:l(n-9)+ 18:4(n-3); d16:4(n-3) + 18:2(n-6)+ 18:3(n-3); VZould not be determined
(dividing by zero).
250 JOHANNE DALSGAARD ETAL.
Honya et al., 1994). Some general trends distinguishing the three classes
have been recognized, which are largely independent of geographical
locations and morphological differences (e.g. Ackman and McLachlan, 1977
(Nova Scotia); Chuecas and Riley, 1966 (Isle of Man); Dembritsky et al.,
1993 (Caspian Sea); Fleurence et al., 1994 (French Brittany coast); Graeve
et al., 2002 (Arctic and Antarctic); Khotimchenko et al., 2002 (Pacific
coast); Li et al., 2002 (Bohai Sea); Pohl and Zurheide, 1982 (Baltic Sea)). Cis
and C2e FA constitute the principal PUFA in macroalgae whereas Cz2
PUFA are more abundant in microalgae (Chuecas and Riley, 1966; Graeve
et al., 2002). Moreover, the (n-6) family (particularly AA) is more prevalent
in macroalgae than in microalgae. High concentrations of AA combined
with insignificant amounts of C is PUFA distinguish rhodophytes from
phaeophytes in which both C is PUFA (particularly 18:4(n-3)) and C2,,
PUFA (principally EPA and AA) are major FA constituents. The FA
pattern of chlorophyte macroalgae is similar to that of chlorophyte
microalgae. Overall, the chlorophytes differ from the rest of the eukaryotic
algae by a FA composition more similar to that of higher plants (Pohl and
Zurheide, 1979; Wood, 1988; Kayama et al., 1989). For example, few of the
constituent FA have more than three double bonds and most species have
only modest amounts of > C20 PUFA, while the proportion of Ci6 and Cis
PUFA is generally high (see Table 1; Wood, 1988; Lechevalier and
Lechevalier, 1988; Volkman et al., 1998; Graeve et al., 2002). A particular
trait of chlorophytes is the high content of 18:3(n-3) regarded as a
characteristic of the phylum Chlorophyta (Li et al., 2002). Furthermore,
several species exhibit a 18:l(n-7)/18:l(n-9) ratio > 1 (Khotimchenko et al.,
2002; Li et al., 2002) which in combination with 18:2(n-3) and 18:3(n-3)
may potentially serve as a biomarker of this algal class.
2.5.1. Bacteria
following major plankton blooms (e.g., Morris, 1984; Mayzaud et al., 1989;
Skerratt et al., 1995; Najdek et al., 2002). As a consequence, the FA
composition of marine bacteria has been studied predominantly by
geochemists seeking to resolve the source and diagenetic state of
POM and of sediments (e.g., Brooks et al., 1976; Haddad et al., 1992;
Harvey, 1994; Colombo et al., 1997; Harvey and Macko, 1997; Volkman
et al., 1980a; Wakeham and Beier, 1991). However, as mentioned in
Section 1.1, heterotrophic bacteria are also very important in areas
dominated by the microbial loop, where they occupy a critical position,
recycling DOM and POM to higher trophic levels (Sherr and Sheer, 2000
and references therein). Unfortunately, very few studies have examined
the FA dynamics of these systems (e.g., Claustre et al., 1988; Ederington
et al., 1995).
Bacteria do not store TAG but incorporate FA chiefly into PL (Fulco,
1983; DeLong and Yayanos, 1986; Parkes, 1987). Fatty acids commonly
biosynthesized by bacteria are within the range C10-C20 and are dominated
by SFA and MUFA, whereas PUFA, with a few exceptions including deep-
sea bacteria and some bacterial strains isolated from fish intestines, are
rarely detected (e.g., Johns and Perry, 1977; DeLong and Yayanos, 1986;
Yazawa et al., 1988; Pond et al., 1997a, 2002; Nichols and McMeekin,
2002). Bacteria, moreover, differ from eukaryotes in biosynthesizing large
amounts of odd-numbered, branched tvans-unsaturated and cyclopropyl
FA such as 15:0, 17:0, 15:1, 17:1, iso and ante&-branched SFAand MUFA,
IO-methylpalmitic acid, trans-16:l(n-7) cy17:O and cy19:O (Perry et al.,
1979; Volkman et al., 1980a; Gillan et al., 1981; Parkes, 1987; Vestal and
White, 1989 and references therein; Rajendran et al., 1994). In the same way
as for microalgae, several combinations of FA have been used to detect the
presence of bacteria (summarized in Table 3).
Bacteria also biosynthesize large amounts of more common FA including
16:l(n-7) and 18:l(n-7) (e.g., Volkman and Johns, 1977 and references
therein; Perry et al., 1979; Gillan et al., 1981; Parkes, 1987 and references
therein; Vestal and White, 1989; Volkman et al., 1998). These particular FA
are, however, also biosynthesized by and used as markers of eukaryotic
organisms, principally diatoms and their entrainment into food webs.
Therefore, unless elevated levels of some of the more specific
bacterial marker FA summarized above are detected, and if PUFA are
present in large amounts, it is in most cases presumably safe to assume that
16:l(n-7) and lS:l(n-7) derive from eukaryotic rather than bacterial
production.
The only controlled laboratory experiment so far to demonstrate the
transfer of bacteria (and diatom) FA markers to higher trophic levels was
carried out by Ederington et al. (1995). In this experiment, cultures of either
bacterivorous ciliates or diatoms were fed to Acartia tonsa for 96 hours and
FATTY ACID TROPHIC MARKERS
253
FATM Reference
Bacterial markers
C Odd carbon numbered Budge and Parrish
+ branched chain FA (1998), Budge et al. (2001)
C Iso- and anteiso-Ct5 Viso and Marty (1993)
and Ct7
18:l(n-7)/18:l(n-9) Volkman et al. (1980b)
Iso+anteiso 15:0/16:0 Mancuso et al. (1990)
Zso+anteiso 15:0/15:0 White er al. (1980)
C 15:0, iso- and anteiso-Cl5 Najdek et al. (2002)
and Ct7, lS:l(n-7)
brCt5/15:0” Najdek ef al. (2002)
Terrestrial markers
18:2(n-6) Napolitano et al. (1997)
18:2(n-6)+ 18:3(n-3) > 2.5 Budge and Parrish
(1998), Budge et al. (2001)
22:0 + 24:0 Budge et al. (2001)
c c24:0-c32.0 Meziane et al. (1997)
are very prevalent and often constitute the major prey of squids, fish, marine
mammals and seabirds (Pond et al., 1993; Virtue et al., 1993; Hagen et al.,
2001; Saito et al., 2002). In contrast, very little information is available on
FA of cyclopoid and poeicilostomatoid “microcopepods”, which usually
dominate in terms of copepod abundance but not in terms of biomass
(Paffenhofer, 1993; Metz, 1998; Bottger-Schnack et al., 2001). This holds
true also for other invertebrate groups of noncommercial interest such as,
e.g., pteropods and amphipods, which nevertheless are essential members
of marine food webs (Joseph, 1989; Kattner et al. 1998; Hagen and
Auel, 2001). However, there is a large body of literature on the general
distribution and composition of lipids in marine invertebrates, and a
comprehensive compilation was provided by Joseph (1982, 1989).
In this next section, we deal predominantly with the dynamics of FA in
calanoid copepods for which most information is available. A discussion
of fatty alcohols is also included, since these are the constituents of WE
accumulated in large amounts by some of the species. Some fatty alcohols
are unique to certain copepods, and therefore, of potential biomarker value.
The lipid and FA dynamics of other zooplankton groups are mentioned
where pertinent, but otherwise confined to Section 5, where they are
discussed in conjunction with major food webs.
Given their central position within the food web, a key aspect of FA
dynamics in copepods and other zooplankton is whether they modify
dietary FA, and if so, to what extent do these modifications take place, and
how might this interfere with the interpretation of FATM? On the basis of
controlled laboratory experiments (Section 3.4) it is generally accepted that
phytoplankton FATM are incorporated largely unaltered by phytopha-
geous species, allowing conclusions to be drawn on the major type of food
ingested. Herbivorous calanoid copepods from higher latitudes are classical
examples of this. They typically accumulate large lipid reserves as an
adaptation to the pronounced seasonality and strongly pulsed supply of
food in these regions (Lee et al., 1971a; Lee and Hirota, 1973). The lipid
reserves consist predominantly of WE, and may contain considerable
amounts of specific FA such as 16:l(n-7), 18:4(n-3) and EPA, presumably
incorporated directly from the consumption of microalgae (e.g., Sargent and
Henderson, 1986; Graeve et al., 1994a). Moreover, calanoid copepods are so
far the only known organisms that biosynthesize de ylovo considerable
FATTY ACID TROPHIC MARKERS 257
enter the biosynthetic pathway (Section 1.5) and are modified to longer-
chain SFA and MUFA (Sargent and Henderson, 1986). The entrainment
of such short-chain SFA probably varies with the dietary regime, e.g.,
throughout a phytoplankton bloom. This may account for some of the
variation observed within the WE fatty alcohol composition of a given
species as well as the differences observed between different developmental
stages of copepods (Sargent and Falk-Petersen, 1988; Tande and
Henderson, 1988). However, dietary 16:l(n-7), which is used as a specific
diatom tracer, does probably not enter this internal biosynthetic pathway as
it may only be elongated to longer-chain (n-7) isomers (Figure 4B), which
are generally not detected in large amounts in calanoid copepods (Sargent
and Falk-Petersen, 198 1, 1988).
The reduction of SFA and MUFA to fatty alcohols is presumably
mediated by a NADPH-fatty acyl coenzyme A oxidoreductase specific
to WE producing animals, and once formed they may subsequently be
esterified to dietary FA by a nonspecific ester synthetase (reviewed by
Sargent and Henderson, 1986). Through these processes, dietary carbohy-
drates, proteins and FA may effectively be converted to WE even in periods
of high intakes of dietary FA. In contrast, this situation usually causes a
feedback inhibition of FA biosynthesis in other organisms such as fish (e.g.,
Sargent et al., 1989; Section 4.3). Hence, the possession of this specific
biosynthetic pathway is presumably largely restricted to higher latitude
herbivorous species. These species have both to accumulate enough energy
reserves during the short feeding season to survive the prolonged periods of
starvation, and to fuel reproductive processes starting prior to the onset of
phytoplankton spring blooms (Sargent and Falk-Petersen, 1988; Hagen and
Schnack-Schiel, 1996). Altogether, these processes sustain the hypothesis
that the FA component of WE in herbivorous calanoid copepods is largely
derived from the diet (i.e., phytoplankton), whereas the fatty alcohols are
derived from the animal’s internal biosynthesis (Sargent and Henderson,
1986).
The conservative incorporation of dietary FA into WE has been
established through controlled laboratory experiments (Section 3.4) even
though it has also been demonstrated that herbivorous marine invertebrates
can modify dietary 18:3(n-3) to EPA and DHA at very slow rates (e.g.,
Moreno et al., 1979; Sargent and Whittle, 1981 and references therein). As
the natural diet of herbivorous copepods is typically rich in EPA and DHA
and relatively poor in Cis PUFA (e.g., Scott et al., 2002) they presumably
do not need to undertake these modifications to sustain their growth
requirements (Sargent and Henderson, 1986). In all circumstances, FATM
are most “applicable” to herbivorous copepods sampled in mid- or late-
summer (Sargent and Henderson, 1986) when they are actively accumulat-
ing lipid reserves, whereas specimens sampled from mid-winter and onwards
FATTY ACID TROPHIC MARKERS 259
0 Calanoides acutus
A Calanus propinquus
Calanus finmarchicus
v Calanus glacialis
P
. a Calanus hyperboreus
‘.a o Euchaeta sp.
l .
. Metridia geriachei
v Rhincalaks gigas
-6 -
I I PCl, 22%
-6 6
0.5 PC3,13%
C. hyperboreus
0.c
-O.! ii
I 1 PCl,lS%
-0.5 0.0 0.5
Fatty acid,
14:o 4.4 l 1.3 3.6 + 0.7 16.9 ct 5.1 9.8 k 4.0 3.7 5 0.5 1.6 xt 0.4 4.4 zt 0.6 0.7 zt 0.1
190 0.2 * 0.3 0.6 * 0.6 0.7 f 0.4 0 0 0.9 i 1.3 0.6 f 0.4 0
16:0 4.5 i 2.1 13.0 f 1.4 12.7 i 2.4 6.9 z!z 1.2 4.3 f 0.8 2.4 f 2.2 12.3 zt 2.1 3.3 k 0.8
16:l(n-7) 7.7 f 2.7 4.3 i 1.2 6.2 + 2.0 25.2 f 6.3 10.6 zt 4.0 20.3 5 4.6 5.6 + 2.9 11.6 f 1.9
16: 1(n-5) 0.2 f 0.2 0.2 f 0.1 0.4 f 0.3 0.7 f 0.3 0 0.3 f 0.5 0.2 f 0.1 0.0 f 0.1
16:2(n-6) 0.6 f 0.1 0.4 f 0.3 0.9 f 0.3 1.0 f 0.2 1.8 IO.6 0.9 f 0.1 1.3 f 0.7 2.7 f 0.5 b
16:3(n-3) 0.2 * 0.3 0.1 + 0.2 0.3 i 0.3 0.9 + 0.4 0.5 f 0.7 0.4 5 0.3 0.2 It 0.2 0.7 i 0.3 g
16:4(n-3) 0.6 f 1.3 0.2 * 0.2 0.0 f 0.1 2.ozk1.2 0 0.1 f 0.2 1.0 f 0.9 3.4zt1.2 ;
18:0 0.1 f 0.2 1.3 f 0.1 1.5 f 0.8 0.4 f 0.3 0.4 f 0.2 0.4 f 0.3 12.8
1.4 f 0.6
3.1 18.2
0.3 f 0.5
1.7 p
18:l(n-9) 4.8 f 1.0 2.9 f 0.6 5.3 * 1.2 3.7 zt 0.8 3.2 k 0.7 37.9 f 12.4 z;
18:l(n-7) 1.5 f 0.4 1.1 f 0.3 0.4 f 0.9 1.0 i 0.2 0.9 f 0.4 1.3 * 0.4 3.4 It 2.0 3.3 It 0.2 0
18:2(n-6) 1.6 i 0.5 1.2 f 0.4 1.8 i 0.6 0.9 f 0.2 1.7 It 0.7 1.5 f 0.3 1.7 rt 0.2 1.7hO.3 z
18:3(n-3) 0.5 f 0.3 0.6 z!z0.2 1.1 f 0.4 0.5 f 0.4 0.7 f 0.4 0.5 f 0.1 0.8 f 0.3 0.9 f 0.2 ?I
18:4(n-3) 4.6 f 5.2 2.8 f 1.7 9.5 f 6.5 3.2 f 2.4 10.3 * 7.3 2.8 f 1.2 5.1 f 1.9 14.6 f 4.2 q
br-
20: 1(n-9) 23.1 f 6.2 2.7 i 0.5 7.7 f 3.8 12.3 xk 3.4 19.8 i 3.2 2.2 It 1.1 1.3 * 0.2 0.6 f 0.3 $
20:l(n-7) 0.8 f 0.2 0.6 i 0.1 1.0 f 0.5 1.0 f 0.1 1.9 f 0.9 0.1 f 0.1 0.1 f 0.1 0.0 f 0.1 3
20:4(n-6) 1.4 i 0.8 0.9 f 0.4 0 0.2 * 0.3 0 1.9 i 1.2 0.9 i 0.3 0.6 f 0.6
20:5(n-3) 17.1 f 4.7 12.4 + 4.6 13.2 f 5.8 16.0 f 7.2 14.1 f 4.5 10.5 f 5.8 20.9 * 2.9 27.4 f 2.1 bo
22:l(n-11) 9.8 zt 2.4 20.1 f 6.4 8.0 * 4.1 7.1 f 1.7 15.0 f 2.5 0.4 f 0.6 0.7 * 0.9 0.8 + 1.9
22:l(n-9) 3.7 f 0.8 19.2 i 6.5 0.3 f 0.3 1.1 i 0.3 3.5 i 1.6 0.4 i 0.3 0.4 i 1.1 0.1 f 0.2 ;:0
22:5(n-3) 0.8 f 0.6 0.8 i 0.1 0.3 f 0.3 0.6 f 0.7 1.0 * 1.3 0.3 f 0.3 0.9 f 0.2 0.3 It 0.2 g
22:6(n-3) 11.8 i 4.5 10.9 f 5.6 11.6 i 6.3 5.2 i 1.5 7.8 xk 1.7 12.8 i 6.2 24.1 zk 3.8 15.5 * 17.5
Alcohols E
14:o 6.2 i 1.7 0 1.7 i 0.7 3.2 + 1.3 2.8 f 1.5 58.9 f 5.7 50.0 It 3.5 45.6 i 2.0 ,^
16:O 8.1 f 2.8 0 9.6 f 4.3 11.2 i 2.6 6.1 i 2.5 37.3 f 5.4 48.1 f 4.1 48.1 f 2.4 g
16:l(n-7) 2.4 zt 1.3 0 3.2 i 2.5 7.1 f 3.1 3.6 f 1.7 3.8 zt 1.2 1.9 f 2.4 4.3 f 0.9
18:O 0 0 1.7 f 1.9 0 0.4 f 0.4 0 0 0
lS:l(n-9) 1.2*0.6 0 2.6 5 1.3 2.1 f 0.5 0.5 f 0.6 0 0 1.9 f 0.3
20: 1(n-9) 55.0 * 4.8 0 36.6 * 4.3 43.4 It 5.9 32.6 zt 3.9 0 0 0
22:l(n-11) 27.2 f 4.8 0 44.6 f 6.2 30.4 f 4.7 55.0 f 7.2 0 0 0
“Based on unpublished data compiled from field trips to the Arctic and Antarctic. Values are mean * one standard deviation
264 JOHANNE DALSGAARD ETAL.
0
I II III IV V Adults
Copepodid stages
Euphausia superba
16
0
Calyptopis Furcilia lmmatures Adults
Euphausia ctystallorophias
25
m 16:l(n-7)
0 18:4(n-3)
cza 20:5(n-3)
m 22:6(n-3)
30
20
10
0
0 10 20 30 40 50 0 10 20 30 40 50
Days Days
10
0
0 10 20 30 0 10 20 30
Days Days
Marine fish use lipids as a chief metabolic energy source (Shul’man, 1960)
fulfilling their energetic requirements primarily through the oxidation of
cellular lipids and proteins rather than carbohydrates (Cowey and Sargent,
1977; Jobling, 1994; Sargent et al., 1993). TAG is the primary mode of lipid
storage in most species whereas WE are usually much less important
(Shul’man, 1960; Love, 1970; Owen et al., 1972; Ackman, 1980; Navarro
and Gutiirrez, 1995; Sargent and Henderson, 1995). Many meso- and
270 JOHANNE DALSGAARD HAL.
Like most other organisms, fish can readily biosynthesize SFA with up to 18
carbon atoms de nova (Ackman, 1980; Henderson and Sargent, 1985) and
desaturate them into monounsaturates following the common lipid pathway
(Section 1.5). However, in contrast to calanoid copepods discussed in
Section 3.2.1, a dietary excess of FA ( > 10%) apparently suppresses de now
272 JOHANNE DALSGAARD EJAL.
4.4.1. Sturvution
4.4.2. Reproduction
and DHA were preferentially deposited in the ovaries, whereas 16:0, 18:0,
18: l(n-7), 18: l(n-9). 22: l(n-1 l), 22:5(n-3) and EPA were catabolized after
mobilization.
These results emphasize that the FA patterns of fish depleting their lipid
reserves are highly distorted, reflecting internal metabolic processes rather
than potential dietary signals. In fish roe, EPA and DHA typically
constitute 50% of the TL, suggesting an essential need of the developing
embryo for the formation of cellular membranes. Interestingly, the FA
composition of fish roe is remarkably similar among species and presumably
optimized nutritionally for the growth of the developing embryo and yolk-
sac larvae until first-feeding (Kaitaranta and Linko, 1984; Tocher and
Sargent, 1984; Klungsoyr et ul., 1989). The dietary FA composition of the
parent fish typically has little impact on the FA composition of the eggs.
However, when comparing the roe of Atlantic and Baltic herring
(Kaitaranta and Linko, 1984) relatively large proportions of 20:l(n-9)
and 22: 1(n- 11) (i.e., 3.1% and 1.5% of total FA, respectively) were detected
in the Atlantic herring roe, whereas these FA were absent in Baltic herring
eggs. Calanoid copepods are much less common in the Baltic Sea compared
to the Atlantic, presumably because the lower salinity in this system
(Ackman, 1980) and this probably explains the absence of these tracers in
Baltic herring roe. In another example, Lasker and Theilacker (1962) found
a relatively close similarity between the FA composition of the ovary of
Pacific sardine (Sardinops caerdenj and the diet of the adult fish, consisting
mostly of Culanus. However, apart from a few such exceptions, it may be
anticipated that FA add a limited amount of information useful for
resolving the trophodynamic processes resulting ultimately in the
production of offspring.
as an intermediary, larval North Sea cod were reared on food webs based
on monocultures of either the diatom Skeletonema costatum or the
dinoflagellate Heterocapsa triquetra, i.e., algae dominating in the mixed
and stratified regions of the North Sea, respectively. The cod larvae
required 8 days on either food type before the tracer lipid signals started to
change from their original values to those similar to the algae at the base
of their respective food webs (Figure 2). After 13 days, the lipid tracer
content in the larvae was no longer significantly different from that of the
cultures of Skeletonema costatum or Heterocapsa triquetra. Subsequently, a
sub-sample of 100 juvenile cod from stratified, mixed and frontal regimes
in the northeastern North Sea was examined for the content of FA tracers
and condition (as determined by the ratio of total lipid content to total
length). Juvenile cod displaying a lipid tracer content indicating utilization
of a diatom-based food web (found in proximity to regions of frontal
mixing) were in significantly better condition (P > 0.05) than those
containing a lipid signal indicative of a flagellate-based food web (found
in stratified regions of the North Sea; Figure 11).
In another laboratory study, Kirsch et al. (1998) examined how the FA
signature of whole adult Atlantic cod changed when offered first a prepared
diet of low-fat squid (Zllex illecebrosus, 2% lipid DM) for six weeks,
followed by a prepared diet of high-fat Atlantic mackerel (Scomber
scombrus, 16% lipid DM) for another eight weeks. Intriguingly, after only 3
weeks on the squid diet, and despite the absence of any mass gain, the FA
composition of the cod had changed significantly toward that of the squid,
changing no further after 6 weeks of feeding. When switched to the mackerel
diet, the overall tissue lipid content of the cod increased from 2% to 4%.
Furthermore, the FA patterns had reversed toward that of the mackerel diet
within 5 weeks of first feeding, with no further changes during the last three
weeks. Applying a classification and regression tree analysis (CART) to the
FA compositional data, the authors showed that the cod treatment groups,
despite the influence of dietary FA, were still readily differentiated from
each other and from their diet.
The results of these two studies demonstrate the relevance of dietary FA
as qualitative markers for resolving trophic interactions in both larval and
adult fish. Moreover, the latter study supports the application of FATM
for assessing the diet of yet higher trophic level predators such as marine
mammals (e.g., Iverson et ul., 1997b).
A series of enclosure studies have been carried out in Loch Ewe, Scotland,
demonstrating the impact of ontogeny and varying dietary regimes on the
FA composition of herring larvae (Clupea harengus). In the first study,
Gatten et ul. (1983) observed that a switch in the diet of herring larvae from
microalgae and nauplii (as determined from gut analyses) to WE rich stages
of copepodites and adult calanoid copepods, was accompanied by a gradual
replacement of typical dinoflagellate and flagellate FATM (18:4(n-3) EPA,
DHA) by calanoid FATM. Considering the condition of the herring
larvae, Fraser et al. (1987) later found that a dietary resemblance was much
more pronounced in well-nourished larvae, which were accumulating
TAG, than in under-nourished larvae. Finally, using 18:4(n-3) as a specific
flagellate tracer, Fraser et ul. (1989) were able to follow a natural succes-
sion in the enclosed microalgal community from dinoflagellates and
flagellates to diatoms, and furthermore, could detect the signal, presumably
through zooplankton, to herring larvae (Figure 12). However, whereas the
zooplankton community closely mirrored the temporal development in the
phytoplankton, the peak in the tracer content was delayed by 23 days in
herring larvae. This delay suggests that the fish larvae either continued
feeding selectively on dinoflagellates and flagellates rather than on diatoms
or zooplankton, or that the turnover rates of the tissue lipid pools decreased
as the larvae grew (see also Section 5.2.5). The authors did not, however,
discuss this.
Apart from the studies summarized above, several studies of natural
fish populations have been carried out, comparing the FA composition of fish
and their potential prey, and assuming simply a conservative transfer of FA
from prey to predators. These studies will be summarized in Section 5.
278 JOHANNE DALSGAARD ETAL.
4
0 10 20 30 40 50 60
TIME (DAYS)
1996) and hence, the basic FATM patterns recognized in higher trophic
levels.
Seasonal patterns of phytoplankton group dominance, driven by
stratification, are most pronounced in high latitude and temperate
systems, and are used here as an example to outline the general
processes, conceptualized in Figure 13. First, as light intensity increases
in early spring, the phytoplankton community is dominated by small
flagellates, typically Phaeocystis spp., with blooms occurring in some
situations. Accompanying such blooms are typical FATM (Section 2.4.1)
available for transfer to higher trophic levels. With the onset of
stratification, the spring diatom bloom is initiated and flagellate FATM
are largely replaced by diatom FATM. Continued and increased
stratification results in a period of nutrient limitation. As a consequence,
the phytoplankton community becomes dominated by flagellates, dino-
flagellates and microbial loop production again with a characteristic
FATM distribution.
Variations in the content of these different group specific FATM in higher
trophic levels during the succession of phytoplankton dominance are
indicative of the importance of the various algal groups for the transfer of
energy up the food webs.
The importance of the different temporal components of this evolution of
phytoplankton dominance, and hence FATM, varies dramatically between
geographic regions (e.g., polar, temperate regions and tropics), and is in
essence based on the dynamics of water column stratification as indicated in
Figure 13A.
A comparison of the dynamics of FATM in these different systems has
not been made. However, based on the processes outlined above, a
continuum of the importance of diatom versus flagellate, microbial loop and
dinoflagellate production to higher trophic levels (dependent upon transfer
efficiencies), coupled to the relative contribution of these different groups
to the total phytoplankton biomass of the system, might be expected
(Figure 13B). For example, in boreal and temperate systems the spring
diatom bloom contributes a higher proportion to the overall phytoplankton
biomass than in tropical systems. The reason for this is that tropical systems
are generally stratified and dominated by flagellate phytoplankton and
microbial loop production. The latter comprises also cyanobacteria,
however, these are more difficult to categorize. They are N-fixers and
may act like diatom blooms, but as they are not necessarily driven by
stability, they are not included in Figure 13.
Phytoplankton group dominance is also influenced by mesoscale features
such as coastal upwelling and tidal mixing processes, which impact on water
column stratification and nutrient availability. These systems in essence
create localized “spring bloom” conditions for phytoplankton communities,
280 JOHANNE DALSGAARD ETAL.
None
I I
4 NoneL 1 Low
joi%
+
Jan May Sep Dee
LL +- Period of stratification
phytoplankton
group dominance and FATM over a seasonal cycle in (A) polar, temperate and
Long
and are also dominated by diatom production (e.g., St. John and Lund,
1996). The dynamics of phytoplankton group production in upwelling
systems is well understood, but the dynamics of FATM has not received
very much attention. On the other hand. in tidal mixing regions the
FATTY ACID TROPHIC MARKERS 281
Light, nutrients and stratification are the major driving forces in the Arctic,
controlling the short but intensive period of primary production with
60-70% of the total annual primary production taking place between mid-
March and early July (Falk-Petersen et al., 1990 and references therein). The
pelagic spring bloom is initiated in fjords (where fresh water run-offs result
in early stratification), followed by blooms in the open water of the marginal
ice zone (MIZ) (Falk-Petersen et al., 1998 and references therein). Ice algae
consist predominantly of diatoms, whereas open water phytoplankton
communities are relatively richer in dinoflagellates and smaller flagellates
(Falk-Petersen et al., 1998; Henderson et al., 1998). In particular,
Phaeocystis spp. often dominate at the onset of the open water spring
bloom (Sargent et al., 1985; Falk-Petersen et al., 1990, 2000 and references
therein; Marchant and Thomsen, 1994; Hamm et al., 2001). The different
phytoplankton communities are accompanied by typical FA signatures
reflecting the dominant algal classes (Section 2.4.1). A notable exception is
Phaeocystis pouchetii in Balsfjord (Sargent et al., 1985; Hamm et al., 2001),
which contained a FA pattern quite different from that observed in other
areas (Section 2.4.3) i.e., high proportions of 18:4(n-3) 18:5(n-3) EPA and
DHA combined with relatively low levels of CL6 PUFA.
The FA signature of size-fractionated plankton samples collected during
the spring and post-plankton bloom off the west coast of Greenland was
recently combined with detailed microscopic analyses of biomass and
species level composition of microalgae (Reuss and Poulsen, 2002). This
study revealed that most of the spring bloom biomass was contained within
the 1 l-300 urn size-fraction and was dominated by diatoms, while 80% of
the biomass in the 6- 11 urn size-fraction was composed of Phaeocystis
pouchetii. The spring plankton bloom was succeeded by flagellates
(Haptophyceae; < 11 urn) with the total biomass of FA being an order of
magnitude lower and significantly different (r = 0.95, P < 0.001) from the
spring bloom. On this basis, specific FATM were coupled with the
phytoplankton species composition. The biomass of diatoms correlated
significantly and positively with 16: 1(n-7)/ 16:0, Xi6/CCi8, 16: 1(n-7) and
282 JOHANNE DALSGAARD ETAL.
140 6
_ 120 1.2
5
Y
2 100 1.0 .o
H 4 .o
$u) 60 0.6 0 H
rr z 3:
0.6 5
c G
ij 6o F 20'
0.4 $
5 4o
1
E 20 0.2
0 0.0 0
2 4 6 6 10 13 14 19 24 26
station
26
0.0 L0
station
EPA and negatively with CCis FA and 18: l(n-9). The temporal develop-
ment in the diatom FATM composition of the particulate matter is shown in
Figure 14.
In contrast, the typical dinoflagellate FATM 18:4(n-3) and DHA did not
correlate with the biomass of either flagellates or dinoflagellates. The
authors emphasized that dinoflagellates are a complex group of organisms
comprising auto-, hetero- and mixotrophs that contain chloroplasts of
diverse endosymbiotic origin. This may explain some of the variation in
specific FATM observed within this group (Table 2) and based on the
FATTY ACID TROPHIC MARKERS
283
results in the study, the authors deduced that X1s FA provide a better
indicator of flagellate contribution than 18:4(n-3).
5.2.2. Copepods
-- A Calanus hyperboreus
I
m 16:l(n-7)
II) I 16:4(d)
z
A B C D
Sampling site
B Calanoides acutus
m 16:l(n-7)
I 16:4(n-3)
IV
Sampling site
Figure I.5 Spatial variation of the dietary FATM 16: l(n - 7) and 18:4(n - 3) in
(A) Calanus hyperboreus (CV stages) collected in the Fram Strait (4”W to 4”E, 78”N
to 80”N), July 1984 and (B) Culanoides acutus (CV stages) collected in the south
eastern Weddell Sea (site I + II: 36”W to 42”W, 77”3O’S to 78’S; site III-V: 18”W to
21”W, 72”s to 73”3O’S), January-February 1985. Redrawn with permission after
Kattner and Hagen (1995).
80
-I%+ l&4(&3), FA
-0- 16:1+20:6(n-3), FA
+ 20:1+22:1, FA Ale
80
0 50
Feb-Mar Jun-Aug Sep-Ott NovJan
Metridia longa and M. okhotensis may, however, deviate from the general
fatty alcohol pattern summarized above, and their WE contain substantial
amounts of CZO and CZ2 monounsaturates. It is not clear whether these
long-chain monounsaturates are biosynthesized de novo or derived from
feeding on calanoid copepods (Falk-Petersen et al., 1987; Albers et al.,
1996). Assuming that the latter is true, Figure 17B indicates an uptake of
calanoid copepods in late winter by M. longa. Similarly, the WE fatty acid
composition of another carnivorous Arctic copepod, Pareuchaeta norvegica,
indicated that this species also feeds on calanoid copepods (Sargent and
McIntosh, 1974). Interestingly, the WE of the Antarctic congeners, Metridiu
gerluchei (Graeve et al., 1994b) and Euchaeta antarctica (Hagen et al., 1995)
were characterized by the near absence of calanoid FATM. E. antarctica
has been observed to prey on Culunus acutus (Oresland, 1991), and Metridia
gerlachei is believed to show similar feeding behavior. Hence, it is not
fully understood why long-chain monounsaturates are apparently entirely
catabolized in these Antarctic species (Hagen et al., 1995), while retained
in the Arctic congeners (Auel, 1999). However, it obviously weakens
the potential of these long-chain monounsaturates as calanoid FATM,
particularly in the southern hemisphere, due to the uncertainty of their
dietary and biosynthetic origin.
286 JOHANNE DALSGAARD ETAL.
Figure 17 Seasonal development in specific WE (A) fatty acids, and (B) fatty
alcohols in Metridia longa collected in Balsfjord, northern Norway. Based on data
from Falk-Petersen et al. (1988) cited and reproduced in Sargent and Falk-Petersen
(1988).
5.2.3. Euphausiids
High levels of 16:l(n-7) Cis PUFA and EPA have been detected in the two
major Arctic euphausiids, Thysanoessa inerrnis and T. raschii (Ackman
et al., 1970; Sargent and Falk-Petersen, 1981; Saether et al., 1986; Falk-
Petersen et al., 2000; Hamm et al., 2001) indicating that these species feed as
herbivores during the Arctic summer. Substantiating this hypothesis, the
ingestion of Phaeocystis pouchetii by Thysanoessa spp. has been verified
both in the field (Balsfjord) and in the laboratory (T. raschii; Hamm et al.,
2001).
In addition, analyses of Thysanoessa inermis sampled in autumn in
Balsfjord and Ullsfjord revealed increasing proportions of calanoid FATM
suggesting a switch in diet to include copepods in the Arctic dark period
(Falk-Petersen et al., 2000). However, whereas a low 18:l(n-7)/18:l(n-9)
ratio in T. inermis from Kongsfjord, Svalbard also suggested an animal
FATTY ACID TROPHIC MARKERS 287
dominant lipid class in all three species. These observations conform to the
hypothesis that TAG represents the recent feeding history of animals
whereas WE integrate over a longer period of time (Hakanson, 1984;
Sargent and Henderson, 1986). Alternatively, Bolinopsis injiindibulum is not
a “true” intermediate link between Calanus glacialis and Beroe cucumis. This
hypothesis is proposed based on a closer examination of the data in the
paper, revealing a quite similar WE fatty acid composition of Calanus
glacialis and Beroe cucumis, i.e., suggesting that Bolinopsis infi4ndibulum is
not part of the food web. Support for this hypothesis may be found in the
paper by Falk-Petersen et al. (2002) where a close coupling of the FA
composition of the NL, mostly WE, between the dominant calanoid
copepods Calanus hyperboreus, C. glacialis and C. finmarchicus, the
ctenophores Mertensia ovum and Beroe cucumis was found. Based on the
presence of calanoid FATM these authors suggested that WE moieties are
transmitted unmodified from Calanus spp. via Mertensia ovum to Beroe
cztcumis.
The chaetognath Sagitta elegans is another active carnivore in the Arctic,
and high abundances have been observed, e.g., in Balsfjord. S. elegans stores
moderate amounts of TAG with a low 18: l(n-7)/18:l(n-9) ratio and high
proportions of calanoid FATM (Falk-Petersen et al., 1987) suggesting that
it is an important predator of these copepods.
Finally, stomach content analyses of different age-groups of the deep-
water prawn Pandalus borealis, collected during spring and summer in
Balsfjord, revealed very clear ontogenetic changes in diet composition
(Hopkins et al., 1993). Age-groups &I were found to consume mostly
calanoid copepods whereas older prawns (II-IV) contained remains of
euphausiids (Thysanoessa spp.) and scales from capelin (probably from fish
discarded by prawn trawlers). These observations were substantiated by
FATM showing that the concentration of calanoid FATM was highest
in the youngest age-classes, whereas in the more mature prawns, higher
proportions of lS:l(n-9), EPA and DHA were found.
5.2.5. Fish
1986b) and eastern Canadian (Ackman et al., 1969; Eaton et al., 1975)
capelin populations (see also summary table by Jangaard, 1974).
Furthermore, in capelin caught off Novaya Zemlya, northern Norway, a
similar FA signature to that seen in Balsfjord was observed (i.e., low levels
of calanoid FATM and elevated proportions of 16:0 and 18:1), consistent
with a mixed diet of calanoid copepods and locally abundant Thysanoessa
inermis (Falk-Petersen et al., 1986b, 1990).
In contrast, large proportions of calanoid FATM were detected in
Maurolicus muelleri and Benthosema glaciale, important members of
the pelagic fish community in Ullsfjord located adjacent to Balsfjord
(Falk-Petersen and Sargent, 1986a; Falk-Petersen et al., 1987). Calanus
finmarchicus, C. hyperboreus and the predatory amphipod Themisto
abyssorum are common species in this fjord (Falk-Petersen et al., 1986a,
1987), and based on FA signatures are all hypothesized to contribute to the
diet of these fish.
In another study including FATM in fish, the 16:1(n-7)/16:0 ratio was
applied as a food web tracer to clarify the impact of food quantity and
quality on the condition of juvenile snail lish (Liparis sp.) off west
Greenland (Pedersen et al., 1999). On the assumption that mesozooplankton
> 400 urn (consisting predominantly of Calanus) constituted the major prey,
the 16:1(n-7)/16:0 tracer was observed to follow the same spatial pattern in
the fish and in the mesozooplankton (Figure 18). Hence, the ratio of the
diatom tracer in the fish increased significantly (T-test, P < 0.001) toward
the northern part of the region, and in addition, correlated significantly
(T-test, P < 0.001) with the condition of the fish. Concurrent analyses of
size-fractionated plankton samples revealed a succession from a hetero-
trophic dinoflagellate and nanoflagellate dominated plankton community in
the south to a late spring bloom, diatom dominated community in the north,
consistent with the withdrawal of sea ice in this area. Intriguingly, the
16:1(n-7)/16:0 ratio did not show a significant south-north trend in
the phytoplankton, and it was therefore deduced that the tracer signal in
the mesozooplankton in the north (high 16:1(n-7)/16:0, low DHA)
originated from a recent diatom bloom (which was not sampled), reflecting
a lower turnover rate of FATM with increasing body size (see also Section
4.5). The effect was carried over to the snail fish, whose FA composition
suggested that they had been feeding on a flagellate-based food web in the
south and a diatom-based food web in the north.
The Arcto-Norwegian cod, Gadus morhua, utilizes the Lofoten area,
northern Norway, as an important spawning site, and FATM of first-
feeding cod larvae have been examined to ascertain the contribution
of phytoplankton to their diet (Klungsoyr et al., 1989). Close similarities
between the FA compositions of the phytoplankton community, composed
primarily of diatoms and Phaeocystis pouchetii, and the cod larvae were
290 JOHANNE DALSGAARD ETAL.
4-
Ynozoopl8nkton B 4OOpm A
.
f J- .
. . .
0 l * ..
(0 . .
. .
G 2-
08
:
. .
g * .
.. l-
.
z
0 . .
8
0, / , 1,, ( , , ( ( , , ( r, r,, , , , / , , , , ( , , , ) , , , , , , ,
1.01 !
0 LIpuk SPP.
F 08-1
G
F!
(0 0.6
I
N i
. i/
1
E 0.4
c
;r
s 02 :
ij
0.0
souul North
FATM have also been used to assess the potential prey of white whales
(Delphinapterus leucas; Dahl et al., 2000). Comparing blubber FA signatures
derived from biopsies of white whales foraging close to Svalbard with FA
compositional data of potential prey species by principal component
analyses, Dahl et al. (2000) deduced that juvenile polar cod, capelin, Calanus
hyperhoreus and Pandalus borealis constituted the most likely prey. From
observations of feeding behavior and stomach content analyses it was,
however, suggested that copepods are not ingested directly but represent a
secondary input via predation on polar cod and capelin.
A large share of primary production in the Antarctic takes place in the sea-
ice. Here, as in the Arctic, the microalgal communities are dominated by
diatoms (Fahl and Kattner, 1993; Nichols et al., 1993) although a variety of
autotrophic flagellates, particularly Phaeocystis, are also present (Marchant
and Thomsen, 1994). Most of the pelagic primary production occurs in the
marginal ice zone rather than in the open ocean (Marchant and Thomsen,
1994, and references therein). Pelagic phytoplankton is also composed
largely of diatoms superimposed on a background of Phaeocystis and
dinoflagellates. Moreover, Phaeocystis spp. typically bloom in the marginal
ice zone in the spring prior to the increase in diatoms (Pond et al., 1993;
Marchant and Thomsen, 1994; Skerratt et al., 1995; Cripps et al., 1999;
Cripps and Atkinson, 2000, and references therein).
As in other regions, the phytoplankton biomass in the Antarctic is typified
by characteristic signature FA reflecting the prevailing class of microalgae
(Section 2.4.3). However, in contrast to the Arctic and as discussed in Section
2.4.3, Phaeocystis spp. in this region are much less rich in (n-3) PUFA, which
together with a low 16:1(n-7)/16:0 ratio and an elevated concentration of
18: l(n-9) distinguishes them from diatoms (Skerratt et al., 1995).
5.3.2. Copepods
these copepods feed throughout the year and have evolved a more
opportunistic feeding strategy than strictly herbivorous species. In WE-
storing copepods, long-chain monounsaturated fatty alcohols are typically
replaced either by short-chain saturated fatty alcohols (e.g., Rhincalanus
gigas, Graeve et al., 1994b) or the concentration of 20:l(n-9) is higher than
that of 22: l(n-11) (Calanoides acutus, Graeve et al., 1994b; Kattner and
Hagen, 1995; Falk-Petersen et al., 1999; Section 3.2). FATM are typically
less evident in Antarctic copepods as compared to Arctic copepods (e.g.,
Calanus propinquus, Kattner et al., 1994; Falk-Petersen et al., 1999;
C. simillimus and Euchirella rostromagna, Hagen et al., 1995; Ward et al.,
1996) though this does not apply to all species. High concentrations of
16: l(n-7) and 18:4(n-3) have, for example, been detected in the dominant
circum-Antarctic species, Calanoides acutus (Graeve et al., 1994b). This is
illustrated in Figure 15B, which was based on samples of C. acutus from the
Weddell Sea. Here, specimens from site III-V contained very high
concentrations of 18:4(n-3) probably as a result of the uptake of
Phaeocystis, which was the dominant microalgae in these areas at the time
of sampling (Kattner and Hagen, 1995, and references therein). Lower levels
of 18:4(n-3) combined with elevated concentrations of 16: l(n-7) at site I
suggested a higher uptake of diatoms there, whereas the resolution of group-
specific phytoplankton contributing to the diet of specimens from site II
was less easy to interpret, suggesting a more mixed diet. These observations
were supported by data from the Lazarev Sea, where the FA signature of
C. acutus indicated extensive feeding on a mixed but probably diatom-
dominated phytoplankton diet, with seasonal differences in the uptake of
dinoflagellates and Phaeocystis (Falk-Petersen et a/., 1999).
Fatty acid trophic markers have also been used to resolve the diet
composition of the Antarctic copepod Rhincalanus gigas revealing char-
acteristics of both a herbivorous and omnivorous feeding behavior (Graeve
et al., 1994b; Ward et al., 1996). Hence, the WE fatty acid composition of
specimens from the Weddell Sea revealed a mixture of 18: l(n-9) typical
of a carnivorous diet, and 16:l(n-7), 18:4(n-3), EPA and DHA indicating
additional uptake of phytoplankton (Graeve et al., 1994b). Omnivorous
feeding behavior by R. gigas on phytoplankton, detritus and zooplankton
has indeed been reported by Arashkevich (1978, in Bathmann et al., 1993).
Hence, utilizing FATM, R. gigas has been suggested to be a facultative
herbivore able to switch to nonphytoplankton food when algae are scarce.
Similarly, the predominance of 16: 1(n-7) and 18: l(n-9) in the WE of older,
lipid-rich specimens of another Southern Ocean species, Pareuchaeta
antarctica, collected in the southeastern Weddell Sea (Hagen et al., 1995),
suggested omnivorous feeding behavior by this species as well. Older stages
of P. antarctica are, however, known to feed as carnivores consuming only
small amounts of phytoplankton (Hopkins, 1987). Hence, whereas the
FATTY ACID TROPHIC MARKERS 293
predominance of 18: l(n-9) agreed with conventional feeding studies, the high
concentrations of 16:l(n-7) may be explained by indirect ingestion via
herbivorous copepods (Hagen et al., 1995). Moreover, FA and fatty alcohols
can be subjected to intense restructuring processes and apparently, P.
antarctica completely catabolizes any long-chain monounsaturated com-
pounds ingested with, e.g., Calanoides acutus or Calanus propinquus
(IZlresland, 1991).
Cripps and Hill (1998) examined the effect of different dietary regimes on
the FA (and hydrocarbon) composition of five common Antarctic copepods
in addition to the krill Euphausia superba, sampled along a transect from
the MIZ to the open water. A principal component analysis of the FA
data grouped the copepods into dinoflagellate-feeders, diatom-feeders
and omnivores, whereas E. superba formed a group of its own. The
dinoflagellate-feeding copepods consisted of Calanoides acutus, Calanus
propinquus and Metridia gerlachei, sampled chiefly under the pack-ice. These
specimens were all characterized by high levels of DHA and a low 16: 1(n-7)/
16:O ratio. In the MIZ, Calanus propinquus and Metridia gerlachei had
apparently switched to a more omnivorous feeding behavior, as specimens
from this sampling location contained higher proportions of 16:0 and
18: l(n-9) while typical microalgal FATM were absent. This was also true of
cyclopoid copepods (Oithona spp.), common in the MIZ as well. Diatom
feeding copepods were confined to the open ocean and comprised specimens
of Calanoides acutus, Metridia gerlachei and Rhincalanus gigas. Diatom
FATM were most evident in Calanoides acutus and Rhincalanus gigas, which
both contained a 16:1(n-7)/16:0 ratio > 1 in addition to high concentrations
of EPA. The FA composition of Metridia gerlachei, on the other hand,
was quite similar to specimens of this species sampled in the pack-ice.
Dinoflagellate markers were indeed present in all three species sampled in
the open ocean, indicating that these microalgae, in addition to diatoms,
contributed to the diet at this location. In contrast to the copepods, there
was no spatial resolution in the FA pattern of Euphausia superba, suggesting
a dietary regime and lipid metabolism distinct from the copepods.
5.3.3. Euphausiids
5.3.5. Fish
copepods may also form part of the diet. Moreover, higher concentrations
of 2O:l than 22:l indicated that Calanoides acutus rather than Calanus
propinquus forms part of the diet, conforming with the vertical distribution
pattern of these copepods (Hagen et al., 2000). Remarkably high levels of
monoenoic fatty alcohols (37-90% of total fatty alcohols) and FA (37-88%
of total FA), comprising mainly 18: l(n-9) 22: 1 and 20: 1, were also found in
lipid rich myctophids (lantern fish) caught in the northern sub-Arctic Pacific
(Saito and Murata, 1996, 1998; Seo et al., 1996) and in the Antarctic
(Phleger et al., 1997a). Consistent with these findings, remains of copepods
and other crustaceans have been recognized in the stomachs of myctophids
from the northern Pacilic (Saito and Murata, 1998) whereas amphipods,
copepods and euphausiids (Thysanoessa macrura) comprise the major prey
of the Antarctic myctophid Electrona antarctica (Phleger et al., 1997a, and
references therein).
Interestingly, it has been suggested that myctophids in general, and in
contrast to northern hemisphere zooplanktivorous species, incorporate
dietary lipids directly, including zooplankton WE (Saito and Murata, 1996,
1998). If that is the case, FATM may prove a very valuable tool for
resolving trophic interactions in these species.
’ CZI 16:l(n-7)/16:0
94 6 - Z16/Xl8
EZT3 16:4wl
g5
5.4.2. Euphausiids
5.4.4. Macrobenthos
FATM (18: l(n-9), 18:4(n-3) and DHA), which were partly replaced during
the spring and post-bloom period by diatom-specific FATM (16: l(n-7)
16: 1(n-4) and EPA).
Similar patterns were recognized in the digestive gland and in the gut
content of the scallop Placopecten magellanicus from Georges Bank, Nova
Scotia (Napolitano and Ackman, 1993). Here, maximum concentrations of
Cl6 PUFA (mostly 16:4(n-1)) and EPA also coincided with the diatom
dominated spring bloom, while a smaller increase in 18:4(n-3) in addition to
a marked increase in the proportion of DHA occurred in the fall, coinciding
with a dinoflagellate and flagellate-dominated fall bloom. These findings
were supported by the trend in the polyunsaturation index (the summed
products of PUFA weight percentages > 1 multiplied by the number of
double bonds) measured in the digestive gland. Hence, this index increased
from summer to fall, consistent with an intensive feeding on particulate
matter rich in AA, EPA and DHA. It was followed by a dramatic decrease
from fall to winter reflecting the mobilization of TAG reserves from the
digestive gland to the maturing gonads. Based on the presence of typical
algal FATM, combined with an overall lack of typical bacterial FATM both
in the gut content and in the digestive gland, it was deduced that the supply
of photosynthetically produced organic matter on Georges Bank was
sufficient to sustain the scallop population throughout the year (Napolitano
and Ackman, 1993).
Comparable temporal patterns in the FA composition were also observed
in the tissue of the blue mussel, Mytilus edulis, from Notre Dame Bay,
Newfoundland (Budge et al., 2001). Here, the level of AA was five-fold
greater than in the phytoplankton, indicating a selective retention of this
FA by the mussels. Moreover, 18:5(n-3) was not detected in mussel tissues
despite significant concentrations in the phytoplankton presumed to
comprise the bulk of their diet. On this basis it was hypothesized that
18:5(n-3) was chain-elongated to EPA, and the potential of employing
18:5(n-3) as a specific dinoflagellate tracer at higher trophic levels was
dismissed. In contrast, Mayzaud (1976) had earlier applied 18:5(n-3) as a
specific dinoflagellate tracer to a natural plankton community in Bedford
Basin, Nova Scotia. In that study, the FA was observed to decrease by
roughly a factor of 10 for each trophic level in a “linear food web”
consisting of microalgae (9% of PL fatty acids) - copepods (2% of TAG
fatty acids) - chaetognaths (0.1% of TAG and WE fatty acids).
5.4.5. Fish
has recently been assessed using FATM (Storr-Paulsen et al., 2003). In this
study, a significant positive correlation between larval condition and the
specific diatom tracer 20:5(n-3)/18:4(n-3) suggested that utilization of a
diatom-based food web contributed to enhanced larval condition. In
contrast, a significant negative correlation between larval condition and the
specific flagellate tracer Cis PUFA/total FA indicated that larvae trapped in
areas of flagellate-dominated primary production experienced sub-optimal
feeding conditions. These observations are consistent with earlier findings
on juvenile cod and sandlance in the North Sea (St. John and Lund, 1996;
Marller et al., 1998; Section 5.5.3) and juvenile snail fish off West Greenland
(Pedersen et al., 1999; Section 5.2.5).
The inter- and intraspecific variability in the FA signature of 28 species
of fish and invertebrates from the Scotian Shelf, Georges Bank and the
southern Gulf of St. Lawrence has also recently been assessed (Budge et al.,
2002). In this study, a CART analysis successfully classified 89% of
the samples, demonstrating that FA, besides containing information on
diets, have the potential to resolve between species based on species-specific
FA compositions. A discriminant analysis separated the 16 species with
sufficient sample sizes into three distinct groups likely to share similar
feeding strategies (Figure 20). The groups separated were the Pleuronectidae
(American plaice, yellowtail flounder, winter flounder), small planktivorous
fish (capelin, herring, northern sandlance) and a third group consisting
mostly of Gadidae (cod, haddock, pollock, silver hake, white hake),
but also including redtish, ocean pout, longhorn sculpin and shrimp.
Shrimps were believed to cluster with Gadidae as they comprise a large
fraction of the diet of this group, resulting in similar FA compositions.
Capelin, herring and northern sandlance separated from the other two
groups by the first discriminant function defined primarily by 22: l(n- 11)
and 20:l(n-9). These results are suggestive of a zooplanktivorous diet
and are supported by previous FA analyses of these species from
the same region (e.g., capelin, Ackman et al., 1969; sandlance, Ackman
and Eaton, 1971; Jangaard, 1974; Eaton et al., 1975; Pascal and Ackman,
1976; capelin, herring and mackerel, Ratnayake, 1979; Ratnayake and
Ackman, 1979).
Significant size-related changes in the FA composition were also observed
in several species from this region, and were consistent with reported
stomach content analyses. Moreover, in all species with statistically
significant sample sizes, there was a significant effect of the sampling
location on the FA signature. As discussed by the authors, such findings
may be attributed to broad-scale differences in prey assemblages in the
northwest Atlantic and ultimately to subtle geographical differences in
primary production (Budge et al., 2002).
JOHANNE DALSGAARD ETAL.
“” n American plaice
5.5.2. Copepodv
pmol C.dms3
L 70
60
. 50
. 40
. 30
. 20
10
(A)
pmol C.drnm3
26.3 3.4 10.4 24.4 30.4 3.5 10.5 22.5 24.5 29.5 5.6 date
(B)
Figure 21 Continued.
also been observed in the closely related but more temperate species,
C. helgofundicus, sampled in the eastern North Sea (Kattner and Krause,
1989). Together, these observations support the hypothesis that the foraging
by C. jikmarchicus and C. helgolandicus is closely coupled to the seasonal
phytoplankton production.
In comparison with Calanus finmarchicus and C. helgolandicus, Kattner
and Krause (1989) found a significantly different FA composition in
Pseudocalunus elongatus. These observations were attributed to a different
308 JOHANNE DALSGAARD ETAL
5.5.3. Fish
5.6.2. Fish
et ul., 1997b, 2002). In one study, 22 common species of forage fish and
invertebrates were sampled within Prince William Sound (PWS) over a four
year period (1994-1998). The species were readily distinguished by their
total FA composition via a CART analysis (92% classified correctly;
Iverson et al., 2002). Species with partly overlapping diets such as walleye
pollock (Theragra chalcogramma), Pacific herring (Clupea harengus pallusi)
and Pacific sandlance (Ammodytes hexapterus) were, however, less success-
fully classified. These observations were supported by a discriminant
analysis in which the three species tended to cluster together on a plot of the
first two discriminant functions. Flatfishes, which presumably also share a
similar diet and life history, constituted another cluster. Ontogenetic
changes in specific dietary FATM (14:0, 20: l(n-1 l), 22: l(n-1 l), EPA,
DHA) were also observed in this system. Hence, herring showed a shift in
FA composition commensurate with a dietary switch from zooplankton in
earlier life stages to a more piscivorous diet as the fish grew larger, an
observation consistent with stomach content analyses. Similar changes have
previously been reported for both herring and pollock in PWS (see below,
Iverson et a/., 1997b), and more lately for several species of fish in the
northwest Atlantic (Section 5.4.5).
Finally, unusually high levels of 20: l(n-11) and 22: l(n-11) were found in
young herring, pollock and sandlance sampled in the spring and summer
1995/1996. These changes in FA composition were attributed to a more
highly stratified ocean surface layer contributing to a reduced biomass of
calanoid copepods in these two years, an occurrence which was hypothe-
sized to have forced a dietary shift in the young fish (Iverson et al., 2002).
In a study of harbor seals (Phoca vitulina) from this system, Iverson et al.
(1997b) employed a CART analysis on blubber FA. The analysis readily
classified the seals according to region (PWS, Kodiak Island, Southeast
Alaska) and even specific haulout sites within PWS, suggesting site-specific
diets (Iverson et a/., 1997b). Moreover, herring and pollock were classified
according to size (length) and sampling location in a CART analysis on the
FA composition of potential prey, and the authors commented: “One result
of these,findings is that given a fatty acid composition of an unknown herring
or pollock, one could essentially determine its size-class and location within the
study area with reasonable certainty... This couldprovide an importunt tool,for
studying .fLwuging ecology and stock structure of ,fish species”.
In a preliminary analysis, the FA data of the seals were subjected to the
classification rules derived from the FA composition of their potential prey.
Intriguingly, the seals separated into two groups suggesting possible prey
310 JOHANNE DALSGAARD ETAL.
differences. Hence, seals from the southern PWS and Kodiak Island
grouped with yellowfin sole and larger herring and pollock, whereas seals
from the northern and eastern part of PWS and southeastern Alaska
grouped with smaller herring and pollock, smelt, sandlance, cod, octopus,
squid and shrimp (Iverson et al., 1997b).
5.7. Mediterranean
60
0
20 40 60 80
March April May DAYS
0 I I
5.7.2. Euphausiids
Clyde Sea and Kattegat, which contained higher diatom signals (a high
16: l(n-7)/16:0 ratio and high EPA). The latter also contained higher
concentrations of calanoid FATM suggesting that they were relying heavily
on copepods. In contrast, M. norvegica from the Ligurian Sea contained
significantly lower concentrations of long-chain MUFA, although copepods
from this area were also relatively deficient in these compounds. Hence, 20: 1
and 22: 1 cannot be used as copepod FATM in this area (Virtue et al., 2000).
Mayzaud et al. (1999) emphasized that one should exercise caution when
interpreting FATM in omnivorous species such as M. norvegica. They
wrote: “To be of practical use under natural conditions, fatty acid tracers in
omnivorous species should at least be present at concentrations higher than 1%
of the total.fiztty acids (below that the tracer is likely to be a contaminantfrom
ingestedgrazers) and display over time a pattern coherent with that of the,food
supply”.
6.1. State-of-the-art
state of our knowledge it would appear that fatty acid analyses represent
a rather blunt tool in defining food chain inter-relationships. Until further
knowledge is accumulated it would appear best to apply fatty acid analyses
as a corroborative method to support prey-predator relationships already
indicated on independent grounds, such as the analyses of stomach
contents”.
Fraser et al. (1989) later added that: “to clartfy the transfer of lipid
biomarkers up the food web, the availability of tracer lipids in algae and the
zooplankton prey of larvalfish must be established before these tracers may be
employed either quantitatively or qualitatively”. St. John and Lund (1996),
while recognizing the statement of Fraser et al. (1989) were more specific
about problems concerning the quantitative application of FATM. They
stated: “these biomarkers may be suitable as a qualitative index of utilization
qf a spectfic food source in field studies, however, quantitative estimates of
transfer between trophic levels in the field may prove to be difficult ,for a
number of reasons. It is evident that a better understanding of the dynamics qf
lipid incorporation and utilization with respect to environmental conditions
such as temperature, light and nutrients in phytoplankton as well as during
ontogeny in zooplankton and larval and juvenile fish is required before these
biomarkers may be used quantitatively”.
With the recommendations of these authors in mind, it is clear that to
quantify relationships using FATM, information would need to be
available on a number of aspects of the dynamics of FA in marine
animals, including not least, time scales for incorporation of new FA
signatures into tissues. This has been examined in a few laboratory
studies of copepods, larval and adult cod (Graeve et al., 1994a; St. John
and Lund, 1996; Kirsch et al., 1998) and in one field experiment on
mussels (Mytilus galloprovincialis; Freites et al., 2002), but there is still a
long way to go, considering the physiological status of the organisms
(i.e., adding or depleting lipid reserves), growth rates and ontogenetic
state of development, the impact of mixed diets, etc. Given the resolution
of FATM, we question whether turning FATM into a quantitative tool is
worth the effort, although in some studies a quantitative approach has
been considered (e.g., Desvilettes et al., 1997).
Resolution of ecological niches is the strength of the FATM approach
and a key to resolving complex trophic interactions. FATM are
incorporated largely unaltered into the NL pool of primary consumers,
especially in periods of low catabolic activity, as when the animals are
accumulating lipid reserves. In particular, 16:l(n-7), Cl6 PUFA and EPA
have been used as indicators of diatom-based diets, whereas 18:4(n-3) Cl 8
PUFA and DHA are used as dietary tracers of dinoflagellates and
prymnesiophytes. Secondary and higher order consumers may also
incorporate dietary FA largely unaltered into their NL reserves, but the
FATTY ACID TROPHIC MARKERS 315
FATM are obviously good tools for assessing trophic interactions in the
marine environment, adding information that is at times difficult, and in
some instances impossible, to derive from more traditional techniques, such
as stomach content analyses. In particular, FA provide information on the
origin of lipid reserves generated over a period of time.
Primary producer communities in the marine environment are
dominated by diatoms, dinoflagellates and prymnesiophytes, which can
be distinguished based on the presence and combinations of particular
FA (as summarized in Table 2; see also Mayzaud et al., 1990). The
spatial and temporal resolution of the various phytoplankton groups,
and hence the basic FA pattern in the marine environment, is largely
determined by macro and mesoscale stratification processes acting on
phytoplankton group dominance through affecting light and nutrient
availability. Large concentrations of phytoplankton biomass, essentially
dominated by diatoms, evolve under spring bloom type conditions and
form the basis for an efficient transfer of energy to higher trophic levels.
Flagellates, on the other hand, typically dominate the phytoplankton
communities before and after diatom bloom events when either light or
nutrients are limiting, establishing the potential for microbial loop food
webs. This coupling between hydrodynamic processes and the transfer of
group-specific phytoplankton production to higher trophic levels has
been established using FATM. For example, St. John and Lund (1996)
showed a coupling between the growth and condition of larval and
juvenile fish to spatial variations in frontal primary production, linking
ultimately with physical frontal mixing processes, using 16: 1(n-7)/1 6:0
as a food web tracer. Hence, applied in this manner, in examination of
the potential impact of mesoscale processes, FATM may provide a tool
for resolving the impact of global change on marine ecosystem
dynamics. To further develop this approach, we suggest that it could
be combined with the analyses of larval fish otolith microstructures.
These allow an indication of the growth history of the individuals,
thereby contributing to the identification of the potential dynamics of
FATM incorporation.
Fatty acids have principally been used as qualitative markers of
trophic interactions in shelf sea ecosystems with an emphasis on higher
latitudes. In contrast, very few FATM studies have been carried out
in upwelling and open ocean, oligotrophic areas, including tropical systems.
Primary producers in oligotrophic systems are composed largely of small,
autotrophic flagellates and cyanobacteria, forming the basis of low biomass,
microbial loop food webs. As microorganisms usually do not accumulate
large lipid reserves, FATM may be of less relevance here. However, despite
FATTY ACID TROPHIC MARKERS
317
ACKNOWLEDGEMENTS
REFERENCES
norvegica and Thysanobsa inermis) and their role in the aquatic food web. Journal
of the Fisheries Research Board of Canada 27, 513-533.
Ahern, T. J., Katoh, S. and Sada, E. (1983). Arachidonic acid production by the red
algae Porphyridiurn cruentum. Biotechnology and Bioengineering 15, 1057-1070.
Ahlgren, G., Blomqvist, P., Boberg, M. and Gustafsson, I.-B. (1994). Fatty acid
content of the dorsal muscle - an indicator of fat quality in freshwater fish. Journal
ofFish Biology 45, 131-157.
Ahlgren, G., Goedkoop, W., Markensten, H., Sonesten, L. and Boberg, M. (1997).
Seasonal variations in food quality for pelagic and benthic invertebrates in Lake
Erken - the role of fatty acids. Freshwater Biology 38, 555-570.
Albano, C., Blomqvist, G., Coomans, D., Dunn III, W. J., Edlund, U., Eliasson, B.,
Hellberg, S., Johansson, E., Norden, B., Sjiistriim, M., Sbderstrom, B., Wold, H.
and Wold, S. (1981). Pattern recognition by means of disjoint principal
components models (SIM,CA). Philosophy and methods. In “Proceedings of the
symposium on applied statistics” (A. Hiiskuldsson, K. Conradsen, B. Sloth-Jensen
and K. Esbesen, eds.), pp. 183-218, Danmarks Tekniske Hsjskole, Lyngby.
Denmark.
Albers, C. (1999). “Lipidstoffwechsel polarer Copepoden: Laborexperimente und
Felduntersuchungen”. Ph.D. Thesis, University of Bremen.
Albers, C. S., Kattner, G. and Hagen, W. (1996). The compositions of wax esters,
triacylglycerols and phospholipids in Arctic and Antarctic copepods: evidence of
energetic adaptations. Marine Chemistry 55, 347-358.
Alonso, D. L., Segura de1 Castillo, C. I., Garcia Sanchez, J. L. Sanchez Perez, J. A.
and Garcia Camacho, F. (1994). Quantitative genetics of fatty acid variation in
Isochrysis galbana (Prymnesiophyceae) and Phaeodactylum tricornutum
(Bacillariophyceae). Journal of Phycology 30, 553-558.
Al-Hasan, R. H., Ali, A. M. and Radwan, S. S. (1990). Lipids, and their constituent
fatty acids, of Phaeocystis sp. from the Arabian Gulf. Marine Biology 105, 9-14.
Anthony, J. A., Roby, D. D. and Turco, K. R. (2000). Lipid content and energy
density of forage fishes from the northern Gulf of Alaska. Journal qf Experimental
Marine Biology and Ecology 248, 53-78.
Armstrong, S. G., Wyllie, S. G. and Leach, D. N. (1994). Effects of season and
location of catch on the fatty acid compositions of some Australian fish species.
Food Chemistry 51, 295-305.
Arts, M. T. and Wainman, B. C. (eds.). (1999). “Lipids in freshwater ecosystems”.
Springer, New York.
Atkinson, A. and Snyder, R. (1997). Krill-copepod interactions at South Georgia,
Antarctica, I. Omnivory by Euphausia superba. Marine Ecology Progress Series
160, 63376.
Auel, H. (1999). The ecology of Arctic deep-sea copepods (Euchaetidae and
Aetideidae). Aspects of their distribution, trophodynamics and effect on the
carbon flux. Berichte zur Polarforschung 319, l-97.
Auel, H., Harjes, M., da Rocha, R., Sttibing, D. and Hagen, W. (2002). Lipid
biomarkers indicate different ecological niches and trophic relationships of the
Arctic hyperiid amphipods Themisto abyssorum and T. libellula. Polar Biology 25,
374383.
Bakes, M. J., Elliott, N. G., Green, G. J. and Nichols, P. D. (1995). Variation in lipid
composition of some deep-sea fish (Teleostei: Oreosomatidae and
Trachichthyidae). Comparative Biochemistry and Physiology lllB, 6333645.
Bathmann, U. V., Makarov, R. R., Spiridonov, V. A. and Rohardt, G. (1993).
Winter distribution and overwintering strategies of the Antarctic copepod species
320 JOHANNE DALSGAARD ETAL.
Castell, J. D., Sinnhuber, R. O., Lee, D. J. and Wales, J. H. (197213). Essential fatty
acids in the diet of Rainbow Trout (Salmo gnirdneri): Physiological symptoms of
EFA deficiency. Journal of Nutrition 102, 87-92.
Castell, J. D., Sinnhuber, R. O., Wales, J. H. and Lee, D. J. (1972~). Essential fatty
acids in the diet of Rainbow Trout (Salmo gairdneri): Growth, feed conversion and
some gross deficiency symptoms. Journal of Nutrition 102, 77-86.
Christie, W. (1982). “Lipid analysis: isolation, separation, identification and
structural analysis of lipids”. Pergamon Press, Oxford.
Christie, W. W. (1989). “Gas chromatography and lipids. A practical guide”. The
Oily Press, Ayr, Scotland.
Christie, W. W. (ed.) (1992). “Advances in lipid methodology - One”. The Oily Press,
Ayr, Scotland.
Christie, W. W. (ed.) (1993). “Advances in lipid methodology - Two”. The Oily
Press, Dundee, Scotland.
Christie, W. W. (ed.) (1996). “Advances in lipid methodology - Three”. The Oily
Press, Dundee, Scotland.
Christie, W. W. (ed.) (1997). “Advances in lipid methodology - Four”. The Oily
Press, Dundee, Scotland.
Chuecas, L. and Riley, J. P. (1966). The component fatty acids of some sea-weed
fats. Journal of the Marine Biological Association of the United Kingdom 46,
1533159.
Chuecas, L. and Riley, J. P. (1969). Component fatty acids of the total lipids of some
marine phytoplankton. Journal of the Marine Biological Association qf the United
Kingdom 49, 97-l 16.
Clarke, A., Holmes, L. J. and Hopkins, C. C. E. (1987). Lipid in an Arctic food
chain: Calanus, Bolinopsis, Beroe. Sarsia 72, 41-48.
Claustre, H., Marty, J. C., Cassiani, L. and Dagaut, J. (1988). Fatty acid dynamics in
phytoplankton and microzooplankton communities during a spring bloom in the
coastal Ligurian Sea: Ecological implications. Marine Microbial Food Webs 3,
51-66.
Claustre, H., Marty, J. C. and Cassiani, L. (1989). Intraspecific differences in
the biochemical composition of a diatom during a spring bloom in Villefranche-
sur-Mer Bay, Mediterranean Sea. Journal of E.uperimental Marine Biology and
Ecology 129, 17-32.
Claustre, H., Poulet, S. A., Williams, R., Marty, J.-C., Coombs, F., Mlih, B.,
Hapette, A. M. and Martin-Jezequel, V. (1990). A biochemical investigation of a
Phaeocystis sp. bloom in the Irish Sea. Journal qf the Marine Biological Association
of the United Kingdom 70, 197-207.
Cobelas, M. A. (1989). Lipid in microalgae. A review II. Environment. Grases J
Aceites 40, 2 13-223.
Cobelas, M. A. and Lechado, J. Z. (1989). Lipid in microalgae. A review I.
Biochemistry. Grases y Aceites 40, 118-l 45.
Cohen, Z., Vonshak, A. and Richmond, A. (1988). Effect of environmental
conditions on fatty acid composition of the red alga Porphyridium cruentum:
correlation to growth rate. Journal qf Phycology 24, 3288332.
Colombo, J. C., Silverberg, N. and Gearing, J. N. (1997). Lipid biogeochemistry in
the Laurentian Trough - II. Changes in composition of fatty acids, sterols and
aliphatic hydrocarbons during early diagenesis. Organic Geochemistry 314.
Conte, M. H., Volkman, J. K. and Eglinton, G. (1994). Lipid biomarkers of the
Haptophyta. IU “The Haptophyte Algae” (J. C. Green and B. S. C. Leadbeater,
eds.), pp. 351-377. Clarendon Press, Oxford.
322 JOHANNE DALSGAARD ETAL.
Hamm, C., Reigstad, M., Riser, C. W., Miihlebach, A. and Wassmann, P. (2001). On
the trophic fate of Phaeacystis pouchetii. VII. Sterols and fatty acids reveal
sedimentation of P. pouchetii-derived organic matter via krill fecal strings. Murine
Ecology Progress Series 209, 55-69.
Harrison, P. J., Thompson, P. A. and Calderwood, G. S. (1990). Effects of nutrient
and light limitation on the biochemical composition of phytoplankton. Journal uf
Applied Phycology 2, 45-56.
Harvey, H. R. (1994). Fatty acids and sterols as source markers of organic matter in
sediments of the North Carolina continental slope. Deep-Sea Reseurch II 41,
783-796.
Harvey, H. R. and Macko, S. A. (1997). Catalysts or contributors? Tracking
bacterial mediation of early diagenesis in the marine water column. Orgunic
Geochemistry 26, 53 l-544.
Harvey, H. R., Eglinton, G.,’ O’Hara, S. C. M. and Corner, E. D. S. (1987).
Biotransformation and assimilation of dietary lipids by Culunu.r feeding on a
dinoflagellate. Geochimicu et Cosmochimica Acta 51, 303 l-3040.
Harwood, J. L. and Jones, A. L. (1989). Lipid metabolism in algae. Advances in
Botanical Research 16, l-53.
Harwood, J. L. and Russell, N. J. (1984). “Lipids in plants and microbes”. George
Allen and Unwin, London.
Henderson, R. J. and Sargent, J. R. (1985). Fatty acid metabolism in fish. In
“Nutrition and feeding in fish” (C. B. Cowey, A. M. Mackie and J. G. Bell, eds.),
pp. 349-364, Academic Press, London.
Henderson, R. J. and Tocher, D. R. (1987). The lipid composition and biochemistry
of freshwater fish. Progress in Lipid Research 26, 281-347.
Henderson, R. J., Sargent, J. R. and Hopkins, C. C. E. (1984). Changes in the
content and fatty acid composition of lipid in an isolated population of the capelin
Mullotus villosus during sexual maturation and spawning. Marine Biology 78,
255-263.
Henderson, R. J., Park, M. T. and Sargent, J. R. (1995). The desaturation and
elongation of 14C-labelled polyunsaturated fatty acids by pike (Es0.u Lucius L.)
in vivo. Fish Physiology and Biochemistry 14, 223-235.
Henderson, R. J., Hegseth, E. N. and Park, M. T. (1998). Seasonal variation in lipid
and fatty acid composition of ice algae from the Barents Sea. Polar Biology 20,
48-55.
Hirche, H.-J. and Kattner, G. (1993). Egg production and lipid content of Calanus
glacialis in spring: indication of a food-dependent and food-independent
reproductive mode. Marine Biology 117, 615-622.
Hirche, H.-J., Hagen, W., Mumm, N. and Richter, C. (1994). The Northeast Water
polynya, Greenland Sea. III: Meso- and macrozooplankton distribution and
production of dominant herbivorous copepods during spring. Polar Biology 14,
49 I-503.
Honya, M., Kinoshita, T., Ishikawa, M., Mori, H. and Nisizawa, K. (1994).
Seasonal variation in the lipid content of cultured Laminaria japonicu: fatty
acids, sterols, p-carotene and tocopherol. Journal of Applied Phycology 6, 25-29.
Hopkins, T. L. (1987). Midwater food web in McMurdo Sound, Ross Sea.
Antarctica. Marine Biology 96, 933106.
Hopkins, C. C. E., Sargent, J. R. and Nilssen, E. M. (1993). Total lipid content, and
lipid and fatty acid composition of the deep-water prawn Pandulus borealis from
Balsfjord, northern Norway: growth and feeding relationships. Marine Ecology
Progress Series 96, 217-228.
FATTY ACID TROPHIC MARKERS 327
Holmer, G. (1989). Triglycerides. In “Marine Biogenic lipids, fats, and oils”, Vol. I
(R. G. Ackman, ed.), pp. 139-173, CRC Press, Boca Raton, Florida.
IUPAC-IUB Commission on Biochemical Nomenclature (1967). The nomenclature
of lipids. The Journal qf Biological Chemistry, 242, 48454849.
IUPAC-IUB Commision on Biochemical Nomenclature (CBN)( 1977). The nomen-
clature of lipids. Recommendations, 1976. European Journal of Biochemistry 79,
11-21.
Iverson, S. J. (1993). Milk secretion in marine mammals in relation to foraging: can
milk fatty acids predict diet? Symposia qf the Zoological Society qf London 66,
263-291.
Iverson, S. J., Arnould, J. P. Y. and Boyd, I. L. (1997a). Milk fatty acid signatures
indicate both major and minor shifts in the diet of lactating Antarctic fur seals.
Canadian Journal of’ Zoology 75, 188-l 97.
Iverson, S. J., Frost, K. J. and Lowry, L. F. (1997b). Fatty acid signatures reveal fine
scale structure of foraging distribution of harbor seals and their prey in Prince
William Sound, Alaska. Marine Ecology Progress Series 151, 255-271.
Iverson, S. J., Frost, K. J. and Lang, S. L. C. (2002). Fat content and fatty acid
composition of forage fish and invertebrates in Prince William Sound, Alaska:
Factors contributing to among and within species variability. Marine Ecolog,v
Progress Series 241, 161-181.
Jangaard, P. M. (1974). The capelin (Mallotus villosus). Biology, distribution,
exploitation, utilization, and composition. Bulletin of the Fisheries Research Board
of Canada 186, l-70.
Jeffries, H. P. (1970). Seasonal composition of temperate plankton communities:
fatty acids. Limnology and Oceanography 15, 419426.
Jezyk, P. F. and Penicnak, A. J. (1966). Fatty acid relationships in aquatic food
chain. Lipids 1, 427429.
Jobling, M. (1994). “Fish bioenergetics”. Chapman and Hall, London.
Johns, R. B. and Perry, G. J. (1977). Lipids of the marine bacterium Flexibacter
polymorphus. Archives qf’ Microbiology 114, 267-271.
Jonasdottir, S. H., Fields, D. and Pantoja, S. (1995). Copepod egg production in
Long Island Sound USA, as a function of the chemical composition of seston.
Marine Ecology Progress Series 119, 87-98.
Joseph, J. D. (1975). Identification of 3, 6, 9, 12, 15-octadecapentaenoic acid in
laboratory-cultured photosynthetic dinoflagellates. Lipids 10, 395403.
Joseph, J. D. (1982). Lipid composition of marine and estuarine invertebrates. Part
II: Mollusca. Progress in Lipid Research 21, 109-153.
Joseph, J. D. (1989). Distribution and composition of lipids in marine invertebrates.
In “Marine biogenic lipids, fats, and oils”, Vol. II. (R. G. Ackman, ed.),
pp. 499143, CRC Press, Boca Raton, Florida.
Kaitaranta, J. K. and Linko, R. R. (1984). Fatty acids in the roe lipids of common
food fishes. Comparative Biochemistry and Physiology 79B, 331-334.
Kanazawa, A., Teshima, S. I. and Ono, K. (1979). Relationship between essential
fatty acid requirements of aquatic animals and the capacity for bioconversion of
linolenic acid to highly unsaturated fatty acids. Comparative Biochemistry and
Physiology 63B, 295-298.
Kattner, G. and Brockmann, U. H. (1990). Particulate and dissolved fatty acids in an
enclosure containing a unialgal Skeletonema costatum (Greve) Cleve culture.
Journal of Experimental Marine Biology and Ecology 141, l-13.
Kattner, G. and Hagen, W. (1995). Polar herbivorous copepods - different pathways
in lipid biosynthesis. ICES Journal qf Marine Science 52, 329-335.
328 JOHANNE DALSGAARD ETAL
Kiwboe, T. (1993). Turbulence, phytoplankton cell size, and the structure of pelagic
food webs. Advances in Marine Biology 29, l-72.
Kiyashko, S. I., Kharlamenko, V. 1. and Imbs, A. B. (1998). Stable isotope ratios and
fatty acids as food source markers of deposit-feeding invertebrates. Russian
Journal qf Marine Biology 24, 170-174.
Klungsayr, J., Tilseth, S., Wilhelmsen, S., Falk-Petersen, S. and Sargent, J. R.
(1989). Fatty acid composition as an indicator of food intake in cod larvae Gadus
morhua from Lofoten, Northern Norway. Marine Biology 102, 183-188.
Kristiansen, A., Moestrup, 0. and Nielsen, H. (1981). “Introduktion til alger og
bakterier”. Nucleus, Arhus, Denmark.
Lasker, R. and Theilacker, G. H. (1962). The fatty acid composition of the lipids of
some Pacific sardine tissues in relation to ovarian maturation and diet. Journal qf
Lipid Research 3, 60-64.
Le Fevre, J. (1986). Aspects of the biology of frontal systems. Advances in Marine
Biology 23, 1633299.
Lechevalier, H. and Lechevalier, M. P. (1988). Chemotaxonomic use of lipids - an
overview. III “Microbial lipids”, Vol. 1 (C. Ratledge and S. G. Wilkinson, eds.).
pp. 869-902, Academic Press, London.
Lee, R. F. (1974). Lipid composition of the copepod Calanus hyperboreus
from the Arctic Ocean. Changes with depth and season. Marine Biology 26,
3133318.
Lee, R. F. (1975). Lipids of Arctic zooplankton. Comparative Biochemistry and
Physiology 51B, 263-266.
Lee, R. F. and Hirota, J. (1973). Wax esters in tropical zooplankton and nekton and
the geographical distribution of wax esters in marine copepods. Limnology and
Oceanography 18, 2277239.
Lee, R. F. and Nevenzel, J. C. (1979). Wax esters in the marine environment: origin
and composition of the wax from Bute Inlet, British Columbia. Journal of’ the
Fisheries Research Board of Canada 36, 15 19-l 523.
Lee, R. F. and Patton, J. S. (1989). Alcohol and waxes. In “Marine biogenic lipids,
fats, and oils”, Vol. I (R. G. Ackman, ed.), pp. 733102, CRC Press, Boca Raton,
Florida.
Lee, R. F., Hirota, J. and Barnett, A. M. (1971a). Distribution and importance of
wax esters in marine copepods and other zooplankton. Deep-Sea Research
18, 1147-l 165.
Lee, R. F., Nevenzel, J. C. and Paffenhiifer, G.-A. (1971b). Importance of wax esters
and other lipids in the marine food chain: phytoplankton and copepods. Marine
Biology 9, 999108.
Legendre, P. and Legendre, L. (1998). “Numerical ecology”. Elsevier Science,
Amsterdam.
Levinsen, H., Turner, J. T., Nielsen, T. G. and Hansen, B. W. (2000). On the trophic
coupling between protists and copepods in Arctic marine ecosystems. Marine
Ecology Progress Series 204, 6577.
Levring, T. (1979). The vegetation in the sea. In “Marine ,algae in pharmaceutical
science”, Vol. 1 (H. A. Hoppe, T. Levring and Y. Tanaka, eds.), pp. 3-23, Walter
de Gruyter, Berlin.
Lewis, R. W. (1969). The fatty acid composition of Arctic marine phytoplankton and
zooplankton with special reference to minor acids. Limnology and Oceanograph\.
14, 3540.
Li. X., Fan, X., Han, L. and Lou, Q. (2002). Fatty acids of some algae from the
Bohai Sea. P/l~,tochcnll~trl, 59, 157--l 61,
330 JOHANNE DALSGAARD ETAL.
Meziane, T., Bodineau, L., Retiere, C. and Thoumelin, G. (1997). The use of lipid
markers to define sources of organic matter in sediment and food web of the
intertidal salt-marsh-flat ecosystem of Mont-Saint-Michel Bay, France. Journal qf
Sea Research 38, 47-58.
Meziane, T., Sanabe, M. C. and Tsuchiya, M. (2002). Role of fiddler crabs of a
subtropical intertidal flat on the fate of sedimentary fatty acids. Journal of
Experimental Marine Biology and Ecology 270, 191-201.
Miller T. J., Crowder L. B., Rice J. A. and Marschall E. A. (1988). Larval size and
recruitment in fishes: toward a conceptual framework. Canadian Journal qf
Fisheries and Aquatic Sciences 45, 1657-1670.
Miyazaki, T. (1983). Compositional changes of fatty acids in particulate matter and
water temperature, and their implications to the seasonal succession of
phytoplankton in a hypereutrophic lake, Lake Kasumigaura, Japan. Archiv /‘Gr
Hydrobiologie 99, 1-14.
Mjaavatten, O., Levings, C. D. and Poon, P. (1998). Variation in the fatty acid
composition of juvenile chinook and coho salmon from Fraser River Estuary
determined by multivariate analysis; role of environment and genetic origin.
Comparative Biochemistry and Physiology 120B, 291-309.
Moller, P., St. John, M., Lund, T. and Madsen, K. P. (1998). Identifying the effect of
frontal regimes on condition in larval and juvenile sand lance (Ammodytes sp.):
Utilisation of food web specific tracer lipids. International Council for the
Exploration of the Sea, Council Meeting 1998/R:8.
Morales, C. E., Bedo, A., Harris, R. P. and Tranter P. R. G. (1991). Grazing of
copepods assemblages in the north-east Atlantic: the importance of the small size
fraction. Journal of Plankton Research 13, 455-472.
Moreno, V. J., Moreno, J. E. A. and Brenner, R. R. (1979). Fatty acid metabolism in
the calanoid copepod Pctracalunus parvus: 1. Polyunsaturated fatty acids. Lipids
14, 313-317.
Morris, I. (1981). Photosynthetic products, physiological state, and phyto-
plankton growth. Canadian Bulletin of Fisheries and Aquatic Sciences 210,
833102.
Morris, R. J. (1984). Studies of a spring phytoplankton bloom in an enclosed
experimental ecosystem. II. Changes in the component fatty acids and sterols.
Journal qf E?cperimental Marine Biology and Ecology 75, 59-70.
Morris, R. J., McCartney, M. J. and Robinson, G. A. (1983). Studies of a spring
phytoplankton bloom in an enclosed experimental ecosystem. I. Biochemical
changes in relation to the nutrient chemistry of water. Journal of’ Experimental
Marine Biology and Ecology 70, 249-262.
Morris, R. J., McCartney, M. J., Joint, I. R. and Robinson, G. A. (1985). Further
studies of a spring phytoplankton bloom in an enclosed experimental ecosystem.
Journal of Experimental Marine Biology and Ecology 86, 151-170.
Mourente, G. and Tocher, D. R. (1993a). Incorporation and metabolism of
“C-labelled polyunsaturated fatty acids in juvenile gilthead sea bream Sparus
uurata L. in vivo. Fish Physiology and Biochemistry 10, 443453.
Mourente, G. and Tocher, D. R. (1993b). Incorporation and metabolism
of ‘“C-labelled polyunsaturated fatty acids in wild-caught juveniles of
golden grey mullet, L&r aurata, in vivo. Fish Physiology and Biochemistry 12,
119-130.
Mourente, G., Lubiin, L. M. and Odriozola, J. M. (1990). Total fatty acid
composition as a taxonomic index of some marine microalgae used as food in
marine aquaculture. Hydrobiologia 203, 147-l 54.
340 JOHANNE DALSGAARD ETAL.
Ward, P., Shreeve, R. S. and Cripps, G. C. (1996). Rhincalanus gigas and Calanus
simillimus: lipid storage patterns of two species of copepod in the seasonally ice-
free zone of the Southern Ocean. Journal of Plankton Research 18, 1439-1454.
Watanabe, T. (1982). Lipid nutrition in fish. Comparative Biochemistry ami
Physiology 73B, 3-15.
White, D. C., Bobbie, R. J., Nickels, J. S., Fazio, S. D. and Davis, W. M. (1980).
Nonselective biochemical methods for the determination of fungal mass and
community structure in estuarine detrital microflora. Botanica Marina 23,
239-250.
Weld, S., Albano, C., Dunn III, W. J., Esbensen, K., Hellberg, S., Johansson, E. and
Sjiistriim, M. (1988). Pattern recognition: finding and using regularities in
multivariate data. In “Food research and data analysis” (H. Martens and H.
Russwurth Jr., eds.), pp. 147-188, Applied Science Publishers, London.
Wood, B. J. B. (1988). Lipids of algae and protozoa. In “Microbial lipids”, Vol. 1
(C. Ratledge and S. G. Wilkinson, eds.), pp. 807-867, Academic Press, London.
Yazawa, K., Araki, K., Watanabe, K., Ishikawa, C., Inoue, A., Kondo, K., Watabe,
S. and Hashimoto, K. (1988). Eicosapentaenoic acid productivity of the bacteria
isolated from fish intestines. Nippon Suisan Gakkaishi 54, 1835-1838.
Yunker, M. B., Macdonald, R. W., Veltkamp, D. J. and Cretney, W. J. (1995).
Terrestrial and marine biomarkers in a seasonally ice-covered Arctic estuary ~
integration of multivariate and biomarker approaches. Marine chemistry 49. I-50.
FATTY ACID TROPHIC MARKERS 333
Nichols, P. D., Jones, G. J., de Leeuw, J. W. and Johns, R. B. (1984). The fatty acid
and sterol composition of two marine dinoflagellates. Phytochemistry 23,
1043-1047.
Nichols, P. D., Palmisano, A. C., Smith, A. and White, D. C. (1986). Lipids
of the Antarctic sea ice diatom Nitzschia cylindrus. Phytochemistry 25,
164991654.
Nichols, P. D., Volkman, J. K., Hallegraeff, G. M. and Blackburn, S. I. (1987).
Sterols and fatty acids of the red tide flagellates Heterosigma akashiwo and
Chattonella antiqua (Raphidophyceae). Phytochemistry 26, 2537-2.541.
Nichols, P. D., Skerratt, J. H., Davidson, A., Burton, H. and McMeekin, T. A.
(199 1). Lipids of cultured Phaeocystis pouchetii: signatures for food-web,
biogeochemical and environmental studies in Antarctica and the Southern
Ocean. Phytochemistry 30, 3209-3214.
Njinkoue, J.-M., Barnathan. G., Miralles, J., Gaydou, E.-M. and Samb, A. (2002).
Lipids and fatty acids in muscle, liver and skin of three edible fish from the
Senegalese coast: Sardinella maderensis, Sardinella aurita and Cephalopholis
taeniops. Comparative Biochemistry and Physiology 131B, 395-402.
IZlresland, V. (1991). Feeding of the carnivorous copepod Euchaeta antarctica in
Antarctic waters. Marine Ecology Progress Series 78, 4147.
Owen, J. M., Adron, J. W., Sargent, J. R. and Cowey, C. B. (1972). Studies on the
nutrition of marine flat&h. The effect of dietary fatty acids on the tissue fatty-
acids of the plaice Pleuronectes platessa. Marine Biology 13, 16O166.
Owen, J. M., Adron, J. W., Middleton, C. and Cowey, C. B. (1975). Elongation and
desaturation of dietary fatty acids in turbot Scophthalmus maximus L., and
rainbow trout, Salmo gairdnerii Rich. Lipids 10, 528-53 1.
Paerl, H. W. and Zehr, J. P. (2000). Marine nitrogen fixation. In “Microbial Ecology
of the Oceans” (D. L. Kirchman, ed.), pp. 387426. Wiley-Liss, New York.
Paffenhiifer, G.-A. (1993). On the ecology of marine cyclopoid copepods (Crustacea,
Copepoda). Journal of Plankton Research 15, 37-55.
Paradis, M. and Ackman, R. G. (1976). Localization of a source of marine odd
chain-length fatty acids. II. Seasonal propagation of odd chain-length mono-
ethylenic fatty acids in a marine food chain. Lipids 11, 871-876.
Parkes. R. J. (1987). Analysis of microbial communities within sediments
using biomarkers. In “Ecology of microbial communities” (M. Fletcher,
T. R. G. Gray and J. G. Jones, eds.), pp. 147-177, Cambridge University Press,
Cambridge.
Parrish, C. C. and Wangersky, P. J. (1990). Growth and lipid class composition of the
marine diatom, Chaeterceros gracilis, in laboratory and mass culture turbidostats.
Journal of Plankton Research 12, 101 l-1021.
Parrish, C. C., Freitas, A., Bodennec, G., Macpherson, E. J. and Ackman, R. G.
(1990). Unusual fatty acid composition of the toxic marine diatom Nitzschia
pungens. Bulletin qf the Aquaculture Association of Canada 90, 15-18.
Parrish, C. C., Bodennec, G. and Gentien, P. (1994). Time courses of intracellular
and extracellular lipid classes in batch cultures of the toxic dinoflagellate,
Gymnodium cf. nagasakiense. Marine Chemistry 48, 71-82.
Parrish, C. C., McKenzie, C. H., MacDonald, B. A. and Hatfield, E. A. (1995).
Seasonal studies of seston lipids in relation to microplankton species composition
and scallop growth in South Broad Cove, Newfoundland. Marine Ecology
Progress Series 129, 151-164.
Parsons, T. R. (1963). Suspended organic matter in sea water. Progress in
Oceanography 1, 205-293.
334 JOHANNE DALSGAARD ETAL.
Pascal, J.-C. and Ackman, R. G. (1976). Long chain monoethylenic alcohol and acid
isomers in lipids of copepods and capelin. Chemistry and Physics of Lipids 19,
219-223.
Pedersen, L., Jensen, H. M., Burmeister, A. D. and Hansen, B. W. (1999). The
significance of food web structure for the condition and tracer lipid content of
juvenile snail fish (Pisces: Liparis spp.) along 65572”N off West Greenland. Journal
of Plankton Research 21, 1593-1611.
Perry, G. J., Volkman, J. K., Johns, R. B. and Bavor Jr., H. J. (1979). Fatty acids of
bacterial origin in contemporary marine sediments. Geochimica et Cosmochimica
Acta 43, 1715-1725.
Phleger, C. F., Nichols, P. D. and Virtue, P. (1997a). The lipid, fatty acid and fatty
alcohol composition of the myctophid fish Electrona antarctica: high level of wax
esters and food-chain implications. Antarctic Science 9, 258-265.
Phleger, C. F., Nichols, P. D. and Virtue, P. (1997b). Lipids and buoyancy in
Southern Ocean pteropods. Lipids 32, 1093-l 100.
Phleger, C. F., Nichols, P. D. and Virtue, P. (1998). Lipids and trophodynamics
in Antarctic zooplankton. Comparative Biochemistry and Physiology 120B,
31 l-323.
Phleger, C. F., Nelson, M. M., Mooney, B. and Nichols, P. D. (1999). Lipids of
abducted Antarctic pteropods, Spongiobranchaea australis, and their hypriid
amphipod host. Comparative Biochemistry and Physiology 124B, 295-307.
Phleger, C. F., Nelson, M. M., Mooney, B. and Nichols, P. D. (2000). Lipids of
Antarctic salps and their commensal hyperiid amphipods. Polar Biology 23,
329-337.
Phleger, C. F., Nelson, M. M., Mooney, B. D. and Nichols, P. D. (2001). Interannual
variations in the lipids of the Antarctic pteropods Clione limacina and Clio
pyramidata. Comparative Biochemistry and Physiology 128, 553-564.
Phleger, C. F., Nelson, M. M., Mooney, B. D. and Nichols, P. D. (2002). Interannual
and between species comparison of the lipids, fatty acids and sterols of Antarctic
krill from the US AMLR Elephant Island survey area. Comparative Biochemistry
and Physiology 131B, 733-747.
Pohl, P. and Zurheide, F. (1979). Fatty acids and lipids of marine algae and the
control of their biosynthesis by environmental factors. In “Marine algae in
pharmaceutical science”, Vol. 1 (H. A. Hoppe, T. Levring and Y. Tanaka, eds.),
pp. 473-523. Walter de Gruyter, Berlin.
Pohl, P. and Zurheide, F. (1982). Fat production in freshwater and marine algae. In
“Marine algae in pharmaceutical science”, Vol. 2 (H. A. Hoppe and T. Levring.
eds.), pp. 65-80. Walter de Gruyter and Co., Berlin.
Pond, D. W. and Sargent, J. R. (1998). Lipid composition of the pelagic tunicate
Dolioletta gegenbauri (Tunicata, Thaliacea). Journal of Plankton Research 20,
169-174.
Pond, D., Priddle, J., Sargent, J. and Watkins, J. L. (1993). Lipid composition of
Antarctic microplankton in relation to the nutrition of krill. Antarctic Special
Topic, 133-139.
Pond, D. W., Dixon, D. R., Bell, M. V., Fallick, A. E. and Sargent, J. R. (1997a).
Occurrence of 16:2(n-4) and 18:2(n-4) fatty acids in the lipids of the hydrothermal
vent shrimps Rimicaris exoculata and Alvinocaris markensis: Nutritional and
trophic implications. Marine Ecology Progress Series 156, 167-l 74.
Pond, D. W., Segonzac, M., Bell, M. V., Dixon, D. R., Fallick, A. E. and
Sargent, J. R. (1997b). Lipid and lipid carbon stable isotope composition of
the hydrothermal vent shrimp Microcaris ,fortunata: evidence for nutritional
FATTY ACID TROPHIC MARKERS 335
pacifica and its role as a source of docosahexaenoic and eicosapentaenoic acids for
higher trophic levels. Marine Chemistry 78, 9-28.
Sargent, J. R. (1976). The structure, metabolism and function of lipids in marine
organisms. In “Biochemical and biophysical perspectives in marine biology”,
Vol. 3 (D. C. Malins and J. R. Sargent, eds.), pp. 15&212, Academic Press,
London.
Sargent, J. R. (1978). Marine wax esters. Science Progress 65, 437458.
Sargent, R. J. (1995). Origins and functions of egg lipids: Nutritional implications. In
“Broodstock management and egg and larval quality” (N. R. Bromage and R. J.
Roberts, eds.), pp. 353-372, Blackwell Science, London.
Sargent, J. R. and Falk-Petersen, S. (1981). Ecological investigations on the
zooplankton community in Balsfjorden, Northern Norway: lipids and fatty acids
in Meganyctiphanes norvegica, Thysanoessa raschi and T. inermis during mid-
winter. Marine Biology 62, 131-137.
Sargent, J. R. and Falk-Petersen, S. (1988). The lipid biochemistry of calanoid
copepods. Hydrobiologia 167/168, 101-l 14.
Sargent, J. R. and Henderson, R. J. (1986). Lipids. In “The biological chemistry of
marine copepods”, Vol. 1 (E. D. S. Corner and S. C. M. O’Hara, eds.), pp. 59-108,
Clarendon Press, Oxford.
Sargent, J. R. and Henderson, R. J. (1995). Marine (n-3) polyunsaturated fatty acids.
In “Developments in oils and fats” (R. J. Hamilton, ed.), pp. 32-65. Blackie
Academic and Professional, London.
Sargent, J. R. and McIntosh, R. (1974). Studies on the mechanism of biosynthesis of
wax esters in Euchaeta norvegica. Marine Biology 25, 271-277.
Sargent, J. R. and Whittle, K. J. (1981). Lipids and hydrocarbons in the marine food
web. In “Analysis of marine ecosystems” (A. R. Longhurst, ed.), pp. 491-533.
Academic Press.
Sargent, J. R., Gatten, R. R. and McIntosh, R. (1977). Wax esters in the marine
environment - their occurrence, formation, transformation and ultimate fates.
Marine Chemistry 5, 573-584.
Sargent, J. R., Eilertsen, H. C., Falk-Petersen, S. and Taasen, J. P. (1985). Carbon
assimilation and lipid production in phytoplankton in northern Norwegian fjords.
Marine Biology 85, 109-l 16.
Sargent, J. R., Parkes, R. J., Mueller-Harvey, I. and Henderson, R. J. (1987). Lipid
biomarkers in marine ecology. In “Microbes in the sea.” (M. A. Sleigh, ed.),
pp. 119-138, Ellis Horwood Ltd., Chichester.
Sargent, J., Henderson, R. J. and Tocher, D. R. (1989). The lipids. In “Fish
nutrition” (J. E. Halver, ed.), pp. 153-218, Academic Press, San Diego.
Sargent, J. R., Bell, J. G., Bell, M. V., Henderson, R. J. and Tocher, D. R. (1990).
Polyunsaturated fatty acids in marine and terrestrial food webs. Comparative
Physiology 5, 1 l-23.
Sargent, J. R., Bell, J. G., Bell, M. V., Henderson, R. J. and Tocher, D. R.
(1993). The metabolism of phospholipids and polyunsaturated fatty acids
in fish. In “Aquaculture: Fundamental and applied research” (B. Lahlou
and P. Vitiello, eds.), pp. 103-124, American Geophysical Union, Washington,
D.C.
Sargent, J. R., Bell, J. G., Bell, M. V., Henderson, R. J. and Tocher, D. R. (1995a).
Requirement criteria for essential fatty acids. Journal qf‘ Applied Ichthyology 11,
1833198.
Sargent, J. R., Bell, M. V., Bell, J. G., Henderson, R. J. and Tocher, D. R. (1995b).
Origins and functions of n-3 polyunsaturated fatty acids in marine organisms.
FATTY ACID TROPHIC MARKERS
337
Ward, P., Shreeve, R. S. and Cripps, G. C. (1996). Rhincalanus gigas and Calanus
simillimus: lipid storage patterns of two species of copepod in the seasonally ice-
free zone of the Southern Ocean. Journal of Plankton Research 18, 1439-1454.
Watanabe, T. (1982). Lipid nutrition in fish. Comparative Biochemistry ami
Physiology 73B, 3-15.
White, D. C., Bobbie, R. J., Nickels, J. S., Fazio, S. D. and Davis, W. M. (1980).
Nonselective biochemical methods for the determination of fungal mass and
community structure in estuarine detrital microflora. Botanica Marina 23,
239-250.
Weld, S., Albano, C., Dunn III, W. J., Esbensen, K., Hellberg, S., Johansson, E. and
Sjiistriim, M. (1988). Pattern recognition: finding and using regularities in
multivariate data. In “Food research and data analysis” (H. Martens and H.
Russwurth Jr., eds.), pp. 147-188, Applied Science Publishers, London.
Wood, B. J. B. (1988). Lipids of algae and protozoa. In “Microbial lipids”, Vol. 1
(C. Ratledge and S. G. Wilkinson, eds.), pp. 807-867, Academic Press, London.
Yazawa, K., Araki, K., Watanabe, K., Ishikawa, C., Inoue, A., Kondo, K., Watabe,
S. and Hashimoto, K. (1988). Eicosapentaenoic acid productivity of the bacteria
isolated from fish intestines. Nippon Suisan Gakkaishi 54, 1835-1838.
Yunker, M. B., Macdonald, R. W., Veltkamp, D. J. and Cretney, W. J. (1995).
Terrestrial and marine biomarkers in a seasonally ice-covered Arctic estuary ~
integration of multivariate and biomarker approaches. Marine chemistry 49. I-50.
TAXONOMIC INDEX
Saccumminu sphaericu 42
Pugettiu productu 170
Sugitta elegans 288
Pugurus bernhardus 156
Surdinops cuerulea 275
spermatophores 107
Scomber scombrus 276
Pcrluemon serratus 159
Scrippsiella trochoideu 267
Pundalus borealis 103, 288, 291
Skeletonemu eostutum 232, 234, 276
Puratelphusa hydrodromus 153
Sphueroidinu bulloides 2 1-3
Parathemisto libellulu 290
Syualus megulops 3 13
Pareuchuetu anturcticu 292, 293
Stuinforthia upertura 30
Pureuchueta norvegicn 285
Stuinforthiu,fus~formis 31
Pavonu 204
Stuinfbrthia SQQ. 21, 35
Penaeus monodon 138
Stenosemella ventricosu 3 10
Phueocystis pouchetii 281. 283.
Stetsoniu hovarthi 23, 43
286, 289
Stygiomedusa gigunteu 296
Phaeocystis SQQ. 248, 25 1, 260.
Stylophora pistillutu 196
267, 291. 292, 305, 306
Symbiodinium
Pheronema carpenteri 4 I, 42
microadriaticum 193
Phocu vitulina 308, 309
Phyllobothrium 170
Plucopecten magellanicus 30 1. 302 Temora 305
Planulinu ariminensis 42 Temora longicornis 25 1 1 305
Pleuragramma antarcticurn 296 Textulurirr krrttegutensis 30
Pocillipora damicornis 204. 205 Textulariida 5
Pocillopora bulbosu 189 Thulassiosira unturcticu 232, 267, 268
Pocilloporu cuespitosu 189 Themisto abyssorum 287, 289, 296
Pocilloporu dumicornis 189. 190. 194, 196 Themisto guudichuudi 296
Pocilloporu eleguns 189 Themisto libellula 287, 296
Pocilloporu meundrinu 189 Therugra chalcogrammu 309
Pocilloporu SQQ. 189, 208 Thysanoessu inermis 286. 287, 288.
Pontctstuacus leptoductylus 289, 300, 3 15
leptodactvhrs 100 Thysunoessu longicaudtztu 3 15
346 TAXONOMIC INDEX