Sie sind auf Seite 1von 24

DeepSea Research II, Vol.44,No.5,pp. 1085-l 108.

1997
Pergamon 0 1997 Elsevicr 8cieaa Ltd
All rights rewved. F’rinted in Great Britain
PII: s0967-0645(96)00111-7 09674%45/!37s17.00+0.00

Organic geochemistry of marine sediments of the Subantarctic Indian


Ocean sector: Lipid classes-sources and fate

J. LAUREILLARD,* L. PINTURIER,* J. FILLAUX* and A. SALIOT*

(Received 19 December 1994; in revtiedform 21 November 1995; accepted 8 December 1995)

Abstract-Nine sections of core KTB16 (473’98 S, W5974 E, 4240 m) taken from the Indian
Ocean sector of the Antarctic Polar Front Zone were analyzed for their lipid class and total chlorin
contents using thin-layer chromatography-flame ionisation detection (TLC/FID). The following
series were included: hydrocarbons, chlorins and chlorin esters, alcohols, sterols, triacylglycerols,
free fatty acids and phospholipids. For these major classes, the distribution of their individual
components was determined to evaluate the source and fate of each lipid class. Relationships between
the lipid composition of overlying particles and of buried material were investigated.
The sediment-water interface was found to be of primary importance in the biogeochemical
transformations occurring in organic matter, such as loss of the major part of the lipids, and
formation of a notable unresolved complex mixture (UCM), of biological origin. Some classes such
as chlorin esters appeared to convey their lipid contents from the upper layers without major
alteration. In contrast, several other classes, such as storage lipids, appeared to be. degraded during
sedimentation, indicating extensive recycling of labile lipids in the water column. The burial efficiency
and bacterial inputs downcore were highly variable depending on the class considered. Several novel
compounds are reported. They consist of alkyl chlorin esters and include a wide variety of saturated
and unsaturated long-chain alcohols.
Owing to the lack of lipid data in the study area, these results provide an insight into the various
biological and chemical processes occurring in open-sea Antarctic sedimentary environments.
0 1997 Elsevier Science Ltd.

INTRODUCTION
There is increasing interest in the relatively unpolluted areas of the world for studying
biogeochemical cycles. Thus, the Southern Ocean appears an attractive region since it is
distant from major sources of human pollution. Apart from characteristic features such as
very cold average temperatures and a strong seasonality, another interesting feature is the
almost complete absence of higher plants, except in a small part of the Antarctic peninsula.
The role of lipids in polar environments is of primary importance for understanding the
cycling of organic carbon and associated elements. However, the study of lipids has been
restricted to nearshore sites, unrepresentative of the Antarctic Ocean, and open-sea
measurements are few. Although lipids represent a small part of biogenic organic matter,
they are essential for marine organisms. Depending on their molecular structure, they play a
key role in membrane composition, energy storage and regulation of metabolic processes.
Particulate matter and sediments consist of both living organisms and detrital matter. The
lipid class distribution can give an insight into their relative proportions. Moreover, lipid

*Laboratoire de Physique et Chimie Marines, Unite de Recherche asso& au CNRS no. 2076, Universite Pierre
et Marie Curie, case 134, tour 25-24,4 Place Jussieu, F-75252 Paris cedex 05, France.

1085
1086 J. Laureillard et al.

classes have been proved to be useful tracers of the source and fate of the organic matter in
particles and sediments. They thus can contribute to further understanding of the transfer of
organic particulate matter to the sediments. Zooplankton fecal pellets and algal floes have
been shown to be the most effective means for the rapid transport of fresh organic matter to
the sea-floor. However, in spite of fast sinking rates of the highly silicified Antarctic diatoms
(Jacques, 1991) and of fecal pellets (Cadee et al., 1992) and the low temperatures (from 3” to
7°C on average at the sea surface, 0-1°C near the sediment), microbial degradation and
successive predations may significantly alter their lipid composition during transit through
the water column.
A comprehensive study of the biogeochemical processes taking place in the Antarctic
Ocean is being conducted in the Indian Ocean sector for the JGOFS France ANTARES
program. The aim of the first cruise, ANTARES 1 (April-May 1993), was to study the
sediments. At the present time, to our knowledge, there is no literature on lipid classes of
open-sea Antarctic sediments and more generally on lipid classes for a whole core. In order
to elucidate the relationship between the lipid composition of the large particles reaching the
ocean floor and the organic matter buried in the sediments, we have undertaken a study of
core KTB16 (47O59.98’S, 5Y59.74 E) and the particles suspended in the closely overlying
water. It is evident that the overlying particles recovered above the core do not represent the
whole annual input whereas sediment is a record of numerous years of organic matter
production and deposition. The main organic flux in the Southern Ocean consists essentially
of diatoms (Jacques, 1991; Riaux-Gobin et al., 1997) with the dominance of Fragiluriopsis
kerguelensis in the area studied (Pichon et al., in press). This site has been proved to contain
recently deposited material (Talbot, et al., personal communication, 1994), which might not
have penetrated deeply into the sediment. No lluff was found at the top of the core, and the
whole core has been shown to be oxic (De Wit et af., 1997). Additional meaningful
information as to the sources of the organic matter reaching the sediment was gained by
analysis of the molecular components of the major lipid classes.
The scope of this paper is twofold: first, to determine whether different lipid imprints
could be identified for primary production, and for autochthonous benthic life, and whether
each lipid class conveys a specific message from a given source; secondly, to provide an
insight into the importance of the bacterial reworking in sinking particles, at the sediment-
water interface, and in buried material. This information may enable organic compounds
that are readily biodegradable to be discriminated from those that are more resistant.

EXPERIMENTAL

Sampling

Sediments were collected in the Indian Sector of the Antarctic Ocean during the RV
“Marion Dufresne” expedition ANTARES 1 in April-May 1993. The location and depth of
the core studied, KTB16, were as follows: 47O59.98 S, 55O59.74’E, 4240 m. It was collected
on 23 April.
A multiple corer was used, which collected approximately the upper 30 cm. The sediment
core was cut horizontally on board at 4°C into samples of different thickness: l-cm sections
down to 5 cm; 2-cm sections down to 11 cm; and finally a 4-cm section (1 l-l 5 cm).
All samples were frozen at the time of extrusion and transported frozen at - 20°C to the
Organic geochemistry of marine sediments 1087

laboratory. Then the frozen sediments were freeze-dried and stored under argon at - 20°C
in the dark until analysis.
The use of a multicorer allowed the sediment-water interface to be recovered with very
limited perturbation (Legeleux et al., 1994). For better understanding of the ultimate state
of the sinking particles before deposition, the overlying water was siphoned from the top of
the corer and the suspended particles collected on a glass fiber filter (Whatman GF/F,
47 mm) that had been heated overnight at 450°C.

Lipid extraction

All chemicals and solvents were analytical grade. The solvents were distilled twice in
glassware, and chemicals were solvent-extracted. Glassware was cleaned by heating
overnight at 45O”C, and rinsed with solvent before use. The standard lipids for GC
identification were purchased from Sigma and Supelco Laboratories.
GF/F filters cut into small pieces or about 10 g of sediment (2 g for the uppermost slice)
were ultrasonically extracted (15 min, room temperature, under argon) by a modified one-
phase method (Bligh and Dyer, 1959), using methylene chloride instead of chloroform. A
solvent volume/dry sediment weight ratio of 8 (vol. dry wt.) was used for extraction of the
core and a ratio threefold higher for overlying suspensions. The solution obtained was
filtered on a fritted no.4 glass filter. The procedure was repeated with fresh solvents.
After phase-separation, the lipid extract was recovered in the lower methylene chloride
layer, dried by solvent-washed magnesium sulfate, and evaporated to dryness in a rotary
evaporator at a temperature lower than 30°C.

Zatroscan quantzjkation

A fraction of the total lipid extract was analyzed for lipid class composition with an
Iatroscan MKV TLC/FID analyzer (Iatron Laboratories, Japan), using a method with four
consecutive developing solvent systems similar to the multistep procedure previously

Table 1. Lipid class composition in pg/g dry sediment from the sections of core KTB16 (TLC/FID data)

Core
KTB16 O-1 cm l-2 cm 2-3 cm 34cm 4-S cm 5-7 cm 7-9 cm 9-llcm 11-15cm

HC 112.0 39.4 3.1 7.3 19.8 14.8 12.3 8.7 0.61


WE - tr - - tr - -
ME - - - - tr - -
FFA 14.6 6.2 5.5 3.3 3.5 2.4 2.4 1.8 0.62
TAG 5.1 0.71 - 4.8 1.3 - - 0.24
ALC 0.79 0.20 - 0.98 0.57 0.95 n.a. 0.30
ST 4.6 1.5 0.80 0.72 0.99 0.80 0.67 0.43 0.15
DAG 0.27 tr tr tr tr tr tr 0.14 -
MAC - tr - - - - - -
DPG + PG 72.1 31.2 38.3 16.7 22.1 46.5 2.0 9.5 3.9
PE 8.1 - 2.2 2.9 - 3.9 6.1 - 0.57
Z lipids 217.6 78.3 50.8 30.9 52.2 70.3 24.5 20.6 6.4
Pigments 1 0.79 0.37 0.45 0.35 0.43 0.36 0.39 0.11

ma. = not available. tr = detected at trace level.


1088 J. Laureillard et al.

Table 2. Lipid class distribution of


the overlying particles from the top
of core KTB16, expressed as a
percentage

Lipid class Percentages

HC 30.9
FFA 2.5
TAG 4.1
DPG + PG 41.2
PE 15.4
PIG tr

tr = detected at trace level

described by Parrish (1987). The data reported in Tables 1 and 2 are the result of three to six
measurements for each sample. Samples were spotted on SIII Chromarods with a l-p1
automatic syringe (SES, Germany). The data were processed using the “Boreal” software
developed by JMBS, Grenoble. The operating conditions were: hydrogen flow 160 ml/min,
air flow 1800 ml/min and scanning time 30 s. The response of the FID was calibrated using
external standards covering the concentration range in the samples (0.1-5.0 pg).
Hydrocarbons (HC), wax esters + steryl esters (WE), fatty acid methyl esters (ME) and
free fatty acids (FFA) were separated with hexane-diethyl ether-formic acid (98:2:0.5).
Triacylglycerols (TAG), alcohols (ALC), sterols (ST) and diacylglycerols (DAG) were
separated with hexane-diethyl ether (87: 13). Chlorins (PIG) and monoacylglycerols (MAG)
were roughly separated with diethyl ether-acetone (57:43). A fourth developing system,
chloroform-methanol-water (57:47: 1.4) permitted separation of diphosphatidylglycerols
(DPG) + phosphatidylgiycerols (PG) and phosphatidylethanolamines (PE) from nonlipid
material that remains at the origin. Separate calibration curves were established for each
lipid class, using the following standards: HC: eicosane (Chemical Sample Company); WE:
cholesteryl palmitate (Sigma Chemical Company); ME: methyl behenate (Applied Science);
FFA: stearic acid (Sigma); TAG: tripalmitin (Sigma); ALC: stearic alcohol (Sigma); ST: fi-
sitosterol (Sigma); DAG: 1,3-dipalmitin (Sigma); PIG: chlorophyll a (Sigma); MAG:
monopalmitoylglycerol (Sigma); DPG: diphosphatidylglycerol sodium salt (Sigma); PG:
phosphatidylglycerol ammonium salt (Sigma); PE: dipalmitylphosphatidylethanolamine
(Sigma). The Iatroscan TLC/FID chromatogram obtained for the 5- to 7-cm fraction is
shown in Fig. 1. Pigments appeared in the Iatroscan pattern as a broad peak coeluting with
an authentic standard of chlorophyll a. MAG appeared as a small rise in the peak tail, and
separate quantitative measurements proved to be impossible. Owing to the sole co-elution
of MAG that was present at low levels, chlorin measurements using a chlorophyll a
calibration curve should not have been greatly in error. However, the solvent mixture used
for the lipid extraction is not appropriate for pigments. Thus, the pigment concentrations in
Table 1 are normalized with respect to the highest value.

Lipid class separation

After TLC/FID analysis, the total lipids were separated into different classes by column
chromatography on activated silica gel (Kieselgel 40, 70-230 mesh, Merck). Sequential
Organicgeochemistryof marinesediments 1089

FID

NLM
DPG+PG
!Ju)

HC

PE

i L

Fig. 1. Iatroscan TLC/FID chromatogram for the 5-7 cm section of core KTB16. HC =
hydrocarbons, FFA = free fatty acids, TAG = triacylglycerols, ALC = alcohols, ST = sterols,
DAG = diacylglycerols, PIG = chlorins, MAG = monoacylglycerols, DPG = diphosphatidyl-
glycerols, PG = phosphatidylglycerols, PE = phosphatidylethanolamines, NLM = non lipid
material. The phospholipid pattern has been reduced by five on the figure.

elutions with solvent systems of increasing polarity were used to obtain the different lipid
classes. The first hexane fraction contained HC. WE and TAG were collected in the second
non-polar fraction using hexantiiethyl ether (9: 1) as eluent. ALC and ST were eluted with
hexane-ether (85: 15). Then, after a diethyl ether-acetone-acetic acid (7:3:0.1) elution of a
fraction containing degradation products such as DAG, MAG and FFA, the polar lipids
were finally collected with methanol.

Isolation of alcohols - sterois and chlorin esters


ALC and ST were separated by preparative thin-layer chromatography of the total lipid
extract on 20 x 20 cm 150 A silica gel plates (Whatman) precleaned twice in methanol.
Double elution with hexane-ether-acetic acid (75:25: 1, v/v), followed by an ether-acetone
(50:50, v/v) 5-cm-elution, separated free alcohols-sterols on one hand and chlorin esters on
the other hand from other lipids. The band supposed to contain chlorin esters was the
chlorophyll-colored one. The two bands were separated by a 4-cm interval. After
transesterification of the esters and purification by thin-layer chromatography,
1090 J. Laureillard rt al.

derivatization of the two mixtures of alcohols-sterols led to the trimethylsilylether (TMS)


derivatives (see below).

Fatty acid and sterol analysis

Natural lipids with fatty acid moieties were transesterified with 14% boron trifluoride in a
methanol-toluene mixture (2:l) under nitrogen for 1 h at 65°C. Analysis of the fatty acid
methyl esters was performed on a Varian 6000 gas chromatograph (GC) fitted with a Ross-
type injector at 250°C and a DB5 30 m x 0.32 mm i.d. non-polar capillary column and a
flame ionization detector. The column was temperature-programmed from 50°C to 100°C at
25”C/min and from 100°C to 300°C at 2”C/min. To obtain further information on the
composition of the fatty acids, GC analysis of their methyl esters was also performed on a
DB23 30 m x 0.25 mm i.d. polar capillary column fitted on a Girdel320 gas chromatograph
equipped with a Ross-type injector and a Ilame ionization detector. The temperature
program was from 50°C to 240°C at 1.S”C/min. Fatty acids are designated as x:y(n-z)
where “x” is the total number of carbon atoms, “y” is the number of double bonds, and “z”
is the position of the double-bond nearest to the terminal methyl group.
The series of acids are characterized by their Carbon Preference Index (CPI) value,
defined as follows:
CPI = C[even n - acids Cl2 - C2s]/C[odd n - acids Czi - C27]
Sterols were converted into their corresponding TMS derivatives by treatment with
bistrimethylsilyltrifluoroacetamide with 1% trimethylchlorosilane (Supelco) for 1 h at
70°C. Sterol TMS derivatives were analyzed using a Carlo-Erba gas chromatograph fitted
with a flame ionization detector and an on-column-type injector. Separations were
performed on a DB5, 30 m x 0.32 mm id. capillary column. The GC was programmed
from 50°C to 100°C at 25”C/min and from 100°C to 300°C at S”C/min. For GC, the carrier
gas was helium with an inlet pressure of 2 bars. Fatty acid and sterol derivatives were
identified by comparison of their GC retention time and mass spectra data with those
obtained from authentic and laboratory standards. Peak areas were quantified with APEX
integrating software.
Procedural blank runs showed minor contaminants that would not interfere with the
analytical data. All samples and lipid extracts were always protected from light, heat and
oxygen and stored under argon at - 20°C.
Owing to the small amount of overlying particles recovered above the core, it was not
possible to obtain a precise weight; the data are thus reported as the relative abundance of
the compounds detected (Table 2).

Gas chromatography/mass spectrometry (GCIMS)

GC/MS was performed using a Girdel 32 gas chromatograph fitted with a Ross-type
injector and a DB5 30 m x 0.25 mm id. column fed directly into the electron impact source
of a Nermag RIO-1OC quadrupole mass spectrometer. The oven temperature was
programmed as described above. Typical GC/MS operating conditions were: ionization
potential 70 eV, source temperature 150°C filament current 200 PA, and scanning rate from
50 to 600 Da/s. Data were processed using a PDP 1l/23 + /Sidar 111 system.
Organic geochemistry of marine sediments 1091

RESULTS AND DISCUSSION

Lipid class origins

Lipid classes - bulk data. The Iatroscan TLC/FID method provides an overview of the
total lipid content of a sample and a useful basis for the selection of further investigations
such as focusing on the detailed composition of a single class of lipids considered as the
appropriate tracer for a given purpose.
The quantitative analysis of lipid classes provides insights into the composition and
physiologic state of organisms in natural samples (Parrish, 1988; Goutx et al., 1990). They
consist of three groups of compounds according to their role in the cell:

- neutral lipids (alcohols, sterols, hydrocarbons and storage lipids: wax + steryl esters
and triacylglycerols) that are constituents of organisms,
- polar lipids, components of membranes, representative of living cells (including
phospholipids (PL) and glycolipids associated with chloroplast membranes),
- degradation products such as free fatty acids, mono- and diacylglycerols.

Lipid class distributions are shown in Table 1. A striking feature is the high concentration
of HC in the O-l cm level, which accounted for 51.5% of the total lipids. Downcore, the
major lipid class varied according to depth, with CPL (DPG + PG + PE) or HC prevailing.
TAG were present occasionally, and practically no WE were detected, even in the
overlying particles (Tables 1 and 2). However, the strong seasonality of primary production
in Antarctic waters with short periods of food abundance alternating with long periods of
food scarcity induces the formation of storage lipids as an efficient form of energy storage
for organisms. Phytoplankton synthesizes TAG almost exclusively, whereas zooplankton
and animals may accumulate either WE (herbivorous copepods: Conover and Huntley,
1991) or TAG (krill: Virtue et al., 1993; Sargent and Falk-Petersen, 1981; Fricke et al., 1984;
mid-water fish and invertebrates: Reinhardt and Van Vleet, 1986). However, a large number
of Antarctic organisms store more lipids as TAG than as WE. The major WE-synthesizing
animals are copepods. Bacteria are generally believed not to produce storage lipids. Our
analyses showed that polar lipids consisted exclusively of PE, DPG and PG; they always
represented a large part of total lipids. As they accounted for a major proportion of bacterial
lipids ( > 90%), in the absence of reserve compounds for some samples, they are probably
produced by live bacteria.
Significant amounts of chlorins were also identified whose presence was also indicated by
the deep green color of the total lipid extracts up to 11 cm. However, glycolipids, another
important component of chloroplasts, were not detected.
As seen in Fig. 2, no strict correlation existed between depth profiles of total lipids and
organic carbon (from Rabouille et al., 1997). The profiles of the main lipid classes, HC and
CPL, varied differently with depth (Table 1). Three groups may be distinguished:
- HC, pigments, sterols, which decreased at 2-3 cm and to a lesser extent at 3-4 cm,
- TAG encountered only at some levels,
- PL and FFA whose overall sediment profiles were quite different from those of the
above compounds.

In order to assess the origins and fate of the various lipid classes displaying different
trends (Table I), their molecular constituents were analyzed and are discussed below.
1092 J. Laureillard et al.

cqaniccarbon
........ +. . ...
iow.l lipids

I I I
5 10 IS

depth (cm)
Fig. 2. Depthprofiles of total lipids and organic carbon from core KTB16.

Hydrocarbons. The first centimeter of sediment was characterized by a high concentration


of hydrocarbons (Table 1): 112.0 pug/gsediment dry wt. For the lower sections, hydrocarbon
concentrations varied from 39.4 to 0.61 pg/g. These levels are considerably higher than the
low natural levels found in Antarctic sediments, generally ranging from 0. l-l ,ug/g (Indian
Sector, Ltitzow-Holm Bay, Matsumoto et al., 1992b; Harada et al., 1995, and Pacific Sector:
61”34’S, 150”23’E, Matsueda and Handa, 1986, respectively) to l-9.3 pg/g in the Weddell
Sea (Danyushevskaya et al., 1989).
The concentration of total non-aromatic hydrocarbons varies from a few tenths of rig/g in
unpolluted marine sediments to a few rig/g.. In the case of very high natural inputs or
autochthonous biological productivity, this concentration may rise to 100 pg/g.
Concentrations higher than 100 pg/g are generally attributed to petroleum inputs
(Bouloubassi and Saliot, 1993). However, the Antarctic has been found to be a pristine
environment (Venkatesan, 1988) even though a petroleum imprint background has been
shown by aerosol particle analysis near the Drake Passage (Simoneit et al., 1991), as well as
some thermogenic hydrocarbon seepage in the Bransfield Strait (Whiticar et al., 1985).
Local pollution has been restricted to rare offshore oil spillages (Cripps and Priddle, 1991).
In order to assess the origin of these HC, further detailed investigations were performed.
The chromatograms (Fig. 3) were overwhelmingly dominated by a broad unimodal hump,
with a maximum at C&&i9 corresponding to an unresolved complex mixture (UCM), with
a series of resolved hydrocarbons. The detailed results of their analysis will be reported
Organic geochemistry of marine sediments 1093

S.

Time
Fig. 3. Gas chromatogram of hydrocarbons for the 0-l cm section of core KTBld IS. = internal
standard.

elsewhere (in preparation). The concentration of this unresolved mixture decreased


downcore and it could not be detected below 11 cm. The presence of a UCM hump in
hydrocarbon chromatograms is often considered the most convincing diagnostic tool of
petroleum contamination. However, the possibility of a biological origin has been suggested
by Venkatesan and Kaplan (1982) for the envelope found in hydrocarbon chromatograms
of sediment samples from the Gulf of Alaska. These authors assigned this UCM to
microbially altered algal detritus; but later, Volkman et al. (1992) contested this assessment.
The sedimentation rates are relatively high in the Indian sector of the Southern Ocean:
from 4 to less than 10 cm per 1000 years (Reyss et al., 1995). Nevertheless, a massive
anthropogenic contamination of surface sediment up to the 1l-cm level is impossible. Thus,
a natural origin undoubtedly can be ascribed to this UCM, probably derived from recent
biological debris sinking through the water column in this pristine environment.
1094 J. Laureillard et al.

Analysis of the HC fraction from the overlying particles gave a chromatogram displaying
a series of very long-chain n-alkanes (VLCHC) up to Car with minor branched homologues
but without any trace of hump. These VLCHC might have an algal origin, since n-alkane
chain lengths reported in sea-ice diatom communities range up to Cs7 (Nichols et al., 1988).
However, animals could be other possible precursors, since Reinhardt and Van Vleet (1986)
and Cripps (1990) reported the occurrence of straight chain and branched VLCHC in
Antarctic zooplankton and fish with chain lengths up to C36C39. Moreover, at least a part
of these VLCHC could be derived from heterotrophic bacteria. In fact, straight-chain HC
up to Cs6, in addition to branched hydrocarbons detected in the sediments from Ltitzov-
Holm Bay (Matsumoto et al., 1992b) and in McMurdo Dry Valleys, Antarctica
(Matsumoto et al., 1992a), have been reported to have a microbial origin.

Pigments. Another important component of the bulk lipid extract was pigments
(chlorins). The depth profile of pigments (Table 1) shows a regular slope with a slight
decrease for the 2-3 cm fraction, then a more or less constant value before a sharp drop
below 11 cm. Chlorins consist of chlorophylls and their degradation products,
phaeopigments. Phaeopigments are formed in the euphotic layer during phytoplankton
senescence or zooplankton grazing, and continue to degrade slowly inside the sinking fecal
pellets and the sediment. Recently, the occurrence of significant amounts of steryl chlorin
esters (up to 20% of the sedimentary chlorins) has been reported in marine (Black and Baltic
Seas) and lacustrine sediments (King and Repeta, 1991a,b; Eckardt et al., 1992). The sterols
are esterified to a single chlorin nucleus, pyrophaeophorbide a. The corresponding
environments of eutrophic- or anoxic-type (Pearce et al., 1993) tend to involve a rapid
transfer of sinking particles and subsequent good preservation. Thus, the Antarctic is likely
to display the same opportunities for the occurrence of chlorin esters in sediments.
This assumption is confirmed by the results shown in Table 3, which gives the free and
esterified hydroxylated lipids for the overlying particles and the 9-l 1 cm sediment sample.
In order to ascertain the esterified nature of the hydroxylated compounds retrieved using
column chromatography in a fraction containing MAG, separation by plate thin-layer
chromatography of the total lipid extracts from the two samples was also performed (see
Experimental). This gave the same results as the column. In addition to sterols, a suite of
predominantly even n-alcohols of carbon chain length up to CX2,along with branched and
monounsaturated homologues, was esterified as chlorin esters.

Sterols. The depth profile of sterol concentrations (Table 1) was similar to that of HC,
with a net decrease for the 2-3 cm and 3-4 cm sections. The main sterols observed are
reported in Table 3. Close correlations between sterol distributions in free and esterified
fractions were found at both levels. Cholesterol predominated in all fractions and its
concentration ranged from 40.7% to 71.9% of total sterols. The next most abundant sterol
was 24-ethylcholesterol. The only 4-methyl sterol detected was dinosterol (4a,23,24-
trimethyl-5a(H)-cholest-22E-en-3/3-01); its presence was significant in the sterols of
overlying particles where it accounted for 33% and 28.1% of the total free and esterified
sterols, respectively.
Since steryl chlorin esters are thought to be formed in the euphotic layer from decaying
phytoplankton (Eckardt et al., 1992) or from zooplankton grazing on algae (King and
Repeta, 1991a,b), the sterol distribution can convey some source-related information.
Sterols are synthesized in algae, in higher plants and animals. The amounts of sterols (% dry
Organic geochemistry of marine sediments 1095

Table 3. Alcohol and sterol distributions in the free and esteriJedfonns in the overlying particles and the 9-l 1 cm
section for core KTB16

Overlying particles 9-llcm

Free % Esterified % Free % Esterified %

Alcohols

14:o 1.6 6.1 4.6 1.9


15:o 0.4 - 4.8 9.2
16:0 14.4 6.9 15.9 9.3
17:0(br) 0.4 - 9.5 5.2
17:o 1.0 - 1.0 30.0
18:l 5.1 - 5.3 1.8
18:l 2.8 tr 2.5
18:0 10.8 21.8 6.0 5.6
Phytol 1.1 18.7 3.1 1.2
19:0(br) - 1.8
19:0(br) 0.6 - 1.5 -
19:o 0.7 3.3 0.4 10.3
2O:l 0.4 - 0.5 0.4
20: 1 - - 0.2 -
20:o 1.3 1.9 1.3 0.8
21:0(br) 1.2 - 0.4 0.1
21:o 0.8 - 0.3 1.0
22:l - - 1.4 0.7
22: I - 0.6 0.1
22:0(br) 2.7 11.8 1.4 4.9
22:o 1.5 - 3.9 1.1
23:0(br) 4.8 - 0.3 0.5
23:0 0.8 - 1.2 1.4
24: 1 - 0.2 1.2
24: I - 1.5 -
24:0 2.8 3.2 0.5 1.4
25:0 - 0.5
26:l - 0.9 3.6
26:0 6.1 7.2 7.9 1.9
27:0 15.4 - 1.7 0.4
28:0 12.3 10.5 11.4 3.3
29:0 1.2 - 5.5 0.6
30:o 7.2 8.5 1.4 1.9
32:0 2.4 - 0.6 0.3
Branched 9.7 0 14.9 10.5
Unsaturated 8.3 0 13.1 7.6
~,4~20/~21c32 0.69 1.42 1.40 3.18

Sterols
22-Dehydrocholesterol 3.1 - 6.6
Cholesterol 40.7 71.9 50.7 58.3
24-Methylcholesta-5,22E_dien-3B-ol 2.9 - 7.0 4.6
24-Methylcholesterol 2.4 - 8.1 3.3
ZCEthylcholesterol 18.0 tr 21.0 30.4
Dinosterol 33.0 28.1 6.6 3.3

tr = detected at trace level; br = branched.


1096 J. Laureillard et al.

wt.) found in the prokaryotes are substantially lower than those of eukaryotes. However, in
the bacterium Methylococcus cupsulatus, grown on methane, sterols are comparatively
abundant, and they essentially consist of 5a-cholesta-8(9),24-dien-3/I-01 derivatives (Bird et
al., 1971). Thus, as a significant terrestrial input can be excluded in this area, sterols can be
used to infer planktonic inputs. Cholesterol is the major sterol in marine fauna and in some
dinoflagellates and diatoms (Tsitsa-Tzardis et al., 1993). Moreover, significant amounts are
present in prymnesiophytes and cyanobacteria (Volkman, 1986). 22-Dehydrocholesterol,
the major sterol in a cultured sea-ice diatom Nitzschia cylindrus (Nichols et al., 1986) and in
situ sea-ice diatoms at Erebus Ice Tongue (Nichols et al., 1993b) also occurs at significant
levels in McMurdo Sound sediments (Smith et al., 1989). 24-Methylcholesta-5,22E-dien-
38-01 is the dominant sterol in sea-ice diatoms at Cape Armitage (Nichols et al., 1993b).
Dinosterol has been considered to be a specific marker of dinoflagellates, although Nichols
et al. (1990) demonstrated it to be a minor constituent in Antarctic diatoms. Campesterol
(24-methylcholesterol) is encountered most frequently in green algae, diatoms,
dinoflagellates and mosses (Matsumoto and Kanda, 1985). 2CEthylcholesterol is
generally considered a higher-plant marker (Volkman, 1986). Despite the minimal aeolian
transport of terrestrial plant debris, this sterol, perhaps with another C24_stereochemistry,
has often been encountered in Antarctic lacustrine and marine samples. This imprint has
been attributed to cyanobacteria (Matsumoto et al., 1991), green algae (Matsumoto et al.,
1983) or mosses. Nevertheless, this sterol is a major component of Antarctic diatoms
(Orcutt et al., 1986; Volkman, 1986; Nichols et al., 1989, 1993b). In the present study, only
low amounts of stanols were detected in particles reaching the sea-floor and in the surface
oxic sediments, indicating that in situ biohydrogenation of sterols was only a minor
transformation process.
As a result, all the sterols found in free and esterified forms occur in diatoms often as
major constituents and appear to be derived from these algae. Other contributors seem to be
minor with the exception of animals, which must contribute the dominant cholesterol, and
to a lesser extent dinoflagellates. A small contribution from prymnesiophytes cannot be
excluded because 24-methylcholesta-5,22E-dien-3/I-ol is commonly present at a very high
level in these algae (Nichols et al., 1991), and they are likely to be encountered in
oligotrophic Antarctic areas.
Our sterol data from open-sea sediments are rather different from Antarctic nearshore
sediment findings where the dominant sterols are campesterol (Brault and Simoneit, 1988)
and dinosterol (Venkatesan and Kaplan, 1987), respectively, in the Bransfield Strait, and
cholesterol and a C&Tdienol in McMurdo Sound (Venkatesan, 1988).

Alcohols. n-Alcohols were found in the carbon range from Ci4 to Cs2 with the major
components varying according to the samples (Table 3). n-Alcohols are generally
encountered in low concentrations in organisms, and a pattern in the shorter molecular
chains maximizing at Cl6 is generally observed for phytoplankton, whereas the maximum
moves to CZOfor zooplankton (Ogura et al., 1990). However, considerable amounts of
alcohols esterified as WE are found in copepods and especially in polar calanoid copepods.
The carbon range of alcohols is a function of the copepod feeding mode (Sargent and Falk-
Petersen, 1981; Graeve et al., 1994), and these alcohols are found in their feces (Prahl et al.,
1984; Harvey et al., 1987; Bradshaw et al., 1989). Besides, during fish predation on copepods
a major part of the dietary WE is assimilated. However, in fish feces a significant increase in
free alcohols is observed, along with a decrease in WE, compared to the composition of the
Organic geochemistry of marine sediments 1097

diet (Prahl et al., 1985). Thus, the main sources of free alcohols in the sea might be hydrolysis
of copepod WE in fish guts and their subsequent transport in the water column, embedded
in their fecal matter.
However, this source cannot account for the very long-chain alcohols. In fact, the
distribution of alcohols present in Antarctic sediments shows strong similarities with the
distribution pattern obtained in fecal pellets of the copepod Calanus helgolandicus fed on the
dinoflagellate Scrippsiella trochoidea (Harvey et al., 1987). These striking similarities in the
fatty alcohol distribution argue for the formation of chlorin esters in animal guts since these
esters involve alcohol moieties only occurring in fecal pellets. Thus, some alcohols were
attributed to a contribution by the copepod and some others to algal diet. However, a large
number of alcohols not present in copepods or algae appeared in significant amounts in
fecal pellets and especially branched and long-chain compounds. The appearance or the
increase in these alcohols also occurs in other marine invertebrate feces (Bradshaw et al.,
1989, 1990, 1991). A bimodal distribution maximizing at Cl6 and C26 was observed for
alcohols in the Bransfield Strait sediments by Venkatesan and Kaplan (1987). The major
monoene alcohols, 22: 1 and 24: 1, have been ascribed to copepods or bacteria. Furthermore,
Volkman et al. (1986a) suggested that long chain alcohols found in Ace Lake (Antarctica)
could originate from sedimentary bacteria and not from water column organisms.
Nevertheless, given their occurrence in chlorin esters, hence a surface origin, enteric
bacteria in animal guts as well as zooplankton WE seem likely sources of long- and medium-
chain alcohols, respectively.
The occurrence of phytol was observed in all samples. In chlorin ester data, it reflected the
presence of coeluting phaeophytins; it was released during ester hydrolysis. In free alcohols,
the low percentage of phytol (1.1 and 3.1%) may result from phaeophytin degradation
before analysis, probably during core recovery and treatment, free phytol being rarely
detected in sediments (Ogura et al., 1990).

Triacylglycerols. Antarctic animals have been shown to contain high lipid levels, mostly
due to the high proportion of neutral lipids for long-term storage of the wintering energy
reserves: WE and TAG. In addition, Antarctic phytoplankton contains large amounts of
TAG (Nichols et al., 1986; Smith et al., 1989). However, this lipid class accounted for only
4.1% of the total lipids in the overlying particles and 2.1% in the surficial sediment section.
Among nine core fractions analyzed, only five contained a detectable amount of TAG
(Table 1). Furthermore, WE were only encountered at trace level in the 3-4 cm and 7-9 cm
sections.
Molecular analysis (Table 4) revealed that these low amounts of triacylglycerols did not
arise from diatoms, even in the first centimeter. Thus, in our samples, TAG contained a very
low proportion of 16:lw7 (0.3-9.7%), with X@Cis ratios rising to 2.0, and the 16:lw7/
16:O ratios to 0.3. Meanwhile, characteristic features of diatoms include high proportions of
16:107, and low proportions of C rs acids, resulting in significant CC1&C1s ratios.
Moreover, the 16: 107/16:0 ratio is usually > 1 (Viso and Marty, 1993). For 23 different
cultured species analyzed by Volkman et al. (1989), Viso and Marty (1993) and Dunstan et
al. (1994) the amount of 16:107 ranged from 13.6 to 34.4%, the CCr&Crs ratio ranged
from 2.6 to 18.3 and the 16: 107/16:0 ratio from 0.54 to 5.1. The fatty acid distribution
patterns of Antarctic diatoms are in this respect very similar to those of the temperate region
species with 35.5% of 16:lw7, a CCi&Cis ratio of 15.3 and a 16:107/16:0 ratio of 2.9, for
the sea-ice alga Nitzschia cylindrus (Nichols et al., 1986). In situ samplings in the Antarctic
1098 J. Laureillard et al.

Table 4. Characteristics of the triacylglycerolfractionsfor core KTB16 and the overlying particles

KTB16 Overlying particles c-1 cm 2-3 cm 4-5 cm 5-7 cm

%ZSaturated 72.5 42.9 60.9 59.7 63.9


%CBranched and cyclic 8.2 13.9 10.0 6.7 11.2
%ZMonounsaturated 14.7 38.3 26.6 31.4 21.1
%ZPUFA 4.6 4.9 2.5 2.3 3.8
CPI 1.5 3.1 2.5 2.2 I.3
Chain length maximum 30 28 26 26 30
%Length > Cl0 9.1 5.5 8.1 10.2 12.3
%16:lw7 0.6 0.3 1.4 9.1 1.0
16:107/16:0 0.02 0.01 0.06 0.3 0.04
~:c16/~‘cl8 2.0 1.3 1.1 1.6 1.8

(Nichols et al., 1993b) corroborated these findings with 19.5 to 34.5% of palmitoleic acid, a
X16/~C18 ratio from 6.9 to 11.1 and a 16:1w7/16:0 ratio from 1.3 to 1.9. The 16:107 level
was even as high as 16.1 to 34.2% with a 16: 107/ 16:Oratio from 1.Oto 1.8 in samples where
diatoms predominated but were contaminated (Fahl and Kattner, 1993). Consequently, in
our samples, TAG might have a benthic animal source, since animals are another type of
organism likely to contain triacylglycerols. Usually the storage lipids mirror the animal diet
and their analysis is used to assess food webs. Here, the important level of the sum of
branched and cyclic fatty acid methyl esters ~-- essentially 15 and 17 iso and anteiso, as well
as 17 and 19 cyclopropyl fatty acid derivatives --- considered to be bacterial biomarkers,
indicated a bacteriophage regime. Moreover, the food choice of some predators towards
energy-rich particles is indicated by a polyunsaturated fatty acid (PUFA) level in
triacylglycerols higher than in free fatty acids or phospholipids.

Phospholipids. The PL profile is different from those of the preceding lipid classes (Table
1). It decreases generally with depth from 80.2 to 4.5 pg/g, with a slight rise at the 2-3 level
and a significant rise to 50.4 pg/g at the 5-7 cm level. PL decay rapidly after death of cells
(Harvey et al., 1986); consequently, PL reflect living organisms. Here, the meiofauna and
bacteria are possible candidates, along with a minute amount of phytoplanktonic cells in
sinking particles or in the first centimeter of sediment. In fact, low values for 16:107 and low
CC&Cis and 16: 107/l 6:0 ratios (Table 5) argue against a significant occurrence of live
diatom cells. The PL of the living cells associated with the particles reaching the sediment
(overlying particles) were found to be very different from the membrane lipids of the
organisms inhabiting the sediment. The major differences were: higher levels of saturated
fatty acids, smaller amounts of monounsaturated and branched + cyclic compounds
including 0.8% of i-25 and i-26 acids, and longer-chain fatty acids. Nevertheless, for long-
chain acids, the distribution in surficial sediment was closer to particles than to deeper
sections, providing evidence of some live populations of the same nature, either bacterial or
animal, present at both levels. Moreover, in the 5-7 cm section, the low percentage of
branched and cyclic compounds associated with the presence of TAG and the highest value
for PUFA (3.9%) suggests the occurrence of meiofauna. Hence some bioturbation might
have occurred, causing some mixing. Nevertheless, the smooth drop shown by various lipid
class profiles (Table 1) argues for a limited perturbation biasing the data.
Organic geochemistry of marine sediments 1099

Table 5. Characteristics of the polar lipidfractionsfor core KTB16 and the overlying particles

KTB16 Overlying
particles &I cm l-2 cm 2-3 cm 3-4cm 4-5 cm 5-7 cm 7-9 cm ell cm 1l-15 cm

%ZSaturated 89.8 55.9 50.6 58.1 48.0 51.6 58.8 52.0 76.3 63.8
O&Branched and cyclic 2.6 7.3 14.7 11.6 16.1 14.0 11.2 10.5 13.0 13.7
%ZMonounsaturated 5.8 34.1 32.3 28.4 34.1 33.4 26.2 36.6 9.8 21.3
%ZPUFA 1.8 2.8 2.4 1.9 1.8 1.1 3.9 1.0 1.0 1.3
CPI 1.9 1.6 1.4 1.8 2.7 1.5 2.2 2.8 3.7 2.1
Chain length maximum 30 30 30 28 26 26 28 30 32 28
%Length > CzO 25.5 19.6 7.0 8.4 8.1 7.9 22.7 15.7 26.1 13.1
%16:107 0.5 8.9 7.3 6.5 8.9 8.0 6.2 6.1 1.2 5.7
16:107/16:0 0.01 0.4 0.3 0.3 0.4 0.3 0.3 0.3 0.04 0.2
=k/~C18 1.7 1.5 1.6 1.5 1.8 1.7 1.5 1.3 2.4 2.0

Some Antarctic bacteria are capable of producing PUFA at low levels (Nichols et al.,
1992, 1993a) in the same way as barophilic bacteria (Delong and Yayanos, 1986)
presumably as an adaptive mechanism to maintain membrane fluidity. These bacteria
could be encountered throughout the core, and decrease proportionately downcore.
However, the level of branched and cyclic bacterial markers remained unaffected by
depth. In addition, some bacteria populations, characterized by the same features as
continental higher plant compounds (high CPI associated with a major proportion of long
chains), increased in abundance with depth. These results are in good agreement with the
data of Volkman et al. (1988) and Matsumoto et al. (1992a) who reported a high abundance
of long-chain acids in Antarctic lake sediments, which they ascribed to bacteria. These
findings suggest a downcore succession of bacterial species with different lipid
characteristics.

Free fatty acids. FFA accounted for 14.6 pug/g in the surficial sediment, decreasing
downcore to 0.62 pg/g (Table 1). The PUFA level was significant although very low for FFA
and ranges from 0.4% (overlying particles) to 4.2% of total acids. This latter value was even
higher than the PUFA level in any PL fraction. In fact, free fatty acids reflect detrital matter
as opposed to living cells. Usually, in most living organisms, FFA are only detected at a low
level. The occurrence of FFA in diatom cultures has been proved to be an artifact. Free acids
result from lipase activity on PL due to cell rupture during isolation of algae (Berge et al.,
1995). The observation of PUFA in FFA seems to be unreliable, since PUFA are rapidly
oxidized and cannot survive unaltered for more than a few days in an oxic environment. So,
the PUFA detected in FFA very probably result from lysis of benthic organism cells during
corer recovery from 4240 m, leading to some extent to the degradation of PL into FFA
containing PUFA. Consequently, the FFA seem to result from mixed sources: detritus and
PL decomposition resulting from cell rupture.
The composition and distribution of the FFA encountered in the overlying particles
exhibited some differences compared to the FFA encountered in the sediments (Table 6),
with higher levels of saturated fatty acids, 75.8% against 4060% in the buried particles.
Moreover, characteristics of particle FFA were close to the features of the corresponding
TAG and PL, pointing to a common origin. The very low level of the 16:1~7 percentage and
1100 J. Laureillard et al.

Table 6. Characferislics of the free fatty acidfractions for core KTB16 and Ihe overlying particles

KTB16 Overlying particles o-1Cm l-2 cm 2-3cm 3-4cm 4-scm

%ZSaturated 75.8 58.2 40.0 50.3 48.8 52.4


%ZBranched and cyclic 6.0 6.9 11.2 10.8 13.3 12.5
%ZMonounsaturated 17.9 30.7 46.7 38.1 36.6 34.1
%ZPUFA 0.4 4.2 2.1 0.9 1.4 1.0
CPI 1.5 1.6 7.0 5.4 6.4 5.1
Chain length maximum 30 30 34 34 34 30
%Length > COO 15.3 7.7 21.8 27.4 24.0 27.9
%16:107 1.6 I.1 15.0 12.8 11.3 10.3
16:lw7/16:0 1 0.05 1.5 1 0.94 0.76
~C&CIR 0.91 1.4 1.6 1.7 1.4 1.5

of the CC1&C1s and 16:107/16:0 ratios argue against diatom detritus. Nevertheless,
despite the overlying particles representing a short period of deposition, in many respects
their composition was similar to that of the 0-l cm fraction of the core with a comparable
CPI value, percentage of long-chain compounds, level of branched + cyclic acids and of
16:107 acid. In addition, there were marked similarities in the surficial sediment between
FFA and PL (Tables 5 and 6). Consequently, this set of findings demonstrates that particles
and surficial sediments are derived from closely related sources, probably of benthic animal
and/or epifaunal and bacterial origin. Downcore, as shown by the sudden variations in the
lipid data (Table 6), the FFA were of a different origin, reflecting essentially older detrital
matter.

Degradation in water column and sediments

Lipid classes. A more pronounced decrease in some lipid classes and especially HC and
sterols (Table 1) was observed in the 2-3 cm and 3-4 cm sections and in some extent for
pigments (2-3 cm sample). No abnormal features in the bacterial profiles were observed,
either for microbial respiration (De Wit et al., 1997), or for proteolytic activity (Talbot and
Bianchi, 1997) in this part of core KTB16. These findings may suggest variations in the
intensity and the composition of the organic matter flux, leading to a decrease in some lipid
classes.
Nevertheless, a sediment slippage or particle flux with inputs of different composition
cannot be excluded.

Hydrocarbons. UCM is considered to be a multitude of branched and cyclic compounds


unresolved by capillary columns. It is well known that n-alkanes are more rapidly degraded
than the branched-cyclic species constituting the petrogenic UCM. Gough and Rowland
(1990) showed that synthesized T-branched alkanes were as resistant as petroleum UCM
and highly branched acyclic isoprenoids, and more resistant than normal and monomethyl-
branched compounds. In that study, only resolved hydrocarbons were observed in the 1l-
15 cm sample. The possibility of variability in the composition of sedimentary flux leading
to the absence of UCM for such a large timespan seems unlikely. Indeed, this complex
Organic geochemistry of marine sediments 1101

mixture accounted for half the concentration of total lipids in the surface sediment (O-1 cm).
Consequently, these findings underlined the labile character of this biogenic UCM, more
prone to bacterial attack than n-alkanes. Thus, the degradability of this natural UCM is
higher than that of a petroleum UCM. Hence, its composition ought to be different. Organic
sulfur compounds, less stable than hydrocarbons, can give rise to a UCM (Volkman et al.,
1986b), but high concentrations are not expected in oxic sediments.
Thus, the UCM probably resulted from complex bacterial reworking of biological debris
at the sediment/water interface. However, GC data showed considerable differences
between the structures of HC from overlying particles and from sediments. No hump
appeared for the former. It can be objected that the overlying particles reflected only the
recent inputs in April, when primary production is weak. They were not representative of
the whole local austral production, and further studies must be undertaken to ascertain
UCM formation exclusively at the sea-floor.
Nevertheless, the transformation pathways leading to such a biogenic UCM may be a
widespread phenomenon in the aquatic environment, masked by the superimposed
chromatographic imprint of ubiquitous petroleum humps.

Pigments. Two processes of rapid transfer may be involved in the transport of fresh algal
material through the water column:
(1) dense, rapidly sinking fecal pellets from macrozooplankton grazing, which have been
found to contain numerous intact cells unaffected by digestion (Silver and Bruland, 1981);
(2) post-bloom floes, observed in surf%zialsediments of the northeast Atlantic down to
4100 m (Lampitt, 1985), and responsible for higher levels of fresh material than fecal pellets
at depths down to 1000 m, up to 20% of the total pigments present in the euphotic layer
(Mejanelle et al., 1995).
Pigments are labile components. Detrital particles remaining in lake surface waters for
more than about 3 days lost nearly all detectable pigments to photodegradation (Carpenter
et al., 1986). However in the Antarctic, free-floating diatom chloroplasts have been found to
maintain their pigments for 10 days at - 1.5”C, despite the sensitivity of chlorins to oxygen,
light and enzymes associated with cell disintegration or bacterial activity (Gieskes and
Elbrachter, 1986). Rapid sedimentation from the euphotic zone also reduces their decay
(Leavitt and Carpenter, 1990). Owing to their high density, Antarctic diatoms can sink
quickly and, as a result, encysted microphytic cells have been evidenced in the deep-sea floor
of the area studied (Riaux-Gobin et al., 1997). This is consistent with the presence of
Iatroscan-measured chlorins in the sediments.
The presence of chlorin esters in the 9-l 1 cm section provided evidence for their stability
and for a widespread occurrence of these compounds, as suggested by Eckardt et al. (1992).
However, to our knowledge, this is the first report of alkyl chlorin ester synthesis involving
very long-chain saturated, branched and monounsaturated alcohols, as well as steryl chlorin
esters, probably due to the prominent occurrence of the former in the surface layers of this
particular environment.

Sterols-alcohols. In contrast to the findings of King and Repeta (1991a,b) and Pearce et
al. (1993), we found that the free and esterified sterols had a similar distribution. These
similar trends for the two series (Table 3) seem to suggest a common origin and
biogeochemical behavior. However, this is unlikely, owing to the sedimentary deposition
1102 .I.Laureillard el al.

of sterol-containing algal and animal material, modified by successive reworking through


the water column, and the easier degradation of free sterols compared to esterified ones,
observed in lacustrine sediments at Loch Clair, Scotland (Cranwell and Volkman, 1981). A
more likely inference would be the formation of free compounds at least in part from the
hydrolysis of the esterified fraction. The differences observed in the distribution of the
overlying particles and the 9-l 1 cm sample might be due to variations in the sedimentary
deposits. The major importance of cholesterol, together with the presence of alkyl chlorins,
argue for the formation of steryl chlorin esters in zooplankton guts, since fecal pellets are
known to be enriched in cholesterol (Prahl et al., 1984; Harvey et al., 1987) or through the
predation of zooplankton or its feces. Moreover, the significant percentage of dinosterol in
overlying particles might reflect the imprint of a temporary development of dinoflagellates
in surface waters, enhanced by concentration in herbivore guts of this non-assimilated sterol
(Harvey et al., 1989).
Such a close similarity between free and esterified forms was not observed for the
alcohols, implying an enhanced preservation of alkyl chlorins, free alcohols, and/or
additional inputs. Cranwell and Volkman (1981) attributed the lower abundance of steryl
esters vs. alkyl esters, found at lo-20 cm depth in lacustrine sediments (compared to the
corresponding surface sediments), to a greater susceptibility of steryl esters to hydrolysis.
However, the higher level of short-chain alcohols in the esterified form, even in the overlying
particles, is consistent with a bacterial reworking of free alcohols, arguing for an upper layer
origin. Indeed, short-chain compounds are known to be a preferential substrate for bacteria
rather than longer ones.
However, despite this apparent degradation of free alcohols, a higher level of branched
and unsaturated compounds more labile than their straight-chain saturated counterparts
were encountered in the free form. This feature points to some benthic input, maybe through
live bacterial cells. Volkman et al. (1986a) have already suggested a bacterial origin for
alcohols in an Antarctic lake. Thus, bacterial inputs to the sediment seem to be both
superficial and benthic.
Consequently, the burial efficiencies for hydroxylated compounds according to our
results are as follows: sterols < short-chain alcohols <long-chain alcohols, as observed by
Madureira (1994) in the eastern North Atlantic.

Triacylglycerols. In contrast, the pattern of the only storage lipids observed,


triacylglycerols, was found to be very irregular and not related to photic zone production
(Table 1). An explanation is provided by Palmisano et aE. (1988) and Nichols et al. (1989)
who described a net decrease in the TAG content of Antarctic diatoms at a late-bloom
period. This finding is rather surprising given the tendency of algae and especially diatoms
to accumulate fats when aging. In any case, TAG as well as WE originating from
phytoplankton or pelagic animal biosynthesis seem to be totally degraded when they sink
down through the water column. Their labile character was pointed out by Cranwell and
Volkman (198 I), Conte and Bishop (1988), and Volkman et al. (1988) who observed the lack
of TAG in the sediments of a shallow Antarctic lake in spite of their occurrence in algae and
fecal pellets. These features provide an insight into the efficiency of the bacterial activity
notwithstanding the very low temperatures and the high sinking rates of particles, due to
highly silicified diatoms (Treguer et al., 1991).

Free jiztly acids. The degradation of TAG, WE and PL is thought to lead to FFA. This
Organic geochemistry of marine sediments 1103

lipid class was relatively abundant. At all levels, the ratio of the sum of the acid lipid classes
vs. FFA (CTAG + DAG + PE + DPG + PG/FFA) varied widely from 26.7 in the overlying
particles to 3.4 at the 7-9 cm level. The accumulation of FFA within the sediment is evidence
of the higher stability of this lipid class compared to the other acid-containing classes.
Below the first centimeter, no relationship was found between the percentage of long-
chain acids and the especially high FFA CPI and the corresponding PL and TAG data. The
higher percentage of long-chain compounds found in FFA underlined the higher
degradability and the potential of short-chain components to be incorporated into the
mineral matrix. Below the surficial sediment, an average CPI of 6 was observed, due to the
preferential microbial-mediated removal of odd-carbon number acids. Thus the deeper
FFA mostly consisted of the better preserved, more refractory, long-chain even compounds
and no longer reflected the endofaunal and bacterial compositions represented by TAG and
PL.

Water column-sediment coupling. In the water column, the various lipids of the organic
matter in large particles were probably differently affected by preburial degradation and
microbial reworking during sinking. Within the sediment, the low level of acid 16:107, a
prominent component in diatoms, supports this view. Depending on the lipid class
considered, a relationship can be observed with the water column lipids, as for chlorins.
However, in spite of an apparent pelagic origin, the sedimentary HC, mainly
chromatographically non-resolved compounds, did not seem to reflect the original input.
Moreover, no relationship with upper ocean layers could be demonstrated from lipid classes
with acid moieties.
In Antarctic environments, despite the velocity of sinking particles and very cold
temperatures, favorable to good preservation of organic matter, the few experiments
comparing particle-sediment composition generally suggest an absence of correlation,
whatever the depth. Thus, in the open-sea, Matsueda and Handa (1986) clearly observed the
imprint of surface plankton in sinking particles throughout the water column, but it
vanished at sea-floor level, at 3750 m. Surprisingly, the same lack of correlation has been
found in shallow Antarctic lakes, Ace Lake (Volkman et al., 1988), Vanda and Fryxell
Lakes (Matsumoto, 1989), highlighting the extensive bacterial degradation of lipids at the
sediment-water interface. In contrast, in McMurdo Sound during November-December,
nearshore shallow marine sediments exhibited a diatom signature, proof of a surface-
sediment correlation (Smith et al., 1986, 1989).
Consequently, a relationship between algae and shallow sediment lipids seems to depend
on the lipid class studied and on the environmental conditions, but lipids of deep-sea surface
sediments show considerable alteration compared to their sources.

CONCLUSION
The study of lipid classes in particles reaching the sea-floor and in the sediments has
shown that simultaneous analysis of multiple classes is a powerful and indispensable tool for
the comprehensive assessment of the various inputs and of the live or detrital nature of the
organic matter constituting the sediments. The determination of their detailed composition
has been shown to be a useful approach for ascertaining their specific sources and fate.
Despite the rapid settling of large particles through the water column and slow
decomposition rates in the cold Antarctic environment, major alterations were observed
1104 J Laureillard er al.

in the composition of lipids associated with the particles near the sea-floor. However, their
characteristic features in some respects were close to those of the top layer, but these
particles do not represent the input of the whole year. The rapid decrease in lipid
concentration within the first centimeter highlighted the fact that degradation processes at
the sediment-water interface overwhelmingly dominate the losses in this recent Antarctic
sediment.
Each lipid class was shown to carry a specific message. In the sediment, the various lipid
classes studied can be divided into two groups according to their probable sources. Chlorins,
sterols and alcohols seem to originate from the surface layer or water column material,
through successive reworking by fauna and bacteria. Conversely, TAG, PL and FFA
appear to reflect essentially a benthic origin, deriving from meiofauna or bacteria. These
findings point to fundamental changes occurring in the water column.
Of special interest is the first report of a major UCM of biological origin in this pristine
Antarctic environment; however, the direct precursors are unknown. The complex
transformations leading to this HC mixture seemed to take place at the sediment-water
interface. The possibility of such a biological UCM should be taken into account in
petroleum contamination studies, based on UCM evaluation.
Finally, chlorin ester biosynthesis was demonstrated, illustrating the widespread
occurrence of this transformation pathway. Long-chain alkyl chlorin esters have never
been observed up to now. These chlorin esters are resistant and constitute a major sink for
chlorins, alcohols and sterols in sediments.

Acknowledgements-We wish to thank Professor P. Treguer, coordinator of the ANTARES arm of the French
JGOFS Program for his advice and constant support. We are grateful to Dr J. F. Gaillard, chief scientist of the
ANTARES 1 program for organic carbon analysis and for assistance in sampling operations, and to the crew of the
R/V “Marion Dufresne” for their help during the cruise. We also thank the IFRTP (Institut Franqais pour la
Recherche et la Technologie Polaires) for logistic and financial support.

REFERENCES
Berge, J.-P., Gouygou, J.-P., Dubacq, J.-P. and Durand, P. (1995) Reassessment of lipid composition of the
diatom Skeletonema costatum. Phytochemistry, 39, 1017-1021.
Bird, C. W., Lynch, J. M., Pirt, F. J., Reid, W. W., Brooks, C. J. W. and Middleditch, B. S. (1971) Steroids and
squalene in Methylococcus capsulatus grown on methane. Nature, 230, 413414.
Bligh, E. G. and Dyer, W. J. (1959) A rapid method of total lipid extraction and purification. Canadian Journal of
Biochemistry and Physiology, 31, 911-917.
Bouloubassi, I. and Saliot, A. (1993) Investigation of anthropogenic and natural organic inputs in estuarine
sediments using hydrocarbon markers. Oceanologica Acta, 16, 145-161.
Bradshaw, S. A., O’Hara. S. C. M., Corner. E. D. S. and Eglinton, G. (1989) Assimilation of dietary sterols and
faecal contribution of lipids by the marine invertebrates Neomysis integer, Scrocularia plana and Nereis
diversicolor. Journal of the Marine Biological Association of the United Kingdom, 69, 891-911.
Bradshaw, S. A., O’Hara, S. C. M., Corner, E. D. S. and Eglinton, G. (1990) Dietary lipid changes during
herbivory and coprophagy by the marine invertebrate Nereis diversicolor. Journal of the Marine Biological
Association of the United Kingdom, 70, 771-187.
Bradshaw, S. A., O’Hara, S. C. M., Corner, E. D. S. and Eglinton, G. (1991) Effects on dietary lipids of the
marine bivalve Scrocularia plana feeding in different modes. Journal of the Marine Biological Association of
the United Kingdom, 71, 6355653.
Brault, M. and Simoneit, B. R. T. (1988) Steroid and triterpenoid distributions in the Bransfield Strait sediments:
hydrothermally-enhanced diagenetic transformations. In: Advances in Organic Geochemistry, 1987, I_.
Mattavelli and L. Novelli, editors, Organic Geochemistry, 13, 697-705.
Organic geochemistry of marine sediments 1105

Cadee, G. C., Gonzales, H. and Schnack-Schiel, S. B. (1992) Krill diet affects faecal string settling. Polar Biology,
12, 75-80.
Carpenter, S. R., Elser, M. M. and Elser, J. J. (1986) Chlorophyll production, degradation and sedimentation:
Implication for paleolimnology. Limnology and Oceanography, 31, 112-124.
Conte, M. H. and Bishop, J. K. B. (1988) Nanogram quantification of non polar lipid classes in environmental
samples by high performance thin layer chromatography. Lipiak, 23, 492500.
Conover, R. J. and Huntley, M. E. (1991) Copepods in ice-covered seas-distribution, adaptations to seasonally
limited food, metabolism, growth patterns and life cycle strategies in polar seas. Journal of MarineSystems,
2, 140.
Cranwell, P. A. and Volkman, J. K. (1981) Alkyl and steryl esters in a recent lacustrine sediment. Chemical
Geology, 32, 29-43.
Cripps, G. C. (1990) Hydrocarbons in the seawater and pelagic organisms of the Southern Ocean. Polar Biology,
10, 393-402.
Cripps, G. C. and Priddle, J. (1991) Review: Hydrocarbons in the Antarctic marine environment. Antarctic
Science, 3, 233-249.
Danyushevskaya, A. I., Smirnov, V. I., Petrova, V. I. and Belyayeva, A. N. (1989) Geochemistry of organic
matter in bottom sediments of the Weddell Sea. Oceanology, 29, 322-327.
Delong, E. F. and Yayanos, A. A. (1986) Biogeochemical function and ecological significance of novel bacterial
lipids in deep-sea procaryotes. Applied and Environmental Microbiology, 51, 730-737.
De Wit, R., Relexans, J. C., Bouvier, T. and Moriarty, D. (1997) Microbial respiration and diffusive oxygen uptake
of deep-sea sediments in the Southern Ocean (ANTARES-I cruise). Deep-Sea Research IZ, 44, 1053-1068.
Dunstan, G. A., Volkman, J. K., Barrett, S. M., Leroi, J.-M. and Jeffrey, S. W. (1994) Essential polyunsaturated
fatty acids from 14 species of diatom (Bacillariophyceae). Phyfochemistry, 35, 155-161.
Eckardt, C. B., Pearce, G. E. S., Keely, B. J., Kowalewska, G., Jaffe, R. and Maxwell, J. R. (1992) A widespread
chlorophyll transformation pathway in the aquatic environment. In: Advances in Organic Geochemistry,
1991, C. B. Eckhardt, S. R. Larter er al., editors, Organic Geochemistry, 19, 217-227.
Fahl, K. and Kattner, G. (1993) Lipid content and fatty acid composition of algal communities in sea-ice and
water from the Weddell Sea (Antarctica). Polar Biology, 13, 405409.
Fricke, H., Gercken, G., Schreiber, W. and Gehlenschlager, J. (1984) Lipid, sterol and fatty acid composition of
Antarctic krill (Euphausia superba Dana). Lipids, 19, 821-827.
Gieskes, W. W. C. and Elbrachter, M. (1986) Abundance of nanoplankton-size chlorophyll-containing particles
caused by diatom disruption in surface waters of the Southern Ocean (Antarctic peninsula region).
Netherlands Journal of Sea Research, 20, 291-303.
Gough, M. A. and Rowland, S. J. (1990) Characterization of unresolved complex mixtures of hydrocarbons in
petroleum. Nature, 344, 648650.
Goutx, M., Germ, C. and Bertrand, J. C. (1990) An application of Iatroscan thin-layer chromatography with
flame ionization detection - Lipid classes of microorganisms as biomarkers in the marine environment.
Organic Geochemistry, 16, 1231-1237.
Graeve, M., Hagen, W. and Kattner, G. (1994) Herbivorous or omnivorous? On the significance of lipid
compositions as trophic markers in Antarctic copepods. Deep-Sea Research Z, 41, 915-924.
Harada, N., Handa, N., Fukuchi, M. and Ishiwatari, R. (1995) Source of hydrocarbons in marine sediments in
Liitzow-Holm Bay, Antarctica. Organic Geochemistry, 23, 229-237.
Harvey, H. R., Fallon, R. D. and Patton, J. S. (1986) The effect of organic matter and oxygen on the degradation
of bacterial membrane lipids in marine sediments. Geochimica et Cosmochimica Acta, 50, 795-804.
Harvey, H. R., Eglinton, G., O’Hara, S. C. M. and Comer, E. D. S. (1987) Biotransformation and assimilation of
dietary lipids by Calanus feeding on a dinoflagellate. Geochimica et Cosmochimica Acta, 51, 3031-3040.
Harvey, H. R., O’Hara, S. C. M., Eglinton, G. and Comer, E. D. S. (1989) The comparative fate of dinosterol
and cholesterol in copepod feeding: Implications for a conservative molecular biomarker in the marine water
column. Organic Geochemistry, 14, 635-641.
Jacques, G. (1991) Is the concept of new production-regenerated production valid for the Southern Ocean?
Marine Chemistry, 35, 273-286.
King, L. L. and Repeta, D. J. (1991a) New degradation pathways for chlorophyll in Black Sea sediments. In:
Organic geochemistry. Advances and applications in energy and the natural environment, D. A. C. Manning,
editor, Manchester University Press, pp. 239-240.
King, L. L. and Repeta, D. J. (1991b) Novel pyrophaeophorbide steryl esters in Black Sea sediments. Geochimica
et Cosmochimica Acta, 55, 2067-2074.
1106 J. Laureillard et al.

Lampitt, R. S. (1985) Evidence for seasonal deposition of detritus to the deep-sea floor and its subsequent
resuspension. Deep-Sea Research, 32, 881-897.
Leavitt, P. R. and Carpenter, S. R. (1990) Aphotic pigment degradation in the hypolimnion: Implications for
sedimentation studies and paleolimnology. Limnology and Oceanography, 35, 52&534.
Legeleux, F., Reyss, J. L. and Schmidt, S. (1994) Particle mixing rates in sediments of the North-East tropical
Atlantic: evidence from 2’0PbXs, 13’Cs, 22sTh Xs and *‘?hXs downcore distributions. Earth and Planetary
Science Letters, 128, 545-562.
Madureira, L. A. S. (1994) Lipids in recent sediments of the Eastern North Atlantic. PhD Thesis. University of
Bristol, 286 pp.
Matsueda, H. and Handa, N. (1986) Source of organic matter in the sinking particles collected from the Pacific
Sector of the Antarctic Ocean by sediment trap experiment. Memoirs of National Institute of Polar Research,
Special Issue, 40, 364-379.
Matsumoto, G. I. (1989) Biogeochemical study of organic substances in Antarctic lakes. Hydrobiologia, 172,265-
289.
Matsumoto, G. I. and Kanda, H. (1985) Hydrocarbons, sterols and hydroxy acids in Antarctic mosses. Antarctic
Record, 87, 23-3 1.
Matsumoto, G. I., Torii, T. and Hanya, T. (1983) Stenols and phytol in lake sediments from Syowa and Vestfold
Oases in the Antarctic. Geochemical Journal, 17. 1-8.
Matsumoto, G. I., Hirota, K., Ohtani, S., Watanuki, K. and Torii, T. (1991) Occurrence of long-chain n-alkanes
and n-alkanoic acids and C 29 sterols, in Antarctica and their geochemical signilicance. In: Organic
geochemistry. Advances and applications in energy and the natural environment, D. A. C. Manning, editor,
Manchester University Press, pp. 536-539.
Matsumoto, G. I., Friedmann, E. I., Watanuki, K. and Ocampo-Friedman, R. (1992a) Novel long-chain anteiso-
alkanes and anteiso-alkanoic acids in Antarctic rocks colonized by living and fossil cryptoendolithic
microorganisms. Journal of Chromatography, 598, 267-276.
Matsumoto, G. I., Matsumoto, E., Sasaki, K. and Watanuki, K. (1992b) Geochemical features of organic matter
in sediment cores from Liltzow-Holm Bay, Antarctica. In: Organic matter: Productivity, accumulation, and
preservation in recent and ancient sediments, J. K. Wheland and J. W. Farrington, editors, Columbia
University Press, New York, pp. 142-175.
Mejanelle. L., Laureillard, J., Fillaux, J., Saliot, A. and Lambert, C. (1995) Winter distribution of algal pigments
in small- and large-size particles in the Northeastern Atlantic. Deep-Sea Research I, 42, 117-I33.
Nichols, D. S., Nichols, P. D. and McMeekin, T. A. (1992) Anaerobic production of polyunsaturated fatty acids
by Shewaneha putrefaciens strain ACAM 342. FEMS Microbiology Letters, 98, 117-122.
Nichols, D. S.. Nichols, P. D. and McMeekin, T. A. (1993a) Polyunsaturated fatty acids in Antarctic bacteria.
Antarctic Science, 5. 1499160.
Nichols, D. S., Nichols, P. D. and Sullivan, C. W. (1993b) Fatty acid, sterol and hydrocarbon composition of
Antarctic sea ice diatom communities during the spring bloom in McMurdo Sound. Antarctic Science, 5,
271-278.
Nichols, P. D., Palmisano, A. C., Smith, G. A. and White, D. C. (1986) Lipids of the Antarctic sea ice diatom
Nitzschia cylindrus. Phytochemistry, 25, 164991653.
Nichols, P. D., Volkman, J. K., Palmisano, A. C., Smith, G. A. and White, D. C. (1988) Occurrence of an
isoprenoid Czs diunsaturated alkene and high neutral lipid content in Antarctic sea-ice diatom communities.
Journal of Phycology. 24, 90-96.
Nichols, P. D.. Palmisano, A. C., Rayner. M. S., Smith, G. A. and White, D. C. (1989) Changes in the lipid
composition of Antarctic sea-ice diatom communities during a spring bloom: an indication of community
physiological status. Antarctic Science, 1, 1333140.
Nichols, P. D.. Palmisano, A. C., Rayner, M. S., Smith, G. A. and White, D. C. (1990) Occurrence of novel C3a
sterols in Antarctic sea-ice diatom communities during a spring bloom. Organic Geochemistry, 15, 503-508.
Nichols. P. D., Skerratt, J. H., Davidson, A., Burton, H. and McMeekin, T. A. (1991) Lipids of cultured
Phaeocystis pouchetii: signatures for food-web, biogeochemical and environmental studies in Antarctica and
the Southern Ocean. Phytochemistry, 30, 3209-3219.
Ogura, K., Machihara, T. and Takada, H. (1990) Diagenesis of biomarkers in Biwa Lake sediments over 1million
years. In: Advances in Organic Geochemistry, 1989, B. Durand and F. Behar, editors, Organic Geochemistry,
16, 805-813.
Orcutt, D. M., Parker, B. C. and Lusby, W. R. (1986) Lipids in blue-green algal mats (modern stromatolites)
from Antarctic oasis lakes. Journal of Phycology, 22, 5233530.
Organic geochemistry of marine sediments 1107

Palmisano, A. C., Lizotte, M. P., Smith, G. A., Nichols, P. D., White, D. C. and Sullivan, C. W. (1988) Changes
in photosynthetic carbon assimilation in Antarctic sea-ice diatoms during spring bloom: variation in
synthesis of lipid classes. Journal of Experimental Marine Biology and Ecology, 116, I-13.
Parrish, C. C. (1987) Separation of aquatic lipid classes by chromarod thin-layer chromatography with
measurement by Iatroscan flame ionisation detection. Canadian Journal of Fisheries and Aquatic Sciences,
44. 722-73 1.
Parrish, C. C. (1988) Dissolved and particulate marine lipid classes: a review. Marine Chemistry, 23, 17-40.
Pearce, G. E. S., Raoux, C., Eckardt, C. B., Harradine, P. J., Keely, B. J. and Maxwell, J. R. (1993) Identification
of steryl chlorin ester pigments in marine and lacustrine sediments. In: Organic Geochemistry. Poster sessions
from the 16th International Meeting on Organic Geochemistry, Stavanger, K. Oygard, editor, Falch
Hurtigtrykk, Oslo, pp. 589-592.
Pichon, J.-J., Sikes, E. L., Hiramatsu, C. and Robertson, L. (in press) Comparison of LJY,and diatom assemblage
SST estimates with atlas derived data in Holocene sediments from the Southern West-Indian Ocean. Journal
of Marine Systems.
Prahl, F. G., Eglinton, G., Comer, E. D. S., O’Hara, S. C. M. and For&erg, T. E. V. (1984) Changes in plant
lipids during passage through the gut of Calanus. Journal of the Marine Biological Association of the United
Kingdom, 64, 3 17-334.
Prahl, F. G., Eglinton, G., Comer, E. D. S. and O’Hara, S. C. M. (1985) Faecal lipids released by fish feeding on
zooplankton. Journal of the Marine Biological Association of the United Kingdom, 65, 547-560.
Rabouille, C., Gaillard, J. F., Relexans, J.-C., Trtguer, P. and Vincendeau, M. (1997) Biogenic silica recycling in
surlicial sediments across the Polar Front of the Southern Ocean (Indian Sector). Deep-Sea Research, 44,
1151-1176.
Reinhardt, S. B. and Van Vleet, E. S. (1986) Lipid composition of twenty-two species of Antarctic midwater
zooplankton and fish. Marine Biology, 91, 149-159.
Reyss, J. L., Organo, C., Le Bet, N. and Schmidt, S. (1995) Sedimentation rate in South Ocean: Evidence from
230TheXcdowncore distributions. International Symposium: Carbon fluxes and dynamic processes in the
Southern Ocean: present and past, Brest, p. 126.
Riaux-Gobin, C., Hargraves, P. E., Neveux, J., Oriol, L. and Vetion, G. (1997) Microphyte pigments and resting
spores at the water-sediment interface in Subantarctic deep sea (Indian sector of the Southern Ocean). Deep-
Sea Research II, 44, 1033-105 1.
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
7’. inermis during mid-winter. Marine Biology, 62, 131-137.
Silver, M. W. and Bruland, K. W. (1981) Differential feeding and fecal pellet composition of salps and pteropods,
and the possible origin of the deep-water flora and olive-green “cells”. Marine Biology, 62, 263-273.
Simoneit, B. R. T., Cardoso, J. N. and Robinson, N. (1991) An assessment of terrestrial higher molecular weight
lipid compounds in aerosol particulate matter over the South Atlantic from about 30-70”s. Chemosphere,
23, 447465.
Smith, G. A., Nichols, P. D. and White, D. C. (1986) Fatty acid composition and microbial activity of benthic
marine sediment from McMurdo Sound, Antarctica. FEMS Microbiology Ecology, 38, 219231.
Smith, G. A., Nichols, P. D. and White, D. C. (1989) Triacylglycerol fatty acid and sterol composition of
sediment microorganisms from McMurdo Sound, Antarctica. Polar Biology, 9, 273-279.
Talbot, V. and Bianchi, M. (1997) Bacterial enzymatic hydrolysis in sediments of Subantarctic Indian sector.
Deep-Sea Research II, 44.
Tsitsa-Tzardis, E., Patterson, G. W., Wilkfors, G. H., Gladu, P. K. and Harrison, D. (1993) Sterols of
Chaetoceros and Skeletonema. Lipids, 28, 465461.
Treguer, P., Lindner, L., Van Bennekom, A. J., Panouse, M., Leynaert, A. and Jacques, G. (1991) The production
of biogenic silica in the Weddell-Scotia Seas measured by using radiotracer 32Si. Limnology and
Oceanography, 36, I2 17-l 221.
Venkatesan, M. 1. (1988) Organic geochemistry of marine sediments in Antarctic region: marine lipids in
McMurdo Sound. Organic Geochemistry, 12, 13-27.
Venkatesan, M. I. and Kaplan, I. R. (1982) Distribution and transport of hydrocarbons in surface sediments of
the Alaskan Outer Continental Shelf. Geochimica et Cosmochimica Acta, 46,2135-2149.
Venkatesan, M. 1. and Kaplan, I. R. (1987) The lipid geochemistry of Antarctic marine sediments: Branstield
Strait. Marine Chemistry, 21, 341-375.
Virtue, P., Nichols, P. D., Nicol, S., McMinn, A. and Sikes, E. L. (1993) The lipid composition of Euphausia
1108 J. Laureillard et 01.

superba Dana in relation to the nutritional value of Phaeocysfis pouchetii (Hariot) Lagerheim. Antarctic
Science, 5, 169-I 71.
Viso, A.-C. and Marty, J. C. (1993) Fatty acids from 28 marine microalgae. Phytochembrry, 34, 1521-1533.
Volkman, J. K. (1986) A review of sterol markers for marine and terrigenous organic matter. Organic
Geochemistry, 9, 83-99.
Volkman, J. K., Eve&t, D. A. and Allen, D. I. (1986a) Some analysis of lipid classes in marine organisms,
sediments and seawater using thin-layer chromatography-flame ionisation detection. Journal of
Chromatography, 356, 147-162.
Volkman, J. K., Allen, D. I., Stevenson, P. L. and Burton, H. R. (1986b) Bacterial and algal hydrocarbons in
sediments from a saline Antarctic lake, Ace Lake. Organic Geochemistry, 10, 671-681.
Volkman, 1. K., Burton, H. R., Eve&t, D. A. and Allen, D. I. (1988) Pigment and lipid compositions of algal and
bacterial communities in Ace Lake, Vestfold Hills, Antarctica. Hydrobiologia, 165,41-57.
Volkman, J. K., Jeffrey, S. W., Nichols, P. D., Rogers, G. 1. and Garland, C. D. (1989) Fatty acid and lipid
composition of 10 species of microalgae used in mariculture. Journal of Experimental Marine Biofogy and
Ecology, 128, 219-240.
Volkman, J. K., Holdsworth, D. G., Neill, G. P. and Bavor, H. J. (1992) Identification of natural, anthropogenic
and petroleum hydrocarbons in aquatic sediments. The Science of the Total Environment, 112, 203-219.
Whiticar, M. J., Suess, E. and Wehner, H. (1985) Thennogenic hydrocarbons in surface sediments of the
Bransfield Strait, Antarctic Peninsula. Nature, 314, 87-90.

Das könnte Ihnen auch gefallen