You are on page 1of 35

8.

15 Alkenone Paleotemperature Determinations


TD Herbert, Brown University, Providence, RI, USA
ã 2014 Elsevier Ltd. All rights reserved.

8.15.1 Introduction 399


8.15.2 Systematics and Detection 401
8.15.3 Occurrence of Alkenones in Marine Waters and Sediments 403
8.15.3.1 Genetic and Evolutionary Aspects of Alkenone Production 404
8.15.4 Function 405
8.15.5 Ecological Controls on Alkenone Production and Downward Flux 406
8.15.5.1 Effects of Water Column Recycling and Sediment Diagenesis on the Alkenone Unsaturation Index 409
0
8.15.6 Calibration of Uk 37 Index to Temperature 411
8.15.6.1 Culture Calibrations 412
8.15.6.2 Particles 413
8.15.6.2.1 Sediment traps 414
8.15.6.3 Core Tops 414
8.15.7 Synthesis of Calibration 416
8.15.8 Paleotemperature Studies Using the Alkenone Method 417
8.15.8.1 Holocene High-Resolution Studies 417
8.15.8.2 Millennial-Scale Events of the Late Pleistocene and Last Glacial Termination 418
8.15.8.3 Marine Temperatures During the LGM 419
8.15.8.4 SST Records of the Late Pleistocene Ice Age Cycles 422
8.15.8.5 SST before the Late Pleistocene 423
8.15.8.6 Comparison with other Proxies: d18O 423
8.15.8.7 Comparison with other Proxies: Microfossils 424
8.15.8.8 Comparison with other Proxies: Mg/Ca 425
8.15.8.9 Comparison with other Proxies: TEX86 and other Glycerol Dialkyl Glycerol Tetraethers Indices 426
8.15.8.10 Intercomparison of SST Proxies: Some Generalities 426
8.15.9 Conclusions 426
References 427

8.15.1 Introduction Alkenone paleothermometry promises a direct estimate of


near-surface ocean temperatures. Alkenones and the related
The alkenone paleotemperature method is the most mature of alkenoates come exclusively from a few species of haptophyte
the recently developed biomarker proxies that supplement pre- algae (Volkman et al., 1980a,b). These organisms require sun-
vious methods of paleo-sea surface temperature (SST) estima- light, and they generally prefer the upper photic zone (Popp
tions that rely on assemblages of fossilized microplankton et al., 2006). The environmental information contained in
(foraminifera, coccolithophorids, radiolaria, and diatoms) or their molecular fossils therefore is quite specific; although, as
isotopic or trace element clues contained in the fossilized re- will be discussed at length in a later section, ambiguities still exist
mains of planktonic organisms. Organic molecules traceable to on the depth and seasonal variations of alkenone-producing
marine planktonic sources, extracted and separated from a ma- species in the ocean. In contrast, many assemblages of plank-
trix of hundreds to thousands of other organic compounds, are tonic organisms such as foraminifera and radiolaria contain
the analytical targets. In most cases, the remnant alkenones and many species known to live well below the surface mixed layer.
alkenoates that are the subject of this review constitute no more, As originally defined by the Bristol organic geochemistry
and often considerably less, than a few percent of their initial group (Brassell, 1986b), the Uk37 index reflects the proportions
flux that left the surface layer of the ocean and fell toward the of the di- (C37:2), tri- (C37:3), and tetra- (C37:4) unsaturated
sediments. Good preservation is thus not a major issue for use of ketones. Subsequent work by Prahl and Wakeham (1987)
the proxy, in contrast to microfossil-based methods. Last, while showed that there was no empirical benefit in including the
many geochemical techniques assume that skeletal material is a C37:4 ketone in a paleotemperature equation. The currently
0
passive recorder of isotopic and trace element composition of accepted Uk 37 index (Prahl and Wakeham, 1987) varies posi-
sea water and that incorporation of paleoenvironmental signals tively with growth temperature and is defined as
follows thermodynamic laws that can be modeled using non-
0
biogenic phases in the laboratory, the alkenone method assumes Uk 37 ¼ C37:2 =ðC37:2 þ C37:3 Þ
that the ratios of biomarkers measured were actively regulated by
the producing organisms in life according to the temperature of where C37:2 represents the quantity of the diunsaturated ketone
the water in which they grew. and C37:3 represents the quantity of the triunsaturated form.

Treatise on Geochemistry 2nd Edition http://dx.doi.org/10.1016/B978-0-08-095975-7.00615-X 399


400 Alkenone Paleotemperature Determinations

The alkenone paleotemperature proxy thus depends only on (Prahl et al., 2000a,b), the temperature equation proposed by
the relative proportions of the common C37 ketones and not Prahl et al. (1988) has proven robust:
on their absolute amounts. Importantly, although the alke- h 0 i
Tð CÞ ¼ Uk 37  0:039 =0:034
nones are produced by calcareous algae, they survive in sedi-
ments where carbonate has dissolved, as first recognized by The first systematic study of core-top sediments by Sikes
Marlowe et al. (1984a,b) and Brassell et al. (1986b). Inspection et al. (1991) found that the unsaturation index in recent sed-
of the equation above shows that it can vary between 0 and 1.0; iments followed a relation to overlying SST very similar to the
it may therefore saturate at either extremely cold or extremely Prahl et al. (1988) calibration. Sikes et al. (1991) also con-
warm temperatures and give no useful information beyond cluded that there appeared to be no discernible effects on the
these limits (Conte et al., 2006; Sikes and Volkman, 1993). unsaturation index over the time of core storage.
The alkenone method has a number of strong points that As with any paleoceanographic proxy, a number of uncer-
compare favorably with other methods of determining past tainties must be evaluated that could affect its accuracy as an
ocean surface temperatures. Extractions and analyses by gas estimate of past SST. The principal biases incorporated into a
chromatography (GC) can be automated to generate much biologically based proxy can be broadly categorized as ecolog-
higher volumes of data more rapidly than for most other ical, physiological, genetic, and diagenetic. Combinations of
0
paleotemperature information. Furthermore, the precision of these factors could cause the Uk 37 index to deviate from a
properly operated gas chromatographic analyses is extremely unique relation to SST. Ecological biases may be derived from
high, yielding very high signal–noise ratios for time-series the fact that alkenone-producing species do not inhabit pre-
analysis. The alkenone unsaturation index appears to obey a cisely the same depth throughout the ocean and that they vary
global relationship to growth temperature (Muller et al., 1998; in abundance seasonally (Popp et al., 2006; Prahl et al., 1993).
with suitable caveats, e.g., Sikes et al., 1997, as discussed later), The alkenone unsaturation parameter recorded by sediments
in contrast to many paleotemperature proxies which require could therefore measure past temperatures very precisely, but
regional calibrations, corrections for carbonate dissolution, at which depths and with what seasonal bias (Prahl et al.,
etc. The proportion of di- and triunsaturated alkenones 1993)? It is also possible that the proportions of alkenones
also appears recalcitrant, if not immune, to diagenesis in synthesized by haptophyte algae vary with growth rate, inde-
the water column and within sediments relative to other pendent of temperature. The present state of ignorance dictates
large macromolecules (Rontani et al., 1997). Indeed, the first that people do not know the growth phase of haptophyte
reported occurrence of alkenones came not from recent material exported out of the photic zone, in particular whether
material but from Miocene age sediments of the Walvis Ridge the products represent the initial exponential growth phase
(Boon et al., 1978). observed in culture or stationary growth, which has been
0
Synthesis of long-chained alkenones was linked to extant shown to affect Uk 37 (Epstein et al., 1998). Natural populations
haptophyte algae, principally Emiliania huxleyi, shortly thereaf- also differ in their genetic composition. Alkenone-producing
ter (de Leeuw et al., 1980; Marlowe et al., 1984a,b; Volkman species are notable for their wide range of environmental tole-
et al., 1980a,b). Reviews of lipid analyses of Deep Sea Drilling rances. The consequences of genetic variations within strains of
Project sediments revealed that most sediments of Pleistocene the same producing species and between the different alke-
0
through mid-Eocene age appeared to contain measurable none-synthesizing species, for the Uk 37 index, are still debated.
quantities of alkenones and alkenoates (Brassell, 1993; Mar- Last, alkenones measured in sediments represent the surviving
lowe et al., 1984a, 1990). Brassell et al. (1986b) provided the molecules of a series of degradational pathways that begin in
seminal study linking alkenone unsaturation to paleotempera- the water column, proceed to the sediment–water interface, and
ture fluctuations in the late Pleistocene. After noting that which may continue into the sediment. A bias in the relative
modern surface sediments differed in their unsaturation ratios lability of the C37:2 and C37:3 ketones would be imparted to
depending on latitude, Brassell et al. (1986a,b) reconstructed paleoceanographic reconstructions of temperature, although
alkenone unsaturation in conjunction with benthic and present evidence suggests that differential preservation of the
planktonic foraminiferal d18O over the last 800 000 years in a two ketones is very small.
0
core from the subtropical North Atlantic. The unsaturation The Uk 37 index appears nevertheless to provide a remark-
index declined during glacial periods, suggesting cooler surface ably faithful estimate of near-sea-surface paleotemperatures. At
ocean temperatures during ice age conditions. The authors the same time, difficulties in matching the space and timescales
further demonstrated that the alkenone index gave a con- of modern process studies to the information contained in
tinuous paleoclimatic curve, even in intervals barren of sediments mean that the caveats raised above remain signifi-
foraminifera due to dissolution. cant. Field studies provide only snapshots of haptophyte abun-
Prahl and Wakeham (1987) and Prahl et al. (1988) proposed dance and alkenone unsaturation parameters, sediment traps
the first quantitative calibration of alkenone unsaturation to provide only a few years of data at only a few locations in the
growth temperature. Unsaturation parameters measured on a global ocean, and it is unclear how well laboratory cultures
strain of E. huxleyi grown in the laboratory at known tempera- replicate the natural environment. The author has endeavored
tures were compared to the unsaturation index on particulate to treat different lines of evidence systematically but found it
material collected from the near-surface ocean in the northeast difficult to discuss each aspect in a purely serial way. The reader
Pacific. Prahl and Wakeham (1987) showed that the labo- will therefore be asked to digest a review in which very diverse
ratory calibration appeared to apply well to the field observa- measurements and paradigms are woven together to answer
tions of unsaturation and the water temperature in which the central question of how to reconstruct past ocean SST with
0
the alkenones apparently were synthesized. As reviewed later the Uk 37 proxy.
Alkenone Paleotemperature Determinations 401

8.15.2 Systematics and Detection of the concentration of C37 ketones in analyses of alkenone-
containing materials (Prahl et al., 2001). Novel C35 and C36
Alkenones occur as a typical suite of 37-, 38-, and 39-carbon- methyl ketones have been reported recently from Black Sea
chained (C37, C38, and C39) ketones in marine particles and sediments (Xu et al., 2001); these apparently come from phy-
sediments. This set of compounds (Figure 1) constitutes a toplankton not commonly found in normal marine waters.
‘fingerprint’ for alkenones extracted from sediments. Their ex- C36 and C37 fatty acid methyl esters (alkenoates) and C37
istence went undetected until chromatographic columns capa- and C38 alkenes also accompany alkenones in alkenone-
ble of sustaining the high temperatures at which long-chained producing species (Conte and Eglinton, 1993; Grossi et al.,
alkenones elute came into existence in the late 1970s 2000; Marlowe et al., 1984b; Mouzdahir et al., 2001; Rosell-
(Volkman et al., 1980b). Although all alkenones are straight- Melé et al., 1994; Sikes et al., 1997; Volkman et al., 1980b). The
chained hydrocarbons, they may differ in the number of dou- relative proportions of alkenones, alkenoates, and alkenes vary
ble bonds (unsaturation) in the chain and in the structure of greatly between different strains of alkenone-containing spe-
the terminal ketone group (terminal carbon in the chain cies grown in culture (Conte et al., 1994a, 1995). The chain
bonded to either a methyl or an ethyl group). The C37 alke- length and degree of unsaturation of alkenoates and alkenes
0
nones used in the Uk 37 index are methyl ketones. 38-carbon- follow the pattern of the long-chained alkenones (Conte and
chained molecules include not only tri- and diunsaturated Eglinton, 1993; Grossi et al., 2000; Marlowe et al., 1984b).
methyl ketone forms but also C38:3 and C38:2 ethyl ketones Unfortunately, alkenoates and alkenes rarely reach more than
(De Leeuw et al., 1980; Volkman et al., 1980a). A C38:4 ethyl 10% of the concentrations of the C37 ketones in sediment
ketone has been reported from sediments underlying cold extracts and, thus, have limited utility for paleoceanographic
waters (Marlowe et al., 1984b). As discussed at greater length reconstructions.
in a later section, the total concentration of the four common The author’s laboratory has found few marine sediment
C38 ketones is generally nearly the same as the sum of the two locations where alkenones cannot be extracted in enough
common C37 ketones in marine particles and sediments quantity for quantification. Exceptions to the rule include
(Conte et al., 2001). Di- and triunsaturated C39 ethyl ketones very oligotrophic oceanic gyre locations, such as the Ontong
(Figure 1) are also commonly observed at about 10–20% Java Plateau, the red clay province of the North Pacific, and

6000
C37:2
Uk⬘37 = 0.908
5000
Temperature = 25.6 ⬚C
FID amplitude

4000

3000 C38:2et
C38:3me
C37:3
2000 C38:3et C38:2me C39:2et
C36:2FAME

1000
44 46 48 50 52 54

8000

7000 Uk⬘37 = 0.274


C37:3
Temperature = 6.9 ⬚C
6000
FID amplitude

5000 C38:3me
C37:2
C38:2et
4000
C38:3et

3000 C37:4 C39:3et C39:2et


C36:2FAME C38:2me

2000
42 44 46 48 50 52
Elution time
Figure 1 Typical GC–FID chromatograms of alkenone-containing sediment extracts.
402 Alkenone Paleotemperature Determinations

0
polar regions, where Uk 37 determinations are exceedingly dif- 1
ficult because of low concentrations of autochthonous alke-
nones. Typical quantities of C37 ketones in marine sediments
ODP 1019
range from 100 ppb to 10 ppm of total sediment (dry weight).
0.8 ODP 1012
Alkenones are typically extracted from sediments and particles LaPaz 21P
as part of a total lipid extract. From 1 to 5 g of dry sediment is
extracted with organic solvent (typically 9:1 methanol:methy-
lene chloride) by Soxhlet apparatus, by repeated sonication at 0.6

Uk38me
room temperature, or by an Accelerated Solvent Extraction
system such as that offered by Dionex Corp. Some laboratories
analyze alkenones as part of the total lipid extract; others prefer 0.4
to run ‘cleaner’ fractions following one of a number of schemes
of fraction purification using silica gel columns or thin-layer
chromatography (Villanueva et al., 1997). Most open-ocean 0.2
sediment extracts do not contain appreciable amounts of in-
terfering lipids; however, continental margin samples with
more complex matrices may benefit from cleanup procedures.
0
GC coupled to one of several detectors separates alkenones 0 0.2 0.4 0.6 0.8 1
from their lipid matrix and permits their quantification. Rela- k⬘
U 37
tive to most sedimentary lipids amenable to GC techniques,
0
alkenones have high molecular weights and high boiling Figure 2 Relationship observed between the Uk 37 index and the Uk38me
points. Elution through a nonpolar gas chromatographic col- index (C38:2me /(C38:2me þ C38:3me)) in a set of cores along the California
umn separates alkenones largely through boiling point and margin (Herbert et al., 1998). See Figure 1 for identification of the
subordinately by chemical interactions with the column film. alkenone peaks. Note that the two indices are closely related, although
0

Boiling point is largely determined by the molecular weight of the Uk38me index has a slope and intercept offset from the Uk 37 index.
the lipid. Because the mass differences between C37:4, C37:3, Scatter in the relationship primarily comes from analytical difficulties in
quantifying the C38me peaks accurately.
and C37:2 ketones are small (molecular weights of 526, 528,
and 530 amu, respectively), careful attention to compound
separation is required to completely resolve the component as some of the most abundant compounds detected in typical
0
alkenones. In analogy with the Uk 37 index, unsaturation ratios lipid extracts but also as one of the few sets of chromatographic
can be defined by the proportions of di- and triunsaturated peaks to elute late in the separation. Indeed, the virtual isola-
C38 ketones (Conte et al., 1998a). These relate linearly to the tion of the alkenone ‘fingerprint’ from the majority of chro-
0
Uk 37 index in both culture (Conte et al., 1998a,b; Yamamoto matographic peaks almost certainly led to the initial
et al., 2000), water column (Conte and Eglinton, 1993; Conte recognition of their paleotemperature significance.
0
et al., 2001) and sediment studies (Figure 2). However, ana- While GC–FID is the analytical workhorse of Uk 37 deter-
lytical challenges are greater for the four C38 ketones, whose minations, other detection techniques offer greater specific-
pairs of ethyl and methyl di- and triunsaturated ketones prove ity. Alkenones can be determined by GC–mass spectrometry
very difficult to separate completely (Figure 1). Unsaturation (MS) (Boon et al., 1978; De Leeuw et al., 1980), which can
ratios based on the C38 ketones therefore rarely appear in additionally identify each compound by the molecular ion and
the alkenone literature. C39 ketones make up only a minor characteristic fragmentation patterns. However, because the
component of an alkenone-containing extract and are rarely ionization efficiency for alkenones is low, GC–MS can have
quantified. lower sensitivity than GC–FID. Chemical ionization coupled
Most alkenone determinations are made with a flame ion- to MS can improve detection limits for alkenones (Chaler et al.,
ization detector (FID). This detector is simple, reliable, and 2000; Rosell-Melé et al., 1995a). Recently developed variants
highly sensitive. However, the FID functions essentially as a of GC, such as multidimensional GC (Thomsen et al., 1998),
carbon detector and does not give diagnostic information on comprehensive GC  GC (Xu et al., 2001), and GC–time-
the structure of the compounds detected. Alkenones are iden- of-flight MS (Hefter, 2008), promise to improve the specificity
tified by FID, by their elution times, and by reference to exter- of alkenone detection while maintaining the high throughput
nal standards added to the injected solvent. Should other and sensitivity of GC–FID.
compounds share a common elution time (‘coelution’) with In the absence of agreed upon alkenone reference stan-
the alkenone chromatographic peaks, they cannot be separated dards, an interlaboratory comparison was run that found sig-
from the signals of interest. The accuracy of GC–FID analysis nificant consistency among laboratories in determining the
therefore depends on the quality of chromatographic separa- alkenone unsaturation parameter in sediments (Rosell-Melé
tions in the GC column and on the presence or absence of et al., 2001). Splits of homogenized unidentified sediment
interfering compounds. Fortunately, experience shows that samples were sent to each participating laboratory. Bristol
interfering high molecular weight compounds are very rare in University also prepared mixtures of synthetic C37:3 and C37:2
extracts of marine particulate matter in sediments. The stability ketones. Laboratories were asked to perform replicate extrac-
of alkenones during microbial degradation, relative to other tions and chromatographic determinations following their
0
high molecular weight lipids, apparently enhances the alke- standard procedures. Interlaboratory variability for the Uk 37
none signal in sediment extracts. Alkenones stand out not only index, the most commonly measured alkenone unsaturation
Alkenone Paleotemperature Determinations 403

parameter, translated to a maximum of 2.1  C between labo- 1994a,b; Volkman et al., 1995). Other coccolithophorid spe-
ratories (95% confidence limit). Individual laboratories rou- cies tested to date do not yield alkenones. In addition, alke-
tinely obtain precision on replicate measurements on the order none production has been reported from some, but not all,
0
of 0.005 Uk 37 units or about 0.15  C. The interlaboratory species of the noncalcifying haptophyte genera Isochrysis and
comparison found no difference between the mean values Chrysotila (Conte et al., 1994a,b; Marlowe et al., 1984b;
determined by laboratories that put lipid extracts through Versteegh et al., 2001; Volkman et al., 1989). Cruises along
cleanup procedures and those that analyzed total lipid extracts. the California margin and the equatorial Pacific suggest that
This result indicates that reliable alkenone determinations can noncalcifying haptophyte algae often dominate over cocco-
be made for most marine sediments with minimal sample lithophorid forms (Thomsen et al., 1994). The possibility
preparation. Care should, however, be used to make sure that therefore exists that other noncalcifying haptophyte species
compounds are correctly identified and resolved in samples that synthesize long-chained alkenones may well have gone
with unusually large amounts of labile compounds, such as undetected (Brassell et al., 1987; Schulz et al., 2000). As far as it
found in continental margin sediments. is known, no marine algal groups other than the haptophytes
Analytical challenges most frequently occur at the extreme synthesize long-chained alkenones.
ends of the unsaturation index, where either the C37:3 or the The coccolithophorid algae, of which the principal alke-
C37:2 alkenone abundances become very small, and in sedi- none producers represent a subset, play an important role in
ments where one encounters a more complex matrix of lipids. the cycles of both organic and inorganic carbons in the ocean
Ternois et al. (1997) showed that use of a short (30 m) column (Balch et al., 1992; Sikes and Fabry, 1994; Thomsen et al.,
0
on Norwegian fjord samples (Uk 37  0.2) did not resolve C37:2 1994). Coccolithophorids surround their cell with a number
ketone from C36:3 FAME and, hence, produced erroneous esti- (in the case of E. huxleyi, approximately 23) of minute (micron-
mates of SST. Using a 50 m column on the same samples sized) calcite platelets. The intact aggregate is referred to as a
obviated these problems. A more subtle error may come from coccosphere; after death, the coccosphere may disintegrate into
irreversible adsorption of alkenones on the chromatographic its component coccoliths. The biogeography of coccolitho-
column (Grimalt et al., 2001; Villanueva and Grimalt, 1996). phorids is quite well understood, thanks to the distinctive
The effect is negligible for most samples but can become sig- liths produced by different species (Winter et al., 1994).
nificant when one is analyzing either the C37:2 or the C37:3 E. huxleyi is the most abundant and ubiquitous extant cocco-
alkenone near its limit of detection. Chromatographic biases lithophore (Okada and Honjo, 1973; Winter et al., 1994). In
0
will be toward warmer apparent temperatures at high Uk 37 many plankton studies, it may compose 60–80% of the cocco-
values and lower apparent temperatures at low values. The lithophorid assemblage. E. huxleyi appears to tolerate a large
author’s laboratory has also found that buildup of nonvolatile range of temperatures, salinities, nutrient levels, and light
compounds on either a guard or a chromatographic column availability (Winter et al., 1994). It therefore occurs in waters
will lead to preferential adsorption of the C37:3 alkenone or a of nearly all temperatures, excluding those of the truly polar
bias toward warmer temperature estimates, even in samples oceans, and in waters ranging from the highly saline (41%)
with large quantities of both ketones (see also Villanueva and Red Sea to the brackish Sea of Azov and Black Sea (Bukry,
Grimalt, 1996). The solution is to trim the chromatographic 1974). Unlike many coccolithophorids, which thrive in the
column at the injection point routinely. The author has also oligotrophic central gyres of the ocean, E. huxleyi can flourish
found that prolonged exposure of sediments in a dry (but not in more eutrophic environments. E. huxleyi blooms appear to
wet) state can lead to preferential removal of the C37:3 alke- be triggered under conditions of a highly stratified upper water
none; the author recommends extracting and running lipid column, a shallow (10–20 m deep) mixed layer, and high light
extracts within a few days after freeze-drying samples. intensities (Nanninga and Tyrrell, 1996). E. huxleyi may have a
competitive advantage over other phytoplankton under condi-
tions of high light intensity and low inorganic phosphate
8.15.3 Occurrence of Alkenones in Marine Waters ability, as concluded by modeling work on mesocosm experi-
and Sediments ments in Norwegian fjords (Aksnes et al., 1994). Doubling
rates can be as fast as 2.8 per day (Brand and Guillard, 1981).
Long-chained alkenones were first identified in Miocene through Under bloom conditions, E. huxleyi occurs in large enough
Pleistocene age marine sediments on the Walvis Ridge (Boon quantities in the surface ocean that scattering of light by the
et al., 1978). A group led by de Leeuw (de Leeuw et al., 1980) platelets can modify the optical properties of the water ob-
in the Netherlands subsequently identified C37 and C38 alke- served by satellite (Balch et al., 1992). Indeed, dense milky
nones in recent sediments of the Black Sea, where E. huxleyi patches of E. huxleyi blooms as large as 100 000 km2 have been
dominates the coccolithophorid flora. Simultaneously, work observed from space in the high-latitude (40–60 N) North
at Bristol University established the connection between Atlantic and North Pacific oceans, off the Falkland Islands in
alkenone synthesis and certain species of extant haptophyte the South Atlantic, and on the north Australian margin (Brown
(coccolithophorid) algae (Marlowe et al., 1984a,b, 1990; and Yoder, 1994).
Volkman et al., 1980a). Researchers at Bristol began to culture G. oceanica, the other known alkenone-synthesizing species
the producing species in the lab and to examine modern sed- of importance, has a more limited oceanic distribution.
iments for the presence of long-chained alkenones. It is now G. oceanica apparently does not occur in waters colder than
clear that two widely distributed species of coccolithophorid about 12  C (Okada and McIntyre, 1979). It commonly occurs
algae, E. huxleyi and Gephyrocapsa oceanica, are the principal in tropical and subtropical waters, in particular the high-
alkenone synthesizers in the modern ocean (Conte et al., fertility regions of the eastern Pacific and the Arabian Sea
404 Alkenone Paleotemperature Determinations

(Houghton and Guptha, 1991; Roth, 1994). This preference Gephyrocapsa, strains evolve with physiologies attuned to the
is consistent with a pulse of G. oceanica observed along the local environment (Paasche, 2002). The extent of genetic ex-
California margin during spring upwelling conditions (Ziveri change between strains is unknown. If as a result of genetic
et al., 1995). However, the abundance of G. oceanica in the isolation the relations between alkenone unsaturation and
western equatorial Pacific warm pool demonstrates that the growth temperature differ from strain to strain, then one
species does not require high nutrient concentrations. It report- would somehow have to account for this variation in paleo-
edly tolerates very high salinities (45–51%; Winter et al., 1994) ceanographic studies or at least attempt to estimate how much
but cannot grow in water of low (15%) salinity (Brand, 1984). uncertainty should be added to paleotemperature determina-
As with E. huxleyi, G. oceanica can often dominate coccolitho- tions due to strain effects (Prahl et al., 2010). At a larger
phorid assemblages where it thrives (Roth, 1994). biogeographic scale, where the relative abundance of E. huxleyi
The alkenone-producing species are typically reduced in and G. oceanica varies widely, one needs to decide whether each
0
abundance or excluded in oceanic provinces that favor diatom species has a significantly different Uk 37–temperature relation,
growth (Brand, 1994). Thus, they do not occur in the truly and if so, how to quantify the proportions of alkenone flux due
polar Arctic waters and in the much broader siliceous province to the two species. A temporal scale problem exists as well. The
in the Southern Ocean, where they taper to zero abundance stratigraphic range of alkenones in sediments far exceeds the
poleward of about 60 S (Nishida, 1986; Sikes and Volkman, paleontological range of the extant alkenone-producing species
1993). The abundance of E. huxleyi and G. oceanica is also (Brassell, 1993; Marlowe et al., 1990). Does one assume that the
0
reduced in regions of high silicate availability, such as in Uk 37 measured before the appearance of E. huxleyi or G. oceanica
many coastal zones and upwelling regions. Where nutrients denotes the same temperature as a calibration derived from the
such as nitrate and phosphate, or trace metals such as iron, are modern organisms?
low, the competitive advantage reverses to favor coccolitho- Genetic differences between strains of E. huxleyi and
phores over the siliceous phytoplankton (Brand, 1991). G. oceanica undoubtedly exist. Genetic variation appears in
A series of reports of alkenone occurrences in sediments growth rate, temperature preference, salinity tolerance, calcite
from brackish and freshwaters indicate that haptophytes production rate, chloroplast pigment composition, and long-
other than the recognized marine producers may generate in- chained lipid composition (Fisher and Honjo, 1989; Paasche,
puts of long-chained ketones in these environments (Harada 2002). As examples, Brand (1982) found that clones isolated
et al., 2008; Schulz et al., 2000). One distinctive feature of from the cold water Gulf of Maine are genetically adapted to
these (presumed) haptophytes is the production of large quan- lower temperatures than those from the warm Sargasso Sea.
tities of the tetraunsaturated (C37:4) ketone, which is also a Brand (1984) also determined that the oceanic genotypes from
signature of lacustrine alkenone producers (Cranwell, 1985; the Sargasso Sea cannot grow in 15% salinity, while coastal
Thiel et al., 1997; Toney et al., 2011; Zink et al., 2001). In some genotypes can. Reproductive rates of G. oceanica clones from
cases, the alkenone unsaturation ratio in coastal settings is open-ocean and neritic strains also differed in cultures (Brand,
consistent with open-marine producers and may point to the 1982). However, genetic variation does not always correspond
extreme tolerance of E. huxleyi to low salinities. For example, to geographical separation. Young and Westbroek (1991) stud-
Ficken and Farrimond (1995) studied a Norwegian fjord re- ied several genetically different morphological forms of E. huxleyi
cently opened to the open ocean by human activity. The sed- whose populations were only partially separated by temperature
iments recorded the transition from fresh to brackish (10%) preferences. Brand (1982) discovered that while strains of
conditions by an abrupt increase in C37 alkenones in the G. oceanica from coastal waters might differ genetically, strains
0
marine section. The Uk 37 temperatures recorded in the recent of E. huxleyi from the same regions did not.
interval lie between 7 and 10  C using the standard marine At the same time, molecular and paleontological evidence
temperature calibration, suggesting to the authors that E. huxleyi is growing to support the strong genetic similarity of all marine
was likely the producing organism. The earlier lacustrine sedi- alkenone producers. Two dominant morphotypes (A and B) of
ment sequence also contained alkenones, which differed from E. huxleyi are recognized. Yet, Medlin et al. (1996) found little
the marine sequence in their high abundance of the C37:4 ketone genetic variation when they compared ssu rRNA from three
(Ficken and Farrimond, 1995). Unusually high concentrations geographically isolated type A clones and a type B clone. In
of the tetraunsaturated ketone seem to signal a ‘brackish’ com- their judgment, the degree of separation did not warrant sep-
ponent to the haptophyte flora in environments such as the arating the A and B morphotypes at the species level. Two
Baltic Sea (Schulz et al., 2000) and Chesapeake Bay (Mercer studies that sequenced portions of the genome in a number
et al., 1999). of coccolithophorid species were also revealing. Fujiwara et al.
(2001) recognized four clades within the coccolithophores
based on the gene for the subunit of the Rubisco enzyme
8.15.3.1 Genetic and Evolutionary Aspects of Alkenone
rbcL. They recognized Emiliania–Gephyrocapsa as one clade,
Production
with identical nucleotide sequences for the rbcL subunit. By
Given that the synthesis of various alkenones and alkenoates is sequencing the spacer region of the Rubisco operon, Fujiwara
evidently regulated closely by the haptophyte producers, the et al. (2001) found that the noncalcifying alkenone-producing
question arises as to whether a universal response of parame- genus Isochrysis is an ingroup of the Emiliania–Gephyrocapsa
0
ters such as the Uk 37 unsaturation index would be expected in clade. The authors suggest that Isochrysis may have secondarily
the face of genetic diversity. Genetic variations could affect the lost the ability to form coccoliths. Ultrastructural similarities in
use of alkenone paleothermometry in at least three different the endoplasmic reticulum of Emiliania, Isochrysis, Gephyro-
scales. In wide-ranging organisms such as E. huxleyi and capsa, and Chrysotila also link the known alkenone-producing
Alkenone Paleotemperature Determinations 405

species (Fujiwara et al., 2001). An independent study of the sediments. The extinct genus Reticulofenestra evolved early in the
sequence of 18S rDNA by Edvardsen et al. (2000) confirmed Eocene and continued through the Pliocene, where it most
the genetic clustering of the Emiliania–Gephyrocapsa clade likely left the genus Gephyrocapsa as its descendant (Rio, 1982).
within the coccolithophores. Reticulofenestra is the only genus of Gephyrocapsaceae to accom-
A morphometric study of the distribution of Gephyrocapsid pany alkenones in Miocene, Oligocene, and Eocene samples
coccoliths hints that the previously recognized species may all (Marlowe et al., 1990). In the Pleistocene, Gephyrocapsa cocco-
belong to one biological species with strong morphological liths always appear where alkenones have been detected
variations along environmental gradients. Bollman (1997) (Marlowe et al., 1990).
looked at a set of 70 globally distributed Holocene sediment Cretaceous precursor alkenones have been reported; their
assemblages to make quantitative determinations of Gephyro- evolutionary relationship to the modern clade of alkenone-
capsid coccolith morphometric parameters such as size, bridge producing organisms is unknown. Cretaceous alkenones
angle, ellipticity, and width of the central collar. Bollman (Farrimond et al., 1986; Yamamoto et al., 1996) differ from
found strong environmental controls on the occurrence of the modern suite in that C41 and C42 ketones dominate, and
morphotypes. Temperature and chlorophyll abundance (in- these consist (to date) only of diunsaturated forms.
ferred from satellite data) proved to be the strongest predictors, In sum, genetic and micropaleontological data suggest very
although there was also a weak influence of salinity on the close evolutionary relationships among the alkenone-producing
occurrence of some morphotypes. The largest morphological species. Alkenone production occurs and occurred in species or
differences between samples corresponded to the greatest en- variants known to be closely related by micropaleontologists.
vironmental differences. Furthermore, within assemblages, Critical questions such as the degree of genetic exchange among
Bollman found gradations in morphotypes between the end species variants and the degree of genetic variability within
members he recognized. Bollman proposed that six major living populations remain to be determined. Ignorance of the
morphotypes would be conventionally associated with differ- role of alkenones in the cells of haptophyte algae and of genetic
ent Gephyrocapsid species. Although Bollman admitted that versus environmental controls on alkenone parameters means
paleontological examination cannot answer the question of that genetic effects in both space and time remain an important
whether his six morphotypes represent six biological species unknown in alkenone paleotemperature research.
or only one species with strong morphological plasticity,
he clearly leaned toward the latter. This result is in accord
with what many nannofossil workers see as a difficulty in
Gephyrocapsid species concepts (Wei, personal communication; 8.15.4 Function
Ziveri et al., 1995).
Results suggesting the close evolutionary relationship be- Alkenones occur in both motile and coccolith-bearing forms of
tween alkenone-producing species are encouraging as paleo- E. huxleyi (Bell and Pond, 1996; Conte et al., 1995; Volkman
ceanographers apply the alkenone method to sediments et al., 1980a,b). However, the role that alkenones play in the
that predate the appearance of modern taxa. E. huxleyi first producing organisms is not understood. That they play a crit-
appeared during marine oxygen isotope stage 8 (Thierstein ical role in E. huxleyi and related species is suggested by the
et al., 1977), at about 280 ka. Its appearance can be shown to large cellular investment accounted for by alkenones and
be globally synchronous to within 5–10 ky by comparing its alkenoates (Conte et al., 1994a,b; Epstein et al., 2001; Prahl
first appearance datum to oxygen isotope data measured in et al., 1988; Versteegh et al., 2001), generally between 5 and
foraminifera in the same sediments. The species apparently 10% of cell carbon. Marlowe et al. (1984a) and Brassell et al.
reached its modern levels of abundance earlier in the tropics (1986a) proposed that alkenone unsaturation helps to regulate
(c.85 ka) than at higher latitudes (Jordan et al., 1996). Emilia- membrane fluidity at different temperatures, analogous with
nia almost certainly descended from a Gephyrocapsid ancestor the known behavior of membrane lipids in many plants. This
(McIntrye, 1970). G. oceanica was a globally dominant nan- model clearly associates the unsaturation index with a physio-
noplankton taxon during some intervals of the Pleistocene logical response to growth temperature.
(Bollman, 1997; Thierstein, et al., 1977). The genus Gephyro- Subsequent work has failed to confirm the hypothesis that
capsa has been reported in sediments as old as Middle Miocene alkenones play a role in controlling membrane fluidity, how-
(South Atlantic: Jiang and Gartner, 1984; equatorial Pacific: ever. Conte and Eglinton (1993) did not detect alkenones in
Pujos, 1987), although its first frequent occurrence is in the fragmented membranes of E. huxleyi. Mouzdahir et al. (2001)
mid-Pliocene ( 3.5 Ma, Bollman, 1997). All the extant mor- interpreted the rapid light-dependent degradation of C31 and
phological variations of Gephyrocapsa have existed since at least C33 alkenones in contrast to the recalcitrance of C37 and C38
620 ka, with the exception of Bollman’s (1997) Gephyrocapsa alkenones as an indication that the shorter-chained alkenes
Cold and Gephyrocapsa Oligotrophic. resided in membranes, while the longer-chained ketones re-
Marlowe et al. (1990) compared the occurrence of alkenones sided in the interior of the cell. Epstein et al. (1998, 2001)
in sediments with micropaleontological data on coccoliths from noted that cell quotas of alkenones increase with decreasing
the same material. Their synthesis indicated that a number of growth rate in cultures grown from an initial stock of nutrients,
morphologically related species in the family Gephyrocapsaceae a finding confirmed by Prahl et al. (2003), who compared
were potential sources of alkenones and alkenoates (genera cellular quotients of alkenones during exponential and station-
Crenalithus, Dictyococcites, Emiliania, Gephyrocapsa, Pseudoemilia- ary phases in cell cultures. In analogy with triacylglycerides in
nia, and Reticulofenestra). The Gephyrocapsaceae was the only cultures of other marine plankton, Epstein et al. (2001) pro-
nannofossil family uniformly associated with alkenone-bearing posed that alkenones may serve as storage molecules for the
406 Alkenone Paleotemperature Determinations

producing organisms. Eltgroth et al. (2005) at last provided best measured by counting the number of intact coccospheres
a look at reservoirs of alkenones in cells of E. huxleyi and per unit volume of water sampled (e.g., Haidar and Thierstein,
Isochrysis galbana. They detected lipid bodies within the cell 2001). E. huxleyi accounts for the majority of the coccolitho-
that showed high abundances of long-chained alkenones. Sim- phorids surveyed in time series acquired off Bermuda (Haidar
ilar bodies also appeared within the chloroplasts. The authors and Thierstein, 2001; 64% on an annual basis), the San Pedro
suggested that the alkenones were synthesized within chloro- Basin off southern California (Ziveri et al., 1995; 30–80%
plasts and then exported to cytoplasmic lipid bodies. during the annual cycle), the northeast Atlantic (Broerse
Because the function that long-chained alkenones play in et al., 2000b; 69 and 72% at two sites surveyed), and off the
the producing organisms is not clear, the mechanism linking coast of northwest Africa (Sprengel et al., 2000). G. oceanica
alkenone unsaturation to growth temperature still lacks a firm was the most important taxon in year-long sediment trap
basis. Prahl et al. (2003) and Pan and Sun (2011) noted that studies of the Arabian Sea upwelling area, followed by E. huxleyi
significant variations in the alkenone quotient per cell and (Andruleit et al., 2000; Broerse et al., 2000c).
alkenone parameters occurred depending on growth phase Depth profiles of cell densities in the photic zone generally
and light–dark conditions in cell cultures. Their findings con- show E. huxleyi to live within the mixed layer. Cortés et al.
firm earlier observations that haptophyte cultures in the sta- (2001) and Popp et al. (2006) studied the seasonal depth
tionary (nutrient-depleted) phase have much higher cellular distribution of coccolithophorid species off Hawaii. Sampling
inventories than populations undergoing exponential growth. showed that the main production occurred in the middle
Both studies also documented the possibility that alkenone photic zone (50–100 m), which lies within the mixed layer
unsaturation ratios can change under isothermal conditions, for most of the year. While the depth of maximum E. huxleyi
depending on growth stage. As Prahl et al. (2003) concluded, density varied during the annual cycle, it usually reached its
these laboratory-based studies raise the interesting question of maximum between the shallowest sampling level (10 m) and
whether the phase of the growth cycle that gets exported by 100 m. Depth profiles off Bermuda (Haidar and Thierstein,
particles falling from the surface ocean to the abyss can bias the 2001) found that maximum densities of E. huxleyi were nearly
0
Uk 37 measured in sediments. always shallower than 100 m and, more commonly, within the
upper 50 m. The highest cell densities for E. huxleyi recorded
were at 1 m depth in March, after the seasonal advection of
8.15.5 Ecological Controls on Alkenone Production nitrate into the mixed layer. Seven years of water column
and Downward Flux particulate data off Bermuda confirm that alkenone concentra-
tions in the surface mixed layer are two to four times higher
Since alkenone-producing haptophytes can live in a range of than in the deep fluorescence maximum at 75–110 m (Conte
depths in the photic zone and vary greatly in their productivity et al., 2001).
over the course of an annual cycle, there is much to be learned A handful of studies have directly estimated alkenone pro-
0
about how the ecology of these organisms can affect the Uk 37 duction as a function of depth habit (Hamanaka et al., 2000;
signal eventually encoded in the sediments. Depending on the Popp et al., 2006; Prahl et al., 2005). Hamanaka et al. (2000)
depth of maximum production relative to the mixed layer, incubated phytoplankton at 0, 5, 10, 25, 40, and 60 m depth
alkenone producers may synthesize biolipids in temperatures and measured growth rates by spiking the water with d13C.
representative of SST or offset to colder temperatures by several Maximum incorporation into alkenones occurred at 5 m
degrees. Phytoplankton production is also quite seasonal in depth, which coincided with the peak in alkenone concentra-
most ocean locations. If alkenone production follows a strong tion in particulate matter. Production rates at the surface and
0
annual cycle, then again the Uk 37 temperature could contain a below 5 m were very low (from 269 ng l1 day1 at 5 m to
bias relative to mean annual conditions (in the latter case, the 1.4 ng l1 day1 at 25 m). The depth of maximum alkenone
bias could be toward either colder or warmer than average synthesis lies well above the depth of the chlorophyll maxi-
temperatures, depending on the season of maximum produc- mum at 25 m, indicating that the alkenone-producing species
tion). Profiles of coccolith and alkenone abundance through were offset vertically from the majority of the phytoplankton
the photic zone and over the yearly cycle give indications of community. An array of shallow sediment traps intercepted the
how ecological biases may vary between different oceanic largest alkenone flux at 15 m, just below the layer most pro-
provinces. Sediment traps intercept falling coccoliths and ductive of alkenones.
lipids to give windows into how the annual cycle in the near There is indirect evidence to suggest that alkenone synthesis
surface is exported downward to the sea floor. A number of may occur at depths below the mixed layer in some environ-
0
such studies will be reviewed below for their implications for ments. Uk 37 values in particulate matter often correspond to
0
the Uk 37 thermometer. Many are micropaleontological, asses- temperatures colder than the mixed layer. For example, Prahl
sing both the absolute vertical fluxes of the alkenone producers et al. (2001) found that the alkenone flux to sediment traps in
E. huxleyi and G. oceanica and the proportions of these species the Wilkinson Basin (coastal northeastern United States)
relative to the entire coccolithophorid assemblage. Unfortu- peaked during summer time when surface waters stratified
nately, very few studies combine both micropaleontological and a subsurface chlorophyll maximum was established in
0
and geochemical approaches. the upper seasonal thermocline. The Uk 37 temperatures of
Nearly all micropaleontological time-series data collected the summer particulate matter corresponded to temperatures
on near-surface samples and sediment traps found either at the base of the upper thermocline, 6–7  C colder than the
E. huxleyi or G. oceanica to dominate the total coccolith flora approximately 5-m-thick summer mixed layer. Similarly,
0
on an annual basis. The density of living coccolithophorids is Ternois et al. (1996) correlated the Uk 37 of sediment trap
Alkenone Paleotemperature Determinations 407

material to a time series of upper water column temperatures in progression of maximum production with latitude, exceptions
the western Mediterranean to deduce that the depth of maxi- to the rule occur. Goni et al. (2001) recorded a two- to threefold
mum alkenone production varies over the annual cycle. The increase in the organic carbon-normalized fluxes of alkenones in
0
Uk 37 temperature during the spring season corresponded to a the Gulf of California from June to October, well past the
maximum production depth of 50 m, while that depth appar- expected spring bloom. They speculated that maximal alkenone
ently shallowed to 30 m during the second phase of high production could be separated from that of other phytoplan-
production in the fall months. Both Ohkouchi et al. (1999) kton groups by competition. Groups such as diatoms that often
0
and Prahl et al. (1993) have further argued that Uk 37 values dominate peak bloom periods (Broerse et al., 2000a,c; Girau-
should be systematically offset to colder temperatures than the deau et al., 1993) also generally show higher seasonal variability
mixed layer in gyre regions, where the deep nutricline tends to than does E. huxleyi (Beaufort and Heussner, 1999, 2001; Harada
produce subsurface chlorophyll maxima. It should be noted et al., 2001). A competitive interaction apparently was observed
0
that in order to use the Uk 37 value to estimate the depth of by Ziveri et al. (1995), who found that the highest coccolith and
alkenone synthesis, one has to assume a reference temperature coccosphere flux occurred in winter in the San Pedro Basin off
calibration. The suggestion of subsurface production is not southern California, preceding the spring upwelling season fa-
purely circular, however, as one can compare the qualitative vored by diatoms. Further evidence of E. huxleyi and/or alkenone
0
features of the Uk 37 values over the annual cycle to the tem- production offset from that of other phytoplankton groups
0
perature time series. Thus, the Uk 37 of falling particles actually can be found in the studies of Ziveri et al. (2000), Harada et al.
decreased from spring into the onset of summer water column (2001), and Muller and Fischer (2001).
stratification in the Wilkinson Basin (Prahl et al., 2001), in A number of oceanic regimes also produce twice yearly flux
direct contrast to the warming of the surface layer. maxima of alkenone production. In the Mediterranean, a fall
Alkenone production also occurs in the context of the an- bloom of alkenone production occurs (Sicre et al., 1999; Ternois
nual cycle of upper water column temperatures (Goñi et al., et al., 1996). This is also true off Hawaii (Cortés et al., 2001), in
2003; Prahl et al., 2005; Sikes et al., 2005). Standing stocks of the central equatorial Pacific (Harada et al., 2001), in the Sea of
alkenone-producing species and their falling products display Okhotsk (Broerse et al., 2000a), and in the Norwegian Sea
strong changes over the course of the year in all studies to date. (Thomsen et al., 1998). A lack of dissolved silica may inhibit
In most cases, the time of maximum abundance of E. huxleyi diatom growth and promote haptophyte production during
and/or G. oceanica and the maximum flux of alkenones into the fall months in such locations (Broerse et al., 2000a).
sediment traps coincide with the dominant period for phyto- The monsoonally driven upwelling system of the Arabian
plankton blooming. Thus, production peaks in spring months Sea offers a special occasion to study the seasonality of alke-
in most subtropical and mid-latitude locations (Antia et al., none production. In this environment, G. oceanica dominates,
2001; Broerse et al., 2000b; Cortés et al., 2001; Haidar and although E. huxleyi still occupies the second position in the
Thierstein, 2001; Harada et al., 2001; Prahl et al., 1993; Sprengel coccolith flora (Andruleit et al., 2000; Broerse et al., 2000c).
et al., 2000, 2002). In the typical cycle, cell densities of E. huxleyi Winds favorable to upwelling occur twice per year. The south-
increase in the upper photic zone after the seasonal advection of west monsoon in late spring/early summer sees the peak in
nitrate that occurs with winter and early spring mixing (Haidar upwelling conditions (Prahl et al., 2000a,b), while a secondary
and Thierstein, 2001). The abundance of E. huxleyi may increase maximum occurs when the winds shift direction during the
by an order of magnitude (Cortés et al., 2001) or more (Haidar northeast monsoon (Andruleit et al., 2000; Prahl et al., 2000a,b).
and Thierstein, 2001) in the upper water column during the Coccolithophore fluxes, consisting of 60–70% G. oceanica and
spring bloom, as does its flux to sediment traps in locations E. huxleyi, increase during the southwest monsoon, at the same
such as the northeast Atlantic (Broerse et al., 2000b). A sediment time as the maximum in silica flux (Broerse et al., 2000c). Flux
trap sample obtained in May in the Norwegian Sea intercepted maxima lasted for nearly 3 months in this sediment trap study.
a nearly monospecific bloom of E. huxleyi (Cadee, 1985). Flux In another sediment trap experiment conducted in the Arabian
maxima are, however, more commonly diffuse, occupying 2–3 Sea, Prahl et al. (2000a,b) determined that distinct alkenone
months of time (Broerse et al., 2000a,b,c). Alkenone fluxes in flux maxima occur at the start and stop of the northeast and
sediment traps generally show a spring peak, indicating that southwest monsoons. A lag in alkenone flux relative to other
surface ecological signals are exported to depth (Sicre et al., phytoplankton biomarkers suggested a successional delay in
1999; Ternois et al., 1997). haptophyte production during the more dramatic southwest
Maximum alkenone production may be shifted to the sum- monsoon (Prahl et al., 2000a,b). During peak monsoonal
mer months in some regions. For example, Prahl et al. (2001) upwelling, alkenone fluxes reached about 25 times those of
found that alkenone flux to sediment traps peaked during the unproductive months of the year.
the early summer months in the Wilkinson Basin (43 N), Because so few sediment trap experiments report data for
after the water column had stratified. At higher latitude loca- much more than one year’s duration, people have only a
tions such as the Norwegian Sea and the Barents Sea in the glimpse at the importance of interannual variability in the
northern hemisphere (Samtleben and Bickert, 1990; Thomsen productivity of alkenone-synthesizing species. Reports do sug-
et al., 1998) and in the Indian ocean sector of the Southern gest large changes in the downward flux of either E. huxleyi or
Ocean (Ternois et al., 1998), the flux of E. huxleyi and alkenones alkenones between years. Three years of sediment trap data
is phased even more strictly to the summer months. Production from the northwestern Mediterranean Sea revealed, in addition
can reach vanishingly small amounts during winter months in to the biannual flux maxima mentioned earlier, considerable
such harsh environments (Broerse et al., 2000a; Ternois et al., interannual variability (Sicre et al., 1999). In general, Sicre
1998). While such a pattern follows the classic seasonal et al. (1999) found that maximum alkenone fluxes coincided
408 Alkenone Paleotemperature Determinations

with maximum total organic carbon (TOC) fluxes. In the year noted the role of marine snow in entangling coccospheres in
1994, however, there were TOC fluxes comparable with other the fall bloom in the Sea of Okhotsk and a complementary role
bloom periods without correspondingly large C37 ketone played by large diatoms in the spring. Coccospheres represent
fluxes. This study also reported large variability in the ampli- only a tiny portion of the coccolith carbonate flux (Broerse
tudes of the spring and fall blooms in alkenones. For example, et al., 2000a; Ziveri and Thunell, 2000; Ziveri et al., 2000);
the spring and fall blooms produced fluxes of C37 ketones of most of the coccolith flux must undergo multiple cycles of
8 and 16 mg m2 day1, respectively, in 1989–90; but in 1994, release and reaggregation (Ziveri et al., 2000).
the fall bloom amounted to a flux of only 0.1 mg m2 day1 to Studies that report alkenone fluxes relative to total organic
the 200 m trap. Muller and Fischer’s (2001) 4 year sediment carbon (normalized to TOC in typical units of 100–
trap study in the upwelling region of North Africa also docu- 1000 mg g1 TOC) also provide evidence that much of the
mented approximately sevenfold variations in the annual flux alkenone flux from the photic zone comes from bloom epi-
of alkenones to sediment traps. These authors found no repeat- sodes. Such normalized alkenone concentrations in sediment
able seasonal cycle in alkenone fluxes at their study location. traps ranged from lows of 29 mg g1 C during unproductive
Strong year-to-year variations in coccolith and coccosphere winter months in the Wilkinson Basin (northeast margin of the
fluxes have also been identified in sediment trap data from United States) to highs of 1054 mg g1 C during the early
the North Atlantic (Ziveri et al., 2000) and in water column summer bloom period (Conte et al., 1998b; Prahl et al.,
censuses off Bermuda (Haidar and Thierstein, 2001). Conte 2001). Similar results were reported in the Southern Ocean
et al. (1998b) reported a particularly interesting short-lived (Ternois et al., 1997), Mediterranean Sea (Sicre et al., 1999),
alkenone flux peak in a Sargasso Sea sediment trap experiment. and Norwegian Sea (Thomsen et al., 1998). Alkenone accumu-
This region produces a classic spring bloom following the lations during times of very low flux may indeed represent the
winter deepening of the mixed layer. The pulses of alkenone sedimentation of ‘relict’ material synthesized during more pro-
production detected by Conte et al. (1998b) preceded this ductive seasons (Conte et al., 1998b; Prahl et al., 2001; Sicre
predictable part of the annual cycle. Several short-lived flux et al., 1999). There are, however, several sediment trap studies
events occurred, usually in December and January, but not for that do not report significant correlations between the alke-
every year studied. During the flux events, falling organic mat- none flux and the total organic flux (Muller and Fischer, 2001;
ter was enriched in labile, phytoplankton-derived debris, Ternois et al., 1996).
including alkenones. Evidence has emerged that some benthic environments
Tiered sediment trap arrays present a picture of how sea- may also receive a diffuse supply of reworked fine-grained
sonal and episodic production works its way toward the sea particles with associated alkenones. Processes such as resus-
floor. Nearly, all such arrays show that the near-surface tem- pension of continental margin and slope sediments during
poral variability is attenuated with depth. The seasonal vari- major storms and benthic currents may thus contribute lateral
ability so evident in many shallow sediment trap time series is supplies of alkenones to the sediments. Fluxes of coccoliths in
reduced by factors of 2 (Broerse et al., 2000a, Sea of Okhotsk; a deep-sea canyon setting in the Bay of Biscay increased signif-
Ziveri et al., 2000, northwest Atlantic; Muller and Fischer, icantly during the fall and winter stormy months (Beaufort and
2001, northwest African margin) to 3 (Harada et al., 2001, Heussner, 1999), as did coccolith and alkenone fluxes during
central equatorial Pacific; Thomsen et al., 1998, Norwegian summer resuspension months in the Norwegian Sea (Andruleit,
0
Sea). This attenuation presumably comes both from the selec- 1997; Thomsen et al., 1998). Core-top Uk 37 estimates can ap-
tive biological degradation of more labile lipids at shallow parently be affected by the input of fossil alkenones in the
depths and from a diffuse supply of fine-grained sediment Norwegian and Barents seas. Thomsen et al. (1998) determined
0
particles at depth that partially masks the variability of surface Uk 37 values in shallow sediment traps consistent with produc-
production (see more below). tion in (cold) SST but showed that deeper traps and surface
0
These observations raise the question of exactly what factors sediments had Uk 37 indices 5–10  C too warm for the locations.
control the downward transport of alkenones. Clearly, as mol- The presence of pre-Quaternary coccoliths in floral assemblages
ecules associated with organisms only 10–20 mm in size (be- unequivocally indicates that some portion of the flux comes
fore decomposition and disaggregation of coccoliths), they from the erosion of older slope and shelf sediments (Beaufort
must have vanishingly low settling rates without the aid of and Heussner, 1999, 2001; Weaver et al., 1999). A number of
processes that cause aggregation. The majority of sediment other sediment trap studies have found higher coccolith fluxes at
trap studies find that the coccolith and/or alkenone flux is depth than in shallow traps, indicating a lateral source of
highly correlated with the total mass flux over the yearly cycle material (Antia et al., 2001; Sprengel et al., 2000, 2002; Ziveri
(Andruleit et al., 2000; Beaufort and Heussner 1999; Broerse et al., 2000).
0
et al., 2000a,b). This may signal the general synchronization of The effect of lateral transport on Uk 37 temperature esti-
E. huxleyi and/or G. oceanica production with the total biogenic mates will depend on the age of the transported material
flux, which then sweeps smaller particles out of the photic zone (e.g., contemporary or fossil), the location from which it
(Thomsen et al., 1998; Ziveri and Thunell, 2000). To support comes (e.g., from similar latitude and SST or from long dis-
this idea, shallower sediment trap collections find that the tance), and the ratio of the advected flux to the alkenones
intact coccosphere flux, which must represent the most re- sedimented in the vertical sense. Deep equatorward benthic
cently produced and least remineralized component of the currents are apparently responsible for transporting alkenones,
haptophyte flux, coincides with the time of highest detached along with the fine fraction, a great distance from their source
coccolith flux and that these peak the same time as the highest in at least two instances. In the southwestern Atlantic, recent
total mass flux (Broerse et al., 2000a). Broerse et al. (2000a) age sediments under the Brazil–Malvinas Confluence and
Alkenone Paleotemperature Determinations 409

0
Malvinas Current produce anomalously cold (by 2–6  C) Uk 37 be exposed to degradation under oxic conditions for thousands
temperature values (Benthien and Müller, 2000; Rühlemann of years. However, where sediment accumulates more rapidly,
and Butzin, 2006). The remainder of the 87 core tops that these as along continental margins, and where oxygen levels in
0
authors analyzed in a large region from 5 N to 50 S had Uk 37 either the bottom water or the sediment pore waters plummet,
values in good agreement with global core-top calibrations to alkenones will be exposed to nonoxic bacterial metabolic path-
local mean annual SST. Benthien and Müller (2000) argue that ways. Comparison of sediment trap fluxes with surface sedi-
sediments in these specific regions were transported northward ments in pelagic regions suggests alkenone preservation factors
0
and offshore by benthic currents and, hence, contain Uk 37 of 0.2% (relative to the deep trap flux) (Muller and Fischer,
signals of their origin in cold waters. Benthic boundary current 2001), 0.6% (Prahl et al., 1989a), and 1% (Prahl et al., 2000a,b).
advection of alkenones has been documented in a different In pelagic environments, degradation of alkenones in surface
way by Ohkouchi et al. (2002) in Bermuda Rise drift sedi- sediments appears approximately an order of magnitude more
ments. Compound-specific accelerator MS (AMS) 14C dating efficient than the reduction in bulk organic carbon (Muller and
of alkenones shows that they may be significantly older than Fischer, 2001; Prahl et al., 2000a,b). However, much better
foraminifera in the same layers. This the authors attribute to preservation occurs in settings with high sediment flux. The
southward advection of the fossil fine fraction material from highest preservation efficiencies of alkenones reported is 44%
the Nova Scotia margin. There may, however, also be environ- in the shallow (300 m), high-deposition rate setting of the
ments with significant lateral transport in which the effect of Wilkinson Basin (Prahl et al., 2001), and nearly 100% in the
allochthonous material does not swamp primary coccolitho- anoxic bottom sediments of the Guaymas Basin (Goni et al.,
phorid ecological signals. For example, Beaufort and Heussner 2001). Prahl et al. (1993) used an onshore–offshore transect
(1999) found that the seasonal cycle of succession in cocco- of sediments and sediment traps to demonstrate that the preser-
lithophorid species was preserved in the Bay of Biscay sediment vation efficiency of alkenones and total organic carbon was high
trap time series, despite the input from resuspended margin and similar ( 25%) at their nearshore site (water depth
0
sediments. The primary Uk 37 signal would presumably be 2717 m) but fell systematically in the direction of the open
preserved as well in such a case. ocean to only 0.25% in the most distal site. The deepest water
site clearly displayed preferential loss of alkenones relative to
bulk organic carbon (Prahl et al., 1993). Preservation efficiency
8.15.5.1 Effects of Water Column Recycling and Sediment
in sediments thus seems to depend strongly on exposure time to
Diagenesis on the Alkenone Unsaturation Index
oxic degradation (Madureira et al., 1995; Prahl et al., 2001).
As they descend through the water column and are incorpo- None of the sediment trap studies report significant shifts in
rated into sediments, alkenone and other lipid biomarkers the alkenone unsaturation index as a consequence of degrada-
encounter different degradational conditions (Koopmans tion. A sediment trap deployment in the northwestern Medi-
et al., 1997; McCaffrey et al., 1990; Prahl et al., 1989b; Rontani terranean Sea found that despite a fivefold loss in alkenones
et al., 1997; Sinninghe Damsté et al., 2002; Sun and Wakeham, between shallow and deep traps, there was no apparent offset
0
1994; Teece et al., 1998). The amount of time that lipids spend in the Uk 37 ratio of the biomarkers. Sawada et al. (1998)
exposed to these metabolic pathways varies substantially, as similarly found no shift in the unsaturation index through
does the degradational efficiency of each pathway. Rapid tran- the enormous vertical path length of the northwest Pacific
sit under oxic conditions generally occurs through the water (traps arrayed at 1674, 4180, 5687, and 8688 m). Further-
0
column. Slow passage, often under suboxic to anoxic condi- more, the Uk 37 composition of underlying sediments agreed
tions, characterizes the entrance of alkenones into the sedi- with sediment trap estimates, indicating that early diagenesis
mentary record. The key question for paleoceanographers is in the sediments did not modify the unsaturation index rela-
whether these steps produce measurable changes in the alke- tive to the incoming composition. Other comparisons of the
0
none unsaturation index. Uk 37 index between sediment traps and core tops find very
The fraction of alkenones buried in sediments represents good agreement as well (Muller and Fischer, 2001; Prahl, et al.,
less than 1% of the initial flux from the photic zone in most 1989a, 2001; Sikes et al., 2005).
cases. This attenuation can be measured by comparing tiered Investigators have conducted laboratory studies of the ef-
sediment trap fluxes to surficial sediment fluxes. The synchro- fects of both oxic and anoxic degradations on the alkenone
nization of alkenone flux peaks between shallow and deep unsaturation index. In the early history of the development of
0
traps shows that the time required for alkenone-containing the Uk 37 index, Volkman et al. (1980a,b) compared the unsa-
particles to reach the sea floor is only 1–2 weeks (Conte turation index in fecal pellets of the copepod Calanus helgolan-
et al., 1998b; Muller and Fischer, 2001). Somewhat slower dicus to that of the E. huxleyi used to feed them. No change in
average sinking velocities pertain to periods of low alkenone the index was observed after passage through the guts of these
and total mass flux (Muller and Fischer, 2001). During their zooplankton. Grice et al. (1998) obtained similar results by
descent, alkenones are strongly recycled, both relative to their feeding the alkenone-synthesizing haptophyte I. galbana to the
initial flux and to bulk organic matter (Muller and Fischer, copepod Temora. Teece et al. (1998) exposed alkenones and
2001; Sicre et al., 1999). Alkenones do appear more resistant other lipids to oxic, sulfate-reducing, and methanogenic mi-
to degradation than most other lipids of planktonic origin crobial degradation for almost 800 days. After rapid initial
(Prahl et al., 2000a,b; Volkman et al., 1980a). Recycling in degradation of lipids, the fate of alkenones varied significantly.
the sediments adds further attenuation to the roughly one About 85% of the initial alkenone inventory had been de-
order of magnitude loss in the water column. In the slow graded under oxic conditions by the end of the experiment.
deposition rate of most marine environments, alkenones will The two anoxic pathways yielded different results. Under
410 Alkenone Paleotemperature Determinations

0
sulfate-reducing conditions, degradation essentially ceased concluded that the shift to higher Uk 37 values could well be
with about 60% of the alkenones remaining. Methanogenic an analytical artifact of attempting to analyze the C37:3 ketone
conditions led to preservation not much better than for oxic at its limit of detection. As noted earlier, irreversible chromato-
conditions (about 80% degradation). Teece et al. (1998) con- graphic column adsorption becomes significant at very low
cluded that the different apparent alkenone degradation rate concentrations of either ketones. In Madeira Abyssal Plain
constants under different anoxic conditions lead to subtleties sediments, the initial (unoxidized) C37:3 ketone concentration
in understanding the preservation of alkenones under varying is quite low relative to the diunsaturated ketone, and it could
redox states. A very important finding, however, was that the well reach its limit of detection under the concentrations ana-
0
Uk 37 index did not shift beyond a very modest increase (0.03 lyzed by Hoefs et al. (1998).
units, equivalent to less than 1  C apparent temperature What is the long-term fate of the alkenone unsaturation
change) reported for the oxic experiment (Teece et al., 1998). index in sediments? The exposure to diagenesis of biomarkers
Zabeti et al. (2010) isolated bacterial strains from E. huxleyi buried in sediments dwarfs the few thousand years of near-
cultures and reported that some aerobic strains preferentially surface exposure. It is difficult to completely answer this question
oxidize the C37:3 alkenone, while others do not. Whether se- since without knowing a priori what correct SST values are for
lective degradation will occur, according to the authors, ancient sediments, authors cannot get a direct estimate of diage-
depends on the pathway of microbial attack on the alkenone netic offsets or lack thereof. Nevertheless, some qualitative lines
molecule. of reasoning suggest that the alkenone index probably remains
A consensus view of laboratory and field studies is that the quite stable on geological timescales. The signature of differential
0 0
autooxidation and bacterial alteration of the Uk 37 index are diagenesis on the Uk 37 index would include long-term trends
small but possible under oxic settings (Kim et al., 2009; toward decreased concentrations of alkenones with greater time
Rontani et al., 2011). In all instances where a change was and burial depth. Most workers would expect that a diagenetic
reported, degradation resulted in a preferential loss of the overprint, if there is one, would be at the expense of the C37:3
C37:3 alkenone, thereby producing a warm bias. Gong and ketone. If one models the degradation as first order with concen-
0
Hollander (1999) compared Uk 37 sediment data acquired tration of the C37 ketones, one can imagine that slightly different
down-core at two sites in the Santa Monica Basin that differed rate constants would be involved. The available evidence sug-
0
in bottom water oxygenation. They attributed a positive Uk 37 gests that, once the alkenones have survived the relatively high
offset at the oxic site to indicated preferential degradation of metabolic activity of the upper few centimeters to tens of centi-
the C37:3 ketone. In order to estimate the offset, Gong and meters of the sediment column, these rate constants must be very
Hollander had to match samples of the same age between the small. This is because the time alkenones and other lipids spend
cores. Some of the apparent temperature offsets of up to 2.5  C below the surficial layer is one order of magnitude longer at
therefore depended on the quality of the age models. Never- 10 000 years and three orders of magnitude longer at 1 My
theless, Gong and Hollander (1999) presented evidence over than the time spent during early diagenesis. Over these long
0
the last two centuries of deposition for an average Uk 37 offset timescales, decay rate constants of perceptible size should pro-
equal to 0.9  C at the oxic site. Another way to study the duce nearly monotonic trends in alkenone loss as a function of
0
possibility of diagenetic alteration of the Uk 37 index is to sediment age, even if there were primary variations in the initial
look for gradients in the index with preservation of initially alkenone concentration of the sediment, because the sediment
homogeneous material. Fine-grained turbidites, the sedimen- age would be many e-folding times of the inverse rate constant.
tary product of near instantaneous emplacement of well-mixed The author’s laboratory has frequently found alkenone
sediment on the sea floor, offer this opportunity. After the concentrations higher in sediments hundreds of thousands to
emplacement of the turbidite layer, a well-developed redox millions of years old than they are in the core-top material of
front moves downward from the sediment–water interface. the same sites (Figure 3). The implication is that variations in
The top of the turbidite layer will experience significant oxida- initial near-surface alkenone concentration persist for very long
tion, while the base may suffer none at all. Hoefs et al. (1998) durations in the deeper sediment column without diagenetic
0
studied Uk 37 indices through oxidation fronts in a number of attenuation. Furthermore, alkenones appear stable when nor-
fossil turbidite layers cored on the Madeira Abyssal Plain. They malized to other products of photosynthesis. Figure 4 displays
reversed the conclusion of Prahl et al. (1989a,b), who had the alkenone content of sediments from the Oman Margin
0
earlier found negligible shift in the Uk 37 index in a relatively (Ocean Drilling Program Site 723) normalized to total organic
young turbidite sequence in the same region. The samples with nitrogen and chlorins, an early diagenetic transformation
0
the lowest alkenone concentrations produced Uk 37 estimates product of chlorophylls, as a function of age. Considered
2.5–3.5  C warmer than samples in the unaltered bases of the individually, all three datasets show high-amplitude variations
turbidites. Hoefs et al. (1998) concluded that diagenesis may with time that are related to changes in production and pre-
indeed selectively degrade the C37:3 ketone and produce signif- servation of these compounds in this dynamic upwelling
icant artifacts for paleoceanography. zone. This variability disappears into near-monotonic trends
Grimalt et al. (2000) mounted a serious criticism of the of increasing normalized alkenone concentration with greater
evidence for differential diagenesis based on the turbidite stud- age (Figure 4), which are best explained as preferential deg-
ies. They noted that the amount of apparent temperature radation of organic nitrogen and chlorophyll relative to
change depended strongly on the reported alkenone concen- alkenones.
trations. The strongest evidence for differential diagenesis in We also found that alkenones appear to be well preserved in
the Hoefs et al. (1998) study comes from samples with extraor- uplifted and lithified marine sections outcropping on land. In
dinarily low alkenone concentrations. Grimalt et al. (2000) that study, they were able to compare the alkenone
Alkenone Paleotemperature Determinations 411

ODP 1012 (S. California)


80 1.8

70 1.6

Total C37 ketones (µg g−1)


60 1.4
C 37 :C 38

Ratio C37:C38
50 1.2

40 1

30 0.8

20 0.6

10 C37 0.4

0 0.2
0 0.5 1 1.5 2 2.5 3
Age (Ma)
Figure 3 Alkenone concentrations (mg per g dry sediment weight) determined over a 2.8 Ma record from ODP Site 1012 off the southern California
coast (Liu et al., 2005). Note the absence of a down-core decrease in alkenone concentrations that would suggest progressive diagenetic loss of
alkenones in older sediments. Evolutionary conservatism in alkenone synthesis is suggested by the stability of the ratio of total C37 alkenone to C38
alkenones (SC37/SC38) over the record.

ODP Site 723B


120 14

C37 ketones / total chlorins


C37 ketones / % organic N

100 12

10
80
C / %N
37 8
60
6
40
C / chlorin 4
37
20
2
0
10 20 30 40 50 60 70 80
Age (ka)
Figure 4 Down-core trends in alkenone-normalized organic nitrogen and total chlorins from ODP Site 722, Oman Margin (Higginson et al., 2004).
The records extend from the late Holocene at the core top to about 70 ka. Note the monotonic decrease of the two organic classes relative to alkenones.
This behavior is consistent with the progressive degradation of these more labile components and the diagenetic stability of the C37 alkenones.

unsaturation estimates from outcrops in Italy to coeval pelagic unsaturation index to the water temperature in which the
sediments recovered by the Ocean Drilling Program and producing organisms grew. The second involves understanding
demonstrate little to no effect of diagenesis. Sediments with how the unsaturation index recorded in sediments, which
elevated pore water temperatures provide at least one excep- represents an enormous integration of growth temperature
tion to this general rule of the diagenetic stability of alkenones. histories of individual organisms, relates to a consistent mea-
Simoneit et al. (1994) measured alkenone abundances in sure of SST (e.g., SST vs. subsurface depth and annual vs.
sediments of the Middle Valley on the Juan de Fuca Ridge sea seasonal temperature). These are not one and the same ques-
floor spreading axis. Hydrothermal alteration of the sediments tion, because E. huxleyi and related species may not always
resulted in the loss of alkenones at temperatures greater than live in the mixed layer and because their production may
200  C. Simoneit et al.’s (1994) study thus suggests caution in vary seasonally. For this reason, in the following section evi-
using the alkenone unsaturation index in other regions of dence is evaluated for how the influences of water column
unusually high geothermal gradients or in very deeply buried habit, seasonality, and particle transport through the ocean
0
sediments. may affect the interpretation of the sedimentary Uk 37 index.
One can approach the question of the quantitative relation
0
of the Uk 37 index to growth temperature at considerably dif-
0
8.15.6 Calibration of Uk 37 Index to Temperature ferent scales of time, genetic, and environmental variability.
Culture studies allow the experimentalist to eliminate many
The calibration of the alkenone unsaturation index actually confounding variables in the natural environment (genetically
resolves into two questions for paleoceanographers to address. mixed populations, varying nutrient availability, different
The first concerns defining an equation, or sets of equations if depth habitats, etc.) to isolate the influence of factors such
one should not prove globally applicable, that relates the as growth temperature, physiological state (exponential vs.
412 Alkenone Paleotemperature Determinations

late-log vs. stationary growth), and nutrient availability on results may reflect the effects of different growth rates of hap-
alkenone and alkenoate systematics. Despite the elegance of tophytes grown in batch culture or the physiological state in
the experimental approach, results from such studies must be which the cells were harvested. Growth of alkenone-producing
viewed as models for what may exist under natural conditions, haptophyte algae in chemostats (Popp et al., 1998) represents
not necessarily as calibrations. By collecting particulate mate- a particularly sophisticated manipulation, as these cultures
rial in the water column, either by filtering material in the grow in a medium of constant (low) nutrient availability. It is
euphotic zone or by collecting falling particles in sediment not clear whether batch or continuous growth models better
traps, and relating their alkenone composition to time series represent natural conditions or whether the sinking flux of
of near ocean temperatures, one moves closer to relating the alkenones and alkenoates in the ocean comes from popula-
growth environment of the alkenone-producing algae to the tions in exponential, late logarithmic, or stationary growth
signals sent to the sediment. This comes at the cost of unknown state.
genetic variability in the natural populations and some ambi- All culture studies confirm the first-order dependence of
0
guity in the actual time and depth of alkenone synthesis, and Uk 37 and other alkenone (Conte et al., 1998a) unsaturation
hence, the appropriate growth temperature–alkenone/alkenoate parameters on growth temperature but produce results that
relations. Near-surface sediments now provide a global database conflict in many ways. The study of Prahl et al. (1988) dem-
to examine the paleoenvironmental information contained in onstrated that haptophytes grown in the laboratory adjust their
the preserved record of alkenones and alkenoates. The time unsaturation to temperature changes on a timescale of days;
averaging inherent in sedimentation means that one is integrat- further culture work confirmed rapid adjustment of alkenoate/
ing temporal, physiological, and genetic influences on scales not alkenone ratios to changes in growth temperature Conte et al.
approached in the laboratory or in field studies. However, cor- (1998a) and light–dark conditions (Prahl et al., 2003). How-
relations from the sediments to environmental variables in the ever, culture calibration studies suggest very large variations in
water column rely on statistical relations, since one cannot the relation of unsaturation to growth temperature that may
0
directly link the sediment Uk 37 index to particular ecological depend on genetic and physiological factors (Conte et al.,
or genetic controls on the unsaturation–temperature relation. 1995, 1998a,b; Epstein et al., 1998). Twenty-four strains of
alkenone-producing species cultured by Conte et al. (1995) at
0
15  C gave Uk 37 values that ranged from 0.3 to 0.55. Volkman
8.15.6.1 Culture Calibrations
et al. (1995) suggested that cultures of G. oceanica produce a
Prahl and colleagues (Prahl and Wakeham, 1987; Prahl et al., significantly different relation of unsaturation to growth tem-
1988) conducted seminal studies that still provide the widely perature; however, their experimental results do not agree with
0
accepted calibration of the Uk 37 index to growth temperature. G. oceanica cultures grown by Sawada et al. (1996) or Conte
k0
These papers report the U 37 values of one strain of E. huxleyi et al. (1998a).
grown at different temperatures. Prahl and his collaborators It also seems clear that factors such as growth phase, light,
argued that the good agreement of their laboratory results to and nutrient levels can significantly influence the unsaturation
0
the results of Uk 37 analyses of water column particulate matter index of haptophytes grown in the laboratory (see summary in
collected at known temperatures in the eastern Pacific suggested Tables 4 and 5 of Versteegh et al., 2001 and in Prahl et al.,
that they had arrived at a valid calibration between 15 and 25  C. 2003). Both Conte et al. (1998a) and Epstein et al. (1998)
It may well be that linking laboratory to field data helped ensure found changes in the unsaturation index between log, late-log,
the robustness of their calibration estimate. The relationship and stationary phases of growth. Epstein et al. (1998) pro-
derived by Prahl and colleagues (Prahl and Wakeham, 1987 posed that nutrient availability, which would control growth
and amended slightly by Prahl et al., 1988) followed a linear rates of cultured and natural populations, could significantly
relationship to temperature. Extrapolated to the limits of the affect the calibration of unsaturation to growth temperature.
0
Uk 37 index of 0 and 1, it suggests a lower temperature limit of Both investigations found increasing alkenone concentrations
about 1  C and an upper limit of about 28  C. Both the very (pg cell1) in late logarithmic and stationary phase growth.
coldest and warmest SST of the ocean would therefore lie outside Conte et al. (1998a) and Prahl et al. (2003) also documented
0
the range of the Uk 37 index. very large ranges in the ratios of alkenoates to C37 and C38
Numerous culture studies of alkenone-producing species ketones (0–2.8) and in the SC37/SC38 ketone ratio depending
and strains followed Prahl’s initial studies, under the theory on growth phase. Comparison of these parameters to field
that cultures uniquely allow the experimentalist to isolate fac- data led Conte et al. (1998a) to conclude that natural popula-
tors that may influence alkenone and alkenoate distributions. tions most closely resemble late-log or stationary populations
Experience shows that rather different results can be obtained grown in the laboratory. In contrast to the batch culture exper-
from using batch or continuous culture methods on the same iments discussed above, Popp et al. (1998) used chemostats
strain of alkenone-producing algae (Popp et al., 1998), from to control steady-state growth rates, which they argue may be
the phase of growth from which alkenones are harvested a better model for natural systems. The latter study found no
0
(Conte et al., 1998a; Epstein et al., 1998 and unpublished; significant dependence of Uk 37 to growth rate at constant
Yamamoto et al., 2000), and from different laboratories cul- temperature.
turing the same strain (e.g., Conte et al., 1998a,b, compared A number of studies have investigated whether genetic and/
to Prahl et al., 1988, or to Sawada et al., 1996). Further- or physiological differences create ‘fingerprints’ in other aspects
more, replicate cultures grown in the same laboratory under of alkenone/alkenoate systematics that might allow investiga-
0
ostensibly similar conditions can yield a spread of Uk 37 values tors to distinguish past variations in species/strain production
(Conte et al., 1995; Versteegh et al., 2001). These inconsistent in sediments. Volkman et al. (1995) and Sawada et al. (1996)
Alkenone Paleotemperature Determinations 413

0
suggested that the proportions of C37 to C38 ketones (‘chain The arguments favoring regional Uk 37–temperature calibra-
length index’), or alkenoate/alkenone ratios, might relate to tions are based on regression estimates of a small number of
the proportions of E. huxleyi to G. oceanica at the time of samples and do not seem to hold up as more water column
production (see also Yamamoto et al., 2000). Further culture data are aggregated. The most extensive synthesis of water
0
work by Conte et al. (1998a) does not support either sugges- column Uk 37–temperature relations argues for the global ap-
tion (see also Section 8.15.6.3 below). plicability of a single calibration equation (Figure 5; Conte
Nearly all culture calibration studies predict higher growth et al., 2006). This study examined 392 samples from all
temperatures for the same unsaturation index than postulated major ocean basins, gathered in the mixed layer (0–30 m) to
by the Prahl et al. (1988) regression, although several (Conte prevent including samples acquired in the seasonal thermo-
et al., 1998a and G. oceanica cultures of Sawada et al., 1996) fall cline, which might be falling from the warmer layer above.
very close to the Prahl relation. If this ensemble of culture data Although the dataset is weighted heavily to the North Atlantic,
is correct, field and sediment studies applying the Prahl et al. it includes provinces dominated by G. oceanica as well as
(1988) calibration might frequently underestimate tempera- E. huxleyi, upwelling zones, and gyres. (See also the study by
tures. As authors evaluate water column, sediment trap, and Bentaleb et al. (2002) for additional water column data in the G.
core-top data, they should assess whether the large range oceanica province of the western equatorial Pacific.) The simi-
0
of possible Uk 37–temperature relations suggested by culture larity of the regional datasets is striking enough for Conte et al.
studies, whether of genetic or physiological origin, demonstra- (2006) to conclude that differences in the genetic makeup of
bly affects the accuracy of a unified calibration relation. One natural alkenone-synthesizing populations and differences in
would expect to find that different haptophyte biogeographic their growth environment (differing nutrient fluxes and/or water
regions produce distinct calibrations of unsaturation to column stability) do not significantly detract from the use of a
temperature and to see the influence of nutrient availability global calibration equation for paleotemperature estimation.
in offsets between upwelling and nonupwelling regions. In this author’s judgement, however, Conte et al. (2006)
overstep their data by arguing that the correct form of a global
0
Uk 37–temperature equation must be nonlinear and that water
8.15.6.2 Particles 0
column calibrations conflict with sediment-based Uk 37–temper-
By studying alkenone parameters in particulate matter col- ature regressions. Following Sikes and Volkman (1993), Conte
lected in the photic zone, one loses the ability to manipulate et al. propose a third-order polynomial equation to describe the
0
potential genetic or physiological influences, but gain the flattening of the Uk 37–temperature relation near the cold and
ability to compare alkenone systematics to temperatures in warm ends of the dataset. As they note, the data suggest more
the natural setting. Calibration equations can be generated variability of the index in relation to temperature at the warm
and tested by comparing the alkenone unsaturation index in and cold extremes of the surface ocean. The scatter may reflect
suspended particles to ambient water temperatures. One gen- some combination of the importance of nonthermal factors on
erally assumes that the measured water temperature is the same alkenone synthesis near the extremes of temperature encoun-
as the temperature in which the alkenones and alkenoates were tered by haptophyte algae (Conte et al., 2006), but it may also
synthesized. This may not always be the case for particles
sinking or mixing through a temperature-stratified water col-
umn. In fact, given a general tendency for particles to sink, the Conte water column data
temperature of alkenone synthesis might be higher than the
temperature at which the particles are collected, but almost 1
certainly not lower (Sicre et al., 2002). A potential temporal
offset also exists between the time of alkenone synthesis and
the measurement. Thus, the alkenone temperatures could be 0.8
set to the temperature of previous ‘bloom’ conditions, rather
than the currently measured water column temperature. Prahl et al. equation
0
In contrast to culture studies, relationships between Uk 37 0.6
Uk⬘37

and temperature in suspended particulate organic carbon gen-


erally show much closer agreement to the Prahl et al. (1988)
calibration equation. Several large-scale compilations of water 0.4
column unsaturation ratios have been presented (Brassell,
1993; Conte et al., 2006; Sikes et al., 1997), as well as regional
studies in the North Atlantic and Mediterranean (Conte and 0.2
Eglinton, 1993; Conte et al., 1992; Sicre et al., 2002; Sikes and
Volkman, 1993; Ternois et al., 1997) and tropical Pacific
(Popp et al., 2006; Prahl et al., 2005). Data have been variously 0
−5 0 5 10 15 20 25 30 35
interpreted as requiring regional calibrations to growth tem-
perature (Conte and Eglinton, 1993; Ternois et al., 1997) or as Temperature (⬚C)
0
requiring modifications of the original Prahl et al. (1988) Figure 5 Compilation of the Uk 37 index of mixed-layer particulate
0
culture-based Uk 37 equation to a more appropriate relation matter in relation to in situ temperature (Conte et al., 2006). The heavy
based on a water column particulate matter equation (Brassell, solid line indicates the linear Prahl et al. (1988) paleotemperature relation
1993; Conte et al., 2006; Sikes and Volkman, 1993). used for sediment estimates.
414 Alkenone Paleotemperature Determinations

include analytical errors, as C37:3 and C37:2 alkenones approach sediment traps, the authors conclude that both the amplitude
0
their detection limits in warm and cold waters, respectively and absolute values of the Uk 37 temperature estimates are in
(cf. Grimalt et al., 2001; Pelejero and Calvo, 2003). In any good agreement with weekly SST estimates.
0
event, the third-order polynomial fit of water column Uk 37 to
2
in situ temperatures improves the r value to 0.97 as compared to
the r2 value of 0.96 for a linear fit and reduces the standard error 8.15.6.3 Core Tops
of estimate from 1.4 to 1.2  C. If the Conte et al. (2006) calibra- Surficial sediments should reflect the weighting function of
0
tion is accurate, then the Uk 37–temperature relationship of core- alkenone production at all seasons and depths throughout
top sediments must be biased in comparison to the relationship the annual cycle – the integrated production temperature con-
0
to in situ growth temperatures. Uk 37 sediment values at mid- cept of Conte et al. (1992). Core-top material also provides the
latitudes would be systematically high compared to mean an- benefit of temporal and spatial averaging of other factors, such
nual growth temperatures. Whether the sediment bias of 2–3  C as genetic variability and variations in growth rate that may
at mid-latitudes inferred by Conte et al. (2006) exists depends influence alkenone systematics. This comes at a cost: the time
critically on the choice of the third-order polynomial description averaging varies with sedimentation rate and bioturbation
0
of the Uk 37–temperature relationship and may be premature. and much information that relates to the original produc-
tion of alkenones in the surface ocean is lost. Furthermore, a
8.15.6.2.1 Sediment traps sediment-based regression compares surficial material repre-
0
Sediment trap material provides a valuable view of the Uk 37 senting centuries to millennia of ocean history to the short
and quantity of alkenones transiting to the sea floor. Few time- period of instrumental temperature data used by ocean atlases
series experiments have been reported to date, although data such as the Levitus (1994) global climatology.
from the Gulf of California (Goni et al., 2001) and off the coast Large datasets (Herbert et al., 1998; Rosell-Melé et al.,
of Angola (Muller and Fischer, 2001) offer reasonable resolu- 1995b; Sikes et al., 1991; Sonzogni et al., 1997) of core-top
0
tion for one and one-half and four year periods, respectively. Uk 37 show strong convergence with the original Prahl et al.
0
Goni et al. (2001) compared the Uk 37 of monthly sediment (1988) temperature calibration, using mean annual SST
collections in the Gulf of California to SST taken from satellite (MAST) (0–10 m) as the reference (Figure 7). A compilation
0
(AVHRR). Their results (Figure 6) show that the Uk 37 index by Muller et al. (1998) synthesized results of over 300 core-top
closely tracks changes in the satellite-derived SST, with little or analyses from the different ocean basins, determined by vari-
no time offset. For most of the year, the temperatures estimated ous laboratories. Muller et al.’s (1998) dataset encompasses the
from the Prahl et al. (1988) calibration agree with SST. Some- entire range of temperatures and biogeographic provinces of
0
what lower than predicted Uk 37 values were obtained during alkenone producers but is biased toward continental margin
0
the warmest summer months (inferred SST greater than 28  C). sediments. Muller et al. (1998) noted that Uk 37 in core tops
These results could be rationalized by some combination of also correlates highly to seasonal temperatures in the upper
subsurface alkenone production during the summer thermal water column but only because these covary with mean annual
stratification, errors in the satellite SST measurements, and temperature. The correlation decreased significantly, however,
0
nonlinearity in the Uk 37 index at the warm extreme of growth if alkenone unsaturation was regressed against temperatures at
conditions. Muller and Fischer’s time-series data from the Cap 20 m and below. Sediments thus provide strong empirical
Blanc upwelling center off the coast of southwest Africa also evidence that alkenone synthesis occurs in the mixed layer in
0
show a strong seasonal cycle in Uk 37 that is consistent with most areas of the ocean. Updated recently to more than 600
changes in SST. After removing a small temporal offset due to samples by M. Conte and colleagues (Conte et al., 2006), the
0
the sinking time of particles from the surface ocean to their core-top calibration of Uk 37 to mean annual SST does not

Gulf of California sediment traps


1.1 35

Uk⬘37
1
30
SST (Satellite)

0.9
Uk⬘37

25
0.8

0.7 20
SST

0.6
15
1996 1996 1997 1998 1998
Date
k0
Figure 6 Comparison of Gulf of California sediment trap U 37 time series versus SST estimated from satellite measurements (Goni et al., 2001).
0 0
Note the rapid response of the Uk 37 index to changes in SST. The Uk 37 index may not record the warmest temperatures accurately; see text for
discussion.
Alkenone Paleotemperature Determinations 415

Global core tops −122.0 −121.5 −121.0 −120.5

1
14.0
13.9
13.9 14.2
14.1 14.0
14.2 35.0
0.8 35.0
13.8

0.6 14.1
35.5 35.5
Uk⬘37

Core top
Uk⬘37temp
0.4

−122.0 −121.5 −121.0 −120.5

0.2
−112.0 −111.0 −110.0

0
0 5 10 15 20 25 30
Levitus 1994 annual SST
0
Figure 7 Global compilation of near-surface sediment Uk 37 values 24.0 23.2 24.0
22.6
(Muller et al., 1998; Herbert, unpublished) plotted versus mean 23.3
annual SST from the closest grid point of the Levitus (1994) World 22.6
23.0
23.1

Ocean Atlas. 23.2

22.7

23.0 23.0

differ statistically from the original Prahl et al. (1988) culture


and water column line. −112.0 −111.0 −110.0
0
The reproducibility of the Uk 37 values in surface sediments
k0
from the same region can be quite extraordinary. Figure 8 Figure 8 U 37 temperature estimates taken from the top centimeter
displays two regions where the author’s laboratory was able of box cores off the California Coast (Herbert et al., 1998). Top panel
shows estimates from the central California coast and the lower panel
to obtain recent sediments from a number of box cores off the
estimates from Baja, California. Note that the temperature estimates are
coast of California. The temperature estimates derived from
0 consistent to nearly the analytical error of the gas chromatographic
Uk 37 analysis not only agree closely with MAST using the technique.
Prahl/Muller equation but they agree with each other to very
nearly within the analytical error.
Little support for the large range in alkenone parameters The success of core-top temperature calibrations indicates
(ratios of C37 to C38 ketones, alkenones to alkenoates, etc.) that physiological state, genetic variability, and depth and
observed in culture experiments comes from the sediment seasonality of production play secondary roles to the control
0
realm. As one example, the author compared (Herbert, 2000) on the sedimentary Uk 37 index exerted by near-SST. In most
the frequency distribution of the ratio of total C37 ketones cases, these factors produce errors at the level of 1.5  C or less
(SC37) to total C38 ketones (SC38) in a core-top dataset in the global core-top calibration. In this observer’s opinion,
generated by the author’s laboratory. The dataset is weighted core-top data cannot be reconciled with the large variations in
0
to samples along the California margin, but also includes the Uk 37 index attributed to genetic or physiological factors
samples from the North Atlantic, equatorial Pacific, central by some culture studies. This does not indicate that the
Pacific gyre, Peru margin, and western Pacific. The tight cluster culture data are wrong in a technical sense, but that the results
of core-top values around a mean of just over 1.0 contrasts with cannot always be extrapolated to the natural environment
the extreme heterogeneity of culture results but is in very good (Popp et al., 1998).
agreement with the average ratio of 1.04 from a large upper Core-top datasets still leave significant room for improve-
water column dataset off Bermuda (Conte et al., 2001). ment in several important regards. The important question of
0
The values found are similar to the mean of about 1.2 deter- nonlinearity at the high and low ends of the Uk 37–temperature
mined by Rosell-Melé et al. (1994) in North Atlantic core tops relation remains unsettled. In particular, several studies sug-
0
and by Sonzogni et al. (1997, 1998) in Indian Ocean gest that the Uk 37–temperature relationship flattens at tem-
sediments and the average of about 1.15 reported by Sawada peratures above 26 or 27  C (Bentaleb et al., 2002; Sikes
et al. (1996) from the Sea of Japan. Core-top data do not et al., 1997; Sonzogni et al., 1997). However, other studies of
therefore encourage the idea that alkenone systematics can sediments underlying tropical waters find that the linear Prahl
‘fingerprint’ regional variations in the fraction of production relationship seems to hold throughout the calibration range
due to E. huxleyi and G. oceanica (cf. Sawada et al., 1996; (Pelejero and Grimalt, 1997; Pelejero and Calvo, 2003) and
Volkman et al., 1995). the global sediment compilation of Muller et al. (1998)
416 Alkenone Paleotemperature Determinations

provides no support for a nonlinear relationship. One would (1998) predicts that physiological factors would lead to a
0
expect that calibrations on the extreme ends of the temperature large offset in the Uk 37 temperatures estimated from alkenones
range become difficult. Analytical difficulties grow as the di- synthesized in upwelling and nonupwelling regions. Core-top
unsaturated ketone becomes a minor peak at the low end of data cover the oceans well enough to state that such an
the index; similar difficulties pertain to the detection of the upwelling/nutrient bias must be no larger than the mean
C37:3 ketone in the face of chromatographic interferences in standard error of the entire regression (1.4  C), if it exists at
sediments under very warm ocean surface waters. In addition, all. Similarly, core-top datasets cover a large span in the prop-
it is also likely that seasonal production biases become large ortions of strains of E. huxleyi and in the contribution of
in high-latitude waters, emphasizing the need to combine G. oceanica to alkenone production, yet no statistical evidence
0
modern day ecological information with core-top data to bet- emerges to treat the Uk 37 of different oceanic biotic provinces
0
ter calibrate the Uk 37 signal in cold waters (Prahl et al., 2010; differently. (See Kienast et al. (2012) for a perspective on the
Sikes, 1997, 2005). eastern equatorial Pacific.)
There is also little consensus on the temperature signifi- The convergence of so much sedimentary data to a simple
cance of the tetraunsaturated C37 ketone. Originally included model suggests that the good relation of alkenone unsatura-
in the original Uk37 index by Brassell et al. (1986b), the C37:4 tion to mean annual SST has an underlying predictability.
ketone appears only in cold waters and in sediments underly- Examining the consequences of seasonal variations in produc-
ing cold waters. Prahl and Wakeham (1987) and Prahl et al. tion and/or remineralization of alkenones on the sedimented
0
(1988) found that including the tetraunsaturated ketone in a Uk 37 is instructive. Conte et al. (2006) modeled the impact
0
paleotemperature equation did not improve the fit to calibra- that such variations would have on the Uk 37 in a time-averaged
tion datasets of temperature and, therefore, omitted it from sediment. They found that seasonality of production produced
0
their Uk 37 index. It remains to be explained why the C37:4 only small offsets (<1  C) from SST in the net (‘integrated
ketone should be common in the high-latitude North Atlantic production temperature’) alkenone signal except at very high
where SST fall below 10  C (Calvo et al., 2002; Conte et al., latitudes. Unless the season of the alkenone-producing bloom
1994a,b; Rosell-Melé et al., 1994, 1995a,b, 1997; Sicre et al., is restricted to the precise time of the coldest or warmest
0
2002) but rare to absent in water column particles and sedi- seasonal temperatures, it is not easy to cause the Uk 37 delivered
ments of the very cold waters of the Southern Ocean (Sikes and to sediments to depart much from mean annual SST. This
Volkman, 1993; Sikes et al., 1997; Ternois et al., 1998). Rosell- lesson was demonstrated by Sonzogni et al. (1997), Muller
Melé (1998) and colleagues (Rosell-Melé et al., 2002) propose et al. (1998), and Sikes et al. (2005) who used satellite chlo-
that the presence of large amounts of the C37:4 ketone is rophyll estimates to make seasonally weighted flux estimates of
associated with low-salinity waters, at least in the North Atlan- alkenone production at core locations, assuming that alkenone
tic. This brackish alkenone producer (or producers) may also fluxes correlate with bulk phytoplankton production. They
be responsible for episodic inputs of C37:4 ketone in the North found that the resulting flux-weighted temperature corrections
Atlantic and North Pacific during particularly cold periods of to mean annual SST were negligible. It may also be that when
the Pleistocene Ice Ages (Bard et al., 2000; Harada et al., 2008; subsurface production occurs, the temperature near the top of
McClymont et al., 2008; Rosell-Melé et al., 2002). In the the seasonal thermocline is not colder by more than 1–2  C
author’s opinion, the C37:4 ketone most likely signals a discrete from mean annual SST.
strain or species of alkenone producer (perhaps closely related Alkenone production in oceanic gyres does appear to give
to lacustrine haptophytes) that is regionally restricted to evidence for a subsurface temperature bias. For example, Prahl
lower salinity waters of the North Atlantic and should not be et al. (1993), Doose et al. (1997), Herbert et al. (1998), and
0
interpreted as a direct physiological response to salinity. Popp et al. (2006) all report core-top Uk 37 values lower than

mean annual SST by 1–2 C in gyre locations in the eastern
North Pacific, consistent with subsurface fluorescence maxima
8.15.7 Synthesis of Calibration (Prahl et al., 1993) and maximal production during the late
winter and early spring in the region. Ohkouchi et al. (1999)
The resemblance of core-top alkenone unsaturation data to argued similarly for a subsurface gyre bias in a survey of core
both mean annual SST and the original Prahl et al. (1988) tops in the central North Pacific Ocean. One can therefore
culture calibration is in some sense fortuitous. Other culture expect that a careful treatment of the core-top database may
0
studies produce results that differ by as much as 5  C from the develop rules for how the Uk 37 index will deviate systemati-
standard Prahl et al. (1988) calibration, and there is no inher- cally (on the order of 1  C) according to distance from the
ent reason to prefer a linear calibration of unsaturation to nearest coastline or some other simple proxy for gyre versus
0
growth temperature to a nonlinear one. Further, it is known margin position. The influence of diagenesis on Uk 37 values
that alkenone producers do not operate at constant rates cannot be excluded, but its impact apparently ranges from
throughout the year in most regions of the ocean and that negligible (most studies) to perhaps a warm bias on the order
they do not always live in the mixed layer. One should there- of 1  C (Gong and Hollander, 1999; Hoefs et al., 1998).
fore keep in mind that the Prahl et al. (1988) calibration and Caution should probably be used in interpreting small
0
the nearly identical Muller et al. (1998) sediment core-top down-core changes in Uk 37 in high- and low-latitude regions.
0
relation of Uk 37 to mean annual SST are idealizations. It can Here, one is in less reliable analytical territory (Grimalt
be noted, however, that at least some of the caveats raised in et al., 2000; Pelejero and Calvo, 2003), and the core-top and
0
the application of alkenone thermometry do not seem borne water column Uk 37 data can be modeled by either linear
out by sedimentary evidence. The hypothesis of Epstein et al. or polynomial fits (Sikes and Volkman, 1993; Sonzogni
Alkenone Paleotemperature Determinations 417

et al., 1997). Even if evidence eventually supports nonlinear is high. Areas along various coastlines where anoxic or dysaero-
0
fits conclusively at the cold and/or warm extremes of the Uk 37 bic bottom water contacts sediments have the additional po-
range, interpreting small changes in these regions quantita- tential to resolve yearly or even seasonal variations through the
0
tively is probably unwise. Flattening of the Uk 37 relationship deposition of laminated sediments. Such highly productive,
to temperature would mean that the index loses sensitivity to highly preserving settings have elevated alkenone concentra-
temperature in very high and low latitudes. At the same time, tions in the sediments. Because so little material (perhaps only
0
analytical errors grow. One therefore needs to be wary of 10 mg) is required for a reliable Uk 37 determination, very
0
generating spuriously large temperature changes from Uk 37 high-resolution sampling becomes feasible.
deviations of dubious reliability. As a practical limit, this Two regions of the eastern Pacific have been tested by
0
author suggests not interpreting Uk 37 deviations quantitatively the alkenone method for details of past El Niño (ENSO) var-
for paleotemperatures at values lower than about 0.20 or iability. A set of early papers by Farrington et al. (1988) and
higher than about 0.98 (5 and 27.7  C, respectively, according McCaffrey et al. (1990) targeted the Peru upwelling zone,
to the Prahl/Muller equations). which experiences large surface warmings during El Niño con-
ditions. The authors used 210Pb-dated box cores to study the
alkenone record of the past few centuries at three locations
0
8.15.8 Paleotemperature Studies Using the along the Peru margin. Uk 37–temperature variations correlated
Alkenone Method in part to historical ENSO indices, but without one-to-one
matches. McCaffrey et al. (1990) attributed the mismatches
The rapidity and high precision of alkenone analysis make the to a combination of dating uncertainties, the difficulty in pre-
technique ideally suited to produce time series of past near- cisely resolving layers for sampling, and redistribution of the
surface ocean temperatures. High signal to noise can be dem- alkenone signal over a broader stratigraphic interval, so that
onstrated in a number of ways. The Ocean Drilling Program ENSO anomalies are smoothed. They also noted that alkenone
acquires offset holes at drilling sites to assure the continuity of records may be biased away from ENSO warm events by
the recovered sedimentary record. High-resolution analyses of the decline in phytoplankton production that accompanies El
offset holes that cover the same stratigraphic interval show that Niño anomalies.
0
even small-scale changes in the Uk 37 index are reproducible Sediments from the Santa Barbara Basin off southern
(Zhao et al., 1993). Alkenone temperature estimates that cover California display layers that can be sampled at annual resolu-
the late Holocene paint a picture of subdued SST changes, in tion. A pioneering study by Kennedy and Brassell (1992)
accord with evidence from polar ice cores that describes this produced annual-resolution data from the core top to an esti-
period as quite stable in comparison to other recent intervals mated basal age of AD 1915, based on a varve chronology. The
of earth history. Alkenone-derived temperatures over the past authors showed that SST varied by 1–2  C on an annual basis
10 ky rarely deviate from modern values by more than 1–2  C, in this time interval. Intervals of major historical El Niño-
even in very densely sampled records (e.g., Schulte et al., 1999; related warmings along the California coast stood out in most
0
Zhao et al., 2000). cases as warm Uk 37 anomalies. The author’s laboratory (Herbert
In the following sections, alkenone paleotemperature et al., 1998) produced a slightly longer record that confirms most
studies are divided roughly by the time period covered. While of the details of the Kennedy and Brassell (1992) record, al-
arbitrary to some degree, this should allow the reader to gauge though the absolute temperature estimates are offset due to
the contributions of the alkenone technique to important interlaboratory differences. Herbert et al. (1998) found that the
paleoclimatic questions, which tend to be arrayed by the age alkenone method detected 80–90% of the known El Niño warm-
and duration of earth history studied. In addition, studies at ings over the last century. They also found that most inferred
different time resolution will have different sets of supporting warm intervals had low alkenone abundances, which would be
information, caveats, and questions to be addressed by future consistent with the historical association of decreased upwelling
work. Nearly all of the studies cited rely on the standard Prahl along the California margin during major El Niño events. Zhao
et al. (1988) temperature scale. However, a few applications of et al. (2000) extended the Santa Barbara record to AD 1440
nonstandard calibrations exist in the literature (Calvo et al., with approximately biannual sample resolution. A significant
2002; Pelejero et al., 1999b; Wang et al., 1999). trend to low SST around the turn of the twentieth century co-
incides with a known period of low SST along the California
margin. Other approximately centennial-scale oscillations are
8.15.8.1 Holocene High-Resolution Studies
evident in the time series as well (Zhao et al., 2000). Signifi-
Because instrumental records of temperature rarely date back cantly, the approximately 500 year record shows no linear tem-
more than a century, the alkenone technique may play an perature trend over time and no dip in SST during the Little Ice
important role in characterizing past ocean SST variability on Age period of the late 1500s to early 1600s. If the Zhao et al.
timescales of a few years to centuries. Alkenone analyses may (2000) alkenone record accurately reflects regional SST, then the
thus complement information derived from geochemical ana- Little Ice Age cooling of the North Atlantic may not have been
lyses of corals, whose usefulness in studying past El Niño cycles expressed along the west coast of North America.
and other phenomena is now well established. Unlike corals, With the intense interest of the paleoclimatic community
alkenone-producing organisms range over nearly the entire in reconstructing temperature trends over the past 2000 years
ocean. Only certain locations will be favorable, however, for to place the human imprint in context, the  2000 years alke-
preserving alkenone signals for high-resolution analysis. These none record generated by Sicre et al. (2008) represents a spec-
generally occur along continental margins where sediment flux tacular example of the potential for alkenone-based methods.
418 Alkenone Paleotemperature Determinations

Sicre et al. (2008) took advantage of a core with a very high sampling along the western margin of North America (Barron
deposition rate from a site in the climatically sensitive region of et al., 2003; Kienast and McKay, 2001; Prahl et al., 1995) shows
northern Iceland to produce a record sampled every 2–5 years. no temperature trend at all during the last 9 ky, although the
A very clear Medieval Warm Period that lasted from about AD alkenone data do suggest millennial oscillations of perhaps
1000 to 1350 is revealed, along with century-scale cooler in- 1  C amplitude. Variability within the Holocene also emerges
tervals of time and decadal to centennial oscillations in tem- from Doose-Rolinski et al.’s (2001) high-resolution study of
perature (Figure 9). The increasing number of studies SST over the last 5 ky in the Arabian Sea. The authors sampled a
examining late Holocene SST change from very high-resolution high-deposition rate core at approximately 20 year intervals
alkenone paleothermometry (Mohtadi et al., 2007; Rodrigues and measured the d18O of planktonic foraminifera in addition
et al., 2010) certainly foreshadows a research trend in the to alkenone unsaturation. Variance within the 5 ky period is
coming years. about 0.6  C (1s), but temperature extremes of up to 3  C are
On somewhat longer timescales, several studies have used recorded. (It should be noted that the authors used a non-
the alkenone index to resolve variations of SST within the standard calibration, the slope of which would increase the
Holocene (last 12 ky). The Holocene represents an interesting estimated temperature changes by nearly 50% compared to the
mix of some of the factors that on longer timescales help to canonical Prahl et al. (1988) equation.) Stronger than average
drive ice age cycles. Northern hemisphere summer insolation northeast monsoon winds were inferred to result in cooler
peaked at about 9 ka and has declined continuously since that than average temperatures, with the reverse occurring when
time as the Earth’s orbital configuration shifted. The Holocene stronger southwest monsoons dominated the system. Variance
has not, however, seen any significant change in ice volume or in SST apparently increased in the last 1500 years, suggesting
atmospheric CO2. A coherent long-term cooling of between high variability of monsoonal winds during the latest Holo-
1 and 2  C over the past 9–10 ky seems to have occurred in the cene in comparison to the middle Holocene.
0
Atlantic Ocean and Mediterranean Sea, according to Uk 37 time
series (Bard et al., 2000; Cacho et al., 2002; Calvo et al., 2002;
8.15.8.2 Millennial-Scale Events of the Late Pleistocene
Marchal et al., 2002; Sachs, 2007; Zhao et al., 1995). Marchal
0 and Last Glacial Termination
et al. (2002) point out that the cooling trend exhibited by Uk 37
in the North Atlantic is consistent with evidence for late Holo- The discovery of very rapid climate anomalies in the high-
cene glacial readvance in Iceland, with borehole temperature latitude North Atlantic region sparked the search for similar
reconstructions from Greenland ice cores and with pollen events in other regions. Among the prominent anomalies are
data in Europe and North America that indicate a southward the Younger Dryas event, Dansgaard–Oeschger (D/O) cycles,
migration of the cool spruce forest. and Heinrich events. Pollen successions in northern Europe
The Holocene cooling trend detected in the North Atlantic identified a brief interval during the last deglaciation when
appears to be a regional rather than global pattern. Alkenone climate returned to very cold conditions. Spectacularly revealed
data from the Indian Ocean (Bard et al., 1997; Cayre and Bard, in the isotopic record of Greenland ice cores (Dansgaard et al.,
1999), the South China Sea (Kienast et al., 2001; Steinke 1989; Grootes et al., 1993; Johnsen et al., 1992), the Younger
et al., 2001; Wang et al., 1999), the western tropical Atlantic Dryas interval lasted from about 13 to 11 ka. Greenland ice cores
(Rühlemann et al., 1999), and eastern equatorial Pacific also demonstrate a highly unstable climate during the last
(Leduc et al., 2010) in fact show very slight sea surface warming glacial interval (marine oxygen isotope stages 2 and 3). Isotopic
from the early Holocene toward the present. High-resolution changes equivalent to 6–8  C change in air temperature occur as
rapid bursts known as the Dansgaard–Oeschger cycles. These
appear to be grouped in units of 3–4 cycles, which terminate in a
Two millennia of N Iceland SST longer period of unusually cold temperatures in Greenland.
11.0 Dansgaard–Oeschger cycles can be recognized one to one in
marine sediment cores from the North Atlantic (Bond et al.,
1993), where regionally coherent pulses of ice-rafted debris
9.5 that originate from Canada, Iceland, and the Norwegian Sea
Temperature (°C)

are termed Heinrich events.


Alkenone SST reconstructions in the North Atlantic and
8.0 Mediterranean show the Younger Dryas, Dansgaard–Oeschger
cycles, and Heinrich events very clearly (Bard et al., 2000; Cacho
et al., 1999, 2002; Calvo et al., 2001; Eglinton et al., 1992;
6.5
Rodrigues et al., 2010; Rosell-Melé et al., 1997). Millennial-scale
temperature changes apparently are exported from the North
5.0 Atlantic by the Canary Current (Bard et al., 2000; Zhao et al.,
0 400 800 1200 1600 2000 1995) or through the atmosphere (Cacho et al., 2002). Recog-
nizing that the Greenland anomalies propagate into the subtrop-
Age in years AD ical North Atlantic helps to explain initially puzzling features
Figure 9 Alkenone-derived SST record sampled at 4 years’ resolution of high-resolution alkenone data taken during the glacial
by Sicre et al. (2008) extended to modern by Sicre et al. (2011). Note interval. Zhao et al. (1995) found that the time of maximum
the prominent warm anomaly centered around AD 1000 and cooling global ice volume (last glacial maximum (LGM) 21–23 ka) did
toward the present. not have the coldest SST of the glacial period off northwest Africa.
Alkenone Paleotemperature Determinations 419

It has since become clearer that the interval of Heinrich event H2 temperatures at 10–11 ka (Kienast and McKay, 2001). The
(18 ka) produced the coldest temperatures of the late glacial Younger Dryas event shows up along the central California
period in the North Atlantic and that this cooling is detected in coast in subdued form at Ocean Drilling Program Site 1017
numerous cores off northwest Africa (Zhao et al., 1995). While (Seki et al., 2002). There, the cooling appears to be about 2  C
0
the LGM period saw cold SST in the subtropical north Atlantic according to Uk 37 estimates. Seki et al. (2002) also demon-
relative to the Holocene, it was sandwiched between even colder strated that this part of the California margin felt the effects of
periods paced by millennial climate instability. D/O and Heinrich perturbations on SST. Substantial (3–4  C)
Alkenone records from the Mediterranean Sea (Cacho et al., temperature anomalies correlate to the Greenland D/O events.
1999, 2002) show dramatic evidence for how millennial The coldest temperatures of the 80 ky record correspond to the
events pervaded the glacial regional climate of the North intervals of North Atlantic Heinrich events (Seki et al., 2002).
Atlantic and western Europe (Figure 10). All of the millennial In the tropics, the Younger Dryas period apparently led to
features of the Greenland ice cores for the last 50 ky are imme- small (0.5–1  C) coolings in many locations in phase with
diately recognizable as large SST changes, some as rapid as 6  C the North Atlantic cooling. Well-dated alkenone evidence for
per century (Cacho et al., 2002). Intervals calibrated as colder a tropical expression of Younger Dryas cooling comes from the
0
than 12  C by the Uk 37 method also have pulses of the polar South China Sea (Kienast et al., 2001; Steinke et al., 2001;
foraminifera G. pachyderma (left coiling variety). The coldest Wang et al., 1999), the Indian Ocean (Bard et al., 1997;
events correspond in each case to the time of Heinrich periods Cayre and Bard, 1999), and the South Atlantic (Kim et al.,
of intense ice rafting in the high-latitude North Atlantic. Be- 2002). Evidence for millennial-scale variability matching the
cause the times of Heinrich events do not stand out as the Greenland ice core record is weaker – the very highly sampled
coldest millennial periods in the Greenland ice core record, 40 ky record of Wang et al. (1999) in the South China Sea does
the SST data imply that these events may have a different not reveal significant changes in SST corresponding to either
spatial pattern and mode of propagation than the D/O anom- D/O or Heinrich events, but Schulte and Müller (2001) found
alies. As found in cores off northwest Africa, the LGM in the small SST anomalies in the Arabian Sea at the same time as the
Mediterranean did not have the coldest temperatures recorded North Atlantic millennial events, as did Jaeschke et al. (2007)
during the last 50 ky. These occurred at approximately 16–18, off the Brazilian coast.
24, 30, 39, and 46 ka (Cacho et al., 2002). Rapid temperature changes detected by alkenone paleother-
Details of millennial SST variability away from the North mometry may actually be anticorrelated to the North Atlantic
Atlantic are still emerging. A Younger Dryas cooling occurred pattern in some places. Records from the western tropical
along the northwest margin of North America (Barron et al., Atlantic and from the southern hemisphere might thus show
2003; Kienast and McKay, 2001; Seki et al., 2002). Its timing is warming at the same time as the coolings observed in Greenland
identical to that of the Younger Dryas in the circum-North ice cores. Resolving this question puts a premium on good
Atlantic region within the error of 14C dating by AMS. Off the dating, since the events in question lasted only centuries to
coast of British Columbia, the Younger Dryas produced a cool- one to two millennia. Rühlemann et al. (1999) produced a
ing of about 4  C from the Allerod–Bolling warm interval and densely dated (by AMS 14C) record of SST from a high-
ended with a warming of almost 6  C to peak Holocene deposition rate core off Grenada in the western tropical
Atlantic. Both d18O of the planktonic foraminifer G. ruber
0
and the Uk 37 index show modest warming during the Youn-
ger Dryas and Heinrich event H1 times. Rühlemann et al.
−34 (1999) concluded that the apparent antiphasing of western
Greenland d 18O tropical and north Atlantic SST supports the thermohaline
−36 model of millennial-scale events. Mazaud et al. (2002)
D/O events
attempted a similar analysis in the southern Indian Ocean.
20
−38 They used a geomagnetic intensity signal in the sediments to
Alkenone SST (⬚C)

synchronize alkenone and paleontological estimates of SST to


Ice core d 18O

−40 the Greenland ice core data over the interval 32–50 ka. Alke-
none analyses suggested that oscillations of 1–2  C did occur
15 −42 at millennial timescales in this region. While several cold
pulses appear to coincide with North Atlantic Heinrich
YD −44
events, longer coolings that may correlate to clusters of D/O
Mediterranean cycles appear anticorrelated to the North Atlantic. Similarly,
SST −46
H1 H2 H3 H4 H5 Barrows et al. (2007) found evidence of warming off the
10
New Zealand margin at the time the North Atlantic cooled
−48
0 10 20 30 40 50 dramatically during the Younger Dryas.
Age (ka)

Figure 10 High-resolution alkenone temperature estimates obtained by


Cacho et al. (2002) from the Mediterranean Sea in relation to oscillations 8.15.8.3 Marine Temperatures During the LGM
in temperature recorded by d18O in Greenland ice cores. Important
millennial events (Younger Dryas ¼ YD, H1 . . . H5 ¼ Heinrich events 1–5, Alkenone paleothermometry arrived at a time of controversy
and D/O ¼ Dansgaard–Oeschger events) line up between the records to on the spatial pattern of cooling at the LGM. Estimates of
the precision of the independent chronologies. tropical cooling derived from the Sr/Ca of fossil corals
420 Alkenone Paleotemperature Determinations

(Beck et al., 1997) challenged the results of the seminal Alkenone LGM reconstruction
CLIMAP effort to map surface ocean temperatures at the last 10
Ice Age (CLIMAP, 1976, 1981). CLIMAP employed a variety of
planktonic microfossil groups, principally foraminifera, to de- 8

Temp. anomaly (⬚C)


rive seasonal anomalies of LGM temperatures relative to pre-
sent day. One of their most significant conclusions was the 6
poleward amplification of glacial cooling. The tropical ocean
temperatures remained very near their present values, while 4
temperatures decreased by as much as 8–10  C in the North
Atlantic and 4–6  C in the high-latitude Southern Ocean 2
(CLIMAP, 1976, 1981). By contrast, coral Sr/Ca thermometry
placed tropical cooling in some places at 4–6  C (Beck et al., 0
1997; Guilderson et al., 1994). Noble gases in continental −1 −0.5 0 0.5 1
ground waters also suggested more cooling at low latitudes Sine (latitude)
than evident in the CLIMAP reconstruction (Stute et al., Figure 11 Compilation of available alkenone estimates of cooling of
1995). At the same time, reevaluation of tropical temperature ocean surface temperatures at the last glacial maximum (LGM) relative to
change via alkenone paleothermometry (Sikes and Keigwin, the late Holocene. The estimates have been projected onto the sine of
1994) tended to confirm the CLIMAP reconstruction. latitude to approximately compensate for the distribution of ocean
The LGM is defined by the period during the last glacial surface area from the equator to the poles. Note that the Ice Age
cycle when ice volume reached its largest extent, at about anomalies are much stronger at mid- and high latitudes than in the
tropics. Scatter at any given latitude reflects variability in the quality of
21–23 ka (EPILOG definition of Mix et al., 2001). Since the
the chronological control used to assign the LGM level in alkenone time
definition relates to ice volume and carries a specific chronos-
series but also includes an important contribution from real
tratigraphic value, it does not necessarily correspond to the heterogeneity of cooling at the LGM.
time of coldest ocean temperatures, as demonstrated in the
high-resolution studies mentioned in the previous section. As
reviewed by Mix et al. (2001), the best criterion for defining the
LGM comes from multiple bracketing AMS 14C dates. Oxygen doubt. For a good compilation of regional features of glacial-
isotopic data provide the next best tool for recognizing the Holocene temperature changes estimated by the alkenone
LGM interval. The temperatures and temperature differences method, see Leduc et al. (2010).
that are reviewed in this chapter follow the chronostratigraphic The impact of a few degrees of cooling in the tropical ocean
definition of the LGM. The reader should also keep in mind is not climatically trivial. Since half of the Earth’s surface lies
that the LGM temperature anomaly depends on the reference within 30 of the equator, any cooling has a significant effect
frame. All anomalies discussed here are expressed as the differ- on the globally integrated surface cooling during the last Ice
0
ences between alkenone SST estimates at the stratigraphic level Age. It also seems likely that more Uk 37 work will resolve
of the LGM and the latest Holocene samples, which, it is spatial patterns of cooling that implicate dynamical adjust-
assumed represent temperatures very close to modern day. ments of the tropical surface ocean to the changed sea level,
Alkenone reconstructions favor rather decisively what orbital, and CO2 boundary conditions at the LGM.
might be termed a ‘modified CLIMAP’ view of ocean surface Evidence in hand already suggests that the ocean cooled
cooling at 21 ka (Figure 11). Cases certainly occur where the differently according to region at the LGM. Although the de-
0
Uk 37 index indicates substantially more cooling at the LGM tails of how SST evolved during the last glaciation and through
than the CLIMAP reconstruction (e.g., Jasper and Gagosian, the glacial termination differ widely, one can distinguish two
1989). The alkenone method typically favors more Ice Age major patterns: those regions where the SST profile looks much
tropical cooling than suggested by CLIMAP but preserves the like the oxygen isotope curve and, hence, sea level and ice
fundamental conclusion that the tropics cooled much less than volume and those in which SST depart significantly from the
did the high latitudes (Rosell-Melé et al., 2004). In one partic- global pattern of ice volume. Many records show that SST ran
ularly telling study, Bard and colleagues (Bard et al., 1997; in parallel with the ice volume cycle, for example, in the high-
Sonzogni et al., 1998) looked at an array of 20 cores in the latitude North Atlantic (Calvo et al., 2001; Villanueva et al.,
0
Indian Ocean between 20 N and 20 S. Uk 37 data showed 1997), the South China Sea (Huang et al., 1997; Pelejero et al.,
more cooling (2–3  C) away from the equator, but only 1999; Steinke et al., 2001), off Hawaii (Lee et al., 2001), and
about 1  C cooling within 5 degrees of the equator. Signifi- parts of the South Atlantic (Kirst et al., 1999; Sachs and
0
cantly, the alkenone temperature estimates weighed in with Lehman, 2001; Schneider et al., 1995, 1996). However, Uk 37
cooling in the northern Indian Ocean where foraminiferal data indicate that another common pattern was for SST to rise
estimates had shown no cooling and even warming at the in advance of deglaciation. SST led ice volume in the equatorial
LGM (Bard et al., 1997; Sonzogni et al., 1998). The difference Pacific (Lyle et al., 1992) and tropical Indian Ocean (Bard
in results between faunal and alkenone techniques might be et al., 1997). Even within one ocean basin, there were strong
due to the superior signal to noise of alkenone analysis, to the regional differences in the timing of maximum cooling. Kirst
fact that foraminiferal faunas in the tropics respond to vari- et al. (1999) analyzed three cores from an onshore–offshore
ables other than temperature, and to the observation that transect along the Namibian upwelling margin in the South
0
glacial faunas frequently have poor analogues in the modern Atlantic. They demonstrated that both the Uk 37 and the C37
ocean, making micropaleontological estimates more open to total ketones, an index of productivity, varied systematically
Alkenone Paleotemperature Determinations 421

along the transect in the timing of maxima and minima. Close California margin some distance from the frontal boundaries
to the African continental margin warming began well before display the close association between cold SST and the LGM
the LGM. At the same time that warming occurred, the con- that one would intuitively expect. Alkenone evidence thus
centration of C37 ketones declined steeply in the sediments. suggests that several of the major eastern boundary current
Offshore, temperatures followed ice volume closely. Kirst and systems retracted poleward at the LGM.
colleagues ascribed the patterns they saw to impingement of A review of the current literature suggests that substantial
warmer waters from the north along the coastal margin at the gaps in coverage of the LGM surface ocean by the alkenone
LGM, when wind systems were reorganized. The incursion of method still exist. Very few records have been obtained from
this warmer subtropical water coincided with less favorable the interior gyres of the major ocean basins. The sediments
conditions for upwelling along the Namibian continental mar- underlying these regions are less favorable for alkenone
gin. The offshore cores are far enough from the front of analysis because they tend to have slow deposition rates and
equatorially derived water that they do not display the early low biomarker contents due to low overlying productivity.
warming and instead sense the basin-wide cooling that reached Nevertheless, the gyres are very broad features in the modern
its peak at the LGM. The author’s own laboratory has found a ocean, and their thermal state at the LGM should not be
similar pattern off the California margin, where core locations neglected. The temperature history of the warm pool of the
near the present-day southern boundary of the cold California western equatorial Pacific also awaits more investigation. How-
0
Current show only small cooling at the LGM (Figure 12; ever, this region is not well suited for the Uk 37 proxy as its
0
Herbert et al., 2001). As off Benguela, cores along the temperature lies at or above the saturation of the Uk 37 index at

16
14 ODP 1019 (42⬚ N)
Temp. (⬚C)

12 Alkenone SST
10
8
6
Benthic d 18O
4
14
12
10
8
6
ODP 1020 (41⬚ N) 4
2
20
18 ODP 893 (34⬚ N)
16
14
12
10
8
20
Early warming
18
16
14
12
ODP 1012 (32⬚ N) 10
8

25 LaPaz 21P (23⬚ N)

20
0 100
Age (ka)
Figure 12 Regional variation in the timing of SST rise at the last two glacial terminations revealed by alkenone determinations along the California
margin (Herbert et al., 2001). Data are arranged from north to south to demonstrate the regional nature of SST response (dark lines). Benthic d18O data
acquired on the same cores (lighter color) demonstrate unequivocally that SST rose early relative to the global sea level rises at each termination at many
sites. The strongest anomalies occur in the southern California borderland region (ODP 893 and ODP 1012), which today represents the boundary
between the cold, equatorward-flowing California Current, and subtropical waters which flow northward seasonally (Herbert et al., 2001).
422 Alkenone Paleotemperature Determinations

around 28  C. In addition, the alkenone dataset from the that the magnitude of the oxygen isotope stage 5e anomaly
southern hemisphere (Ikehara et al., 1997; Pahnke and Sachs, declines if it is referenced to the Holocene insolation maximum
2006; Pelejero et al., 2006) lacks the data density now available at 9 ka because many alkenone records show 1–2  C warming
for much of the northern hemisphere. in the early Holocene compared with the modern day.
Alkenone data do show characteristic regional patterns of
SST change over the 100 ky glacial cycles of the late Pleistocene.
8.15.8.4 SST Records of the Late Pleistocene Ice Age Cycles
Records from the North Atlantic tend to show the classical
Legitimate questions arise about how faithfully SST proxies will ‘sawtooth’ pattern of long decline in SST, followed by a rapid
work on the timescales of multiple Pleistocene glacial cycles. rise at glacial terminations (Calvo et al., 2001; Sikes and
0
That question is sharp in the case of the Uk 37 proxy, because Keigwin, 1996; Villanueva and Grimalt, 1996). Similar asym-
one of the principal modern producing species, E. huxleyi, first metrical cooling and warming also appear in the higher latitude
appeared in the geological record during marine oxygen isotope South Atlantic (Kirst et al., 1999) and South China Sea (Pelejero
stage 8, at about 280 ka (Thierstein et al., 1977). Indeed, the et al., 1999). In many places, however, the ocean reached its
micropaleontological record shows that E. huxleyi rose to dom- coldest temperatures well before maximum glaciation. Thus, re-
inance at different times during the last glacial cycle in different cords from the Indian Ocean consistently show that the coldest
regions (Jordan et al., 1996). Gephyrocapsids dominated the temperatures in that region coincided with oxygen isotope stage
coccolithophorid assemblage until the Holocene in regions 4 (Bard et al., 1997; Emeis et al., 1995; Rostek et al., 1993). This
such as the Benguela upwelling system (Summerhayes et al., author’s own data from off the California margin showed that
1995). Several studies that document the sequence of cocco- the onset of warming preceded the onset of deglaciation not only
lithophorids in late Pleistocene records fail to find any indica- at the end of the last Ice Age but at each of the glacial termina-
tion that the proportions of different alkenone-producing tions for the past five glacial cycles (Herbert et al., 2001). In some
0 0
species affect SST reconstructions by the Uk 37 method. For cases, the positive relation between colder Uk 37 estimates and
example, Doose-Rolinski et al. (2001) monitored the propor- increased concentrations of C37 ketones in the sediment
tions of Gephyrocapsa to Emiliania at high resolution in the late implicates enhanced upwelling as the cause of cold Ice Age
Holocene in an Arabian Sea core. High-frequency SST changes temperatures (Kirst et al., 1999). However, in many locations,
did not correspond to changes in the ratio of these species. the pattern of coccolithophorid productivity is completely
Jordan et al. (1996) also failed to find a relation between recon- decoupled from SST or in inverse relation to the expected
structed SST changes and changes in the abundance of E. huxleyi increase with colder temperatures if upwelling was the cause.
in a 130 ky record from the tropical Atlantic. Other work dem- In these cases, the temperature changes must involve changes
onstrates that the proportions of C37 ketones to C38 ketones do in ocean heat transport that occur at a larger scale than the
not vary down-core over long time periods, consistent with the regional changes in wind field.
evolutionary conservatism of alkenone synthesis (Muller et al., The spectral content of the alkenone signal has been
1997; Rostek et al., 1997; Yamamoto et al., 2000). assessed in only a few records. In one noteworthy example,
Alkenone paleotemperatures appear to give reliable infor- Schneider et al. (1999) provided evidence that tropical Atlantic
mation on orbital-scale changes in SST over the past 500 ky. SST has a strong imprint from the precessional (c.21 ky) cycle
The 100 ky cycle of glaciation dominates all alkenone SST re- of insolation. Precessional variability drives contrasts in sum-
cords that cover multiple late Pleistocene glacial cycles mer and winter heating, particularly at low latitudes. An
(Eglinton et al., 1992; Herbert et al., 2001; Martinez-Garcia expected result is that precession regulates the monsoonal
et al., 2011; Schneider et al., 1995, 1996; Sonzogni et al., cycle of the tropics. This result is consistent with the strength
1998). There is no indication of significant linear trends in of precession and the weakness of obliquity (41 ky) compo-
0
SST over the late Pleistocene, which might warn of systematic nents in the spectra of equatorial Uk 37 records. Both Schneider
diagenetic or evolutionary artifacts. Rather, the results of alke- et al. (1999) and the author’s own unpublished work off the
none work to date show remarkably consistent patterns of SST California margin shows as well that the 41 ky component in
change in relation to oxygen isotope evidence for glacial– SST is larger at higher latitudes in late Pleistocene time series
interglacial climate state. Alkenone data suggest that a number than it is at lower latitudes.
of previous interglacial intervals have been warmer by up to One puzzle that remains is the persistent evidence from
3  C from the late Holocene. In particular, alkenone estimates alkenones that the previous glacial period (oxygen isotope
for temperatures at the peak of the previous interglacial period stage 6) was not nearly as cold as the most recent glacial
(oxygen isotope stage 5e) fall consistently above the late maximum in many locations. Relatively warm estimates for
Holocene. This result is consistent with the higher northern oxygen isotope stage 6 typically come from tropical and
hemisphere insolation at 125 ka as compared to the Holocene subtropical regions (Herbert et al., 2010; Pelejero et al., 2006;
and with evidence for higher sea level than modern during the Rostek et al., 1993, 1997; Schneider et al., 1996; Villanueva
last interglacial period. The alkenone results (Herbert et al., et al., 1997). However, many other tropical locations provide
2010; Ikehara et al., 1997; Pelejero and Calvo, 2003) contrast SST estimates as cold as the most recent cycle for the previous
with the CLIMAP reconstruction for isotope stage 5e, which glacial period (Pelejero et al., 1999; Sicre et al., 2000; Wang
showed no systematic departures in SST from modern day et al., 1999). At high latitudes, the temperatures recorded by
values. The alkenone estimates of a warm oxygen isotope alkenones for oxygen isotope stage 6 are in line with those of
stage 5e are consistent, however, with the warm temperatures the last glacial period (Calvo et al., 2001; Herbert et al., 2001
inferred from Mg/Ca in planktonic foraminifer (Lea et al., and unpublished; Kirst et al., 1999; Schneider et al., 1996;
2000; Nuernberg et al., 2000). One should note, however, Villanueva et al., 1997). Furthermore, the stage 6 warm
Alkenone Paleotemperature Determinations 423

anomaly does not repeat itself in previous glacial intervals Although reports of alkenones in sediments of Pliocene,
0
where Uk 37 analysis has been performed. The only reason to Miocene, Oligocene, and even Eocene age are common (see
dismiss the alkenone result, therefore, is that it does not square Brassell, 1993; Lichtfouse et al., 1992; Rinna et al., 2002; van
with oxygen isotope evidence for a glacial climate as extreme as der Smissen and Rullkotter, 1996a,b), the interval of late Mio-
the most recent episode. If they have interpreted them cene through Pliocene is a time frame where data have been
correctly, the alkenone data appear to show that the tropical generated in enough continuity to see the evolution of SST as
0
ocean responded in different ways to glacial maxima over the recorded by the Uk 37 index. The alkenone method appears to
course of time. resolve major cooling in both the North and South Atlantic
during the onset of northern hemisphere glaciation (between
about 3 and 2.5 Ma). Warm temperatures prevailed in the
8.15.8.5 SST before the Late Pleistocene
Pliocene before the time when sizable ice sheets grew in
The long-term prospect for using alkenones to deduce the large- the northern hemisphere, particularly in eastern boundary up-
scale thermal evolution of the oceans in the Cenozoic carries welling zones (Dekens et al., 2007). Furthermore, variability on
enormous potential. Unlike the d18O proxy, the alkenone unsa- the orbital scale was low (about 1  C). Larger amplitude SST
turation index depends in theory only on temperature and, thus, changes appeared at the time that northern hemisphere ice vol-
isolates temperature changes from other influences such as ume began to grow (Herbert et al., 2010). The changes presum-
salinity and global ice volume that mingle in the d18O record ably record the onset of large, orbitally driven variations in ice
of both benthic and planktonic foraminifera. And unlike volume, along with greenhouse gas amplification (Herbert et al.,
micropaleontological-based methods for SST reconstruction, 2010). An even more substantial change in the pattern of SST
which rely on knowledge of the present-day preferences of extant over this critical paleoclimatic interval comes from the South
species, the alkenone paleothermometer may range well back Atlantic. Marlow et al. (2000) followed SST off the Benguela
in time, provided that it has behaved in an evolutionarily con- margin from 4.5 Ma to the present, again at 40–50 ky resolution.
servative manner and provided that diagenesis does not over- This cooling apparently is the second step, following a 10–12  C
print the original SST signal. Although one would clearly prefer drop in SST in the late Miocene in the South Atlantic
0
that the modern calibration of the Uk 37 index applies accurately (Rommerskirchen et al., 2011). Put together, ‘deep time’ alke-
throughout the span of alkenone records, all is not lost if none investigations of SST demonstrate that the proxy will play a
this assumption is not correct. There should still be at minimum useful role in understanding the long-term Cenozoic evolution
a role for a proxy that can unambiguously record spatial into the modern ice-house world.
and temporal variability of SST in long paleoceanographic
time series.
8.15.8.6 Comparison with other Proxies: d18O
It is one of the ironies of the paleoceanographic record that
alkenones are frequently more abundant in sediments before It should be possible to test the fidelity of alkenone SST
the appearance of E. huxleyi than they are in the period in reconstructions with an independent paleothermometer
which this coccolithophorid dominates the floral assemblage. such as d18O. There are a number of reasons why this com-
Three studies which relate the concentrations of C37 ketones to parison is not straightforward, however. First, many plank-
the abundance of E. huxleyi found that the ratio of alkenones to tonic foraminifera live below the mixed layer for part or all of
total organic carbon actually increased in sediments older than their life cycle. Comparing their isotopic temperatures to the
the E. huxleyi acme (Muller et al., 1997; Rostek et al., 1997; unsaturation index of the alkenone producers, which must
Sicre et al., 2000). Down-core profiles of alkenone concentra- live in the photic zone, may not yield consistent estimates of
tion often increase with age, the opposite pattern to what past temperature changes. Many foraminifera have seasonal
would be expected from long-term diagenetic control on their cycles of production far more pronounced than those of
abundance (Muller et al., 1997). At the same time, time series E. huxleyi and G. oceanica, so one may expect offsets in paleo-
of alkenone parameters such as the chain length ratio (SC37/ temperatures due to seasonal biases of either foraminiferal or
SC37 ketones) show no change with time over the past few alkenone production (Sikes et al., 2005). The ideal compari-
0
glacial cycles (Muller et al., 1997; Rostek et al., 1997). The son of foraminiferal isotopic and coccolithophorid Uk 37
author’s laboratory too has produced longer records that lead estimates could therefore come from the tropics, where the
people to conclude that alkenones are very refractory once seasonal amplitude in temperature is small and where some
incorporated into the sediment. Alkenone data from a 2.8 My species of planktonic foraminifera spend most of their life in
record off the coast of Southern California (Liu et al., 2005) the mixed layer. Globigerinoides ruber, which contains algal
show higher concentrations in the older part of the record than photosymbionts and consistently shows mixed-layer isotopic
in the late Pleistocene (Figure 3). Furthermore, there is no temperatures in recent sediments, makes the ideal foramini-
suggestion of a monotonic increase in SST going backward in feral species for comparison, with G. sacculifer the next best
time, as might be expected if diagenesis biased the unsatura- choice.
0
tion index. Changes in the dominant wavelength of the SST The analytical signal-to-noise ratio favors the Uk 37 as an
18
oscillations match the spectral evolution of oxygen isotopes SST proxy over d O by four to five times (e.g., a typical
from 41 ky dominance in the early Pleistocene to the rise of the analytical uncertainty of 0.1% in d18O corresponds to a tem-
100 ky cycle in the later Pleistocene (cf. Ruddiman et al., perature uncertainty of 0.4–0.5  C). Planktonic foraminiferal
1986). The stability of the chain length ratio over this 3 My isotopic values also incorporate the effects of changing global
interval, which argues for evolutionary conservatism in alke- ice volume and regional evaporation/precipitation on the iso-
none synthesis, should also be noted. topic composition of the water in which the foraminifera grow.
424 Alkenone Paleotemperature Determinations

0
Indeed, one of the promises of SST proxies such as the Uk 37 Several studies have in fact attempted to deduce regional
index is to constrain temperatures so that the other compo- changes in salinity by pairing foraminiferal d18O data with
nents of foraminiferal d18O can be studied. alkenone paleotemperature estimates. In a seminal paper,
The amplitudes of temperature change at the LGM inferred Rostek et al. (1993) subtracted the isotopic temperature effect
by both alkenones and planktonic d18O seem largely consis- from an Indian Ocean d18O curve obtained from G. ruber using
0
tent. Broecker (1986) pointed out that the magnitude of trop- the Uk 37 index. Their reconstruction put glacial salinity of the
ical sea surface cooling at the LGM was strongly constrained by tropical Indian Ocean as 1–2 psu higher than today; they fur-
the modest amount of isotopic enrichment in foraminifera ther suggested that salinities were fresher than modern during
beyond that required by the global ice volume effect. Estimat- the previous interglacial peak at oxygen isotope stage 5e. Re-
0
ing the ice volume effect precisely has been difficult. The ca- gional reconstructions of both d18O and Uk 37 from the Medi-
nonical ice volume effect of 1.2% comes from a sea level/d18O terranean also appear to resolve glacial–interglacial shifts in
calibration (Fairbanks and Matthews, 1978). More recent esti- salinity (Cacho et al., 2002; Emeis et al., 2000). The isotopic
mates based on pore-water deconvolution put the ice volume shift immediately outside the Mediterranean in the Gulf of
effect at 0.9–1% (Schrag et al., 1996). As discussed previously, Cadiz is modest (1.8%), as is cooling (3.5  C). Glacial age
alkenone SST estimates support tropical cooling of 1–2  C in isotopic anomalies grow progressively in the Mediterranean
most locations. This cooling is enough to favor the recent from 2.5% in the Alboran Sea to over 3% in the eastern Med-
smaller estimates for the global ice volume effect, because iterranean (Cacho et al., 2002; Emeis et al., 2000). Much of the
it would correct the typical glacial isotopic enrichment of greater isotopic amplitude can be explained by intense surface
1–1.5% by several tenths per mil (see Figure 13). Alternatively, ocean cooling in the interior Mediterranean (5–8  C, depend-
one could infer a relative freshening of many tropical locations ing on location), away from the Atlantic connection. However,
from the comparison of temperature and isotopic change. the residual isotopic enrichment within the Mediterranean in-
Pelejero et al. (1999) provide a regional dataset of G. ruber dicates that this region was even more saline than today during
isotopic values with corresponding alkenone SST estimates in the last glacial period, especially in the eastern basin. Another
the South China Sea. Cores from the northern part of the sea regionally coherent pattern of salinity change comes from the
display amplitudes in d18O of 1.4–1.6% between the late South China Sea (Wang et al., 1999). Surface waters in this
Holocene and LGM. The corresponding alkenone temperature region, in contrast to the Mediterranean, apparently freshened
anomalies are about 3.2–3.5  C, equivalent to about 0.7% in during the LGM in comparison to the present day.
d18O for the temperature effect. The lower latitude sites from One danger, of course, for alkenone deconvolution of
the South China Sea produce smaller glacial–interglacial dif- planktonic d18O records comes from the fact that the proxies
0
ferences in d18O (1.3–1.4%) and in Uk 37 temperatures (2.3– originate from two different types of organisms. Some of the
2.6  C). Cores collected off Hawaii yield glacial–interglacial differences in amplitude and timing of the two signals may
isotopic differences from G. ruber of 1.1–1.3% and alkenone- result from ecological changes, rather than changes in the d18O
based cooling of 2–3  C (Lee et al., 2001). On the other hand, of the water in which the foraminifera grew. Such ambiguities
Muller et al. (1997) compared isotopic and alkenone data in a are present in all multiproxy paleoenvironmental work; resolv-
tropical (11 S) core from the South Atlantic and found higher ing them is one of the chief challenges of paleoceanography.
isotopic amplitude (1.9%) accompanied by an approximately
0
3.5  C temperature change deduced from the Uk 37 index. Such
8.15.8.7 Comparison with other Proxies: Microfossils
evidence certainly suggests that regional changes in the d18O of
seawater in the surface ocean at the LGM have to be considered CLIMAP (1976, 1981) reconstructed temperatures at the LGM
in addition to the global ice volume effect. by using the modern day ecological preferences of microfossil

GeoB-1016
−2.5
−2 Uk⬘37temperature

−1.5
−1
18
Planktic d O
d 18O

−0.5
0
0.5
Temperature-deconvolved d 18O
1
1.5
0 100 200 300 400
Age (ka)
Figure 13 Alkenone data acquired with planktonic foraminiferal d18O can be used to remove the temperature effect from oxygen isotopic signals in
order to assess global (ice volume) and regional (evaporation/precipitation balance) contributions to isotopic change. Data come from an equatorial
Atlantic core studied by Schneider et al. (1996); the isotopic deconvolution was made by the present author. Note that the structure and amplitude
of the temperature-corrected planktonic d18O record are consistent with current thinking on the global ice volume signal at the last glacial maximum
of about þ1% (Schrag et al., 1996).
Alkenone Paleotemperature Determinations 425

groups to estimate past SST. Since the CLIMAP effort, micro- Lea et al., 1999, 2000; Rosenthal et al., 2000). The alkenone
paleontologists have developed new methods of reconstruct- temperature proxy may also have the advantage in several other
ing SST in addition to the transfer-function method pioneered aspects. Synthesis of alkenones is tied to the photic zone, unlike
by CLIMAP investigators. Rather different sets of glacial– that of foraminiferal calcite. The signal to noise of alkenone
interglacial temperature anomalies can result from applying measurements is superior for the paleotemperature range
different microfossil calibrations to the same dataset (Mix 5–25  C, where the exponential dependence of Mg/Ca results
et al., 2001; Waelbroeck and Steinke, 2002). It is therefore in a relatively low-temperature sensitivity (Elderfield and
simplistic to compare alkenone methods to a particular mode Ganssen, 2000; Lea et al., 1999, 2000). At higher temperatures,
of micropaleontological reconstruction as a general test of the same exponential sensitivity to temperature increases the
either the geochemical or the microfossil approach. Experience signal to noise of the Mg/Ca proxy. Unlike the alkenone proxy,
shows that spatially sparse datasets can often produce apparent the Mg/Ca proxy should not lose sensitivity to temperature at
conflicts at the LGM that come simply from extrapolating the warm end of ocean temperatures. The Mg/Ca proxy should
results from one proxy over too great a distance (Broccoli and therefore be the tool of choice in determining past changes in
Marciniak, 1996). SST in warm pool regions of the ocean.
0
Temperature estimates produced from alkenones and fora- Two direct comparisons of Uk 37 and Mg/Ca estimates of
miniferal assemblages generally agree quite well for the Holo- glacial–interglacial SST changes come from the tropical Atlantic.
cene (but see Marchal et al., 2002, for some discrepancies in Elderfield and Ganssen (2000) compared alkenone-estimated
the North Atlantic and Mediterranean). Alkenone tempera- temperatures with the Mg/Ca and d18O of five species of
tures usually fall between the warm and cold season estimates planktonic foraminifera at core site BOFS31K, the same core
from foraminifera (Chapman et al., 1996; Perez-Folgado et al., analyzed by Chapman et al. (1996) for foraminiferal SST
2003; Wang et al., 1999). Since both methods are in some estimates. Elderfield and Ganssen (2000) refer to Mg/Ca tem-
sense optimized for late Holocene conditions, this is perhaps peratures as ‘calcification temperatures’ to account for the ob-
not surprising. Estimates of SST at the LGM more commonly servation that many planktonic foraminifera calcify significantly
diverge (Chapman et al., 1996; Huang et al., 1997; Perez- below the mixed layer. Mg/Ca data produced two down-core
Folgado et al., 2003). It is not evident how to reconcile changes clusters of species: a cold group (G. inflata, N. pachyderma, and G.
0
in the Uk 37 index relative to seasonal SST estimates derived menardii) and a warm group with Mg/Ca temperatures consis-
from foraminiferal faunas. In one of the best-studied examples, tently 3–3.5  C warmer (G. ruber and G. bulloides). Core-top
0
Chapman et al. (1996) found that Uk 37 estimates for glacial temperatures estimated from the warm group fell about 2  C
temperatures followed summer temperatures estimated from colder than modern SST, while the alkenone estimate falls right
planktonic foraminifera at a subtropical North Atlantic site. at mean annual modern SST (Elderfield and Ganssen, 2000).
Chapman et al. (1996) argued that the alkenone producers Temperature changes estimated from G. ruber tracked alkenone
must have changed their season of production from late winter temperatures very closely, while maintaining the approximately
and spring (present day) to summer during glacial and early 2  C offset toward colder temperatures. Mg/Ca measurements
Holocene times. This ecological switch, according to the au- on G. bulloides showed almost no cooling for most of the glacial
thors, explains why the alkenone thermometer would under- period. Elderfield and Ganssen (2000) concluded that the fora-
estimate glacial SST cooling at the study location. Chapman miniferal Mg/Ca temperature estimates at the LGM follow
et al. (1996) do note that the temperatures inferred from the alkenone estimate (about 2  C) more closely than the fora-
alkenones match the isotopic temperatures derived from the miniferal transfer-function estimate (cooling of almost 5  C).
planktonic foraminifer G. bulloides very well. Precisely the op- Their work also hints at the ecological uncertainties (diverging
0
posite temporal dichotomy between Uk 37 and foraminiferal depth and seasonal habitats) that enter into interpreting temp-
estimates was obtained from the South China Sea, where gla- erature records from different planktonic organisms over the
cial alkenone temperatures are consistently colder than the course of glacial to interglacial climate change.
cold season foraminiferal estimate (Huang et al., 1997). Nuernberg et al. (2000) compared the Mg/Ca and alkenone
As noted earlier, most alkenone estimates from the tropics approaches over three glacial–interglacial cycles at a core at 1 S
support more cooling than inferred by CLIMAP (see Bard et al., in the central South Atlantic. The foraminifer used was
1997; Lee et al., 2001; Pelejero et al., 1999; Sonzogni et al., G. sacculifer. Core-top temperature estimates differed by about
1998). It is not clear if the differences are larger than overlap of 2  C. Once again, the Mg/Ca estimate fell consistently cooler
the standard error of the techniques. In most cases, the alkenone than the alkenone record down-core. In most cases, the temper-
method has a clear advantage in signal to noise for resolving ature curves were offset by 2–3  C. Occasionally, as at the last
smaller temperature changes (see e.g., Lee et al., 2001). glacial termination, the two temperature profiles coincided.
The authors noted strong qualitative similarities in the two
paleotemperature datasets. Both gave similar cooling at the
0
LGM (3.5  C for the Uk 37 index and 3.4  C for Mg/Ca), a
8.15.8.8 Comparison with other Proxies: Mg/Ca
warmer-than-Holocene stage 5e and a warmer previous glacial
0
As with the Uk 37 proxy, Mg/Ca measurements in foraminiferal cycle than the last glacial interval. Both geochemical proxies
calcite may yield paleotemperature measurements without the gave smaller SST changes at the LGM than planktonic foraminif-
ambiguities of d18O or micropaleontological assemblage ana- eral transfer-function estimates, which gave 9  C cooling during
lyses. In contrast to the alkenone proxy, preservation markedly the cold season and 5  C cooling for the warm season. They also
affects the Mg/Ca ratio in foraminiferal shells and, hence, agreed on the greater warmth of oxygen isotope stage 5e relative
the paleotemperature estimate (Brown and Elderfield, 1996; to the Holocene, while the foraminiferal transfer function
426 Alkenone Paleotemperature Determinations

0
did not. But the correlation coefficients for the Mg/Ca and Uk 37 1.5  C. Using multiple proxy techniques to reconstruct past
proxies were not strong: 0.49 over the 270 ky record and 0.78 SST contains the promise of improved signal to noise but also
over the last 90 ky (Nuernberg et al., 2000). comes with pitfalls – what happens when proxies disagree? In
particular, how do we interpret disagreements larger than the
ostensible absolute temperature errors of each method (e.g.,
8.15.8.9 Comparison with other Proxies: TEX86 and other
Saher et al., 2009)? To be clear, this is a case of looking at the
Glycerol Dialkyl Glycerol Tetraethers Indices
glass as half empty, for many comparisons show quite good
Another SST proxy based on sediment biomarkers has been agreement between proxies (Dekens et al., 2008). Nevertheless,
developed in the last decade: the TEX86 (Schouten et al., 2002) the question of resolving conflicting estimates of past SST using
index and variants (Kim et al., 2010), all based on glycerol multiple proxies is a real and vexing one.
dialkyl glycerol tetraethers (GDGT). The open-ocean source of The primary causes of proxy discrepancy would seem to lie
GDGT appears to be free-living crenarchaeota. Unlike the hap- in variations over time in patterns of seasonal production of
tophyte alkenone producers, crenarchaeota do not phot- the producing organisms, changes in their average depth of
osynthesize and are free to range up and down the water production, and issues related to preservation of the tempera-
column. Nevertheless, evidence suggests that crenarcheota are ture signals during early diagenesis. These sources of variability
most abundant in the upper water column (Karner et al., almost certainly play out regionally; one would not expect
2001). Recent compilations of sediment core-top GDGT anal- global rules. In the case of multiple proxy estimates, should
yses produce correlations to mean annual SST comparable to an average of all available information be preferred or is one
alkenone and Mg/Ca proxies, although the best formulation proxy a better estimation of SST? What rules should apply? To
of GDGT indices is still in flux (Kim et al., 2010). the author, consistency provides a powerful lens to make this
Using GDGT-based proxies to estimate SST has some ad- judgment. If one looks at the modern ocean, SST patterns
vantages over the alkenone method. Perhaps the strongest (whether as zonal and meridional gradients or as temporal
argument in its favor is the fact that GDGT indices do not anomalies) play out on broad (>1000 km) length scales.
saturate at warm temperatures. Organic methods can therefore One would expect that faithful SST proxies should show sim-
be extended into warm pool regions of the modern ocean or to ilar patterns for the past.
‘deep time’ investigations of globally warm climates. GDGT In its ability to yield consistent regional trends of SST over
proxies also appear even more resistant to alteration by micro- time, the alkenone method fares well. This seems particularly
bial processes than the alkenone unsaturation index (Kim true for the Holocene epoch, where temperature changes have
et al., 2009). almost certainly been much smaller than over the course of ice
However, as the lens on the GDGT proxy has continued to age cycles. In contrast to foraminiferal faunal and d18O esti-
focus, evidence has mounted that GDGT-based estimates of mates, the alkenone technique apparently has the reliability
temperature will more frequently reflect subsurface, rather necessary to define late Holocene cooling trends of only 0.27
than mixed layer, temperatures than either the alkenone or to  0.15  C ky1 in a coherent manner in the northeast Atlan-
Mg/Ca methods. A modern sediment trap study in the Santa tic and Mediterranean Sea (Marchal et al., 2002). Leduc et al.
Barbara Basin found GDGT production temperatures substan- (2010) similarly used a large alkenone and Mg/Ca dataset to
tially below SST over the annual cycle (Huguet et al., 2007). assess the regional coherence of Holocene and marine oxygen
Lee et al. (2008) looked at the spatial gradients in alkenone, isotope stage 5e SST trends. They found that alkenone-inferred
TEX86, and foraminiferal-based SST proxies in surface particu- trends varied by region, but that they did produce regionally
late matter, at 100 m depth, and in modern sediments along consistent patterns. Mg/Ca-inferred trends were much noisier,
the Benguela margin. TEX-based estimates corresponded to perhaps because the Mg/Ca estimate incorporates information
0
conditions below the mixed layer. Uk 37 estimates from core from a relatively small number (30–40) of individual forami-
tops corresponded closely to mean annual temperature, de- nifera. Leduc et al. (2010) conclude, “Yet it remains difficult to
spite the highly seasonal production in the Benguela upwelling assess if Mg/Ca first-order SST trends capture regional-scale
system. Production depths of 100–150 m would be consistent features or very small-scale hydrographic conditions. The
not only in this location but are supported by geological time most puzzling result is likely the lack of reproducibility of the
series that show TEX-based temperatures systematically lower basin-wide SST picture drawn by alkenones in individual
than alkenone-based estimates across glacial–interglacial cy- Mg/Ca-based records.” It is safe to say that much more work
cles in the eastern tropical Atlantic (dos Santos et al., 2010). on these interesting intercomparisons is needed before they
However, it does not seem likely that a simple cold bias will can confidently produce integrated assessments of SST using
always explain differences between alkenone and GDGT multiple proxies.
methods of paleotemperature estimation: Huguet et al.
(2011) find warm offsets of several degrees of the TEX86
0
index relative to Uk 37 in a time series covering the penultimate 8.15.9 Conclusions
glacial cycle from the western Mediterranean.
This review paints an optimistic picture of the ability of alke-
none unsaturation indices recorded in sediments to capture
8.15.8.10 Intercomparison of SST Proxies: Some
past SST changes on a wide variety of timescales. The author
Generalities
believes this optimism is justified on a number of grounds.
Different methods of estimating marine paleotemperatures have First, the analytical precision of the estimate is outstand-
similar RMS errors in modern calibration datasets – about ing. One can anticipate that interlaboratory differences will
Alkenone Paleotemperature Determinations 427

diminish over time as a result of intercalibration experiments References


(e.g., Rosell-Melé et al., 2001) and the maturation of analytical
protocols. The precision of alkenone analyses is meaningful, in Aksnes DL, Egge JK, Rosland R, and Heimdal BR (1994) Representation of Emiliania
huxleyi in phytoplankton simulation models. A first approach. Sarsia 79: 291–300.
the sense that core-top sediments from the same region yield
0 Andruleit H (1997) Coccolithophore fluxes in the Norwegian-Greenland Sea:
very similar values for Uk 37 and very similar down-core esti- Seasonality and assemblage alterations. Marine Micropaleontology 31: 45–64.
mates of cooling at time horizons such as the LGM. Unlike the Andruleit H, von Rad U, Bruns A, and Ittekkot V (2000) Coccolithophore fluxes from
SST proxies derived from foraminifera, the alkenone index sediment traps in the northeastern Arabian Sea off Pakistan. Marine
survives extensive degradation in the water columns and sedi- Micropaleontology 38: 285–308.
Antia AN, Maaßen J, Herman P, et al. (2001) Spatial and temporal variability of particle
ments. And it seems unlikely that the calibration of alkenone flux at the N.W. European continental margin. Deep Sea Research Part II: Topical
unsaturation to growth temperature will change with addi- Studies in Oceanography 48: 3083–3106.
tional data. Balch WM, Holligan PM, and Kilpatrick KA (1992) Calcification, photosynthesis and
The relationship observed in sediments closely mimics that growth of the bloom-forming coccolithophore, Emiliania huxleyi. Continental Shelf
Research 12: 1353–1374.
measured between unsaturation in photic zone particles (Conte
Bard E, Rostek F, and Sonzogni C (1997) Interhemispheric synchrony of
et al., 2006) and in situ temperature. It should be noted, how- the last deglaciation inferred from alkenone paleothermometry. Nature
0
ever, that the calibration of sedimentary Uk 37 to mean annual 385: 707–710.
SST represents a statistical relationship which should not be Bard E, Rostek F, Turon J-L, and Gendreau S (2000) Hydrological impact of Heinrich
interpreted in too simplistic a manner. It has been seen that Events in the subtropical northeast Atlantic. Science 289: 1321–1324.
Barron JA, Heusser L, Herbert T, and Lyle M (2003) High-resolution climatic evolution
this relation holds for sediments because alkenone synthesis of coastal northern California during the past 16,000 years. Paleoceanography
generally occurs in the mixed layer and because alkenone pro- 18(1): 1020.
duction by selected haptophyte algae is less seasonal than for Barrows TT, Lehman SJ, Fifield LK, and De Deckker P (2007) Absence of cooling in
many other planktonic groups. Furthermore, temperatures dur- New Zealand and the adjacent ocean during the Younger Dryas Chronozone.
Science 318: 86–89.
ing the season of maximal alkenone production generally come
Beaufort L and Heussner S (1999) Coccolithophorids on the continental slope of
close to mean annual SST. The index therefore is a quite robust the Bay of Biscay – Production, transport and contribution to mass fluxes. Deep Sea
proxy for mean annual temperature. It should be kept in mind Research Part II: Topical Studies in Oceanography 46: 2147–2174.
that regional variations in the factors that control the depth and Beaufort L and Heussner S (2001) Seasonal dynamics of calcareous nannoplankton
seasonality of alkenone production can cause the index to devi- on a West European continental margin: The Bay of Biscay. Marine
Micropaleontology 43: 27–55.
ate from mean annual SST. As is the case for many other paleo- Beck JW, Recy J, Taylor F, Edwards RL, and Cabioch G (1997) Abrupt changes in early
temperature proxies, uncertainties grow at the extreme ends of Holocene tropical sea surface temperature derived from coral records. Nature
0
the temperature range sampled by the Uk 37 index. These come 385: 705–707.
in part from analytical problems and in part from the lack of a Bell MV and Pond D (1996) Lipid composition during growth of motile and coccolith
forms of Emiliania huxleyi. Photochemistry 41: 465–471.
theoretical basis for preferring a linear or curvilinear calibration 0
Bentaleb I, Fontugne M, and Beaufort L (2002) Long-chain alkenones and Uk 37
at very cold and warm temperatures. Proxies such as Mg/Ca and/ variability along a south-north transect in the Western Pacific Ocean. Global and
or GDGT-based estimates are almost certainly preferable at the Planetary Change 34: 173–183.
warm end (>28  C) of the paleotemperature scale. Bentaleb I, Grimalt JO, Vidussi F, et al. (1999) The C37 alkenone record of seawater
The association of alkenones with fine-grained particles temperatures during seasonal thermocline stratification. Marine Chemistry
64: 301–313.
may cause more significant problems for the proxy than the Benthien A and Müller PJ (2000) Anomalously low alkenone temperatures caused by
calibration uncertainty. Recent evidence for advection of alke- lateral particle and sediment transport in the Malvinas Current region, western
nones by deep currents (Benthien and Müller, 2000; Ohkouchi Argentine Basin. Deep Sea Research Part I: Oceanographic Research Papers
et al., 2002) suggests that paleoceanographers ignore sedimen- 47: 2369–2393.
Bollman J (1997) Morphology and biogeography of Gephyrocapsa coccoliths in
tology at their peril. In particular, the spectacular evidence
0 Holocene sediments. Marine Micropaleontology 29: 319–350.
from Uk 37 for rapid SST changes in some drift sediments Bond GC, Broecker WS, Johnsen S, et al. (1993) Correlations between climate records
(Sachs and Lehman, 1999) may need to be reinterpreted as a from North Atlantic sediments and Greenland ice. Nature 366: 552–554.
signal of pulses of input from high latitudes. New tools such as Boon JJ, van der Meer FW, Schuyl PJ, de Leeuw JW, Schenck PA, and Burlingame AL
14
C AMS dating of alkenones will lead to a clearer picture of (1978) Organic geochemical analyses of core samples from Stie 362, Walvis Ridge,
DSDP Leg 40. Deep Sea Drilling Project, Initial Reports 38, 39, 40, 41 Supplement:
which sedimentary environments are the most favorable for 627–637.
0
applying the Uk 37 index as a measure of past SST. Sedimento- Brand LE (1982) Genetic variability and spatial patterns of genetic differentiation in the
logical expertise in choosing appropriate sites for alkenone- reproductive rates of the marine coccolithophores Emiliania huxleyi and
based work is certainly suggested. Gephyrocapsa oceanica. Limnology and Oceanography 27: 236–245.
Brand LE (1984) The salinity tolerance of forty-six marine phytoplankton isolates.
Alkenone paleotemperature estimates have already weighed
Estuarine, Coastal and Shelf Science 18: 543–556.
in significantly on the controversy of tropical SST at the LGM. Brand LE (1991) Minimum iron requirements of marine phytoplankton and the
The ‘modified CLIMAP’ view they support seems consistent implications for the biogeochemical control of new production. Limnology and
with tropical cooling of 2–3  C estimated from Mg/Ca of Oceanography 36: 1756–1771.
planktonic foraminifera (Lea et al., 2000; Nuernberg et al., Brand LE (1994) Physiological ecology of marine coccolithophores. In: Winter A and
Siesser WG (eds.) Coccolithophores, pp. 39–49. Cambridge, UK: Cambridge
2000; Stott et al., 2002; Visser et al., 2003). SST change over University Press.
deeper timescales awaits more investigation by the alkenone Brand LE and Guillard RRL (1981) The effects of continuous light and light intensity on
method. All indications point to the diagenetic stability of the reproduction rates of twenty-two species of marine phytoplankton. Journal of
alkenones over millions of years. The long evolutionary history Experimental Marine Biology and Ecology 50: 119–132.
Brassell SC (1993) Applications of biomarkers for delineating marine paleoclimate
and apparent conservatism of alkenone synthesis suggest that
fluctuations during the Pleistocene. In: Engel MH and Macko SA (eds.)
the proxy will play a useful role in solving major paleoclimatic Organic Geochemistry: Principles and Applications, pp. 699–738. New York:
questions over most of the Cenozoic. Plenum Press.
428 Alkenone Paleotemperature Determinations

0
Brassell SC, Brereton RG, Eglinton G, et al. (1986a) Palaeoclimatic signals recognized the alkenone production temperature recorded by Uk 37 in sediments with
by chemometric treatment of molecular stratigraphic data. Organic Geochemistry overlying sea surface temperature. Geochemistry, Geophysics, Geosystems
10: 649–660. 7: Q02005.
Brassell SC, Eglinton G, and Howell VJ (1987) Palaeo-environmental assessment for Conte MH, Thompson A, and Eglinton G (1994a) Primary production of lipid biomarker
marine organic-rich sediments using molecular organic geochemistry. In: Fleet AJ compounds by Emiliania huxleyi: Results from an experimental mesocosm
and Brooks J (eds.) Marine Petroleum Source Rocks, pp. 79–98. London: study in fjords of southern Norway. Sarsia 79: 319–332.
Blackwell. Conte MH, Thompson A, Eglinton G, and Green JC (1995) Lipid biomarker diversity in
Brassell SC, Eglinton G, Marlowe IT, Pflaumann U, and Sarnthein M (1986b) Molecular the coccolithophorid Emiliania huxleyi (Prymnesiophyceae) and the related species
stratigraphy: A new tool for climatic assessment. Nature 320: 129–133. Gephyrocapsa oceanica. Journal of Phycology 31: 272–282.
Broccoli AJ and Marciniak EP (1996) Comparing simulated glacial climate and Conte MH, Thompson A, Lesley D, and Harris RP (1998a) Genetic and physiological
paleodata: A reexamination. Paleoceanography 11: 3–14. influences on the alkenone/alkenoate versus growth temperature relationship in
Broecker WS (1986) Oxygen isotope constraints on surface ocean temperatures. Emiliania huxleyi and Gephyrocapsa oceanica. Geochimica et Cosmochimica Acta
Quaternary Research 26: 121–134. 62: 51–68.
Broecker WS (1994) Massive iceberg discharges as triggers for global climate change. Conte M, Volkman JK, and Eglinton G (1994b) Lipid biomarkers of the Haptophyta.
Nature 372: 421–424. In: Green JC and Leadbeater BSC (eds.) The Haptophyte Algae, pp. 351–377.
Broecker WS (1998) Paleocean circulation during the Last Deglaciation: A bipolar Oxford: Clarendon Press.
seesaw? Paleoceanography 13: 119–121. Conte MH, Weber JC, King LL, and Wakeham SG (2001) The alkenone temperature
Broerse ATC, Brummer G-JA, and Van Hinte JE (2000c) Coccolithophore export signal in the western North Atlantic surface waters. Geochimica et Cosmochimica
production in response to monsoonal upwelling off Somalia (northwestern Indian Acta 65: 4275–4287.
Ocean). Deep Sea Research Part II: Topical Studies in Oceanography Conte MH, Weber JC, and Ralph N (1998b) Episodic particle flux in the deep Sargasso
47: 2179–2205. Sea – An organic geochemical assessment. Deep Sea Research Part I:
Broerse ATC, Ziveri P, and Honjo S (2000a) Coccolithophore (-CaCO3) flux in the Sea Oceanographic Research Papers 45: 1819–1841.
of Okhotsk: Seasonality, settling and alteration processes. Marine Cortés MY, Bollmann J, and Thierstein HR (2001) Coccolithophore ecology at the HOT
Micropaleontology 39: 179–200. station ALOHA, Hawaii. Deep Sea Research Part II: Topical Studies in Oceanography
Broerse ATC, Ziveri P, Van Hinte JE, and Honjo S (2000b) Coccolithophore export 48: 1957–1981.
production, species composition, and coccolith-CaCO3 fluxes in the NE Atlantic Cranwell PA (1985) Long-chain unsaturated ketones in recent lacustrine sediments.
(34 N 21 W and 48 N 21 W). Deep Sea Research Part II: Topical Studies in Geochimica et Cosmochimica Acta 49: 1545–1551.
Oceanography 47: 1877–1905. Dansgaard W, White JWC, and Johnsen SJ (1989) The abrupt termination of the
Brown S and Elderfield H (1996) Variations in Mg/Ca and Sr/Ca ratios of planktonic Younger Dryas climate event. Nature 339: 532–534.
foraminifera caused by postdepositional dissolution – Evidence of shallow de Leeuw JW, Meer FWVD, Rijpstra WIC, and Schencck PA (1980) On the occurrence
Mg-dependent dissolution. Paleoceanography 11: 543–551. and structural identification of long chain unsaturated ketones and hydrocarbons in
Brown CW and Yoder JA (1994) Coccolithophorid blooms in the global ocean. Journal sediments. In: Douglas AG and Maxwell JR (eds.) Advances in Organic
of Geophysical Research 99(C4): 7467–7482. Geochemistry 1979. Physics and Chemistry of the Earth, vol. 12, pp. 211–217.
Bukry D (1974) Coccoliths as paleosalinity indicators – Evidence from the Black Sea. Oxford: Pergamon Press.
In: Degens ET and Ross DA (eds.) The Black Sea – Geology, Chemistry and Biology. Dekens PS, Ravelo AC, and McCarthy MD (2007) Warm upwelling regions in the
AAPG Memoir, vol. 20, pp. 353–363. Tulsa, OK: American Association of Petroleum Pliocene warm period. Paleoceanography 22: PA3211.
Geologists. Dekens PS, Ravelo AC, McCarthy MD, and Edwards CA (2008) A 5 million year
Cacho I, Grimalt JO, and Canals M (2002) Response of the Western Mediterranean Sea comparison of Mg/Ca and alkenone paleothermometers. Geochemistry,
to rapid climatic variability during the last 50,000 years: A molecular biomarker Geophysics, Geosystems 9: Q10001.
approach. Journal of Marine Systems 33–34: 253–272. Doose H, Prahl FG, and Lyle MW (1997) Biomarker temperature estimates from modern
Cacho I, Pelejero C, Grimalt JO, Calafat AM, and Canals M (1999) C37 alkenone and last glacial surface waters of the California Current system between 33 and
measurements of sea surface temperature in the Gulf of Lions (NW Mediterranean). 42 N. Paleoceanography 12: 615–622.
Organic Geochemistry 30: 557–566. Doose-Rolinski H, Rogalla U, Scheeder G, Luckge A, and von Rad U (2001)
Cadee GC (1985) Macroaggregates of Emiliania huxleyi in sediment traps. Marine High-resolution temperature and evaporation changes during the late Holocene
Ecology Progress Series 24: 193–196. in the northeastern Arabian Sea. Paleoceanography 16: 358–367.
Calvo E, Grimalt J, and Jansen E (2002) High resolution Uk37 sea surface temperature dos Santos RA, Lopes PM, Castaneda IS, et al. (2010) Glacial-interglacial
reconstruction in the Norwegian Sea during the Holocene. Quaternary Science variability in Atlantic meridional overturning circulation and thermocline adjustments
Reviews 21: 1385–1394. in the tropical North Atlantic. Earth and Planetary Science Letters 300: 407–414.
Calvo E, Villanueva J, Grimalt JO, Boelaert A, and Labeyrie L (2001) New insights into Edvardsen B, Eikrem W, Green JC, Andersen RA, Moon-van der Staay SY, and
0
the glacial latitudinal temperature gradients in the North Atlantic. Results from Uk 37 Medlin LK (2000) Phylogenetic reconstructions of the Haptophyta inferred from 18S
sea surface temperatures and terrigenous inputs. Earth and Planetary Science ribosomal DNA sequences and available morphological data. Phycologia
Letters 188: 509–519. 39: 19–35.
Cayre O and Bard E (1999) Planktonic foraminiferal and alkenone records of the last Eglinton G, Bradshaw SA, Rosell A, Sarnthein M, Pflaumann U, and Tiedemann R
deglaciation from the Eastern Arabian Sea. Quaternary Research 52: 337–342. (1992) Molecular record of secular sea surface temperature changes on 100-year
0
Chaler R, Grimalt JO, Pelejero C, and Calvo E (2000) Sensitivity effects in Uk 37 timescales for glacial terminations I, II and IV. Nature 356: 423–426.
paleotemperature estimation by chemical ionization mass spectrometry. Analytical Elderfield H and Ganssen G (2000) Past temperature and d18O of surface ocean waters
Chemistry 72: 5892–5897. inferred from foraminiferal Mg/Ca ratios. Nature 405: 442–445.
Chapman MR, Shackleton NJ, Zhao M, and Eglinton G (1996) Faunal and alkenone Eltgroth ML, Watwood RL, and Wolfe G (2005) Production and cellular localization
reconstructions of subtropical North Atlantic surface hydrography and of neutral long-chain lipids in the haptophyte algae Isochrysis galbana and
paleotemperature over the last 28 kyr. Paleoceanography 11: 343–357. Emiliania huxleyi. Journal of Phycology 41: 1000–1009.
CLIMAP Project Members (1976) The surface of the Ice-Age Earth. Science Emeis K-C, Anderson DM, Doose H, Kroon D, and Schulz-Bull D (1995)
191: 1131–1137. Sea-surface temperatures and the history of monsoon upwelling in the
CLIMAP Project Members (1981) Seasonal reconstruction of the earth’s surface at the northwest Arabian Sea during the last 500,000 years. Quaternary Research
last glacial maximum. Geological Society of America, Map and Chart Series 36. 43: 355–361.
Conte MH and Eglinton G (1993) Alkenone and alkenoate distributions within the Emeis K-C, Struck U, Schulz H-M, et al. (2000) Temperature and salinity variations of
euphotic zone of the eastern North Atlantic: Correlation with production temperature. Mediterranean Sea surface waters over the last 16,000 years from records of
Deep Sea Research Part I: Oceanographic Research Papers 40: 1935–1961. planktonic stable isotopes and alkenone unsaturation ratios. Palaeogeography,
Conte MH, Eglinton G, and Madureira LAS (1992) Long-chain alkenones and alkyl Palaeoclimatology, Palaeoecology 158: 259–280.
alkenoates as palaeotemperature indicators: Their production, flux, and early Epstein BL, D’Hondt S, and Hargraves PE (2001) The possible metabolic role of C37
sedimentary diagenesis in the Eastern North Atlantic. Organic Geochemistry alkenones in Emiliania huxleyi. Organic Geochemistry 32: 867–875.
19: 287–298. Epstein BL, D’Hondt S, Quinn JG, Zhang J, and Hargraves PE (1998) An effect of
Conte MH, Sicre M-A, Rühlemann C, et al. (2006) The global temperature calibration of dissolved nutrient concentrations on alkenone-based temperature estimates.
0
the alkenone unsaturation index (Uk 37) in surface waters and a comparison of Paleoceanography 13: 122–126.
Alkenone Paleotemperature Determinations 429

Fairbanks RG and Matthews RK (1978) The marine oxygen isotopic record in Higginson MJ, Altabet MA, Wincze L, Herbert TD, and Murray DW (2004) A solar
Pleistocene coral, Barbados, West Indies. Quaternary Research 10: 181–196. (irradiance) trigger for millennial-scale abrupt changes in the southwest monsoon?
Farrimond PG, Eglinton PG, and Brassell SC (1986) Alkenones in Cretaceous black Paleoceanography 19: PA3015.
shales, Blake-Bahama Basin, western North Atlantic. Organic Geochemistry Hoefs MJL, Versteegh GJM, Ripstra WIC, de Leeuw JW, and Sinninghe Damsté JS
10: 897–903. (1998) Postdepositional oxic degradation of alkenones: Implications for the
Farrington JW, Davis AC, Sulanowski J, et al. (1988) Biogeochemistry of lipids in measurement of palaeo sea surface temperatures. Paleoceanography 13: 42–59.
surface sediments of the Peru upwelling area at 15 S. Organic Geochemistry Houghton SD and Guptha MVS (1991) Monsoonal and fertility controls on recent
19: 277–285. marginal sea and continental shelf coccolith assemblages from the western Pacific
Ficken KJ and Farrimond P (1995) Sedimentary lipid geochemistry of Framvaren: and the northern Indian oceans. Marine Geology 97: 251–259.
Impacts of a changing environment. Marine Chemistry 51: 31–43. Huang CY, Wu SF, Zhao M, et al. (1997) Surface ocean and monsoon climate variability
Fisher NS and Honjo S (1989) Intraspecific differences in temperature and salinity in the South China Sea since the last glaciation. Marine Micropaleontology
responses in the coccolithophore Emiliania huxleyi. Biological Oceanography 32: 71–94.
6: 355–361. Huguet C, Kim JH, de Lange GJ, Sinninghe Damsté JS, and Schouten S (2009)
0
Freeman KH and Wakeham SG (1992) Variation in the distribution and isotopic Effects of long term oxic degradation on the U37K , TEX86 and BIT organic proxies.
composition of alkenones in Black Sea particles and sediments. Organic Organic Geochemistry 40: 1188–1194.
Geochemistry 19: 277–285. Huguet C, Martrat B, Grimalt JO, Sinninghe Damsté JS, and Schouten S (2011)
0
Fujiwara S, Tsuzuki M, Kawchi M, Minaka N, and Inouye I (2001) Molecular phylogeny Coherent millennial-scale patterns in U37K and TEX86H temperature records
of the Haptophyta based on the rbcL gene and sequence variation in the spacer during the penultimate interglacial-to-glacial cycle in the western Mediterranean.
region of the RUBISCO operon. Journal of Phycology 37: 121–129. Paleoceanography 26: PA2218.
Giraudeau J, Monteiro MS, and Nikodemus K (1993) Distribution and malformation of Huguet C, Schimmelmann A, Thunell R, Lourens LJ, Sinninghe Damsté JS, and
living coccolithophores in the northern Benguela upwelling system off Namibia. Schouten S (2007) A study of the TEX86 paleothermometer in the water column
Marine Micropaleontology 22: 93–110. and sediments of the Santa Barbara Basin, California. Paleoceanography 22: 3203.
Gong C and Hollander DJ (1999) Evidence for differential degradation of alkenones Ikehara M, Kawamura K, Ohkouchi NJ, et al. (1997) Alkenone sea surface temperature in
under contrasting bottom water oxygen conditions: Implication for paleotemperature the Southern Ocean for the last two deglaciations. Geophysical Research Letters
reconstruction. Geochimica et Cosmochimica Acta 63: 405–411. 24: 679–682.
Goñi MA, Aceves HL, Thunell RC, et al. (2003) Biogenic fluxes in the Cariaco Basin: Jaeschke A, Ruehlemann C, Arz H, Heil G, and Lohmann G (2007) Coupling of
A combined study of sinking particles and underlying sediments. Deep Sea millennialscale changes in sea surface temperature and precipitation off
Research Part I: Oceanographic Research Papers 50: 781–807. northeastern Brazil with high-latitude climate shifts during the last glacial period.
Goñi MA, Hartz DM, Thunell RC, and Tappa E (2001) Oceanographic considerations Paleoceanography 22: PA4206. http://dx.doi.org/10.1029/2006PA001391.
0
in the application of the alkenone-based paleotemperature Uk 37 index in the Gulf of Jasper J and Gagosian RB (1989) Alkenone molecular stratigraphy in an oceanic
California. Geochimica et Cosmochimica Acta 65: 545–557. environment affected by glacial freshwater events. Paleoceanography 4: 603–614.
Grice K, Klein Breteler WCM, Schoten S, Grossi V, de Leeuw JW, and Sinninge Jiang MJ and Gartner S (1984) Neogene and Quaternary calcareous nannofossil
Damsté JS (1998) Effects of zooplankton herbivory on biomarker proxy records. biostratigraphy of the Walvis Ridge. Deep Sea Drilling Project Initial Reports
Paleoceanography 13: 686–693. 74: 561–595.
0
Grimalt JO, Calvo E, and Pelejero C (2001) Sea surface paleotemperature errors in Uk 37 Johnsen SJ, Clausen HB, Dansgaard W, et al. (1992) Irregular glacial interstadials
estimation due to alkenone measurements near the limit of detection. recorded in a new Greenland ice core. Nature 359: 311–313.
Paleoceanography 16: 226–232. Jordan RW, Zhao M, Eglinton G, and Weaver PPE (1996) Coccolith and alkenone
Grimalt JO, Rullkotter J, Sicre M-A, et al. (2000) Modifications of the C37 alkenone stratigraphy at a NW African upwelling site (ODP 658C) over the last 130,000 years.
and alkenoate composition in the water column and sediment: Possible implications In: Moguilevsky AL and Whateley R (eds.) Microfossils and Oceanic Environments,
for sea surface temperature estimates in paleoceanography. Geochemistry, pp. 111–130. University of Wales-Aberystweth Press.
Geophysics, Geosystems 1(11): 1031. Karner M, DeLong EF, and Karl DM (2001) Archaeal dominance in the mesopelagic
Grootes PM, Stuiver M, White JWC, Johnsen S, and Jouzel J (1993) Comparison of zone of the Pacific Ocean. Nature 409: 507–510.
oxygen isotope records from the GISP2 and GRIP Greenland ice cores. Nature Kennedy J and Brassell SC (1992) Molecular records of twentieth century El Nino
366: 552–554. events in laminated sediments from Santa Barbara Basin. Nature 357: 62–64.
Grossi V, Rapel D, Auber C, and Rontani J-F (2000) The effect of growth temperature on Kienast M, MacIntyre G, Dubois N, et al. (2012) Alkenone unsaturation in
the long-chain alkenes composition in the marine coccolithophorid Emiliania surface sediments from the eastern equatorial Pacific: Implications for SST
huxleyi. Photochemistry 54: 393–399. reconstructions. Paleoceanography 27: PA1210.
Guilderson TP, Fairbanks RG, and Rubenstone JL (1994) Tropical temperature Kienast M, Steinke S, Stattegger K, and Calvert SE (2001) Synchronous tropical South
variations since 20,000 years ago: Modulating interhemispheric climate change. China Sea SST change and Greenland warming during deglaciation. Science
Science 263: 663–665. 291: 2132–2134.
Haidar AT and Thierstein HR (2001) Coccolithophore dynamics off Bermuda Kienast SS and McKay JL (2001) Sea surface temperatures in the subarctic northeast
(N. Atlantic). Deep Sea Research Part II: Topical Studies in Oceanography Pacific reflect millennial-scale climate oscillations during the last 16 kyr.
48: 1925–1956. Geophysical Research Letters 28: 1563–1566.
Hamanaka J, Sawada K, and Tanoue E (2000) Production rates of C37 alkenones Kim J-H, Huguet C, Zonneveld KAF, et al. (2009) An experimental field study to test the
0
determined by 13C-labeling technique in the euphotic zone of Sagami Bay, Japan. stability of lipids used for the TEX86 and U37K palaeothermometers. Geochimica et
Organic Geochemistry 31: 1095–1102. Cosmochimica Acta 73: 2888–2898.
Harada N, Handa N, Harada K, and Matsuoka H (2001) Alkenones and particulate Kim J-H, Schneider RR, Müller PJ, and Wefer G (2002) Interhemispheric comparison of
fluxes in sediment traps from the central equatorial Pacific. Deep Sea Research deglacial sea-surface temperature patterns in Atlantic eastern boundary currents.
Part I: Oceanographic Research Papers 48: 891–907. Earth and Planetary Science Letters 194: 383–393.
Harada N, Sato M, and Sakamoto T (2008) Freshwater impacts recorded in Kim J-H, van der Meer J, Schouten S, et al. (2010) New indices and calibrations derived
tetraunsaturated alkenones and alkenone sea surface temperatures from the Okhotsk from the distribution of crenarchaeal isoprenoid tetraether lipids: Implications for
Sea across millennial-scale cycles. Paleoceanography 23: PA3201. past sea surface temperature reconstructions. Geochimica et Cosmochimica Acta
Hefter J (2008) Analysis of alkenone unsaturation indices with fast gas chromatography/ 74: 4639–4654.
time-of-flight mass spectrometry. Analytical Chemistry 80: 2161–2170. Kirst G, Schneider RR, Muller PJ, von Storch I, and Wefer G (1999) Late Quaternary
Herbert TD (2000) Review of alkenone calibrations culture, water column, and temperature variability in the Benguela Current system derived from alkenones.
sediments. Geochemistry, Geophysics, Geosystems 2(2): 1005–1010. Quaternary Research 52: 92–103.
Herbert TD, Peterson LC, Lawrence KT, and Liu Z (2010) Tropical ocean temperatures Koopmans MP, Schaeffer-Reiss C, de Leeuw JW, et al. (1997) Sulphur and oxygen
over the past 3.5 Myr. Science 328: 1530–1534. sequestration of n-C37 and n-C38 unsaturated ketones in an immature kerogen and
Herbert TD, Schuffert JD, Andreasen D, et al. (2001) Collapse of the California Current the release of their carbon skeletons during early stages of thermal maturation.
during glacial maxima linked to climate change on land. Science 293: 71–76. Geochimica et Cosmochimica Acta 61: 2397–2408.
Herbert TD, Schuffert JD, Thomas D, Lange K, Weinheimer A, and Herguera J-C (1998) Lea DW, Mashiotta TA, and Spero HJ (1999) Controls on magnesium and strontium
Depth and seasonality of alkenone production along the California margin uptake in planktonic foraminifera determined by live culturing. Geochimica et
inferred from a core-top transect. Paleoceanography 13: 263–271. Cosmochimica Acta 63: 2369–2379.
430 Alkenone Paleotemperature Determinations

Lea DW, Pak DK, and Spero HJ (2000) Climate impact of late Quaternary equatorial Mouzdahir A, Grossi V, Bakkas S, and Rontani J-F (2001) Visible light-dependent
Pacific sea surface temperature variations. Science 289: 1719–1724. degradation of long-chained alkenes in killed cells of Emiliania huxleyi and
Leduc G, Schneider R, Kim J-H, and Lohmann G (2010) Holocene and Eemian sea Nannocchlorpsis salina. Photochemistry 56: 677–684.
surface temperature trends as revealed by alkenone and Mg/Ca paleothermometry. Muller PJ, Cepek M, Ruhland G, and Schneider RR (1997) Alkenone and
Quaternary Science Reviews 29: 989–1004. coccolithophorid species changes in late Quaternary sediments from the Walvis
Lee KE, Kim JH, Wilke I, Helmke P, and Schouten S (2008) A study of the alkenone, TEX86, Ridge: Implications for the alkenone paleotemperature method. Palaeogeography,
and planktonic foraminifera in the Benguela Upwelling System: Implications for past Palaeoclimatology, Palaeoecology 135: 71–96.
sea surface temperature estimates. Geochemistry, Geophysics, Geosystems 9: Q10019. Muller PJ and Fischer G (2001) A 4-year sediment trap record of alkenones from the
Lee KY, Slowey N, and Herbert TD (2001) Glacial sea surface temperatures in the filamentous upwelling region off Cape Blanc, NW Africa and a comparison with
0
subtropical North Pacific: A comparison of Uk 37, d18O, and foraminiferal distributions in underlying sediments. Deep Sea Research Part I: Oceanographic
assemblage temperature estimates. Paleoceanography 16: 268–279. Research Papers 48: 1877–1903.
Levitus S (1994) Climatological Atlas of the World Ocean. NOAA Professional Paper 13. Muller PJ, Kirst G, Ruhland G, von Storch I, and Rosell-Melé A (1998) Calibration of the
0
Washington, DC: US Government Printing Office. alkenone paleotemperature index Uk 37 based on core-tops from the eastern South
Lichtfouse E, Littke R, Disko U, Willshc H, Ruldotter J, and Stein R (1992) Geochemistry Atlantic and the global ocean (60 N-60 S). Geochimica et Cosmochimica Acta
and petrology of organic matter in Miocene to Quaternary deep sea sediments 62: 1757–1772.
from the Japan Sea (Sites 798 and 799). Proceedings of the Ocean Drilling Program, Nanninga HJ and Tyrrell T (1996) The importance of light for the formation
Scientific Results 127/128: 667–675. of algal blooms by Emiliania huxleyi. Marine Ecology Progress Series
Liu Z, Altabet MA, and Herbert TD (2005) Glacial-interglacial modulation of eastern 136: 195–203.
tropical North Pacific denitrification over the last 1.8-Myr. Geophysical Research Nishida S (1986) Nannoplankton flora in the Southern Ocean, with special reference to
Letters 32: L23607. siliceous varieties. Memoirs of National Institute of Polar Research Special Issue
Lyle M, Prahl FG, and Sparrow MA (1992) Upwelling and productivity changes inferred 40: 56–68.
from a temperature record in the central equatorial Pacific. Nature 355: 812–815. Nuernberg D, Muller A, and Schneider RR (2000) Paleo-sea surface temperature
Madureira LAS, Conte MH, and Eglinton G (1995) The early diagenesis of lipid calculations in the equatorial east Atlantic from Mg/Ca ratios in planktic
0
biomarker compounds in North Atlantic sediments. Paleoceanography foraminifera: A comparison to sea surface temperature estimates from Uk 37, oxygen
10: 627–642. isotopes, and foraminiferal transfer functions. Paleoceanography 15: 124–134.
Marchal O, Cacho I, Stocker TF, et al. (2002) Apparent long-term cooling of the sea Ohkouchi N, Eglinton TI, Keigwin LD, and Hayes JM (2002) Spatial and temporal offsets
surface in the northeast Atlantic and Mediterranean during the Holocene. Quaternary between proxy records in a sediment drift. Science 298: 1224–1227.
Science Reviews 21: 455–483. Ohkouchi N, Kawamura K, Kawahata H, and Okada H (1999) Depth ranges of
Marlow JR, Lange CB, Wefer G, and Rosell-Melé A (2000) Upwelling intensification as alkenone production in the central Pacific Ocean. Global Biogeochemical Cycles
part of the Pliocene-Pleistocene climate transition. Science 290: 2288–2291. 13: 695–704.
Marlowe IT, Brassell SC, Eglinton G, and Green JC (1984a) Long chain unsaturated Ohkouchi N, Kawamura K, Nakamura T, and Taira A (1994) Small changes in the sea
ketones and esters in living algae and marine sediments. Organic Geochemistry surface temperature during the last 20,000 years: Molecular evidence from the
6: 135–141. western tropical Pacific. Geophysical Research Letters 21: 2207–2210.
Marlowe IT, Brassell SC, Eglinton G, and Green JC (1990) Long-chain alkenones and Okada H and Honjo S (1973) The distribution of oceanic coccolithophorids in the
alkyl alkenoates and the fossil coccolith record of marine sediments. Chemical Pacific. Deep Sea Research and Oceanographic Abstracts 20: 355–374.
Geology 88: 349–375. Okada H and McIntyre A (1979) Seasonal distribution of modern coccolithophores in
Marlowe IT, Green JC, Neal AC, Brassell SC, Eglinton G, and Course PA (1984b) the western North Atlantic Ocean. Marine Biology 54: 319–328.
Long chain (n-C37-C39) alkenones in the Prymnesiophyceae. Distribution of Paasche E (2002) A review of the coccolithophorid Emiliania huxleyi
alkenones and other lipids and their taxonomic significance. British Phycological (Prymnesiophyceae) with particular reference to growth, coccolith formation, and
Journal 19: 203–216. calcification-photosynthesis interactions. Phycologia 40: 503–529.
Martinez-Garcia A, Rosell-Melé A, Jaccard SL, Geibert W, Sigman DM, and Haug GH Pan H and Sun M-Y (2011) Variations of alkenone based paleotemperature index (UK’37)
(2011) Southern Ocean dust-climate coupling over the past four million years. during Emiliania huxleyi cell growth, respiration (auto-metabolism) and microbial
Nature 476: 312–315. degradation. Organic Geochemistry 42: 678–687.
Mazaud A, Sicre MA, Ezat U, et al. (2002) Geomagnetic-assisted stratigraphy and sea Pahnke K and Sachs JP (2006) Sea surface temperatures of southern midlatitudes
surface temperature changes in core MD94-103 (Southern Indian Ocean): 0-160 kyr B.P. Paleoceanography 21: PA2003.
0
Possible implications for North-South climatic relationships around H4. Earth and Pelejero C and Calvo E (2003) The upper end of the Uk 37 temperature calibration
Planetary Science Letters 201: 159–170. revisited. Geochemistry, Geophysics, Geosystems 4: 1014.
McCaffrey MA, Farrington JW, and Repeta DJ (1990) The organic geochemistry of Peru Pelejero C, Calvo E, Barrows TT, Logan GA, and De Deckker P (2006) South Tasman
margin surface sediments: 1. A comparison of the C37 alkenone and historical El Sea alkenone palaeothermometry over the last four glacial/interglacial cycles.
Nino records. Geochimica et Cosmochimica Acta 54: 1671–1682. Marine Geology 230: 73–86.
0
McClymont EL, Rosell-Melé A, Haug GH, and Lloyd JM (2008) Expansion of subarctic Pelejero C and Grimalt JO (1997) The correlation between the Uk 37 index and sea
water masses in the North Atlantic and Pacific oceans and implications for surface temperatures in the warm boundary: The South China Sea. Geochimica et
mid-Pleistocene ice sheet growth. Paleoceanography 23: PA4214. Cosmochimica Acta 61: 4789–4797.
0
McIntrye A (1970) Gephyrocapsa protohuxleyi sp. n. as possible phyletic link and index Pelejero C, Grimalt JO, Heilig S, Kienast M, and Wang L (1999) High-resolution Uk 37
fossil for the Pleistocene. Deep Sea Research and Oceanographic Abstracts temperature reconstructions in the South China Sea over the past 220 kyr.
17: 187–190. Paleoceanography 14: 224–231.
Medlin LK, Barker GLA, Campbell L, et al. (1996) Genetic characterization of Emiliania Perez-Folgado M, Sierro FJ, Flores JA, et al. (2003) Western Mediterranean planktonic
huxleyi (Haptophyta). Journal of Marine Systems 9: 13–31. foraminifera events and millennial climatic variability during the last 70 kyr.
Mercer JL, Zhao M, and Coleman SM (1999) Alkenone evidence of sudden changes in Marine Micropaleontology 48: 49–70.
Chesapeake Bay Conditions ca. 300 years ago. Eos, Transactions American Popp BN, Kenig F, Wakeham SG, and Bidigare RR (1998) Does growth rate affect ketone
Geophysical Union 80(supplement): S185. unsaturation and intracellular carbon isotopic variability in Emiliania huxleyi?
Mix AC, Bard E, and Schneider R (2001) Environmental processes of the Ice Age: Land, Paleoceanography 13: 35–41.
ocean, glaciers (EPILOG). Quaternary Science Reviews 20: 627–657. Popp BN, Prahl FG, Wallsgrove RJ, and Tanimoto J (2006) Seasonal patterns of
Mix AC, Morey AE, Pisias NG, and Hostetler SW (1999) Foraminiferal faunal estimates alkenone production in the subtropical oligotrophic North Pacific.
of paleotemperature: Circumventing the no-analog problem yields cool ice age Paleoceanography 21: PA1004.
tropics. Paleoceanography 14: 350–359. Prahl FG, Collier RB, Dymond J, Lyle M, and Sparrow MA (1993) A biomarker
Mohtadi M, Oppo DW, Lueckge A, et al. (2011) Reconstructing the thermal structure of perspective on prymnesiophyte productivity in the northeast Pacific Ocean. Deep
the upper ocean: Insights from planktic foraminifera shell chemistry and alkenones Sea Research Part I: Oceanographic Research Papers 40: 2061–2076.
in modern sediments of the tropical eastern Indian Ocean. Paleoceanography Prahl FG, de Lange GJ, Lyle M, and Sparrow MA (1989a) Post-depositional stability of
26: PA3219. long-chain alkenones under contrasting redox conditions. Nature 341: 434–437.
Mohtadi M, Romero OE, Kaiser J, et al. (2007) Cooling of the southern high latitudes Prahl FG, Dymond J, and Sparrow MA (2000a) Annual biomarker record for
during the Medieval Period and its effect on ENSO. Quaternary Science Reviews export production in the central Arabian Sea. Deep-Sea Research Part II: Topical
26: 1055–1066. Studies in Oceanography 47: 1581–1604.
Alkenone Paleotemperature Determinations 431

Prahl F, Herbert T, Brassell SC, et al. (2000b) Status of alkenone paleothermometer Rosenthal Y, Lohmann GP, Lohmann KC, and Sherrell RM (2000) Incorporation
calibration: Report from Working Group 3. Geochemistry, Geophysics, and preservation of Mg in Globigerinoides sacculifer: Implications for
Geosystems 1(11): 1034. reconstructing the temperature and 18O/16O of seawater. Paleoceanography
Prahl FG, Muelhausen LA, and Lyle M (1989b) An organic geochemical assessment of 15: 135–145.
oceanographic conditions at MANOP Site C over the past 26,000 years. Rostek F, Bard E, Beaufort L, Sonzogni C, and Ganssen G (1997) Sea surface
Paleoceanography 5: 495–510. temperature and productivity records for the past 240 kyr in the Arabian Sea. Deep
Prahl FG, Muelhausen LA, and Zahnle DL (1988) Further evaluation of long-chain Sea Research Part II: Topical Studies in Oceanography 44: 1461–1480.
alkenones as indicators of paleoceanographic conditions. Geochimica et Rostek F, Ruhland G, Bassinot FC, et al. (1993) Reconstructing sea surface temperature
Cosmochimica Acta 52: 2303–2310. using d18O and alkenone records. Nature 364: 319–321.
Prahl FG, Pilskaln CH, and Sparrow MA (2001) Seasonal record for alkenones in Roth PH (1994) Distribution of coccoliths in oceanic sediments. In: Winter A and
sedimentary particles from the Gulf of Maine. Deep Sea Research Part I: Siesser WG (eds.) Coccolithophores, pp. 199–218. Cambridge, UK: Cambridge
Oceanographic Research Papers 48: 515–528. University Press.
Prahl FG, Pisias N, Sparrow MA, and Sabin A (1995) Assessment of sea-surface Ruddiman WF, Raymo M, and McIntyre A (1986) Matuyama 41000 year cycles: North
temperature at 42 N in the California Current over the last 30,000 years. Atlantic ocean and northern hemisphere ice sheets. Earth and Planetary Science
Paleoceanography 10: 763–773. Letters 80: 117–129.
Prahl FG, Popp BN, Karl DM, and Sparrow MA (2005) Ecology and biogeochemistry of Rühlemann C and Butzin M (2006) Alkenone temperature anomalies in the
alkenone production at station ALOHA. Deep Sea Research Part I: Oceanographic Brazil-Malvinas Confluence area caused by lateral advection of suspended
Research Papers 52: 699–719. particulate material. Geochemistry, Geophysics, Geosystems 7: Q10015.
Prahl FG, Rontani J-F, Zabeti N, Walinsky SE, and Sparrow MA (2010) Systematic Rühlemann C, Mulitza S, Muller PJ, Wefer G, and Zahn R (1999) Warming of the
0
pattern in UK 37 – Temperature residuals for surface sediments from high tropical Atlantic Ocean and slowdown of thermohaline circulation during the last
latitude and other oceanographic settings. Geochimica et Cosmochimica Acta deglaciation. Nature 402: 511–514.
74: 131–143. Sachs JP (2007) Cooling of Northwest Atlantic slope waters during the Holocene.
Prahl FG and Wakeham SG (1987) Calibration of unsaturation patterns in long-chain Geophysical Research Letters 34: L03609.
ketone compositions for paleotemperature assessment. Nature 330: 367–369. Sachs JP and Lehman SJ (1999) Subtropical North Atlantic temperatures 60,000 to
Prahl FG, Wolfe GV, and Sparrow MA (2003) Physiological impacts on alkenone 30,000 years ago. Science 286: 756–759.
paleothermometry. Paleoceanography 18: 1025. Sachs JP and Lehman SJ (2001) Glacial surface temperatures of the southeast Atlantic
Pujos A (1987) Late Eocene to Pleistocene medium-sized and small-sized Ocean. Science 293: 2077–2079.
‘Reticulofenestrids’. Abhandlungen der Geologischen Bundensant Austria Saher MH, Rostek F, Jung SJA, et al. (2009) Western Arabian Sea SST during the
0
39: 239–277. penultimate interglacial: A comparison of Uk 37 and Mg/Ca paleothermometry.
Rinna J, Warning B, Meyers PA, Brumsack H-J, and Rullkötter J (2002) Combined Paleoceanography 24: PA2212.
organic and inorganic geochemical reconstruction of paleodepositional conditions Samtleben C and Bickert T (1990) Coccoliths in sediment traps from the Norwegian
of a Pliocene sapropel from the eastern Mediterranean Sea. Geochimica et Sea. Marine Micropaleontology 16: 39–64.
Cosmochimica Acta 66: 1969–1986. Sawada K, Handa N, and Nakatsuka T (1998) Production and transport of long-chain
Rio D (1982) The fossil distribution of coccolithophore genus Gephyrocapsa Kamptner alkenones and alkyl alkenoates in a sea water column in the northwestern Pacific off
and related Plio-Pleistocene chronostratigraphic problems. Deep Sea Drilling central Japan. Marine Chemistry 59: 219–234.
Project Initial Reports 68: 325–343. Sawada K, Handa N, Shiraiwa Y, Danbara A, and Montani S (1996) Long-chain
Rodrigues T, Grimalt JO, Abrantes F, Naughton F, and Flores J-A (2010) The last alkenones and alkyl alkenoates in the coastal and pelagic sediments of the
glacial-interglacial transition (LGIT) in the eastern mid-latitudes of the North northwest North Pacific, with special reference to the reconstruction of
Atlantic: Abrupt sea surface temperature change and sea level implications. Emiliania huxleyi and Gephyrocapsa oceanica ratios. Organic Geochemistry
Quaternary Science Reviews 29: 1853–1862. 24: 751–764.
Rommerskirchen F, Tegan C, Mollenhauer G, Dupont L, and Schefuss L (2011) Schneider RR, Muller PJ, and Acheson R (1999) Atlantic alkenone sea surface
Miocene to Pliocene development of surface and subsurface temperatures in the temperature records. In: Abrantes F and Mix AC (eds.) Reconstructing Ocean
Benguela Current system. Paleoceanography 26: PA3216. History: A Window into the Past, pp. 33–55. New York: Academic Press.
Rontani J-F, Cuny P, Gossi V, and Beker B (1997) Stability of long-chain alkenones Schneider RR, Muller PJ, and Ruhland G (1995) Late Quaternary surface circulation in
in senescing cells of Emiliania huxleyi: Effect of photochemical and aerobic the east equatorial Atlantic: Evidence from alkenone sea surface temperatures.
0
microbial degradation on the alkenone unsaturation ratio (Uk 37). Organic Paleoceanography 10: 197–220.
Geochemistry 26: 503–509. Schneider RR, Muller PJ, Ruhland G, Meinecke G, Schmidt H, and Wefer G (1996) Late
Rontani J-F, Zabeti N, and Wakeham SG (2011) Degradation of particulate organic Quaternary surface temperatures and productivity in the east-Equatorial South
matter in the equatorial Pacific Ocean: Biotic or abiotic? Limnology and Atlantic: Response to changes in trade/monsoon wind forcing and surface water
Oceanography 56: 333–349. advection. In: Wefer G, Berger WH, Siedler G, and Webb DJ (eds.) The South
Rosell-Melé A (1998) Interhemispheric appraisal of the value of alkenone indices as Atlantic: Present and Past Circulation, pp. 527–551. Berlin: Springer.
temperature and salinity proxies in high-latitude locations. Paleoceanography Schouten S, Hopmans EC, Schefuss E, and Sinninghe Damsté JS (2002)
13: 694–703. Distributional variations in marine crenarchaeotal membrane lipids: A new tool
Rosell-Melé A, Bard E, Emeis KC, et al. (2004) Sea surface temperature anomalies in the for reconstructing ancient sea water temperatures? Earth and Planetary
0
oceans at the LGM estimated from the alkenone-UK37k index: Comparison with Science Letters 204: 265–274.
GCMs. Geophysical Research Letters 31: L03208. Schrag DP, Hampt G, and Murray DW (1996) Pore fluid constraints on the
Rosell-Melé A, Bard E, Emeis E-C, et al. (2001) Precision of the current methods to temperature and oxygen isotopic composition of the Glacial Ocean. Science
0
measure the alkenone proxy U37k and absolute alkenone abundance in sediments: 272: 1930–1932.
Results of an interlaboratory comparison study. Geochemistry, Geophysics, Schulte S and Müller PJ (2001) Variations in sea-surface temperature and primary
Geosystems 2(7): 1046. productivity during Heinrich and Dansgaard–Oeschger events in the northeastern
Rosell-Melé A, Carter J, and Eglinton G (1994) Distribution of long-chain alkenones Arabian Sea. Geo-Marine Letters 21: 168–175.
and alkyl alkenoates in marine surface sediments from the North East Atlantic. Schulte S, Rostek F, Bard E, Rullkotter J, and Marchal O (1999) Variations of
Organic Geochemistry 22: 501–509. oxygen-minimum and primary productivity recorded in sediments of the Arabian
Rosell-Melé A, Carter J, Parry AT, and Eglinton G (1995a) Determination of the Uk37 Sea. Earth and Planetary Science Letters 173: 205–221.
index in geological samples. Analytical Chemistry 67: 1283–1289. Schulz H, Emeis K-C, Erlenkeuser H, von Rad U, and Rolf C (2002) The Toba volcanic
Rosell-Melé A, Eglinton G, Pflaumann U, and Sarnthein M (1995b) Atlantic core-top event and interstadial/stadial climates at the marine isotopic stage 5 to 4 transition
calibration of the U37k index as a sea-surface temperature indicator. Geochimica et in the northern Indian Ocean. Quaternary Research 57: 22–31.
Cosmochimica Acta 59: 3099–3107. Schulz H-M, Schoener A, and Emeis K-C (2000) Long-chain alkenone patterns in the
Rosell-Melé A, Jansen E, and Weinelt M (2002) Appraisal of a molecular approach Baltic Sea – An ocean-freshwater transition. Geochimica et Cosmochimica Acta
to infer variations in surface ocean freshwater inputs into the North Atlantic during 64: 469–477.
the last glacial. Global and Planetary Change 34: 143–152. Seki O, Ishiwatari R, and Matsumoto K (2002) Millennial climate oscillations in NE
Rosell-Melé A, Maslin MA, Maxwell JR, and Schaeffer P (1997) Biomarker evidence for Pacific surface waters over the last 82 kyr: New evidence from alkenones.
“Heinrich” events. Geochimica et Cosmochimica Acta 61: 1671–1678. Geophysical Research Letters 29(23): 2144.
432 Alkenone Paleotemperature Determinations

Sicre M-A, Bard E, Ezat U, and Rostek F (2002) Alkenone distributions in the North Ternois Y, Sicre M-A, Boireau A, Beaufort L, Migquel J-C, and Jeandel C (1998)
Atlantic and Nordic sea surface waters. Geochemistry, Geophysics, Geosystems Hydrocarbons, sterols, and alkenones in sinking particles in the Indian Ocean sector
3: 1013. of the Southern Ocean. Organic Geochemistry 28: 489–501.
Sicre M-A, Hall IR, Mignot J, et al. (2011) Sea surface temperature variability in the Ternois Y, Sicre M-A, Boireau A, Conte MH, and Eglinton G (1997) Evaluation of
subpolar Atlantic over the last two millennia. Paleoceanography 26: PA4218. long-chain alkenones as paleotemperature indicators in the Mediterranean Sea.
Sicre M-A, Jacob J, Ezat U, et al. (2008) Decadal variability of sea surface temperatures Deep Sea Research Part I: Oceanographic Research Papers 44: 271–286.
off North Iceland over the last 2000 years. Earth and Planetary Science Letters Ternois Y, Sicre M-A, Boireau A, Marty J-C, and Miquel J-C (1996) Production pattern
268: 137–142. of alkenones in the Mediterranean Sea. Geophysical Research Letters
Sicre M-A, Ternois Y, Miquel J-C, and Marty J-C (1999) Alkenones in the 23: 3171–3174.
Mediterranean sea: Interannual variability and vertical transfer. Geophysical Thiel V, Jenisch A, Landmann G, Reimer A, and Michaelis W (1997)
Research Letters 26(12): 1735–1738. Unusual distributions of long-chain alkenones and tetrahymanol from the highly
Sicre M-A, Ternois Y, Paterne M, et al. (2000) Biomarker stratigraphic records over the alkaline Lake Van, Turkey. Geochimica et Cosmochimica Acta 61: 2053–2064.
last 150 Kyrs off the NW African coast at 25 N. Organic Geochemistry 31: 577–588. Thierstein HR, Geitzenauer KR, Molfino B, and Shackleton NJ (1977) Global
Sikes CS and Fabry VJ (1994) Photosynthesis, CaCO3 deposition, coccolithophorids synchroneity of late Quaternary coccolith datum levels: Validation by oxygen
and the global carbon cycle. In: Tolbert NE and Preiss J (eds.) Regulation of isotopes. Geology 5: 400–404.
Atmospheric CO2 and O2 by Photosynthetic Carbon Metabolism, pp. 217–233. Thomsen HA, Buck KR, and Chavez FP (1994) Haptophytes as components of marine
New York: Oxford University Press. phytoplankton. In: Green JC and Leadbeater BSC (eds.) The Haptophyte Algae,
Sikes EL, Farrington JW, and Keigwin LD (1991) Use of the alkenone unsaturation ratio pp. 187–208. Oxford: Clarendon Press.
Uk37 to determine past sea surface temperatures: Core-top SST calibrations and Thomsen C, Schulz-Bull DE, Petrick G, and Duinker JC (1998) Seasonal variability
0
methodology considerations. Earth and Planetary Science Letters 104: 36–47. of the long-chain alkenone flux and the effect on the Uk 37 index in the Norwegian
Sikes EL and Keigwin LD (1994) Equatorial Atlantic sea surface temperatures for the last Sea. Organic Geochemistry 28: 311–323.
16 kyr: A comparison of Uk37, d18O, and foraminiferal assemblage temperature Toney JL, Leavitt PR, and Huang Y (2011) Alkenones are common in prairie lakes of
estimates. Paleoceanography 9: 31–45. interior Canada. Organic Geochemistry 42: 707–712.
Sikes EL and Keigwin LD (1996) A reexamination of northeast Atlantic sea surface van der Smissen JH and Rullkotter J (1996) Organofacies variations in sediments
temperature and salinity over the last 16 kyr. Paleoceanography 11: 327–342. from the continental slope and rise of the New Jersey continental margin (Sites 903
Sikes EL, Nodder SD, O’Leary T, and Volkman JK (2005) Alkenone temperature and 905). Proceedings of the Ocean Drilling Program, Scientific Results
records and biomarker flux at the subtropical front on the Chatham Rise, SW 150: 329–344.
0
Pacific Ocean. Deep Sea Research Part I: Oceanographic Research Papers Versteegh GJM, Riegman R, de Leew JW, and Jansen JHF (2001) Uk 37 values for
52: 721–748. Isochrysis galbana as a function of culture temperature, light intensity and nutrient
Sikes EL and Volkman JK (1993) Calibration of alkenone unsaturation ratios (Uk37) for concentrations. Organic Geochemistry 32: 785–794.
paleotemperature estimation in cold polar waters. Geochimica et Cosmochimica Villanueva J and Grimalt JO (1996) Pitfalls in the chromatographic determination of
0
Acta 57: 1883–1889. the alkenone Uk 37 index for paleotemperature estimation. Journal of
Sikes EL, Volkman JK, Robertson LG, and Pichon J-J (1997) Alkenones and alkenes in Chromatography A 723: 285–291.
surface waters and sediments of the Southern Ocean: Implications for Villanueva J, Pelejero C, and Grimalt JO (1997) Clean-up procedures for the unbiased
paleotemperature estimation in polar regions. Geochimica et Cosmochimica Acta estimation of C37 alkenone sea surface temperatures and terrigenous n-alkane
61: 1495–1505. inputs in paleoceanograpy. Journal of Chromatography A 757: 145–151.
Simoneit BRT, Prahl FG, Leif RN, and Mao S-Z (1994) Alkenones in sediments of Visser K, Thunell R, and Stott L (2003) Magnitude and timing of temperature change in
Middle Valley. Proceedings of the Ocean drilling Program, Scientific Results the Indo-Pacific warm pool during deglaciation. Nature 421: 152–155.
139: 479–484. Volkman JK, Barrett SM, Blackburn SI, and Sikes EL (1995) Alkenones in Gephyrocapsa
Sinninghe Damsté JS, Rijpstra WIC, and Reichart G-J (2002) The influence of oxic oceanica: Implications for studies of paleoclimate. Geochimica et Cosmochimica
degradation on the sedimentary biomarker record II. Evidence from Arabian Sea Acta 59: 513–520.
sediments. Geochimica et Cosmochimica Acta 66: 2737–2754. Volkman JK, Eglinton G, Corner EDS, and Forsberg TEV (1980a) Long-chain alkenes
Sonzogni C, Bard E, and Rostek F (1998) Tropical sea surface temperatures during the and alkenones in the marine coccolithophorid Emiliania huxleyi. Phytochemistry
last glacial period: A view based on alkenones in Indian Ocean sediments. 19: 2619–2622.
Quaternary Science Reviews 17: 1185–1201. Volkman JK, Eglinton G, Corner EDS, and Sargent JR (1980b) Novel unsaturated
Sonzogni C, Bard E, Rostek F, Dollfus D, Rosell-Melé A, and Eglinton G (1997) straight-chain C37-C39 methyl and ethyl ketones in marine sediments and a
Temperature and salinity effects on alkenone ratios measured in surface sediments coccolithophore Emiliania huxleyi. In: Douglas AG and Maxwell JR (eds.) Advances
from the Indian Ocean. Quaternary Research 47: 344–355. in Organic Geochemistry 1979. Physics and Chemistry of the Earth, vol. 12,
Sprengel C, Baumann KH, Henderiks J, Henrich R, and Neuer S (2002) Modern pp. 219–227. Oxford: Pergamon Press.
coccolithophore and carbonate sedimentation along a productivity gradient in Volkman JK, Jeffer SW, Nichols PD, Rogers GI, and Garland CD (1989) Fatty acid and
the Canary Islands region: Seasonal export production and surface lipid composition of 10 species of microalgae used in mariculture. Journal of
accumulation rates. Deep Sea Research Part II: Topical Studies in Oceanography Experimental Marine Biology and Ecology 128: 219–240.
49: 3577–3598. Waelbroeck C and Steinke S (2002) Comment on “A high-resolution sea-surface
Sprengel C, Baumann KH, and Neuer S (2000) Seasonal and interannual variation of temperature record from the tropical South China Sea (16,500-3000 yr BP)” by
coccolithophore fluxes and species composition in sediment traps north of Gran Steinke et al. Quaternary Research 57: 432–433.
Canaria (29 N 15 W). Marine Micropaleontology 39: 157–178. Wang L, Sarnthein M, Erlenkeuser H, et al. (1999) East Asian monsoon climate
Steinke S, Kienast M, Pflaumann U, Weinelt M, and Stattegger K (2001) A high- during the Late Pleistocene: High-resolution sediment records from the South
resolution sea-surface temperature record from the tropical South China Sea China Sea. Marine Geology 156: 245–284.
(16,500-3000 yr B.P.). Quaternary Research 55: 352–362. Weaver PPE, Chapman MR, Eglinton G, Zhao M, Rutledge D, and Read G (1999)
Stott L, Poulsen C, Lund S, and Thunell R (2002) Super ENSO and global climate Combined coccolith, foraminiferal, and biomarker reconstruction of
oscillations at millennial time scales. Science 297: 222–226. paleoceanographic conditions over the past 120 kyr in the northern North Atlantic
Stute M, Forster M, Frischkorn H, et al. (1995) Cooling of tropical Brazil (5 C) during (59 N, 23 W). Paleoceanography 14: 336–349.
the last Glacial Maximum. Science 269: 379–383. Winter A, Jordan R, and Roth P (1994) Biogeography of living coccolithophores in
Summerhayes CP, Kroon D, Rosell-Melé A, et al. (1995) Variability in the Benguela ocean waters. In: Winter A and Siesser WG (eds.) Coccolithophores, pp. 161–177.
Current upwelling system over the past 70,000 years. Progress in Oceanography Cambridge, UK: Cambridge University Press.
35: 207–251. Xu L, Reddy CM, Farrington JW, et al. (2001) Identification of a novel alkenone in Black
Sun M-Y and Wakeham SG (1994) Molecular evidence for degradation and Sea sediments. Organic Geochemistry 32: 633–645.
preservation of organic matter in the anoxic Black Sea Basin. Geochimica et Yamamoto M, Ficken K, Baas M, Bosch H-J, and de Leuw JW (1996) Molecular
Cosmochimica Acta 58: 3395–3406. paleontology of the earliest Danian at Geulhemmerberg (the Netherlands). Geologie
Teece MA, Getliff JM, Leftley JW, Parkes RJ, and Maxwell JR (1998) Microbial en Mijnbouw 75: 255–267.
degradation of the marine prymnesiophyte Emiliania huxleyi under oxic and anoxic Yamamoto M, Shiraiwa Y, and Inouye I (2000) Physiological responses of lipids in
conditions as a model for early diagenesis: Long chain alkadienes, alkenones and Emiliania huxleyi and Gephyrocapsa oceanica (Haptophyceae) to growth status and
alkyl alkenoates. Organic Geochemistry 29: 863–880. their implications for alkenone paleothermometry. Organic Geochemistry 31: 799–811.
Alkenone Paleotemperature Determinations 433

Young JR and Westbroek P (1991) Genotypic variation in the coccolithophorid species Northeast Atlantic. Palaeogeography, Palaeoclimatology, Palaeoecology
Emiliania huxleyi. Marine Micropaleontology 18: 5–23. 103: 57–65.
Zabeti N, Bonin P, Volkman JK, Jameson ID, Guasco S, and Rontani J-F (2010) Zink KG, Leythaeuser D, Melkonian M, and Schwark L (2001) Temperature dependency
0
Potential alteration of Uk 37 paleothermometer due to selective degradation of of long-chain alkenone distributions in Recent to fossil limnic sediments and in
alkenones by marine bacteria isolated from the haptophyte Emiliania huxleyi. FEMS lake waters. Geochimica et Cosmochimica Acta 65: 253–265.
Microbiology Ecology 73: 83–94. Ziveri P, Broerse ATC, van Hinte JE, Westbroek P, and Honjo S (2000) The fate of
Zhao M, Beveridge NAS, Shackleton NJ, Sarnthein M, and Eglinton G (1995) Molecular coccoliths at 48 N 21 W, northeastern Atlantic. Deep Sea Research Part II: Topical
stratigraphy of cores off northwest Africa: Sea surface temperature history over the Studies in Oceanography 47: 1853–1875.
last 80 ka. Paleoceanography 10: 661–675. Ziveri P and Thunell R (2000) Coccolithophore export production in Guaymas Basin,
0
Zhao M, Eglinton G, Read G, and Schimmelmann A (2000) An alkenone (Uk 37) Gulf of California: Response to climate forcing. Deep Sea Research Part II: Topical
quasi-annual sea surface temperature record (A.D. 1440 to 1940) using varved Studies in Oceanography 47(9–11): 2073–2100.
sediments from the Santa Barbara Basin. Organic Geochemistry 31: 903–917. Ziveri P, Thunell R, and Rio D (1995) Export production of coccolithophore in an
Zhao M, Rosell A, and Eglinton G (1993) Comparison of two Uk37 sea surface upwelling region: Results from San Pedro Basin, Southern California Bight. Marine
temperature records for the last climatic cycle at ODP Site 658 from the sub-tropical Micropaleontology 24: 335–358.

You might also like