You are on page 1of 17

GDGT THERMOMETRY:

LIPID TOOLS FOR RECONSTRUCTING PALEOTEMPERATURES

JESSICA E. TIERNEY

Woods Hole Oceanographic Institution, Woods Hole, MA USA 02543


tierney@whoi.edu

ABSTRACT.—Microbial communities adjust the chemical structure of their cell membranes in response to
environmental temperature. This enables the development of lipid-based paleothermometers such as the
glycerol dialkyl glycerol tetraether (GDGT) proxies described here. Surface-sediment calibrations estab-
lish a strong empirical relationship between the relative distribution of GDGTs and temperature. GDGT
proxies can be used in marine, lacustrine, and paleosol sequences as long as the organic material is not
thermally mature. Thus far, GDGT proxies have been applied to sediments dating back to the middle Ju-
rassic. Many of the key uncertainties of these proxies are related to our emerging understanding of ar-
chaeal (and for the branched GDGTs, bacterial) ecology.

INTRODUCTION paleo sea-surface temperature (SST) proxy (UK


37) can be applied widely in marine environments

THE LAST twenty-five years have witnessed the that are mid-Paleogene in age or younger, and has
growth and expansion of a new technique for es- been a mainstay of paleoclimate reconstruction
timating paleotemperatures based on the use of since its development in the late 1980s (Herbert,
biomarkers: organism-specific, fossil lipids (fats) 2003). There are, however, some limitations to the
found in sedimentary organic matter. Biomarkers alkenone proxy. Extant haptophyte species that
enhance our ability to reconstruct paleoenviron- the modern calibrations of the alkenone proxy are
ments in geologic history, in part because they are based on evolved relatively recently (e.g., 0.28
subject to a different set of diagenetic alterations Ma for Emiliania huxleyi, Thierstein et al., 1977),
than hard parts of fossils. They may be well pre- so alkenones found in older sedimentary se-
served in sedimentary environments where tradi- quences can only be presumed to be produced by
tional skeletal temperature proxies (e.g., calcare- ancient relatives with a similar temperature sensi-
ous foraminifera) typically are not, thereby en- tivity. Also, the alkenone calibrations to SST are
hancing the range of depositional conditions from limited to ~28°C, such that the proxy cannot be
which paleotemperature information can be ex- used in some areas of the tropical oceans, nor in
tracted. ancient hothouse climates.
Biomarker-based paleothermometers are built In the last decade, a new biomarker paleo-
upon the biochemical principle that microorgan- temperature approach based on the relative abun-
isms adjust the rigidity of their cell membrane dances of glycerol dialkyl glycerol tetraethers
structures by altering the number of double bonds, (GDGTs) has been developed that complements
rings or branches in response to environmental and extends the capabilities of the alkenone proxy.
temperature (e.g., Ray et al., 1971; Kita et al., GDGTs are relatively large (up to 86 carbon at-
1973; Nozawa et al., 1974). The first biomarker- oms) membrane lipids produced by Archaea and
based paleotemperature proxy was based on the some Bacteria. They are common, typically abun-
relative distribution of unique long-chain (C37) tri- dant, easy-to-analyze lipids found in a variety of
and di-unsaturated ketones (alkenones) produced environments (lakes, soils, oceans). Such attrib-
by marine haptophyte algae, such as the ubiqui- utes make GDGTs a promising universal tool for
tous coccolithophorid Emiliania huxleyi (Volkman estimating temperatures in the geologic past, es-
et al., 1980; Marlowe et al., 1984). The alkenone pecially in deep geologic time.

In Reconstructing Earth’s Deep-Time Climate—The State of the Art in 2012, Paleontological Society Short Course,
November 3, 2012. The Paleontological Society Papers, Volume 18, Linda C. Ivany and Brian T. Huber (eds.),
pp. 115–131. Copyright © 2012 The Paleontological Society.
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

678!9:4;./04<=15!)*)+: 6-8!>.1023/=!)*)+:

)*)+,$
!"#!"#$% brGDGT-III !"#!"$I$

)*)+,"
!"#!"#$$ +H;/!999 brGDGT-IIIb !"#!"$J'

)*)+,% brGDGT-IIIc !"#!"$J(


!"#!"%&'

)*)+,# brGDGT-II !"#!"$#(


!"#!"%&(

-./01.231/45 +H;/!99 brGDGT-IIb !"#!"$#J

!"#!"%&%
brGDGT-IIc !"#!"$#%
-./01.231/45
./B<4<:4?/.
!"#!"%&%G
brGDGT-I !"#!"$%%
6>8!-4??40!@451.!A/1=B.4C;:!,!9:4;./04<=15!)*)+:
+H;/!9 brGDGT-Ib
!"#!"$%$

brGDGT-Ic !"#!"$"'

@34:;31D/ E4043/F4:/ *<3/F4:/

FIGURE 1.—A) Core structures of the isoprenoidal glycerol dialkyl glycerol tetraethers (isoGDGTs), with mass-to-
charge ratios (m/z) on the right associated with analysis via HPLC/positive ion APCI. B) Common polar head-
groups of intact isoGDGTs, which would replace one or both of the terminal hydroxyl groups indicated in (A). C)
Core structures of branched glycerol dialkyl glycerol tetraethers (brGDGTs).

WHAT ARE GDGTs? GDGT structures remain in ancient sediments.


IsoGDGTs are characteristic lipids of Ar-
There are two main types of GDGTs: the iso- chaea, and are produced by a number of methano-
prenoidal GDGTs (isoGDGTs) and the non- genic, hyperthermophilic, and mesophilic species,
isoprenoidal, branched GDGTs (brGDGTs). The although not all species produce each kind of
isoGDGTs consist of two head-to-head C40 iso- isoGDGT shown in Figure 1. The isoGDGT pa-
prenoid (i.e., made of isoprene units) chains with leothermometer rests on the assumption that the
a varying number of cyclopentane (five-sided) dominant producers of isoGDGTs are members of
and cyclohexane (six-sided) rings, connected by the recently renamed phylum Thaumarchaeota
ether bonds to two terminal glycerol groups (Fig- (formerly a division of the Crenarchaeota;
ure 1). The brGDGTs are structurally similar, but Brochier-Armanet et al., 2008; Spang et al.,
have branched C30 alkyl chains containing 4–6 2010). In many environments, this is a reasonable
methyl groups instead of the C40 isoprenoidal assumption because other Archaea that produce
chains (Figure 1). Additional polar headgroups, GDGTs (e.g., thermophile Archaea) are restricted
including hexose and phosphate moieties, are at- to distinct environmental niches. In contrast,
tached to the core GDGT structure when the Thaumarchaeota are a ubiquitous mesophilic
membrane lipid is intact (Sturt et al., 2004; group in the global oceans, possibly representing
Schouten et al., 2008a; Pitcher et al., 2010; and up to 20% of the ocean’s picoplankton (Karner et
see Fig. 1). Similar polar headgroups accompany al., 2001). Known species of Thaumarchaeota are
the brGDGTs (Liu et al., 2010; Peterse et al., nitrifiers: they convert ammonium to nitrate and
2011). These polar headgroups are typically fix bicarbonate autotrophically to form their bio-
quickly lost during diagenesis; only the “core” mass (Könneke et al., 2005). Nitrifying Thaumar-

116
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

chaeota and their GDGTs are also found in soils *+,-.#%/0/'#-12345+24.567-'89:"


and lakes (Ochsenreiter et al., 2003; Leininger et 48"&98:E9"%F '<=>?-@-ABCAD
al., 2006; Llirós et al., 2010; Schouten et al.,  
2012). Concentrations of isoGDGTs in soils are
low relative to marine levels (Weijers et al.,
2006b), but concentrations in lakes are often quite
high (Tierney et al., 2010b) and have enabled the
application of the isoGDGT paleothermometer to
certain lake environments (see below).
Thaumarchaeota are the only Archaea known A
to make the distinctive GDGT crenarchaeol, I
H 48"&98:E9"%F-
which has a cyclohexyl ring in addition to the cy- 8"G(%(#%)"8
clopentyl rings (Figure 1), and is a diagnostic J ;8/0/'#
biomarker for these species in both moderate- and
high-temperature environments (Sinninghe Dam- *;,-;8/0/'-12345+24.567-'89:"
sté et al., 2002b; Pearson et al., 2004; de la Torre
et al., 2008). Thaumarchaeota also produce 6;'-@-ABDI ..

!"#$%&#"
smaller amounts of a crenarchaeol regioisomer 4;'-@-AB?L ...
(Figure 1; oft-abbreviated as cren’) that plays a   
role in the temperature predictability of the
isoGDGT paleothermometer: relatively more of .
the regioisomer is observed at higher tempera-
tures.
BrGDGTs share many structural characteris-
tics with the isoGDGTs, including the ether bonds
to terminal glycerol groups, a feature in mem-
brane lipids typically seen only in Archaea (with .K
some exceptions, e.g., see Weijers et al., 2006a, .:
and references therein). Yet both the chain archi- '()"
tecture (branched vs. isoprenoidal) and the stereo-
chemistry of the glycerol groups on the brGDGTs FIGURE 2.—A) HPLC trace of isoGDGTs from a
tropical marine sediment. TEX86 value is calibrated to
are diagnostic of Bacteria (Weijers et al., 2006a).
SST using the Kim, et al. (2010) TEXH86 calibration.
BrGDGTs are commonly found in soils, peats, (B) HPLC trace of brGDGTs from a freshwater lake.
lakes, and marginal/deltaic environments, and are MBT and CBT values are calibrated to MAAT using
not present in pelagic marine environments the Tierney et al. (2010b) lakes calibration. All num-
(Hopmans et al., 2004), suggesting that they are bers correspond to the structures in Figure 1.
associated with terrestrial organisms. However, in
spite of their ubiquity, efforts to identify the or- al., 1997), but were difficult to analyze because
ganism(s) responsible for these compounds have intact GDGTs are too large for gas chromatogra-
not been highly successful (e.g., Weijers et al., phy. It was not until the relatively recent devel-
2009). Recently, Sinninghe Damsté et al. (2011) opment of a high-performance liquid-
demonstrated that a strain of Acidobacteria pro- chromatography mass-spectrometry method
duces one type of brGDGT (brGDGT-I in Figure (Hopmans et al., 2000) that GDGTs became a fo-
1), confirming suspicions that this phylum of Bac- cus of both organic geochemical and paleoclima-
teria, which is diverse, commonly found in peats tological research. Figure 2 illustrates some typi-
in soils, but not well characterized, may be a cal chromatograms of both iso- and br-GDGTs
source of brGDGTs (Weijers et al., 2006a, 2009). using high-performance liquid chromatography-
However, the other eight brGDGTs utilized in atmospheric pressure chemical ionization-mass
brGDGT-based temperature proxies are, at pre- spectrometry (HPLC-APCI-MS), which separates
sent, “orphan” compounds. compounds by mass and polarity. The run time for
GDGTs—or chemical remnants of them— a single LC/MS analysis is 60–70 minutes, and
have been observed in both modern and ancient preparative procedures are relatively minimal (ex-
sedimentary deposits and oils for some time (e.g., traction followed by column chromatography for
Chappe et al., 1980; Brassell et al., 1981; Hoefs et

117
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

purification), making GDGT analysis sufficiently tion given above, efforts have been made to ex-
rapid for use in paleoclimate studies. Details of pand the collection of modern core-top sediments
the GDGT LC/MS methodology may be found in available to constrain the TEX86 proxy and better
Schouten et al. (2007b). understand its predictive ability (Kim et al., 2008;
Liu et al., 2009; Kim et al., 2010). The current
HOW DO GDGTS RECORD calibration dataset includes 350+ core tops from
PALEOTEMPERATURES? around the global ocean, though the distribution
of those core tops is not uniform (Figure 3). The
isoGDGTs: the TEX86 paleothermometer relationship between TEX86 and SST remains
In response to temperature variations, the mi- largely linear (Figure 4), although non-linear rela-
crobial producers of GDGTs produce compounds tionships to SST have been proposed (Liu et al.,
with different numbers of rings in their isoprenoid 2009; Kim et al., 2010) based on observations that
chains: a GDGT with more rings has a higher TEX86 is more sensitive to SST changes at the
melting point and is more stable at warm tempera- high end of the temperature range and less sensi-
tures (Gliozzi et al., 1983). Although mesophilic tive at the lower end of the calibration range (e.g.,
Thaumarchaeota have not yet been cultured in Schouten et al., 2003). Both linear and non-linear
different temperature environments, thermophile regressions through the modern core-top dataset
relatives of Thaumarchaeota that are more readily perform comparably (Figure 4).
manipulated in culture produce more cyclic The core-top studies also indicate that there
isoGDGTs at higher temperatures (de Rosa et al., are certain areas in the marine realm where TEX86
1980; Uda et al., 2001). Similarly, mesocosm is unreliable, or necessitates a different calibra-
studies (experiments in which natural seawater is tion. These areas include the polar oceans, where
manipulated in the laboratory) show that more of the relationship between TEX86 and SST is poor
the cyclic isoGDGTs are present in seawater at (Kim et al., 2008, 2010). This could reflect a re-
higher temperatures (Wuchter et al., 2004; duced sensitivity of TEX86 in cold waters, or con-
Schouten et al., 2007a). Schouten et al. (2002) versely, inaccurate mean annual SST estimates at
developed an index to mathematically represent high latitudes. Kim et al. (2010) proposed an al-
this degree of cyclization called TEX86 (the Tet- ternate functional form of TEX86, called “TEXL86”
raEther indeX of 86 carbons; numbers correspond (L for low-temperature) that has better predict-
to compounds labeled in Figure 1; cren′ stands for ability at low temperatures, but is less precise
the crenarchaeol regioisomer): given the entire global SST range (Table 1). Inter-
estingly, TEXL86 excludes the regioisomer:

Schouten et al. (2002) empirically calibrated


TEX86 to SST using a set of 43 modern surface IsoGDGT distributions are also quite different in
sediments, yielding a strong linear relationship: the semi-enclosed Red Sea, necessitating a differ-
ent calibration (Trommer et al., 2009, Table 1)
TEX86 = 0.015 × SST + 0.27 (r2 = 0.92, n = 43) and perhaps indicating the contribution of unique
archaeal species (Eder et al., 2002; Ionescu et al.,
The paucity of pure Thaumarchaeotal cultures, 2009) to the sedimentary isoGDGT pool.
along with the surprising observations that Thau- TEX86 may be used in some (typically large)
marchaeota seem to produce less of the crenar- lakes to infer lake-surface temperature (LST)
chaeol isomer in mesocosms than in the open (Powers et al., 2004, 2005; Tierney et al., 2008;
ocean (Wuchter et al., 2004; Schouten et al., Blaga et al., 2009; Powers et al., 2010; Tierney et
2007a) and that the isomer may have a different al., 2010a); core tops used to constrain the lake
source (Shah et al., 2008), have thwarted attempts calibrations are shown in Figure 3. Unfortunately,
to calibrate TEX86 to SST experimentally. This TEX86 does not appear to be universally applica-
leaves the TEX86 proxy reliant on empirical core- ble in lakes. Reasons for this are not well under-
top calibration, a practice that is commonly stood, but it may be that smaller lakes do not have
adopted for other paleo-SST proxies (e.g., alke- the chemistry or limnology that supports large
none UK′37 and foraminiferal Mg/Ca). Since populations of lacustrine Thaumarchaeota. In an-
Schouten et al. (2002) developed the linear equa- oxic lakes or in anoxic sediment columns, Ar-

118
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

!"#$%&'()$*+,-.+/$*0123,0.2+4$/+24.5
o
80 N

40oN

0o

40oS
70,24-:$!;<=>
?0@-5:$!;<<

80oS
180oW 120oW 60oW 0o 60oE 120oE 180oW

!6#$76%896%$*+,-.+/$*0123,0.2+4$/+24.5
o
80 N

40oN

0o

40oS
A+215:$!;>B
?0@-5:$!;CDC

80oS
180oW 120oW 60oW 0o 60oE 120oE 180oW

FIGURE 3.—A) Locations of sediment core-top calibration points for TEX86, including data from Kim et al. (2010),
Ho et al. (2011), Trommer et al. (2009), Chazen (2011), Leider et al. (2010), Seki et al. (2009), Castañeda et al.
(2010), Shintani et al. (2010), Powers et al. (2010), Blaga et al. (2009) and Tierney et al. (2010a) and excluding
duplicates. (B) Locations of sediment core-top calibration points for MBT/CBT, including data from Tierney et al.
(2010b), Blaga et al. (2010), Zink et al. (2010), Sun et al. (2011), and Pearson et al. (2011).

chaea performing methanogenesis or anaerobic rithimic transform of TEX86) for marine SST, and
oxidation of methane (AOM) may produce the Powers et al. (2010) calibration for lake LST.
isoGDGTs (see below), compromising the Thau-
marchaeotal TEX86 signature (Blaga et al., 2009; brGDGTs: the MBT/CBT paleothermometer
Powers et al., 2010). In small lakes receiving Because the producers of most of the
large amounts of soil organic matter, some of the brGDGTs have not been identified, and no species
isoGDGTs may derive from soil Thaumarchaeota, that produce any of these lipids other than
again complicating the sedimentary signal (Blaga brGDGT-I have been identified in culture, the
et al., 2009). As with the marine TEX86 calibra- physiological relationship between brGDGT cyc-
tion, the relationship between TEX86 and LST is lization and/or methylation to temperature is not
generally linear, although a non-linear form is understood. We might expect that such a relation-
also possible. Tierney et al. (2010a) proposed a ship exists, but to date, it relies solely on empiri-
“warm lakes” calibration with a shallower linear cal evidence in the form of core-top/topsoil stud-
slope (see Table 1) to account for the apparent ies. Weijers et al. (2007c) studied a global collec-
increased sensitivity of lacustrine TEX86 at higher tion of topsoils and identified two major environ-
LSTs, indicating that a nonlinear model could be mental parameters that had a statistical relation-
appropriate (Figure 4). ship to brGDGT distributions: soil temperature
Presently, there are a number of calibrations (approximated by mean annual air temperature;
for TEX86, reflecting ongoing attempts to refine MAAT) and soil pH. The degree of cyclization of
the relationship between TEX86 and temperature the brGDGTs seemed to be related only to soil pH
(Table 1). The most common calibrations used to in a non-linear fashion, implying a power law re-
date are the Kim et al. (2010) “TEXH86” calibra- lationship. Thus, Weijers et al. (2007c) devised an
tion (where TEXH86 is simply the base-10 loga- index representing the cyclization of the

119
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

TABLE 1.—Sediment core top TEX86 calibrations. *= calibration applies just to the Red Sea, where TEX86 distribu-
tions are different from the open ocean, ‡=TEXL86 is an alternate functional form of TEX86 arguably better for pre-
diction at low SSTs, see main text and Kim et al. (2010) for more details, †=lake calibrations.

Range Equation n r2 RMSE Reference

0–30 T E X86 = 0.015 × T + 0.27 43 0.92 N/A Schouten et al. (2002)

22–30 T E X86 = 0.027 × T + 0.016 24 0.78 N/A Schouten et al. (2003)

5–30 T = −10.78 + 56.2 × T E X86 223 0.94 1.7 Kim et al. (2008)

25–28* T E X86 = 0.035 × T − 0.09 11 0.90 N/A Trommer et al. (2009)

-3–30 T = 50.475 − 16.332 × (1/T E X86) 287 0.82 3.7 Liu et al. (2009)

-3–30‡ T = 49.9 + 67.5 × T E X L86 396 0.86 4.0 Kim et al. (2010)

5–30 T = 38.6 + 68.4 × log(T E X86) 255 0.86 2.5 Kim et al. (2010)

4–30† T = −14.0 + 55.2 × T E X86 12 0.86 3.6 Powers et al. (2010)

10–30† T = 3.50 + 38.9 × T E X86 13 0.92 2.1 Tierney et al. (2010a)

brGDGTs called the CBT (Cyclization of and CBT (as a surrogate for the pH influence) to
Branched Tetraethers) index: predict MBT, which could be inverted to solve for
MAAT:

MBT = 0.122 + 0.187 × CBT + 0.020 × MAAT


that could be linearly related to pH:
This relation has a slightly higher r2 of 0.77.
CBT = 3.33 − 0.38 × pH However, when the regression is recalibrated with
MAAT as the dependent variable (necessary for
The degree of methylation of soil brGDGTs was prediction of MAAT), the r2 statistic drops to 0.62
found to relate to both temperature and pH. (Figure 5).
Weijers et al. (2007c) devised the MBT (Methyla- The fact that brGDGTs are abundant in a wide
tion of Branched Tetraethers) index to represent variety of lake basins (Tierney and Russell, 2009;
the degree of methylation (numbers refer to struc- Sinninghe Damsté et al., 2009; Blaga et al., 2010;
tures in Figure 1; denominator indicates the sum Bechtel et al., 2010; Tierney et al., 2010b; Tyler et
of all nine brGDGT compounds): al., 2010; Zink et al., 2010; Sun et al., 2011) of-
fers the tantalizing possibility that these com-
pounds could be used to develop a universal la-
custrine temperature proxy. Early applications of
the Weijers et al. (2007c) equation to lacustrine
MBT is reasonably correlated with MAAT alone brGDGTs (driven by the assumption that all
(r2 =0.62), but is better constrained when both pH brGDGTs are soil-derived) revealed that the re-
and MAAT are used as predictors (r2 = 0.82), sulting temperatures were 10°C or greater differ-
therefore, Weijers et al. (2007c) used both MAAT ent from observed MAAT (Tierney and Russell,

120
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

#! #!

)% )%

;3<2415/6372*82092/385/2
123415/6372*82092/385/2

)! )!

(% (%

(! (!

!*+*#%,-*")*+*!".$-*/012*+*)". !*+*##-*")*+*!".!-*/012*+*#"'
% %
!*+*#%,-*" *+*!".%-*/012*+*)".
)
!*+*##-*")*+*!"'.-*/012*+*#",
!"# !"$ !"% !"& !"' !") !"# !"$ !"% !"& !"' !".
82:.& 82:.&
FIGURE 4.—TEX86 regressions to SST and LST respectively, with best-fit linear and non-linear models. The marine
TEX86 data shown exclude data from the Red Sea (Trommer et al., 2009) and the polar oceans (SST<2, Kim et al.,
2010) where the relationship between TEX86 and SST is different and/or poor. RMSE = root mean square error.

2009; Zink et al., 2010; Tyler et al., 2010); thus, EXAMPLE APPLICATIONS OF GDGT
the soil calibration was not applicable. This may PROXIES
suggest that lacustrine brGDGTs are produced in
situ with a different sensitivity to lake tempera- GDGTs have been used to reconstruct both air
ture—a hypothesis also supported by the presence and water temperatures on a variety of geologic
of intact polar brGDGT in lakes (Tierney et al., timescales, but have been especially useful for
2012). However, recent attempts to develop lake- estimating temperatures in the Mesozoic and early
specific brGDGT distributions are promising and Cenozoic, where there are few applicable methods
demonstrate that lake sediments perform similar for estimating paleotemperature. In one of the
to or better than soils in predicting MAAT (Figure earliest applications of the TEX86 proxy, Schouten
5). et al. (2003) reconstructed low-latitude SSTs dur-
For both soils and lakes, the MBT/CBT tem- ing select intervals in the middle Cretaceous (Al-
perature proxy is considerably less accurate in bian–Cenomanian–Turonian), a hothouse period
predicting temperature than TEX86, with a root in Earth history where temperatures are relatively
mean square error near 5°C. Alternate functional poorly constrained. They estimated low-latitude
forms representing the distribution of brGDGTs SSTs of 32–36°C, in general agreement with
may improve predictive performance and are be- δ18O-based estimates from well-preserved Ceno-
ginning to be explored. Tierney et al. (2010b) manian foraminifera (Schouten et al., 2003).
proposed a calibration based on the fractional GDGTs also have been used to provide both
abundance (relative to the sum of all nine low- and high-latitude constraints on temperature
GDGTs) of the three major non-cyclic brGDGTs excursions during the Paleocene–Eocene Thermal
that provides more precise prediction of MAAT in Maximum (PETM), a brief interval when massive
tropical East African lakes (Figure 5, Table 2). amounts of greenhouse gases were injected into
Similarly, Pearson et al. (2011) proposed a cali- the atmosphere ~55 million years ago, causing
bration based on the fractional abundances of rapid and extreme warming. Zachos et al. (2006)
brGDGT-III, brGDGT-II, and brGDGT-Ib for a applied TEX86 to a section from Wilson Lake,
more globally distributed set of lakes (Table 2). A New Jersey (paleolatitude of 36°N), and found
list of brGDGT-based calibrations is provided in that SSTs were ~25–30°C prior to and after the
Table 2. PETM, but reached a maximum of ~35°C during
the event (Figure 6). Sluijs et al. (2006) and
Weijers et al. (2007b) applied TEX86 and the

121
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

=8>&.?@A&-123412B2 =1>&CDEF&-123412B2 =4>&8GH@IDJ&CDEF&DAK*&LHM6M2B2


%! %! %!
!&'&(!)&&"$&'&!*+$)&,-./&'&0*( !&'&#"<)&&"$&'&!*+%)&,-./&'&"*# !&'&0#)&&"$&'&!*(0)&,-./&'&#*(
$" $" $"
:1./,;/6&2/-5/,829,/

:1./,;/6&2/-5/,829,/

:1./,;/6&2/-5/,829,/
$! $! $!

#" #" #"

#! #! #!

" " "

! ! !

! " #! #" $! $" %! ! " #! #" $! $" %! ! " #! #" $! $" %!


-123412&5,/6742/6&2/-5/,829,/ -123412&5,/6742/6&2/-5/,829,/ -123412&5,/6742/6&2/-5/,829,/

FIGURE 5.—brGDGT regressions to MAAT for (A) soils (data from Weijers et al., 2007c), (B) lakes (data from
Tierney et al., 2010b; Blaga et al., 2010; Zink et al., 2010; Sun et al., 2011), and (C) East African lakes, using an
alternative functional form (Tierney et al., 2010b, see Table 2 for the equation).

MBT/CBT proxy, respectively, to the PETM sec- anthropogenic climate change has had a strong
tion of a drill core taken from the Lomonosov impact on the thermal structure of Lake Tangany-
Ridge (paleolatitude of 85°N), and reconstructed ika.
surprisingly warm temperatures for the high Arc- BrGDGTs have not been used as widely as
tic (Figure 6). Taken at face value, these PETM TEX86 for paleoclimate reconstruction because
GDGT data suggest an extremely warm world their dependence on temperature was only de-
with a surprisingly minimal latitudinal gradient in scribed in 2007 and, to date, is not well con-
temperature (~10°C) compared to present-day strained. In addition to the PETM study, Weijers
gradients (20°C). However, it should be noted that et al. (2007a) used MBT/CBT to reconstruct
because the TEX86 values at Wilson Lake exceed MAAT in the Congo Basin, using deltaic sedi-
those observed in the modern ocean, the inferred ments from the Congo Fan. Their data indicate a
temperatures rely on an extrapolation of the cali- LGM–present change of ~3.5–4°C, a reasonable
bration regression lines; thus, the absolute values finding for the deep tropics, and in agreement
are highly uncertain. Nevertheless, such data pro- with independent proxy data from Africa. Fawcett
vide a hypothesis of how Earth systems responded et al. (2011) used MBT/CBT to reconstruct tem-
to what were likely very high levels of atmos- perature within a Pleistocene lacustrine deposit
pheric carbon dioxide during the PETM. from the Valles Caldera, and estimated glacial/
GDGT proxies also have utility on more re- interglacial thermal variability on the order of
cent timescales, either as a complement to other ~8°C, which is in general agreement with pollen-
paleotemperature proxies or (as is sometimes the based estimates. Neither brGDGTs nor isoGDGTs
case with deep-time studies) because it is the only have been used, as of yet, to reconstruct tempera-
proxy method available in a given region or depo- tures in ancient lake basins.
sitional setting. For example, Tierney et al.
(2010a) used TEX86 in Lake Tanganyika to recon- LIMITATIONS ON THE USE OF GDGT
struct temperatures for the past 1500 years in East PALEOTHERMOMETERS
Africa, a region where there are few options for
recent temperature reconstruction (there are no Preservation
developed tree-ring chronologies, and stable iso- GDGTs are relatively refractory compounds,
topes generally reflect hydroclimatic variability). but they do degrade over time, with isoGDGTs
TEX86 shows evidence for warm conditions dur- apparently degrading faster than brGDGTs (Hu-
ing the Medieval Warm Period, followed by cool/ guet et al., 2008). However, diagenetic degrada-
variable temperatures during the Little Ice Age tion does not affect the TEX86 index within ana-
(Figure 7). In the last 150 years, TEX86 indicates a lytical error (Sinninghe Damsté et al., 2002a;
warming of ~2°C, an event unprecedented in the Schouten et al., 2004; Kim et al., 2009; Huguet et
last 1500 years (Figure 7). These data suggest that al., 2009). When sedimentary sequences undergo

122
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

TABLE 2.—Topsoil (Weijers et al., 2007c) or lacustrine (all others) core top brGDGT calibrations. *=pH calibration
from CBT, †=uses African lakes data only, ‡=uses New Zealand lakes only, §=This calibration predicts mean sum-
mer temperature (MST) not MAAT.

Range Equation n r2 RMSE Reference


-5–30 MBT = 0.122 + 0.187 × CBT + 0.020 × MAAT 90 0.77 N/A Weijers et al. (2007c)

3–8* CBT = 3.33 − 0.38 × pH 90 0.70 N/A Weijers et al. (2007c)

0–26† MAAT = 11.84 + 32.54 × MBT − 9.32 × CBT 39 0.89 3.0 Tierney et al. (2010b)

0–26† MAAT = 50.47 − 74.18 × fbrGDGT III − 39 0.94 2.2 Tierney et al. (2010b)
31.60 × f brGDGT II − 34.69 × f brGDGT I
6–15‡ MAAT = 55.01 × MBT − 6.055 10 0.74 N/P Zink et al. (2010)

-5–30 MAAT = 3.949 − 5.593 × CBT + 38.213 × MBT 100 0.73 4.3 Sun et al. (2011)

5–30§ MST = 20.9 − 20.5 × f brGDGT III − 12 × 85 0.88 2.0 Pearson et al. (2011)
fbrGDGT II + 98.1 × f brGDGT Ib

catagenesis, GDGTs break down more rapidly and (2010) observed that anoxic lakes yield a “colder”
do so preferentially: hydrous pyrolysis experi- MBT/CBT signature than oxic lakes, indicating
ments suggest that they degrade between 240– more methylated brGDGTs in the sediments.
300°C, and that TEX86 values decline, indicating These observations may suggest preferential dia-
preferential destruction of the more cyclic com- genetic loss of Type III brGDGTs upon exposure
pounds (Schouten et al., 2004). This experimental to oxygen, but alternatively, could reflect a sensi-
evidence suggests that TEX86 is generally not ap- tivity of the microbial producers to anoxia
plicable in more mature sediments (i.e., those (Tierney et al., 2012). Further research is needed
with high amounts of 22S hopanes; 17α,21β (H)- to assess the relative diagenetic stability of the
hopane 22S/(22S+22R) ratios > 0.1). Thus, while nine brGDGTs.
isoGDGT likely have an ancient biosynthetic ori-
gin (Thaumarchaeota have been extant for billions Ecological constraints
of years, Spang et al., 2010, although this may or As mentioned previously, the TEX86 proxy is
may not apply to the lipids), and theoretically predicated on members of the phylum Thaumar-
TEX86 could be used throughout much of geo- chaeota acting as the primary producers of
logic history, the application of the proxy is lim- isoGDGTs. This is a reasonable assumption in the
ited to rock and sediment sequences that have global ocean where Thaumarchaeota dominate,
been minimally thermally altered. although the contribution of isoGDGTs from
Diagenetic/catagenic effects on the brGDGT mesophilic Euryarchaeota, also abundant in ma-
distributions (MBT/CBT) are not as well con- rine environments, remains a matter of some un-
strained due to the novelty of the proxy. In a study certainty and debate (Turich et al., 2007; Schouten
of brGDGTs in a suburban lake that became sea- et al., 2008b). Even if the source of isoGDGTs in
sonally anoxic and eutrophic as a result of human the modern ocean can be reasonably constrained
disturbance, Tierney et al. (2012) found that there to Thaumarchaeota, it is unlikely that isoGDGTs
were more “Type III” (see Figure 1) brGDGTs in are predominantly produced by a single species
lake sediments that were recently deposited in the (unlike alkenones that select marine species pro-
anoxic environment, and less in deeper sediments duce). The lack of regioisomer in cultures of a
that were deposited in the pre-anthropogenic oxy- dominant marine archaeon, Nitrosopumilus mari-
genated environment. Similarly, Bechtel et al. timus (Schouten et al., 2008a), and the observa-

123
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

Paleocene-Eocene such cases, TEX86 cannot be used reliably. Meth-


Thermal Maximum anogenic Archaea and Archaea performing an-
aerobic oxidation of methane (AOM) produce
30 TEX86 isoGDGTs 0–3 (especially GDGT-2, Blumenberg
et al., 2004; Zhang et al., 2011), and are common
in organic-rich, anoxic environments, such as
25 stratified freshwater lakes or marine environments
(e.g., the Black Sea, Wakeham et al., 2003). Diag-
nostic signs of GDGTs originating from
20 methanogenic/AOM archaeal species include high
concentrations of isoGDGTs 0–3 relative to
crenarchaeol and its regioisomer, as well as char-
15 IODP 302 acteristically depleted stable carbon isotope signa-
Lomonosov Ridge tures for biphytanes originating from GDGTs 0–3
MBT/CBT Paleolatitude: 85˚N (Wakeham et al., 2003).
10 The nitrifying metabolism of Thaumarchaeota
380 382 384 386 388 390 also introduces some ambiguity regarding the in-
Core Depth (m) terpretation of TEX86-based SSTs. Because they
40 do not require sunlight and may be outcompeted
TEX86 by faster-growing photosynthetic algae and bacte-
ria, marine Thaumarchaeota tend to reside in the
mesopelagic zone (Massana et al., 1997; Karner et
35 al., 2001; Herndl et al., 2005; Wuchter et al.,
2005; Turich et al., 2007). It follows that Thau-
marchaeota are more likely to be most active and
30 to synthesize most of their biomass in the subsur-
face. Several lines of evidence suggest that this is
the case, including the transcription of genes in-
Wilson Lake, NJ volved in ammonia oxidation (Church et al.,
25 2010) and the natural radiocarbon (14C) content of
Paleolatitude: 36˚N isoGDGTs in both the water column and sedi-
95 100 105 110 ments (Pearson et al., 2001; Ingalls et al., 2006;
Shah et al., 2008), which is more similar to ther-
Core Depth (m) mocline dissolved inorganic carbon (DIC; 14C-
depleted) than surface water DIC (14C-enriched,
FIGURE 6.—TEX86 and MBT/CBT-inferred tempera- due to the presence of “bomb” 14C).
tures during the Paleocene-Eocene Thermal Maximum However, despite these physiological obser-
(PETM) from the high arctic (Lomonosov Ridge, at vations, TEX86 clearly has a robust correlation
top, Sluijs et al., 2006;Weijers et al., 2007b) and the with SST. Several explanations have been pro-
mid-low latitudes (Wilson Lake, Zachos et al., 2006). posed to account for this apparent discrepancy.
TEX86 data are re-calibrated with the Kim et al. (2010) One theory argues that GDGTs from the surface
TEXH86 calibration. Error bars represent the calibra-
ocean are more effectively grazed and exported to
tion root mean square error.
the deep ocean via fecal pellets, thus sedimentary
TEX86 inordinately represents SST (Wakeham et
tion that the genetic sequences of active Archaea al., 2003; Wuchter et al., 2006). This argument is
change as seawater temperatures are manipulated inconsistent with the existing (quite limited)
in mesocosm experiments (Schouten et al., sedimentary 14C data (Pearson et al., 2001), al-
2007b) suggest that the empirical relationship be- though grazing almost certainly plays an impor-
tween TEX86 and temperature represents a com- tant role in the export of GDGTs because archaeal
munity effect rather than a single-species adapta- cells are small and do not sink quickly on their
tion. own (Wakeham et al., 2003). An alternative ex-
In some environments, non-Thaumarchaeotal planation is correlation: if TEX86 actually is re-
archaeal species may contribute a sizable, if not cording subsurface ocean temperatures, this may
dominant amount of sedimentary isoGDGTs. In be difficult to statistically distinguish from SST;

124
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

are reflected in the GDGT export to, and preser-


26
vation in, the sedimentary column. For example,
25.5
!-#"*4( Wuchter et al. (2006) found that the seasonal cy-
LAKE SURFACE TEMPERATURE

)&*+$43 cle in TEX86 observed in the Arabian Sea ho-


25
!"#$"%&'( mogenizes in a deep sediment trap, and Rueda et
)&*+(,"*$-#
al. (2009) found that in spite of a highly seasonal
24.5 GDGT production in the water column (Pitcher et
.$//'"(01"(23"
al., 2011), sedimentary TEX86 in the North Sea
24
best reflects mean annual SST. Some paleocli-
23.5 matic studies have argued in favor of a seasonality
bias to explain unreasonable TEX86 temperatures,
23 such as the very warm temperatures derived for
high latitudes during the Early Eocene (e.g., Fig.
22.5 6 and Sluijs et al., 2006, 2011). However, this is
22
largely a post-hoc interpretation. It may be that
500 1000 1500 2000 such warm temperatures reflect the inherent diffi-
YEAR AD
culties in calibrating TEX86 at high latitudes dur-
ing a hothouse interval, for which we have no
FIGURE 7.—TEX86-inferred Lake Tanganyika lake
surface temperatures for the past 1500 years (Tierney modern analog (Pancost et al., 2011). Further-
et al., 2010a). Error bars represent the 95% confi- more, as TEX86 is explicitly calibrated to mean
dence intervals based on a leave-one-out estimation of annual SST, invocations of a seasonal bias need to
the error in the calibration slope. be accompanied by an appropriate seasonal cali-
bration (e.g., Shevenell et al., 2011). Further stud-
ies aimed at constraining possible seasonal biases
temperatures at 100–200 m are highly correlated in GDGT export and sedimentary TEX86 expres-
with SSTs spatially across the global ocean (Hu- sion will be critical towards resolving this issue.
ber, 2010). Everywhere the correlation is robust, The lack of an identifiable producer funda-
TEX86 will predict SST accurately. mentally limits the application of brGDGT prox-
In certain oceanographic environments, such ies. Until the producer(s) is/are better character-
as near-shore and upwelling areas, TEX86 seems ized, it is difficult to identify ecological con-
to be overtly biased towards subsurface tempera- straints that may interfere with the fidelity of
tures (Huguet et al., 2007; Lee et al., 2008; Lopes brGDGTs as temperature recorders (e.g., nutri-
Dos Santos et al., 2010; Chazen, 2011). For ex- ents, oxygen, community effects). Currently, we
ample, Lopes Dos Santos et al. (2010) found that are limited to empirical observations from core-
the time evolution of TEX86-SST was substan- top studies. Weijers et al. (2007c) detected a sig-
tially different from that of alkenone-SST along nificant relationship between the MBT index and
the Guinea Margin in the Late Quaternary, and precipitation amount in soils, but because precipi-
argued that TEX86 reflected subsurface variability. tation co-varied with temperature in their dataset,
Uncertainties surrounding where and why TEX86 they were not able to determine whether this rela-
reflects subsurface temperatures certainly have tionship was meaningful. In lakes, although tem-
relevance for temperature estimates in deep geo- perature is the dominant environmental variable
logical time, where the paleoceanography often is affecting brGDGT distributions there is some in-
not well constrained. fluence of pH, lake depth (Tierney et al., 2010b;
Seasonal changes in Thaumarchaeotal abun- Pearson et al., 2011), and possibly degree of
dance and/or production of isoGDGTs may also bottom-water anoxia (see discussion above;
influence sedimentary TEX86. Water-column stud- Tierney et al., 2012). Interestingly, brGDGTs
ies indicate that Thaumarchaeota and their seem to be absent or have unusual distributions in
isoGDGTs exhibit pronounced seasonal varia- soda lakes with a high pH (Tierney et al., 2010b;
tions, although the timing of peak abundance var- Sun et al., 2011), and brGDGT concentrations are
ies substantially by environment and geographic low in high-pH soils (Weijers et al., 2007c). This
location (Massana et al., 1997; Wuchter et al., is circumstantial evidence that the producers pre-
2006; Galand et al., 2010; Hollibaugh et al., 2010; fer neutral and lower pH environments, and is
Pitcher et al., 2011). Based on limited studies it is supportive of the proposed association with Aci-
not yet clear how or whether these seasonal cycles dobacteria.

125
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

A potentially severe complicating factor for straining absolute temperature error is less impor-
application of brGDGT paleothermometers in tant than relative changes, error can be estimated
lakes is the likelihood that brGDGTs are associ- on the calibration slope (e.g., Tierney et al.,
ated with both allochthonous soil OM and in-situ 2010a). In sum, the RMSE values shown in Ta-
production, each of which appear to have a differ- bles 1 and 2 should not be viewed as immutable
ent relationship to temperature (Tierney and Rus- estimates of the error of GDGT-based temperature
sell, 2009). However, brGDGT concentrations inferences. As with any proxy, error needs to be
(per gram of organic matter) tend to be much rigorously assessed within the geological context
higher in lakes than in soils—sometimes by as at hand, and in light of the climatological hy-
much as an order of magnitude (Tierney and Rus- potheses being investigated.
sell, 2009; Tierney et al., 2010b)—and the newest
lake calibrations based on alternate functional CONCLUSIONS
forms (i.e., those that are not based on the MBT
or CBT indices) have remarkably good predict- Although new to the paleotemperature proxy
ability (Tierney et al., 2010b; Pearson et al., 2011, scene, GDGTs significantly expand the toolkit
and see Table 2). Detailed down-core studies will available to geologists for reconstructing past
help address these issues. temperatures. Since GDGTs are exceptionally
ubiquitous compounds found in marine sedi-
A note on calibration uncertainty ments, ancient sedimentary sequences, lakes,
Both TEX86 and brGDGT paleothermometers soils, hot springs and even speleothems (Yang et
rely on surface-sediment calibrations to convert al., 2011), they offer opportunities to reconstruct
relative abundances to temperature. Global sur- temperatures in geographic places and geological
face calibrations have an advantage in that they time intervals where other proxies may not be
cover a wide diversity of natural environments available. Many uncertainties, especially ecologi-
and thus incorporate a large range of sources of cal uncertainties, need to be better constrained so
variability leading to error. However, the implicit that GDGT proxies can be applied with more con-
assumption of this approach is that the Earth sys- fidence. Nevertheless, as the data in Figure 6 il-
tem is ergodic, having the same type of variability lustrate, GDGTs have already fundamentally in-
in modern geographical space as it does in geo- fluenced our thinking about climate sensitivity
logical time. Depending on the paleoenvironment, during ancient hothouse worlds, highlighting their
this assumption may or may not always be valid. utility as paleothermometers.
For example, the TEX86 temperatures estimated
for the Cretaceous and for the mid to low latitudes REFERENCES
during the PETM (discussed above) rely on an
extrapolation of the modern calibration equations BECHTEL, A., R. SMITTENBERG, S. BERNASCONI, AND
to temperatures higher than observed in the mod- C. SCHUBERT. 2010. Distribution of branched and
ern oceans. They are also contingent on assuming isoprenoid tetraether lipids in an oligotrophic and
that Thaumarchaeota living in (e.g.) the Creta- a eutrophic Swiss lake: Insights into sources and
GDGT-based proxies. Organic Geochemistry,
ceous oceans experienced a similar sensitivity to
41:822–832.
temperature and non-temperature effects as in the BLAGA, C., G. REICHART, S. SCHOUTEN, A. LOTTER, J.
modern oceans. As the latter is very hard to con- WERNE, S. KOSTEN, N. MAZZEO, N., G. LACEROT,
strain with confidence, the error associated with AND J. S. SINNINGHE DAMSTÉ. 2010. Branched
the absolute temperature estimates in ancient hot- glycerol dialkyl glycerol tetraethers in lake sedi-
house environments is almost certainly quite a bit ments: Can they be used as temperature and pH
larger than the root mean square error (RMSE) of proxies? Organic Geochemistry, 41:1225–1234.
the modern core-top calibrations. Conversely, a BLAGA, C. I., G.-J. REICHART, O. HEIRI, AND J. S. SIN-
Late Quaternary TEX86 application in a single NINGHE DAMSTÉ. 2009. Tetraether membrane lipid
tropical locale is unlikely to incorporate sources distributions in water-column particulate matter
of error associated with the polar oceans (and vice and sediments: a study of 47 European lakes along
a north-south transect. Journal of Paleolimnology.
versa), in which case the global core-top calibra-
41 (3), 523–540.
tion error may be an overestimate. In this situa- BLUMENBERG, M., R. SEIFERT, J. REITNER, T. PAPE,
tion, local calibrations can be devised to better AND W. MICHAELIS. 2004. Membrane lipid pat-
constrain sources of error relevant to the region of terns typify distinct anaerobic methanotrophic
interest (e.g., Shevenell et al., 2011) or, if con-

126
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

consortia. Proceedings of the National Academies coastal NW Mediterranean Sea (Blanes Bay Mi-
o f S c i e n c e s U S A , 1 0 11 : 11111 – 1111 6 . crobial Observatory). Limnology and Oceanogra-
doi:10.1073/pnas.0401188101 phy, 55:2117–2125.
BRASSELL, S., A. WARDROPER, I. THOMSON,J. MAX- GLIOZZI, A., G. PAOLI, M. DE ROSA, AND A. GAMBA-
WELL, AND G. EGLINTON. 1981. Specific acyclic CORTA. 1983. Effect of isoprenoid cyclization on
isoprenoids as biological markers of methanogenic the transition temperature of lipids in thermophilic
bacteria in marine sediments. Nature, 290:693. archaebacteria. Biochimica et Biophysica Acta
BROCHIER-ARMANET, C., B. BOUSSAU, S. GRIBALDO, (BBA)—Biomembranes, 735:234–242.
AND P. FORTERRE. 2008. Mesophilic Crenar- HERBERT, T. D., 2003. Alkenone paleotemperature de-
chaeota: proposal for a third archaeal phylum, the terminations. Treatise on Geochemistry, 6:391–
Thaumarchaeota. Nature Reviews Microbiology, 432.
6:245–252. HERNDL, G., T. REINTHALER, E. TEIRA, H. VAN AKEN,
CASTAÑEDA, I., E. SCHEFUß, J. PÄTZOLD, J. S. SIN- C. VETH, A. PERNTHALER,AND J. PERNTHALER.
NINGHE DAMSTÉ, S. WELDEAB, AND S. SCHOUTEN. 2005. Contribution of Archaea to total prokaryotic
2010. Millennial-scale sea surface temperature production in the deep Atlantic Ocean. Applied
changes in the eastern Mediterranean (Nile River Environmental Microbiology, 71:2303–2309.
Delta region) over the last 27,000 years. Paleo- HO, S., M. YAMAMOTO, G. MOLLENHAUER, AND M.
ceanography, 25:PA1208. MINAGAWA. 2011. Core top TEX86 values in the
CHAPPE, B., W. MICHAELIS, AND P. ALBRECHT. 1980. south and equatorial Pacific. Organic Geochemis-
Molecular fossils of archaebacteria as selective try, 42: 94–99.
degradation products of kerogen. Physics and HOEFS, M., S. SCHOUTEN, J. DE LEEUW, L. KING, S.
Chemistry of the Earth, 12:265–274. WAKEHAM, AND J. S. SINNINGHE DAMSTÉ. 1997.
CHAZEN, C., 2011. Holocene climate evolution of the Ether lipids of planktonic archaea in the marine
eastern tropical Pacific told from high resolution water column. Applied Environmental Microbiol-
climate records from the Peru margin and equato- ogy, 63:3090–3095.
rial upwelling regions. Ph.D. thesis, Brown Uni- HOLLIBAUGH, J., S. GIFFORD, S. SHARMA, N. BANO,
versity. AND M. MORAN. 2010. Metatranscriptomic analy-
CHURCH, M., B. WAI, D. KARL, AND E. DELONG. 2010. sis of ammonia-oxidizing organisms in an estuar-
Abundances of crenarchaeal amoA genes and tran- ine bacterioplankton assemblage. The ISME Jour-
scripts in the Pacific Ocean. Environmental Mi- nal, 5:866–878.
crobiology, 12:679–688. HOPMANS, E. C., S. SCHOUTEN, R. D. PANCOST, M. T. J.
DE LA TORRE, J., C. WALKER, A. INGALLS, M. KÖN- VAN DER MEER, AND J. S. SINNINGHE DAMSTÉ.
NEKE, AND D. STAHL. 2008. Cultivation of a ther- 2000. Analysis of intact tetraether lipids in ar-
mophilic ammonia oxidizing archaeon synthesiz- chaeal cell material and sediments by high per-
ing crenarchaeol. Environmental Microbiology, formance liquid chromatography/atmospheric
10:810–818. pressure chemical ionization mass spectrometry.
DE ROSA, M., E. ESPOSITO, A. GAMBACORTA, B. NICO- Rapid Communications in Mass Spectrometry
LAUS, AND J. BU’LOCK. 1980. Effects of tempera- 14:585–589.
ture on ether lipid composition of Caldariella aci- HOPMANS, E. C., J. W. H. WEIJERS, E. SCHEFUß, L.
dophila. Phytochemistry, 19:827–831. HERFORT, J. S. SINNINGHE DAMSTÉ, AND S.
E DER , W., M. S CHMIDT , M. K OCH , D. G ARBE - SCHOUTEN. 2004. A novel proxy for terrestrial
SCHÖNBERG, AND R. HUBER. 2002. Prokaryotic organic matter in sediments based on branched
phylogenetic diversity and corresponding geo- and isoprenoid tetraether lipids. Earth and Plane-
chemical data of the brine–seawater interface of tary Sciences Letters, 224:107–116.
the Shaban Deep, Red Sea. Environmental Micro- HUBER, M., 2010. Why improving molecular and iso-
biology, 4: 758–763. topic proxy paleoclimate records is the most im-
FAWCETT, P., J., J. P. WERNE, R. S. ANDERSON, J. M. portant problem in science today: A climate mod-
HEIKOOP, E. T. BROWN, M. A. BERKE, S. J. SMITH, eling perspective. In Gordon Research Confer-
F. GOFF, L. DONOHOO-HURLEY, L. M. CISNEROS- ence. Holderness, NH, USA.
DOZAL, S. SCHOUTEN, J. S. SINNINGHE DAMSTÉ, Y. HUGUET, C., G. J. DE LANGE, O. GUSTAFSSON, J. J.
HUANG, J. TONEY, J. FESSENDEN, G. WOLDEGAB- MIDDELBURG, J. S. SINNINGHE DAMSTÉ, AND S.
RIEL, V. ATUDOREI, J. W. GEISSMAN, AND C. D. SCHOUTEN. 2008. Selective preservation of soil
ALLEN. 2011. Extended megadroughts in the organic matter in oxidized marine sediments (Ma-
southwestern United States during Pleistocene deira Abyssal Plain). Geochimica et Cosmo-
interglacials. Nature, 470:518–521. chimica Acta, 72: 6061–6068.
GALAND, P., C. GUTIÉRREZ-PROVECHO, R. MASSANA, HUGUET, C., J. KIM, G. DE LANGE, J. S. SINNINGHE
J. GASOL, AND E. CASAMAYORA. 2010. Inter- DAMSTÉ, AND S. SCHOUTEN. 2009. Effects of long
annual recurrence of archaeal assemblages in the term oxic degradation on the TEX86 and BIT or-

127
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

ganic proxies. Organic Geochemistry, 40:1188– VERSTEEGH. 2010. Core-top calibration of the
1194. lipid-based UK0 37 and TEX86 temperature prox-
HUGUET, C., A. SCHIMMELMANN, R. THUNELL, L. ies on the southern Italian shelf (SW Adriatic Sea,
LOURENS, J. S. SINNINGHE DAMSTÉ, AND S. Gulf of Taranto). Earth and Planetary Sciences
SCHOUTEN. 2007. A study of the TEX86 paleo- Letters, 300:112–124.
thermometer in the water column and sediments of LEININGER, S., T. URICH, M. SCHLOTER, L. SCHWARK,
the Santa Barbara Basin, California. Paleoceanog- J. QI, G. NICOL, J. PROSSER, S. SCHUSTER, AND C.
raphy, 22:PA3203. doi:10.1029/2006PA001310 SCHLEPER. 2006. Archaea predominate among
INGALLS, A., S. SHAH, R. HANSMAN, L. ALUWIHARE, ammonia-oxidizing prokaryotes in soils. Nature,
G. SANTOS, E. DRUFFEL, AND A. PEARSON. 2006. 442:806–809.
Quantifying archaeal community autotrophy in the LIU, X., A. LEIDER, A. GILLESPIE, J. GRÖGER, G. VER-
mesopelagic ocean using natural radiocarbon. Pro- STEEGH, AND K. HINRICHS. 2010. Identification of
ceedings of the National Academies of Sciences polar lipid precursors of the ubiquitous branched
USA 103:6442–6447. GDGT orphan lipids in a peat bog in Northern
IONESCU, D., S. PENNO, M. HAIMOVICH, B. RIHT- Germany. Organic Geochemistry, 41:653–660.
MAN,A. GOODWIN,D. SCHWARTZ, L. HAZANOV, LIU, Z., M. PAGANI, D. ZINNIKER,R. DECONTO, M.
M. CHERNIHOVSKY, A. POST, AND A. OREN. 2009. HUBER, H. BRINKHUIS, S. SHAH, R. LECKIE, AND
Archaea in the Gulf of Aqaba. FEMS Microbiol- A. PEARSON. 2009. Global cooling during the
ogy Ecology, 69: 425–438. Eocene-Oligocene climate transition. Science,
KARNER, M., E. DELONG, AND D. KARL. 2001. Ar- 323:1187–1190.
chaeal dominance in the mesopelagic zone of the LLIRÓS, M., F. GICH, A. PLASENCIA, J. AUGUET, F.
Pacific Ocean. Nature, 409:507–510. DARCHAMBEAU, E. CASAMAYOR, J. DESCY, AND C.
KIM, J., J. VAN DER MEER, S. SCHOUTEN, P. HELMKE, BORREGO. 010. Vertical distribution of ammonia-
V. WILLMOTT, F. SANGIORGI, N. KOC, E. HOP- oxidizing crenarchaeota and methanogens in the
MANS, J. S. SINNINGHE DAMSTÉ. 2010. New indi- epipelagic waters of Lake Kivu (Rwanda-
ces and calibrations derived from the distribution Democratic Republic of the Congo). Applied En-
of crenarchaeal isoprenoid tetraether lipids: Impli- vironmental Microbiology, 76:853–6863.
cations for past sea surface temperature recon- LOPES DOS SANTOS, R., M. PRANGE, I. CASTAÑEDA,E.
structions. Geochimica et Cosmochimica Acta, SCHEFUß, S. MULITZA, M. SCHULZ, E. NIEDER-
74:4639–4654. MEYER, J. SINNINGHE DAMSTÉ, AND S. SCHOUTEN.
KIM, J.-H., C. HUGUET, K. A. F. ZONNEVELD, G. J. M. 2010. Glacial-interglacial variability in Atlantic
VERSTEEGH, W. ROEDER, J. S. SINNINGHE DAM- meridional overturning circulation and thermo-
STÉ, AND S. SCHOUTEN. 2009. An experimental cline adjustments in the tropical North Atlantic.
field study to test the stability of lipids used for the Earth and Planetary Sciences Letters, 300:407–
TEX86 and UK0 37 palaeothermometers. Geo- 414.
chimica et Cosmochimica Acta, 73:2888–2898. MARLOWE, I., J. GREEN, A. NEAL,S. BRASSELL, G. EG-
KIM, J.-H., S. SCHOUTEN, E. C. HOPMANS, B. DONNER, LINTON, AND P. COURSE. 1984. Long chain (n-
AND J. S. SINNINGHE DAMSTÉ. 2008. Global sedi- C37–C39) alkenones in the Prymnesiophyceae.
ment core-top calibration of the TEX86 paleother- Distribution of alkenones and other lipids and their
mometer in the ocean. Geochimica et Cosmo- taxonomic significance. British Phycological
chimica Acta72 (4), 1154–1173. Journal, 19:203–216.
KITA, M., S. AIBARA, M. KATO, M. ISHINAGA, AND T. MASSANA, R., A. MURRAY, C. PRESTON, AND E. DE-
HATA. 1973. Effect of changes in fatty acid com- LONG. 1997. Vertical distribution and phylogenetic
position of phospholipid species on the b -galacto- characterization of marine planktonic Archaea in
side transport system of Escherichia coli K-12. the Santa Barbara Channel. Applied Environmen-
Biochimica et Biophysica Acta (BBA)- tal Microbiology63 (1), 50–56.
Biomembranes 298:69–74. NOZAWA, Y., H. IIDA, H. FUKUSHIMA, K. OHKI, AND S.
KÖNNEKE, M., A. E. BERNHARD, J. R. DE LA TORRE, C. OHNISHI. 1974. Studies on Tetrahymena mem-
B. WALKER, J. B. WATERBURY, AND D. A. STAHL. branes: Temperature-induced alterations in fatty
2005. Isolation of an autotrophic ammonia- acid composition of various membrane fractions in
oxidizing marine archaeon. Nature, 437:543–546. Tetrahymena pyriformis and its effect on mem-
LEE, K., J. KIM, I. WILKE, P. HELMKE, AND S. brane fluidity as inferred by spin-label study. Bio-
SCHOUTEN. 2008. A study of the alkenone, TEX86, chimica et Biophysica Acta (BBA)-
and planktonic foraminifera in the Benguela Up- Biomembranes, 367:134–147.
welling System: Implications for past sea surface OCHSENREITER, T., D. SELEZI, A. QUAISER, L. BONCH-
temperature estimates. Geochemistry Geophysics OSMOLOVSKAYA, AND C. SCHLEPER. 2003. Diver-
Geosystems, 9:Q10019. sity and abundance of Crenarchaeota in terrestrial
LEIDER, A., K. HINRICHS, G. MOLLENHAUER, AND G. habitats studied by 16S RNA surveys and real time

128
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

PCR. Environmental Microbiology, 5: 787–797. Geochemistry, 41:404–413.


PANCOST, R., K. TAYLOR, L. HANDLEY, M. HUBER, M., RAY, P., D. WHITE, AND T. BROCK. 1971. Effect of
AND C. HOLLIS. 2011. A Critical Evaluation of temperature on the fatty acid composition of
High TEX86-derived Sea Surface Temperatures Thermus aquaticus. Journal of Bacteriology
from the Early Eocene. In: AGU Fall Meeting Ab- 106:25.
stracts, 1:6. RUEDA, G., A. ROSELL-MELÉ, M. ESCALA, R. GYL-
PEARSON, A., Z. HUANG, A. INGALLS, C. ROMANEK, J. LENCREUTZ, AND J. BACKMAN. 2009. Comparison
WIEGEL, K. FREEMAN, R. SMITTENBERG, AND C. of instrumental and GDGT-based estimates of sea
ZHANG. 2004. Nonmarine crenarchaeol in Nevada surface and air temperatures from the Skagerrak.
hot springs. Applied Environmental Microbiology, Organic Geochemistry, 40:287–291.
70: 5229. SCHOUTEN, S., A. FORSTER, E. PANOTO, AND J. S. SIN-
PEARSON, A., A. P. MCNICHOL, B. C. BENITEZ- NINGHE DAMSTÉ. 2007a. Towards calibration of
NELSON, J. M. HAYES, AND T. I. EGLINTON. 2001. the TEX86 palaeothermometer for tropical sea sur-
Origins of lipid biomarkers in Santa Monica Basin face temperatures in ancient greenhouse worlds.
surface sediment: A case study using compound- Organic Geochemistry, 38:1537–1546.
specific 14C analysis. Geochimica et Cosmo- SCHOUTEN, S., E. HOPMANS, M. BAAS, H. BOUMANN,
chimica Acta, 65:3123–3137. S. STANDFEST, M. KÖNNEKE, D. STAHL, AND J.
PEARSON, E., S. JUGGINS, H. TALBOT, J. WECKSTÖRM, SINNINGHE DAMSTÉ. 2008a. Intact membrane lip-
P. R OSÉN , D. R YVES , S. R OBERTS , AND R. ids of “Candidatus Nitrosopumilus maritimus,” a
SCHMIDT. 2011. A lacustrine GDGT-temperature cultivated representative of the cosmopolitan
calibration from the Scandinavian Arctic to Ant- mesophilic group I crenarchaeota. Applied Envi-
arctic: Renewed potential for the application of ronmental Microbiology, 74: 2433.
GDGTpaleothermometry in lakes. Geochimica et SCHOUTEN, S., E. C. HOPMANS, A. FORSTER, Y. VAN
Cosmochimica Acta, 75:6225–6238. BREUGEL, M. M. M. KUYPERS, AND J. S. SIN-
PETERSE, F., E. HOPMANS, S. SCHOUTEN, A. METS, W. NINGHE DAMSTÉ. 2003. Extremely high sea-
RIJPSTRA, AND J. SINNINGHE DAMSTÉ. 2011. Iden- surface temperatures at low latitudes during the
tification and distribution of intact polar branched middle Cretaceous as revealed by archaeal mem-
tetraether lipids in peat and soil. Organic Geo- brane lipids. Geology, 31: 1069–1072.
chemistry, 42:1007–1015. SCHOUTEN, S., E. C. HOPMANS, E. SCHEFUß, E., AND J.
PITCHER, A., N. RYCHLIK, E. HOPMANS, E. SPIECK, W. S. SINNINGHE DAMSTÉ. 2002. Distributional varia-
RIJPSTRA, J. OSSEBAAR, S. SCHOUTEN, M. WAG- tions in marine crenarchaeotal membrane lipids: a
NER, AND J. SINNINGHE DAMSTÉ. 2010. Crenar- new tool for reconstructing ancient sea water tem-
chaeol dominates the membrane lipids of Candida- peratures? Earth and Planetary Sciences Letters.
tus Nitrososphaera gargensis, a thermophilic 204:265–274.
Group I. 1b Archaeon. The ISME Journal, 4:542– SCHOUTEN, S., E. C. HOPMANS, AND J. S. SINNINGHE
552. DAMSTÉ. 2004. The effect of maturity and deposi-
P ITCHER , A., C. W UCHTER , K. S IEDENBERG , S. tional redox conditions on archaeal tetraether lipid
SCHOUTEN, AND J. S. SINNINGHE DAMSTÉ. 2011. palaeothermometry. Organic Geochemistry,
Crenarchaeol tracks winter blooms of ammonia- 35:567–571.
oxidizing Thaumarchaeota in the coastal North SCHOUTEN, S., C. HUGUET, E. C. HOPMANS, M. V. M.
Sea. Limnology and Oceanography, 56:2308– KIENHUIS, AND J. S. SINNINGHE DAMSTÉ. 2007b.
2318. Analytical methodology for TEX86 paleother-
POWERS, L. A., T. C. JOHNSON, J. P. WERNE, I. S. CAS- mometry by high-performance liquid
TAÑEDA, E. C. HOPMANS, J. S. SINNINGHE DAM- chromatography/atmospheric pressure chemical
STÉ, AND S. SCHOUTEN. 2005. Large temperature ionization-mass spectrometry. Analytical Chemis-
variability in the southern African tropics since the try 79: 2940–2944.
Last Glacial Maximum. Geophysical Research SCHOUTEN, S., W. RIJPSTRA, E. DURISCH-KAISER, C.
Letters 32: L08706. doi:10.1029/2004GL022014. SCHUBERT, AND J. SINNINGHE DAMSTÉ. 2012. Dis-
POWERS, L. A., J. P. WERNE, T. C. JOHNSON, E. C. tribution of glycerol dialkyl glycerol tetraether
H OPMANS , J. S. S INNINGHE D AMSTÉ , AND lipids in the water column of Lake Tanganyika.
SCHOUTEN, S. 2004. Crenarchaeotal membrane Organic Geochemistry, in press.
lipids in lake sediments: A new paleotemperature SCHOUTEN, S., M. VAN DER MEER, E. HOPMANS, AND J.
proxy for continental paleoclimate reconstruction? SINNINGHE DAMSTÉ. 2008b. Comment on “Lipids
Geology, 32:613–616. of marine Archaea: Patterns and provenance in the
POWERS, L. A., J. P. WERNE, A. J. VANDERWOUDE, J. S. water column and sediments” by Turich et al.
SINNINGHE DAMSTÉ ,E. C. HOPMANS, AND S. (2007). Geochimica et Cosmochimica Acta, 72:
SCHOUTEN. 2010. Applicability and calibration of 5342–5346.
the TEX86 paleothermometer in lakes. Organic SEKI, O., T. SAKAMOTO, S. SAKAI, S. SCHOUTEN, E.

129
THE PALEONTOLOGICAL SOCIETY PAPERS, VOL. 18

HOPMANS, J. SINNINGHE DAMSTÉ, AND R. PAN- ARMANET, T. RATTEI, P. TISCHLER, E. SPIECK, W.


COST. 2009. Large changes in seasonal sea ice STREIT, D. STAHL, M. WAGNER, AND C. SCHLEPER.
distribution and productivity in the Sea of Okhotsk 2010. Distinct gene set in two different lineages of
during the deglaciations. Geochemistry Geophys- ammonia-oxidizing archaea supports the phylum
ics Geosystems, 10:Q10007. Thaumarchaeota. Trends In Microbiology,
SHAH, S. R., G. MOLLENHAUER, N. OHKOUCHI, T. I. 18:331–340.
EGLINTON, AND A. PEARSON. 2008. Origins of STURT, H. F., R. E. SUMMONS, K. SMITH, M. ELVERT,
archaeal tetraether lipids in sediments: Insights AND K.-U. HINRICHS. 2004. Intact polar membrane
from radiocarbon analysis. Geochimica et Cosmo- lipids in prokaryotes and sediments deciphered by
chimica Acta, 72:4577–4594. high-performance liquid chromatography/
SHEVENELL, A., A. INGALLS, E. DOMACK, AND C. electrospray ionization multistage mass spec-
KELLY. 2011. Holocene Southern Ocean surface trometry—new biomarkers for biogeochemistry
temperature variability west of the Antarctic Pen- and microbial ecology. Rapid Communications in
insula. Nature, 470:250–254. Mass Spectrometry, 18:617–628.
SHINTANI, T., M. YAMAMOTO, AND M. CHEN. 2010. SUN, Q., G. CHU, M. LIU, M. XIE, S. LI, Y. LING, X.
Paleoenvironmental changes in the northern South WANG, L. SHI, G. JIA, AND H. LÜ. 2011. Distribu-
China Sea over the past 28,000 years: A study of tions and temperature dependence of branched
TEX86-derived sea surface temperatures and ter- glycerol dialkyl glycerol tetraethers in recent la-
restrial biomarkers. Journal of Asian Earth Sci- custrine sediments from China and Nepal. Journal
ences, 40:1221–1229. of Geophysical Research–Biogeoscience,
SINNINGHE DAMSTÉ, J., W. RIJPSTRA, E. HOPMANS, J. 116:G01008.
WEIJERS, B. FOESEL, J. OVERMANN, AND S. DE- THIERSTEIN, H., K. GEITZENAUER, AND B. MOLFINO.
DYSH. 2011. 13, 16-Dimethyl octacosanedioic acid Shackleton, N., 1977. Global synchroneity of late
(iso-diabolic acid), a common membrane-spanning Quaternary coccolith datum levels: Validation by
lipid of Acidobacteria subdivisions 1 and 3. Ap- oxygen isotopes. Geology, 5:400.
plied Environmental Microbiology, 77:4147–4154. TIERNEY, J., M. MAYES, N. MEYER, C. JOHNSON, P.
SINNINGHE DAMSTÉ, J., W. RIJPSTRA, AND G. REI- SWARZENSKI, A. COHEN, AND J. RUSSELL. 2010a.
CHART. 2002a. The influence of oxic degradation Late-twentieth-century warming in Lake Tangany-
on the sedimentary biomarker record II. Evidence ika unprecedented since AD 500. Nature Geosci-
from Arabian Sea sediments. Geochimica et Cos- ence, 3:422–425.
mochimica Acta, 66:2737–2754. TIERNEY, J., S. SCHOUTEN, A. PITCHER, E. HOPMANS,
S INNINGHE D AMSTÉ , J. S., E. C. H OPMANS , S. AND J. SINNINGHE DAMSTÉ. 2012. Core and intact
SCHOUTEN, A. C. T. VAN DUIN, AND J. A. J. polar glycerol dialkyl glycerol tetraethers
GEENEVASEN. 2002b. Crenarchaeol: the character- (GDGTs) in Sand Pond, Warwick, Rhode Island
istic core glycerol dibiphytanyl glycerol tetraether (USA): Insights into the origin of lacustrine
membrane lipid of cosmopolitan pelagic crenar- GDGTs. Geochimica et Cosmochimica Acta,
chaeota. Journal of Lipid Research, 43:1641– 77:561–581.
1651. TIERNEY, J. E., AND J. M. RUSSELL. 2009. Distributions
SINNINGHE DAMSTÉ, J. S., J. OSSEBAAR, B. ABBAS, S. of branched GDGTs in a tropical lake system: Im-
SCHOUTEN, AND D. VERSCHUREN. 2009. Fluxes plications for lacustrine application of the MBT/
and distribution of tetraether lipids in an equatorial CBT paleoproxy. Organic Geochemistry, 40:1032–
African lake: Constraints on the application of the 1036.
TEX86 palaeothermometer and BIT index in lacus- TIERNEY, J. E., J. M. RUSSELL, H. EGGERMONT, E. C.
trine settings. Geochimica et Cosmochimica Acta, HOPMANS, D. VERSCHUREN, AND J. S. SINNINGHE
73:4232–4249. DAMSTÉ. 2010b. Environmental controls on
SLUIJS, A., P. K. BIJL, S. SCHOUTEN, U. RÖHL, G.-J. branched tetraether lipid distributions in tropical
REICHART, AND H. BRINKHUIS. 2011. Southern East African lake sediments. Geochimica et Cos-
ocean warming, sea level and hydrological change mochimica Acta, 74:4902–4918.
during the Paleocene-Eocene thermal maximum. TIERNEY, J. E., J. M. RUSSELL, Y. HUANG, J. S. SIN-
Climate of the Past 7:47–61. NINGHE DAMSTÉ, E. C. HOPMANS, AND A. S. CO-
SLUIJS, A., S. SCHOUTEN, M. PAGANI, M. WOLTERING, HEN. 2008. Northern Hemisphere controls on
H. BRINKHUIS, J. S. SINNINGHE DAMSTÉ, G. R. tropical southeast African climate during the past
DICKENS, M. HUBER, G.-J. REICHART, R. STEIN, J. 60,000 years. Science, 322:252–255.
MATTHIESSEN, L. J. LOURENS, N. PEDENTCHOUK, TROMMER, G., M. SICCHA, M. VAN DER MEER, S.
J. BACKMAN, AND K. MORAN. 2006. Subtropical SCHOUTEN, J. S. SINNINGHE DAMSTÉ, H. SCHULZ,
arctic ocean temperatures during the Palaeocene/ C. HEMLEBEN, C. AND M. KUCERA. 2009. Distri-
Eocene thermal maximum. Nature, 441:610–613. bution of Crenarchaeota tetraether membrane lip-
S PANG , A., R. H ATZENPICHLER , C. B ROCHIER - ids in surface sediments from the Red Sea. Or-

130
TIERNEY: RECONSTRUCTING PALEOTEMPERATURES USING GDGT THERMOMETRY

ganic Geochemistry, 40:724–731. S. S INNINGHE D AMSTÉ . 2004. Temperature-


TURICH, C., K. H. FREEMAN,M. A. BRUNS, M. H. dependent variation in the distribution of tetraether
CONTE, A. D. JONES, AND S. G. WAKEHAM. 2007. membrane lipids of marine Crenarchaeota: Impli-
Lipids of marine archaea: Patterns and provenance cations for TEX86 paleothermometry. Paleocean-
in the water-column and sediments. Geochimica et ography, 19:PA4028.
Cosmochimica Acta, 71:3272–3291. WUCHTER, C., S. SCHOUTEN, S. G. WAKEHAM, AND J.
TYLER, J., A. NEDERBRAGT, V. JONES, AND J. THUROW. S. SINNINGHE DAMSTÉ. 2005. Temporal and spa-
2010. Assessing past temperature and soil pH es- tial variation in tetraether membrane lipids of ma-
timates from bacterial tetraether membrane lipids: rine crenarchaeota in particulate organic matter:
Evidence from the recent lake sediments of Implications for TEX86 paleothermometry. Paleo-
Lochnagar, Scotland. Journal of Geophysical Re- ceanography, 20:PA3013.
search–Biogeoscience, 115:G01015. WUCHTER, C., S. SCHOUTEN, S. G. WAKEHAM, AND J.
UDA, I., A. SUGAI, Y. ITOH, AND T. ITOH, T. 2001. Varia- S. SINNINGHE DAMSTÉ. 2006. Archaeal tetraether
tion in molecular species of polar lipids from membrane lipid fluxes in the northeastern Pacific
Thermoplasma acidophilum depends on growth and the Arabian Sea: Implications for TEX86 pa-
temperature. Lipids, 36:103–105. leothermometry. Paleoceanography, 21:PA4208.
VOLKMAN, J., G. EGLINTON, E. CORNER, AND T. FORS- YANG, H., W. DING, C. ZHANG, X. WU, X. MA, G. HE,
BERG. 1980. Long-chain alkenes and alkenones in J. HUANG, AND S. XIE. 2011. Occurrence of tetra-
the marine coccolithophorid Emiliania huxleyi. ether lipids in stalagmites: Implications for
Phytochemistry, 19:2619–2622. sources and GDGT-based proxies. Organic Geo-
WAKEHAM, S. G., C. LEWIS, E. C. HOPMANS, S. chemistry, 42:108–115.
SCHOUTEN, AND J. S. SINNINGHE DAMSTÉ. 2003. ZACHOS, J. C., S. SCHOUTEN, S. BOHATY, T. QUATTLE-
Archaea mediate anaerobic oxidation of methane BAUM, A. SLUIJS, H. BRINKHUIS, S. J. GIBBS, AND
in deep euxinic waters of the Black Sea. Geo- T. J. BRALOWER. 2006. Extreme warming of mid-
chimica et Cosmochimica Acta67:1359–1374. latitude coastal ocean during the Paleocene-
WEIJERS, J. W. H., E. PANOTO, J. VAN BLEIJSWIJK, S. Eocene Thermal Maximum: Inferences from
SCHOUTEN, W. I. C. RIJPSTRA, M. BALK,A. J. M. TEX86 and isotope data. Geology, 34:737–740.
STAMS, AND J. S. SINNINGHE DAMSTÉ.2009. Con- ZHANG, Y., C. ZHANG, X. LIU, L. LI, K. HINRICHS, AND
straints on the Biological Source(s) of the Orphan J. NOAKES. 2011. Methane Index: A tetraether
Branched Tetraether Membrane Lipids. Geomi- archaeal lipid biomarker indicator for detecting the
crobiology Journal, 26:402–414. instability of marine gas hydrates. Earth and
WEIJERS, J. W. H., E. SCHEFUß, S. SCHOUTEN, AND J. S. Planetary Science Letters, 307:525–534.
SINNINGHE DAMSTÉ. 2007a. Coupled thermal and ZINK, K., M. VANDERGOES, K. MANGELSDORF, A.
hydrological evolution of tropical Africa over the DIEFFENBACHER-KRALL, AND L. SCHWARK, L.
last deglaciation. Science, 315:1701–1704. 2010. Application of bacterial glycerol dialkyl
WEIJERS, J. W. H., S. SCHOUTEN, E. C. HOPMANS, J. A. glycerol tetraethers (GDGTs) to develop modern
J. GEENEVASEN, O. R. P. DAVID, J. M. COLEMAN, and past temperature estimates from New Zealand
R. D. PANCOST, AND J. S. SINNINGHE DAMSTÉ. lakes. Organic Geochemistry 41:1060–1066.
2006a. Membrane lipids of mesophilic anaerobic
bacteria thriving in peats have typical archaeal
traits. Environmental Microbiology, 8:648–657.
WEIJERS, J. W. H., S. SCHOUTEN, A. SLUIJS, H.
BRINKHUIS, AND J. S. SINNINGHE DAMSTÉ. 2007b.
Warm arctic continents during the Palaeocene-
Eocene thermal maximum. Earth and Planetary
Sciences Letters, 261:230–238.
WEIJERS, J. W. H., S. SCHOUTEN, O. C. SPAARGAREN,
AND J. S. SINNINGHE DAMSTÉ. 2006b. Occurrence
and distribution of tetraether membrane lipids in
soils: Implications for the use of the TEX86 proxy
and the BIT index. Organic Geochemistry,
37:1680–1693.
WEIJERS, J. W. H., S. SCHOUTEN, J. C. VAN DEN
DONKER, E. C. HOPMANS, AND J. S. SINNINGHE
DAMSTÉ. 2007c. Environmental controls on bacte-
rial tetraether membrane lipid distribution in soils.
Geochimica et Cosmochimica Acta, 71:703–713.
WUCHTER, C., S. SCHOUTEN, M. J. L. COOLEN, AND J.

131

You might also like