Hydrogen isotope ratios of aliphatic and diterpenoid hydrocarbons

Organic
Geochemistry
Organic Geochemistry 37 (2006) 165–176
www.elsevier.com/locate/orggeochem
Hydrogen isotope ratios of aliphatic and
diterpenoid hydrocarbons in coals and carbonaceous
mudstones from the Liaohe Basin, China
Jincai Tuo
a,*
, Mingfeng Zhang
a,b
, Xianbin Wang a, Chuanlun Zhang
c
a
Key Laboratory of Gas Geochemistry, Institute of Geology and Geophysics, Chinese Academy of Sciences, No. 382 Donggang
West Road, Lanzhou 730000, Gansu, PR China
b
Graduate University of the Chinese Academy of Sciences, 100039, PR China
c
Savannah River Ecology Laboratory and Department of Marine Sciences, University of Georgia, P.O. Box Drawer E, Aiken,
SC 29802, USA
Received 21 June 2005; accepted 1 October 2005
(returned to author for revision 8 September 2005)
Available online 1 December 2005
Abstract
Hydrogen-isotope compositions of the aliphatic and diterpenoid hydrocarbons were determined for five coal and carbonaceous mudstone samples collected from drilling cores (1531–1767 m depths) in the Liaohe Basin, China. The bulk
organic materials were mainly derived from terrestrial higher plants. dD values for most of the n-alkanes varied from
150& to 220&, and were not significantly different among the samples. Pristane was 34–69& depleted in D relative
to phytane; both pristane and phytane, however, had the same trend of variation in dD from sample to sample. Diterpenoids were on average 49–81& depleted in D relative to the n-alkanes. Variations in dD also occurred between different
diterpenoids, indicating a different source for these compounds. An enrichment process for the heavy hydrogen isotope was
observed as expected when a compound was progressively altered through diagenesis (especially the dehydrogenation process). Overall, dD and d13C showed distinct patterns between structurally different lipid classes, although possible hydrogen
exchange cannot be completely excluded during maturation. Our results further support the notion that hydrogen isotopes
of lipid biomarkers from ancient sediments can be used to assess the origin of the organic matter, to determine oil-source
rock correlation, and perhaps to reconstruct the paleoenvironment under which the organic material was deposited.
2005 Elsevier Ltd. All rights reserved.
1. Introduction
Lipid biomarkers are biochemicals that derive
from a restricted range of organisms and thus pro*
Corresponding author. Tel.: +86 0931 4960854; fax: +86 0931
8278667.
E-mail address: [email protected] (J. Tuo).
vide a highly selective means of isolating material
of specific origin (Volkman et al., 1998; Sauer
et al., 2001; Schouten et al., 2001a,b, 2003; Sinninghe Damste´ et al., 2002; Chikaraishi and Naraoka,
2003; Van der Meer et al., 2003; Pancost and Sinninghe Damste´, 2003; Pancost and Boot, 2004).
For example, terrestrial plants are characterized by
a strong odd predominance in the C25–C35
0146-6380/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2005.10.001
166
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
n-alkanes (Castillo et al., 1967; Rieley et al., 1991;
Collister et al., 1994; Chikaraishi and Naraoka,
2003), whereas aquatic plants are characterized by
C23 and C25 n-alkanes (Baas et al., 2000; Ficken
et al., 2000; Zhang et al., 2004). Algae and cyanobacteria are characterized by short-chain n-alkanes
(C15, C17 and C19) (Han et al., 1968; Gelpi et al.,
1970). Tricyclic diterpenoids such as 18-norpimarane, pimarane, dehydroabietane and simonellite
are thought to be derived from resins formed by
higher plants, primarily gymnosperms (in particular
conifers), but also some angiosperms, pteridophytes, and bryophytes (Noble et al., 1985). Tetracyclic diterpenoids such as phyllocladane are
primarily derived from conifers (Karrer, 1958; Aplin
et al., 1963; Erdtman and Norin, 1966; Karrer et al.,
1977; Alexander et al., 1987; Sukh Dev, 1989; Otto
et al., 1997; Otto and Simoneit, 2001; Bechtel et al.,
2002; Otto et al., 2005).
Compound-specific hydrogen isotopes of organic
compounds are emerging as a new proxy for paleoclimatic, paleohydrological, and paleoenvironmental
studies (Andsersen et al., 2001; Sauer et al., 2001;
Huang et al., 2002; Yang and Huang, 2003; Sachse
et al., 2004a,b; Sessions et al., 2004; Dawson et al.,
2004, 2005; Xiong et al., 2005; Liu and Huang,
2005; Sun et al., 2005). Hydrogen in lipid biomarkers
is commonly bound to carbon and nonexchangeable,
whereas it is often exchangeable in kerogen, cellulose, and phenolic substances (Schimmelmann
et al., 1999; Sauer et al., 2001). Hydrogen isotopes
commonly show a large range of fractionation
(Bigeleisen, 1965; Li et al., 2001; Chikaraishi and
Naraoka, 2003). It is known that n-alkanes from
algae are in general depleted in D relative to growth
water by 160&, while sterols are depleted in D relative to growth water by 200& (Sessions et al.,
1999; Sauer et al., 2001). Current knowledge suggests
that the fractionation of H isotopes associated with
biosynthesis is constant and mostly controlled by
the biochemical pathways used by living organisms
(Sessions et al., 1999; Sachse et al., 2004b). The
hydrogen-isotope composition of lipids is controlled
by three factors: isotope composition of biosynthetic
precursors, fractionation and exchange accompanying biosynthesis, and hydrogenation during biosynthesis (Sessions et al., 1999; Xiong et al., 2005).
Variations in the natural abundance of the deuterium (D) covalently bound to carbon may record
both environmental conditions (Yapp and Epstein,
1982; Sternberg, 1988) and biochemical effects
(Estep and Hoering, 1980; Yakir and DeNiro,
1990; Xiong et al., 2005; Sun et al., 2005). Therefore, dD ratios of lipid biomarkers have the potential to record the sources of hydrogen in a
particular environment (Sachse et al., 2004b). Most
studies involving compound-specific hydrogen isotopes are restricted to samples from Holocene to
Miocene (e.g., Xie et al., 2000; Sauer et al., 2001;
Yang and Huang, 2003). Recently, however, dD
ratios of lipid biomarkers have been reported from
source rocks of Oligocene to Upper Permian (Xiong
et al., 2005) and from oil samples derived from
source rocks of Oligocene to Ordovician (Li et al.,
2001; Sun et al., 2005). These studies have proved
that older fossil material preserved in sedimentary
deposits can retain primary D/H compositions in
their lipid biomarkers and thus hydrogen-isotope
ratios of lipid compounds from ancient sediments
can be applied to paleoecological and paleoenvironmental studies at the geological time scale.
Tuo et al. (2003) determined carbon-isotope
compositions for n-alkanes and diterpenoids in
coals and carbonaceous mudstones, in which the
organic materials were mainly derived from terrestrial higher plants. Terrigenous tricyclic and tetracyclic diterpenoid hydrocarbons are about 4–6&
enriched in 13C compared to n-alkanes in these samples. In the current study, the hydrogen-isotope
compositions were determined for the aliphatic
and diterpenoid hydrocarbons in the same samples.
The purpose for this study was to identify differences in hydrogen-isotope compositions between
different lipid classes, which were derived from similar precursor organisms and to validate the use of
hydrogen-isotope signatures for assessing the origin
of the organic matter for oil-source rock correlation
and paleoenvironmental reconstruction.
2. Materials and methods
2.1. Geological setting
The Liaohe Basin, located in Liaonin Province, is
one of the most important Cenozoic sedimentary
basins in northeastern China. The basin was developed from a rift during the Cenozoic and accumulated lacustrine sediments of the Early Tertiary.
Since the discovery of oil from the Xinglongtai Field
in 1975, the Liaohe Basin has become the third largest oil field in China.
The Liaohe Basin can be divided into seven structural units (Tuo et al., 2003; Tuo and Philp, 2005):
western uplift, western depression, central uplift,
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
0.49
0.48
0.48
0.48
0.54
c
d
a
b
E, eogene.
Bit.‘‘A’’, chloroform soluble extract; Sat., saturated hydrocarbons; Arom., aromatic hydrocarbons; NSO, N, S, O compounds (‘‘resins’’); Asph., asphaltenes.
Precent of bitumen ‘‘A’’ extract.
H/C, hydrogen/carbon; O/C, oxygen/carbon.
0.32
0.31
0.33
0.31
0.34
1.01
1.10
1.11
1.16
1.07
0.4
2.0
1.2
1.7
1.8
1.5
5.8
3.4
4.8
6.2
42.70
37.84
36.45
36.22
40.34
28.58
26.81
28.96
28.31
30.07
12.24
15.03
12.66
13.90
13.73
16.40
19.92
21.69
21.17
15.47
0.17
0.39
0.s8
0.60
0.66
0.59
1.11
0.52
1.71
2.26
39.37
19.09
15.12
35.71
36.80
Coal
Carbonaceous mudstone
Carbonaceous mudstone
Coal
Coal
NSO
Arom.
1531
1535
1637
1642
1767
E
E
E
E
E
O/C
H/C
HC/TOC
A/TOC
Sat.
Lithology
Stratuma
Xiao13
Xiao13
Xiao12
Xiao12
Ou 15
Methods for extraction, fractionation, and purification of the soluble organic materials have been
Depth (m)
2.3. Experimental
Well no.
The five samples used in this study were selected
from the Shahejie Formation of the Eocene age, in
the eastern depression of the Liaohe Basin (Wells
Xiao 12, Xiao 13, and Ou 15) (Tuo et al., 2003;
Tuo and Philp, 2005). The geochemical characteristics of these samples are listed in Table 1. The distribution patterns of the aliphatic, diterpenoid and
triterpenoid hydrocarbons and carbon-isotope compositions of aliphatic and diterpenoid hydrocarbons
have been discussed in Tuo et al. (2003) and Tuo
and Philp (2005).
Table 1
Basic geochemical parameters for the samples under study
2.2. Samples
Asph.
Conversion rate
(%)
TOC (wt%)
Bit.‘‘A’’ (%)b
HC (%)c
Bitumen ‘‘A’’ composition
(%)b
Atom
ratios of
kerogend
Ro (%)
eastern depression, eastern uplift, Damintun depression, and Shenbei depression. The Paleogene
sequence can be divided into three formations:
Fangshenpiao, Shahejie and Dongyin. The Shahejie
Formation (Es) is widely distributed and contains
the most important source rock and reservoir units
in the entire basin. Based on lithology and fossil
assemblages, it can be further subdivided into four
members (Es4, Es3, Es2, and Es1, oldest to youngest) (Huang et al., 2003). Previous studies have
shown that the third and fourth members (Es4,
Es3) of the Shahejie Formation in the lower Eogene
(Eocene-Oligocene) are the main source rocks in
this basin. Most of the source rocks are mudstone
and sandy mudstone with types II–III kerogen.
Coal and carbonaceous mudstone are found
mostly in the eastern depression of the basin. They
were deposited mainly in fresh water lacustrine bog
and flood plain facies. Vertically, coal and carbonaceous mudstone are generally present in the third
member of Shahejie Formation and the first member
also contains some carbonaceous mudstone. The
coal beds are distributed in the southern and central
parts of the eastern depression and generally occur in
layers from 1 to 5 m thick, with a maximum singlebed thickness of 25 m (Rong 28 well) and total
cumulative thickness of 111 m (Rong 60 well). The
carbonaceous mudstones are found mostly in the
northern parts of the eastern depression with an
overall cumulative thickness of 344 m (Long 40 well).
Coal and carbonaceous mudstone also constitute the
source rocks for the oil and natural gas produced in
eastern depression of the Liaohe Basin.
167
168
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
described earlier (Tuo et al., 2003; Tuo and Philp,
2005). Briefly, the coal and carbonaceous mudstone-samples were powdered to less than 120 mesh
and Soxhlet-extracted with chloroform for 72 h.
The extracts (chloroform asphalt ‘‘A’’) were concentrated and deasphaltened by addition of excess hexane. Saturated and aromatic hydrocarbons and
nonhydrocarbons (resin) were separated from the
deasphaltened samples by column chromatography
on a column of neutral alumina over silica gel
(approx. 4 g of each). Saturated fractions were
eluted with hexane (150 ml), aromatic fractions with
methylene chloride (150 ml), and nonhydrocarbon
fractions with methanol (30 ml).
Analyses of hydrogen isotopes of individual compounds were performed on a Delta Plus XP gas
chromatography-pyrolysis-isotope ratio mass spectrometer. The gas chromatography was performed
using a Thermo Finnigan GC COMBUSTION III
system equipped with an HP-5 fused silica capillary
column (50 m · 0.32 mm) and helium used as carrier gas with a flow rate of 1 ml/min. The oven temperature was isothermal for 5 min at 80 C and then
programmed from 80 to 300 C at 3 C/min. GCseparated compounds passed through an alumina
tube, which was heated to 1450 C to convert
organic H to H2, which was then introduced to a
Delta Plus XP isotope ratio mass spectrometer.
The reproducibility and accuracy of the analysis
were evaluated routinely using laboratory standards
of known dD values (C14, C16, C18, C23, C28, C32 nalkanes). Typically, one injection of laboratory
standard was performed for every eight sample
analyses. The isotope values are given with respect
to the V-SMOW standard. Samples were analyzed
four times and the results are presented as the average value with a standard deviation. For most of the
lipid compounds, the standard deviations were
below 3&. Greater standard deviations were probably caused by coeluting peaks or small peaks. They
were included in order to provide an unbiased view
of the results.
3. Results and discussion
3.1. Variation of dD in n-alkanes
Table 2 displays measured D/H values for the three
classes of lipid compounds (n-alkanes, isoprenoid
alkanes, and diterpenoids) from five coal- and carbonaceous mudstone-samples. For the n-alkanes,
several features are noticeable in Fig. 1, which
graphically summarizes data from Table 2. First,
for any of the five samples examined, n-alkanes
exhibit restricted ranges in dD, with individual lipids
in each sample generally falling within a range of
<50&. The dD values for most of the measured
n-alkanes in all of the analyzed samples vary from
150& to 220&. Similar dD values for the n-alkanes have also been reported in Jurassic lacustrine
mudstones inter-bedded with coal (Xiong et al.,
2005). Chikaraishi and Naraoka (2003) reported
similar results in n-alkanes from C4 plants
( 171 ± 12&) and from aquatic freshwater plants
( 187 ± 16&). The long-chain n-alkanes of higher
plant leaf waxes isolated from a Chinese loess profile also have a similar dD range ( 140& to
200&) (Liu and Huang, 2005). The measured
dD values for the n-alkanes from all the analyzed
samples in this study are about 10–40& depleted
in D relative to n-alkanes extracted from C3 plants
( 152 ± 26& for angiosperms and 149 ± 16&
for gymnosperms) and from seaweeds ( 155 ±
34&) (Chikaraishi and Naraoka, 2003). Second,
C20 to C30 n-alkanes are depleted in D by 30–50&
relative to both shorter (<C19) and longer (>C31)
chain n-alkanes. Third, in n-alkanes of C20 to C33,
the even carbon numbers are more enriched in D
than they are in their neighboring odd-numbered
carbon homologues. This zigzag distribution pattern for dD value of n-alkanes has been noticed by
Yang and Huang (2003) in the n-alkanes extracted
from some Miocene lacustrine sediments and plant
fossils and by Xiong et al. (2005) in the n-alkanes
from some swamp environments. This distribution
of dD of n-alkanes has been attributed to waxes
from terrestrial higher plants (Xiong et al., 2005).
Finally, n-alkanes with the same chain-lengths from
the different samples exhibit no substantial difference in dD values for four of the five analyzed
samples. The H isotopic compositions of mid-chainlength individual n-alkanes (C19 to C30) from the
sample of Xiao 13 (1535 m) are about 10–20& more
enriched in D than they are from the other four analyzed samples (Table 2, Fig. 1), but this difference is
much smaller than that within each sample for different chain-length n-alkanes. This restricted difference in dD values of n-alkanes with the same
chain-lengths from the different samples may be
attributed to either different rates of D-exchange
due to lithology differences or differences in the
sub-types of organic matter in these sediments,
e.g., difference in the amount of organic matter
derived from angiosperm (Tuo and Philp, 2005).
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
169
Table 2
Hydrogen-isotopic compositions of alkanes and diterpenoids in the coal and carbonaceous mudstone-samples of Liaohe Basin
Well no. (depth)
Xiao 13 (1531 m)
n-Alkanes
dD(&)
r(&)a
dD(&)
r(&)a
dD(&)
nC13
nC14
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
150
162
165
192
168
174
184
207
165
187
196
189
213
194
217
199
221
188
201
162
165
NAb
NAb
2
2
6
2
3
2
3
2
1
4
3
3
3
4
3
NAb
3
NAb
4
166
158
152
166
177
169
171
175
146
175
177
175
180
168
179
162
181
164
165
ndc
162
NAb
3
3
9
6
2
6
19
4
7
2
1
7
NAb
8
NAb
2
NAb
4
1
ndc
158
174
178
175
162
180
202
ndc
207
209
210
213
202
214
210
209
ndc
175
ndc
ndc
Isoprenoid alkanes
Pristane
Pr
Phytane
Ph
243
209
3
3
217
171
15
12
264
213
305
319
198
233
284
3
5
3
2
1
288
297
152
185
280
6
3
2
NAb
3
186
2
168
267
3
240
Diterpenoids
Norpimarane
Pimarane
Simonellite
Dehyaroabietane
16a(H)Phyllocladane
Average (nalkanes)
Average
(diterpenoids)
a
b
c
Xiao13 (1535 m)
Xiao12 (1637 m)
r(&)a
Xiao12 (1642 m)
dD(&)
r(&)a
Ou15 (1767 m)
dD(&)
r(&)a
NAb
NAb
NAb
NAb
2
3
3
NAb
2
5
5
3
1
2
2
8
2
NAb
2
168
160
155
170
152
145
ndc
ndc
ndc
215
206
244
207
196
214
ndc
216
ndc
170
ndc
155
2
4
2
4
2
5
5
158
147
150
165
171
174
192
217
167
200
211
202
206
176
203
185
202
159
173
ndc
160
3
7
245
183
2
14
264
195
3
1
293
296
174
186
273
3
1
2
4
1
292
299
152
234
274
1
1
4
6
3
269
306
176
143
259
4
2
3
2
1
4
192
4
185
3
181
2
3
244
5
250
2
230
2
9
1
3
2
1
5
19
5
1
4
2
2
1
2
3
10
7
4
4
6
4
3
1
2
4
Standard deviation of 4 replicate analyses.
Only one analysis available.
No data.
No substantial differences in dD have been noticed
among the n-alkanes from all other analyzed samples. According to Tuo et al. (2003) and Tuo and
Philp (2005), all the samples used in this study were
collected from similar depositional environments
and the organic matter in all the samples was at a
similar thermal evolution stage (Table 1). So the
generally consistent dD values for the n-alkanes
within each sample and from different samples
probably reflect similar sources of hydrogen for
these saturated hydrocarbons.
3.2. Variation of dD in isoprenoid alkanes (pristane
and phytane)
Values of dD vary from 217& to 264& in pristane (Pr) and from 171& to 213& in phytane
(Ph) (Table 2). Thus, phytane has dD values similar
to the n-alkanes, whereas the pristane is about 50&
depleted in D relative to the n-alkanes. Interestingly,
pristane and phytane follow a similar trend of variation in dD from sample to sample (Fig. 2). The generally lighter dD values of pristane relative to
170
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
-100
δD(‰ , SMOW)
Xiao13(1531)
Xiao13(1535)
Xiao12(1637)
Xiao12(1642)
Ou 15(1767)
-150
-200
-250
13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33
Carbon number
Fig. 1. Hydrogen-isotope ratios of n-alkanes.
-150
δD(‰, SMOW)
Pr
Ph
-200
-250
-300
Xiao13(1531)
Xiao13(1535)
Xiao12(1637)
Xiao12(1642)
Ou 15(1767)
Samples
Fig. 2. Comparison of dD for pristane and phytane among different samples.
phytane was also noticed by Li et al. (2001) in oil
samples derived from source rocks deposited in a
variety of environments. The acyclic isoprenoid
hydrocarbons pristane and phytane are ubiquitous
in sedimentary rocks, crude oils and coals (Koopmans et al., 1999). The phytol side chain of chlorophyll a is known to be the precursor for pristane
and phytane. Alternative precursors for pristane
may be from tocopherols (Goossens et al., 1984;
Koopmans et al., 1999) and lipids from archaea
(Rowland, 1990; Navale, 1994), Pr and its mono-,
di-, and tri-unsatured counterparts in zooplankton
(Blumer et al., 1963, 1969; Blumer and Thomas,
1965), diphytanyl glyceryl ether, S- and O bound
Pr, the (unknown) precursor of prist-2-ene (Koopmans et al., 1999), and bound methylated 2-methyl-
2-(4,8,12-trimethyltridecyl) chromans (Li et al.,
1995). Other precursors for phytane include ether lipids from archaea (Risatti et al., 1984; Rowland, 1990;
Navale, 1994), S-bound Ph (probably originating
from phytol or its diagenetic products) and diphytanyl glyceryl ether (Koopmans et al., 1999).
The multiple sources for pristane and phytane
and associated hydrogen isotope fractionations
may be far more complicated than currently understood. However, for samples that formed in similar
environments, the precursors for pristane and phytane perhaps can be generally characterized. One
such generalization is the similar trend of variation
in dD from sample to sample. For example, Li
et al. (2001) and Schimmelmann et al. (2004) also
observe that phytane is enriched in D relative to
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
pristane in sediments and crude oils. The reason for
the different dD values between pristane and phytane may be attributed to either different origins
of pristane and phytane or different isotope effects
during their derivation from a common phytol precursor (Li et al., 2001; Dawson et al., 2004, 2005).
With increasing maturity, pristane and phytane
become enriched in D while the n-alkanes generally
remain at a constant isotopic composition. The
preferential enrichment of D in isoprenoids over nalkanes with increasing maturity suggests that
hydrogen isotopic exchange occurs more readily
with isoprenoids than with n-alkanes (Dawson
et al., 2004, 2005). The offset between the dD values
of n-alkanes and isoprenoids has been observed in
modern biological samples as well as in ancient oil
and sediment samples (Li et al., 2001; Schimmelmann et al., 2004; Dawson et al., 2004, 2005), indicating that their indigenous dD signatures can be
preserved in geological samples, although possible
hydrogen exchange cannot be completely excluded.
3.3. Variation of dD in diterpenoids
Values of dD measured for diterpenoids are listed
in Table 2 and graphically summarized in Fig. 3.
The major feature is a significant depletion (49–
81&) in D in the diterpenoids relative to the n-alkanes. To the best of our knowledge, no published
work has demonstrated the difference of the hydrogen-isotope ratios between n-alkanes and diterpenoid compounds in sediments or modern plants.
But comparing lipid classes (Estep and Hoering,
1980) or individual compounds within a single
-100
δD(‰, SMOW)
-150
171
organism (Sessions et al., 1999; Sauer et al., 2001),
polyisoprenoid lipids are generally depleted in D relative to acetogenic (n-alkyl) lipids. Therefore, it is
likely that the isotopic difference between n-alkanes
and diterpenoids may reflect biosynthetic control of
initial values of dD in living organisms that are preserved in the sediments (Sauer et al., 2001).
Hydrogen isotopic variations can also occur
between different diterpenoid compounds (Table 2
and Fig. 3). Values of dD varied from 292& to
305& for norpimarane and from 296& to
319& for pimarane. The difference in dD (3–
37&) is insignificant between pimarane and norpimarane. This result is expected since previous study
has proved that both norpimarane and pimarane
are derived from same starting precursor pimaric
acid (Simoneit, 1986; Simoneit et al., 1986; Tuo
and Philp, 2005).
However, simonellite is 120–150& enriched in D
relative to norpimarane or pimarane and has dD
values ranging from 152& to 198&. These values indicate that simonellite has a different source
of hydrogen than norpimarane or pimarane.
Dehydroabietane is about 12–82& depleted in D
relative to simonellite in the Xiao 13 and Xiao 12
samples but about 33& enriched in D relative to
simonellite in the Ou 15 sample (Table 2). Values
of dD are also more variable for dehydroabietane
than for simonellite, ranging from
143& to
234& in the analyzed samples. The diterpenoid
structures encountered in this study may be derived
diagenetically from tricyclic precursors based on the
abietane and pimarane skeletons (Wakeham et al.,
1980; Simoneit, 1986; Simoneit et al., 1986; Ellis
Samples
Xiao13(1531)
Xiao13(1535)
Xiao12(1637)
Xiao12(1642)
Ou 15(1767)
-200
-250
-300
-350
Norpimarane
Pimarane
Simonellite
Dehydroabietane
16α(H)-
Phyllocladane
Fig. 3. Hydrogen-isotope ratios of diterpenoids.
172
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
et al., 1996; Otto et al., 1997; Otto and Simoneit,
2001; Otto et al., 2005; Tuo and Philp, 2005). Both
simonellite and dehydroabietane may be derived
from the same starting precursor, abietic acid. During the formation of tricyclic aromatic hydrocarbons from abietane-type diterpenoid precursors,
however, dehydroabietane is usually present at a relatively earlier diagenetic stage than simonellite
(Wakeham et al., 1980; Simoneit, 1986; Simoneit
et al., 1986; Ellis et al., 1996; Li et al., 1990; Otto
et al., 1997; Otto and Simoneit, 2001; Tuo and
Philp, 2005). If that is the case, our results of hydrogen isotopes for simonellite and dehydroabietane
suggest that a D-enrichment process occurred during diagenesis of the diterpenoids with later-derived
byproducts such as simonellite being enriched in D
relative to the earlier-derived byproduct such as
dehydroabietane. Dehydrogenation may be a major
pathway for such enrichment, which preferentially
removes light hydrogen as diagenesis progresses.
Norpimaranes are slightly enriched in D relative to
pimaranes, which is probably the result of decarboxylation or dehydrogenation from a common
precursor during diagenesis.
The dD for 16a(H)-phyllocladane varies narrowly from 259& to 284&. These values are
about 10–20& enriched in D relative to norpimarane and pimarane but about 50–100& depleted in
D relative to simonellite and dehydroabietane. Such
a comparison also indicates a different source for
these compounds. Tetracyclic diterpenoids such as
phyllocladane are primarily derived from conifers
(Alexander et al., 1987), including Podocarpaceae,
Araucariaceae, Cupressaceae (Karrer, 1958; Aplin
et al., 1963; Erdtman and Norin, 1966; Karrer
et al., 1977; Sukh Dev, 1989; Otto et al., 1997), Taxodiaceae (Otto et al., 1997; Bechtel et al., 2002), and
probably pteridophytes (Cheng et al., 1997). So a
predominance of diterpenoid hydrocarbons with
phyllocladane as the major peak suggests that the
organic material in five samples originates primarily
from higher plants and more specifically, conifers
(Tuo et al., 2003; Tuo and Philp, 2005).
Phyllocladane has a more limited geological distribution than the tricyclic diterpenoids and is found
mostly in Devonian (Sheng et al., 1991; Cheng et al.,
1997), Carboniferous (Grantham et al., 1983; Schulze and Michaelis, 1990) and Tertiary (Dai and Mei,
1988) deposits. The phyllocladanes have fewer precursors compared to the tricyclic diterpenoids (Otto
and Wilde, 2001; Otto et al., 2005), which may be a
reason that phyllocladanes have a relatively con-
strained distribution in their hydrogen-isotopic
compositions. On the contrary, the broader relative
precursor input to the tricyclic diterpenoids (Otto
and Wilde, 2001; Otto et al., 2005) may result in a
slightly greater variation of their hydrogen-isotopic
compositions. This finding is consistent with the
carbon-isotopic compositions measured for the
same compounds in the same samples (Tuo et al.,
2003).
It cannot be excluded that the hydrogen-isotope
signature has been altered to some extent by diagenetic transformation in the subsurface (Dawson
et al., 2004, 2005), even though these sediments
are relatively immature. It has been shown that
hydrogen exchange occurs between organic matter
and water during hydrous pyrolysis (Alexander
et al., 1982, 1983; Schimmelmann et al., 1999,
2004; Leif and Simoneit, 2000). Consequently, the
dD values of diterpenoids in our samples might have
been affected to an unknown extent by hydrogen
exchange reactions between organic matter and formation water during maturation (Dawson et al.,
2004, 2005). But the results in this study suggest that
hydrogen exchange processes have not significantly
altered the indigenous isotopic signatures. Furthermore, different compounds can have different
hydrogen isotopes even after extensive D-exchange
(Dawson et al., 2004, 2005). Theoretical calculations
also indicate that, for polyisoprenoid compounds
(such as steranes) in immature sediments, the D/H
ratio imparted by biosynthesis is largely preserved
in spite of significant structural change (Sessions
et al., 2004).
3.4. Application of lipid-hydrogen isotopes in
geological studies
Figure 4 shows the relationship between dD and
d13C for the lipid compounds in the same samples
(details of carbon-isotope analysis were reported
in Tuo et al., 2003). The plot includes average values
of structurally-similar compound classes (e.g., n-alkanes) as well as individual compounds (pimarane,
norpimarane, 16a(H)-phyllocladane, dehydroabietane, and simonellite). The dD and d13C are distinct
among n-alkanes, dehydroabietane and simonellite,
and pimarane, norpimarane and 16a(H)-phyllocladane (Fig. 4). The differences in dD and d13C among
these groups are likely due to isotope effects associated with different biosynthetic pathways within the
organisms that synthesized these molecules. This
conclusion is based on the following reasons:
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
173
-21
Simonellite
-22
16α(H)-Phyllocladane
-24
Pimarane
-25
Norpimarane
-26
Dehyaroabietane
13
δ C(‰, PDB)
-23
-27
-28
-29
-30
-350
Xiao13 (1531)
Xiao13 (1535)
Xiao12 (1637)
Xiao12 (1642)
Ou15 (1767)
-300
n-Alkanes(average)
-250
-200
-150
-100
δ D(‰, SMOW)
Fig. 4. dD versus d13C of the same lipid compounds (values of d13C for the lipid compounds are from Tuo et al., 2003).
(1) the samples were formed in a similar depositional environment; (2) the biomarkers were mainly
derived from terrestrial higher plants; and (3) the
organic matter in all the samples is at a similar thermal evolution stage (Table 1) (Tuo et al., 2003; Tuo
and Philp, 2005).
It has been a concern that hydrogen-isotope
exchange occurs between organic molecules preserved in sedimentary rock and the surrounding
environment during diagenesis over an extended
geological time scale (Andsersen et al., 2001; Sessions, 2001). A systematic study of compound-specific hydrogen isotopes of major genetic oil
families in the Western Canada Sedimentary Basin
has shown that oils derived from source rocks of
Cambrian age still retain a strong signature of the
hydrogen isotopic compositions of source organic
matter and source water (Li et al., 2001). Based on
their study, Li et al. (2001) proposed that dD values
are useful for oil-source correlation and for paleoenvironmental reconstructions. A study on the lipidhydrogen isotope ratios in Miocene lacustrine
sediments and plant fossils in Clarkia has also
proven that an extensive H-exchange between environmental waters and fossils is unlikely and the lipid
hydrogen isotopes can be used for paleoecological
studies over an extended geological time period
(Yang and Huang, 2003). Xiong et al. (2005)
characterized the hydrogen-isotopic compositions
of individual n-alkanes in terrestrial source rocks.
They observed a trend of depletion in dD for n-alkanes from saline water, to freshwater paralic lacustrine, and to swamp deposits. Our results indicate
that ancient fossil material preserved in sedimentary
deposits can retain large variations in both hydrogen- and carbon-isotope compositions between different lipid biomarkers, which can be applied to
assess the origin of the organic matter, for oil-source
correlation and for paleoenvironmental reconstructions at the geological time scale.
4. Conclusions
Our study on hydrogen-isotope compositions of
aliphatic and diterpenoid hydrocarbons has the following conclusions:
1. The n-alkanes exhibit restricted ranges in dD values, which varied from 150& to 220&. No
significant difference existed in dD values among
the n-alkanes from the analyzed samples.
2. Pristane was 34–69& depleted in D relative to
phytane; the variation of dD values for both pristane and phytane follow a similar trend and
appear as parallel lines among the analyzed samples. The reason for the different dD values
between pristane and phytane may be attributed
to either different origins of pristane and phytane
or different isotopic effects for their derivation
from a common phytol precursor.
3. Values of dD measured for diterpenoid compounds are 49–81& depleted in D relative to nalkanes. Hydrogen isotopic variations also occur
between different diterpenoid compounds, indicating a different source for these compounds.
Based on the comparison of dD values of the
174
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
different classes of diterpenoid compounds,
which may be derived from the same biosynthetic
precursors, a D-enrichment process (especially
dehydrogenation) will be expected when a compound is diagenetically altered from a natural
precursor structure to a geological structure.
4. The dD and d13C were distinct between structurally different compound classes. This observation
indicates that hydrogen-exchange processes have
not altered the indigenous isotopic signature of
aliphatic and diterpenoid hydrocarbons to a large
extent, although possible hydrogen exchange cannot be completely excluded during maturation.
These observations suggest that hydrogen isotopes of lipids can be applied to assess the origin
of the organic matter, oil-source correlation, and
depositional environments in the geological past.
Acknowledgements
This research was supported by the China National Natural Science Foundation (Grant Number
40073022), the China 973 National Key Research
and Development Program (Grant Number:
2003CB214606), the American Chemical Society
Petroleum Research Fund (CLZ), and the US
Department of Energy Financial Assistance Award
DE-FC09-96SR18546 to the University of Georgia
Research Foundation (CLZ). We thank Dr. Maowen Li and an anonymous reviewer for valuable
comments that improved an earlier version of the
manuscript. Drs. S. Schouten and L.R. Snowdon
are acknowledged for their constructive comments
and for handling our paper in the review and publication process.
Associate Editor—Stefan Schouten
References
Alexander, G., Hazai, I., Grimalt, J.O., Albaiges, J., 1987.
Occurrence and transformation of phyllocladanes in brown
coals from Norgad Basin, Hungary. Geochimica et Cosmochimica Acta 51, 2065–2073.
Alexander, G., Kagi, R.I., Larcher, A.V., 1982. Clay catalysis of
aromatic hydrogen-exchange reactions. Geochimica et Cosmochimica Acta 46, 219–222.
Alexander, G., Kagi, R.I., Larcher, A.V., Woodhouse,
G.W., 1983. Aromatic hydrogen exchange in petroleum
source rocks. In: Bjorøy, M., Albrecht, P., Cornford, C.,
de Groot, K., Eglinton, G., Galimov, E., Leythaeuser,
D., Pelet, R., Rullko¨tter, J., Speers, G. (Eds.), In
Advances in Organic Geochemistry 1981. Wiley, New
York, pp. 69–71.
Andsersen, N., Paul, H.A., Bernasconi, S.M., McKenzie, J.A.,
Behrens, A., Schaeffer, P., Albrecht, P., 2001. Large and rapid
climate variability during the Messinian salinity crisis:
Evidence from deuterium concentrations of individual biomarkers. Geology 29, 799–802.
Aplin, R.T., Cambie, R.C., Rutledge, P.S., 1963. The taxonomic
distribution of some diterpene hydrocarbons. Phytochemistry
2, 205–214.
Baas, M., Pancost, R., van Geel, B., Sinninghe Damste`, J.S.,
2000. A comparative study of lipids in Sphagnum species.
Organic Geochemistry 31, 535–541.
Bechtel, A., Sachsenhofer, R.F., Gratzer, R., Lu¨cke, A., Pu¨ttmann, W., 2002. Parameters determining the carbon isotopic
composition of coal and fossil wood in the Early Miocene
Oberdorf lignite seam (Styrian Basin, Austria). Organic
Geochemistry 33, 1001–1024.
Bigeleisen, J., 1965. Chemistry of isotopes. Science 147, 463–471.
Blumer, M., Thomas, D.W., 1965. ÔZameneÕ, isomeric C19
monoolefins from marine zooplankton, fishes and mammals.
Science 148, 370–371.
Blumer, M., Mullin, M.M., Thomas, D.W., 1963. Pristane in
zooplankton. Science 140, 974.
Blumer, M., Robertson, J.C., Gordon, J.E., Sass, J., 1969.
Phytol-derived C19 di- and triolefinic hydrocarbons in marine
zooplankton and fishes. Biochemistry 8, 4067–4074.
Castillo, J.B.D., Brooks, C.J.W., Cambie, R.C., Eglinton, G.,
1967. The taxonomic distribution of some hydrocarbons in
gymnosperms. Phytochemistry 6, 391–398.
Cheng, D.S., Han, D.X., Sheng, G.Y., 1997. The biomarker
composition and origin of tetracyclic diterpanes in Devonian
coal, China. Petroleum Exploration and Development 24, 42–
46.
Chikaraishi, Y., Naraoka, H., 2003. Compound-specific dD–d13C
analyses of n-alkanes extracted from terrestrial and aquatic
plants. Phytochemistry 63, 361–371.
Collister, J.W., Rieley, G., Stern, B., Eglinton, G., Fry, B., 1994.
Compound-specific d13C analyses of leaf lipids from plants
with differing carbon dioxide metabolisms. Organic Geochemistry 21, 619–627.
Dai, Q.L., Mei, B.W., 1988. Diagenetic products and thermotransformation of resinous diterpenoids. Oil and Gas Geology
9, 115–124.
Dawson, D., Grice, K., Alexander, R., 2005. Effect of
maturation on the indigenous dD signatures of individual
hydrocarbons in sediments and crude oils from the Perth
Basin (Western Australia). Organic Geochemistry 36(1),
95–104.
Dawson, D., Grice, K., Wang, S.X., Alexander, R., Radke, J.,
2004. Stable hydrogen isotopic composition of hydrocarbons
in torbanites (Late Carboniferous to Late Permian) deposited
under various climatic conditions. Organic Geochemistry 35,
189–197.
Ellis, L., Singh, R.K., Alexander, R., Kagi, R.I., 1996. Formation
of isohexyl alkylaromatic hydrocarbons from aromatizationrearrangement of terpenoids in the sedimentary environment:
A new class of biomarker. Geochimica et Cosmochimica Acta
60, 4747–4763.
Erdtman, H., Norin, T., 1966. The chemistry of the order
Cupressales. Progress in the Chemistry of Organic Natural
Products 24, 207–287.
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
Estep, M.F., Hoering, T.C., 1980. Biogeochemistry of the stable
hydrogen isotopes. Geochimica et Cosmochimica Acta 44,
1197–1206.
Ficken, J.J., Li, B., Swain, D.L., Eglinton, G., 2000. An n-alkanes
proxy for the sedimentary input of submerged floating
freshwater aquatic macrophytes. Organic Geochemistry 31,
745–749.
Gelpi, E., Schneider, H., Mann, J., Oro`, J., 1970. Hydrocarbons
of geochemical significance in microscopic algae. Phytochemistry 9, 603–612.
Goossens, H., de Leeuw, J.W., Schenck, P.A., Brassell, S.C.,
1984. Tocopherols as likely precursors of pristane in ancient
sediment and crude oils. Nature 312, 440–442.
Grantham, P.J., Posthuma, J., Baak, A., 1983. Triterpanes in a
number of far Eastern crude oils. In: Bjorøy, M. et al. (Eds.),
Advances in Organic Geochemistry. Wiley, Chichester, pp.
675–683.
Han, B.J., McCarthy, E.D., van Hoeven, W., Calvin, M.,
Bradley, W.H., 1968. Organic geochemical studies, II. A
preliminary report on the distribution of aliphatic hydrocarbons in algae, in bacteria, and in a recent lake sediment.
Proceedings of the National Academy of Sciences 59, 29–33.
Huang, H., Bowler, B.F.J., Zhang, Z., Oldenburg, T.B.P., Larter,
S.R., 2003. Influence of biodegradation on carbazole and
benzocarbazole distributions in oil columns from the Liaohe
basin, NE China. Organic Geochemistry 34, 951–9699.
Huang, Y., Shuman, B., Wang, Y., Webb, T., 2002. Hydrogen
isotope ratios of palmitic acid in lacustrine sediments record
late quaternary climate variations. Geology 30, 1103–1106.
Karrer, W., 1958. Konstitution und Vorkommen der organischen
Pflanzenstoffe (exclusive Alkaloide). Birkha¨user, Basel.
Karrer, W., Cherbuliez, E., Eugster, C.H., 1977. Konstitution,
und Vorkommen der organischen Pflanzenstoffe (exclusive
Alkaloide)Erga¨nzungsband, vol. 1. Birkha¨user, Basel.
Koopmans, M.P., Rijpstra, W.I.C., Klapwijk, M.M., de Leeuw,
J.W., Lewan, M.D., Sinninghe Damste´, J.S., 1999. A thermal
and chemical degradation approach to decipher pristane and
phytane precursors in sedimentary organic matter. Organic
Geochemistry 30, 1089–1104.
Leif, R.N., Simoneit, R.T., 2000. The role of alkenes produced
during hydrous pyrolysis of a shales. Organic Geochemistry
31, 1189–1208.
Li, M., Huang, Y., Obermajer, M., Jiang, C., Snowdon, L.R.,
Fowler, M.G., 2001. Hydrogen isotopic compositions of
individual alkanes as a new approach to petroleum correlation: case studies from the Western Canada Sedimentary
Basin. Organic Geochemistry 32, 1387–1399.
Li, M., Johns, R.B., Mei, B.W., 1990. A study in early diagenesis:
biomarker composition of a suite of immature coals and coaly
shales. Organic Geochemistry 16, 1067–1075.
Li, M., Larter, S.R., Taylor, P., Jones, D.M., Bowler, B., Bjorøy,
M., 1995. Biomarkers or not biomarkers? A new hypothesis
for the origin of pristane involving derivation from methyltrimethyltridecylchromans (MTTCs) formed during diagenesis from chlorophyll and alkylphenols. Organic Geochemistry
23, 159–167.
Liu, W., Huang, Y., 2005. Compound specific D/H ratios and
molecular distributions of higher plant leaf waxes as novel
paleoenvironmental indicators in the Chinese Loess Plateau.
Organic Geochemistry 36, 851–860.
Navale, V., 1994. Comparative study of low and high temperature hydrous pyrolysis products of monoglyceryl diether
175
lipid from archaebacteria. Journal of Analytical and Applied
Pyrolysis 29, 33–43.
Noble, R.A., Alexander, R., Kagi, R.I., Knox, J., 1985. Tetracyclic diterpenoid hydrocarbons in some Australian coals,
sediments and crude oils. Geochimica et Cosmochimica Acta
49, 2141–2147.
Otto, A., Simoneit, B.R.T., 2001. Chemosystematics and diagenesis of terpenoids in fossil conifer species and sediment from
the Eocene Zeitz Formation, Saxony, Germany. Geochimica
et Cosmochimica Acta 65, 3505–3527.
Otto, A., Simoneit, B.R.T., Rember, W.C., 2005. Conifer and
angiosperm biomarkers in clay sediments and fossil plants
from the Miocene Clarkia Formation, Idaho, USA. Organic
Geochemistry 36, 907–922.
Otto, A., Walther, H., Pu¨ttmann, W., 1997. Sesqui- and
diterpenoid biomarkers preserved in Taxodium rich Oligocene oxbow lake clays, Weisselster basin, Germany. Organic
Geochemistry 26, 105–115.
Otto, A., Wilde, V., 2001. Sesqui-, di-, and triterpenoids as
chemosystematic markers of extant conifers-a review. The
Botanical Review 67, 141–238.
Pancost, R.D., Boot, C.S., 2004. The palaeoclimatic utility of
terrestrial biomarkers in marine sediments. Marine Chemistry
92, 239–261.
Pancost, R.D., Sinninghe Damste´, J.S., 2003. Carbon isotopic compositions of prokaryotic lipids as tracers of
carbon cycling in diverse settings. Chemical Geology
195, 29–58.
Rieley, G., Collier, R.J., Jones, D.M., Eglinton, G., Eakin, P.A.,
Fallick, A.E., 1991. Source of sedimentary lipids deduced
from stable carbon-isotope analyses of individual compounds.
Nature 352, 425–427.
Risatti, J.B., Rowland, S.J., Yon, D.A., Maxwell, J.R., 1984.
Stereochemical studies of acyclic isoprenoids. XII. Lipids of
methanogenic bacteria and possible contributions to sediment. In: Schenck, P.A., de Leeuw, J.W., Lijmbach, G.W.M.
(Eds.), Advances in Organic Geochemistry 1983, Organic
Geochemistry, vol. 6, pp. 93–104.
Rowland, S.J., 1990. Production of acyclic isoprenoid hydrocarbons by laboratory maturation of methanogenic bacteria.
Organic Geochemistry 15, 9–16.
Sachse, D., Radke, J., Gaupp, R., Schwark, L., Lu¨niger, G.,
Gleixner, G., 2004a. Reconstruction of palaeohydrological
conditions in a lagoon during the 2nd Zechstein cycle through
simultaneous use of dD values of individual n-alkanes and
d18O and d13C values of carbons. International Journal of
Earth Sciences 93, 554–564.
Sachse, D., Radke, J., Gleixner, G., 2004b. Hydrogen isotope
ratios of recent lacustrine sedimentary n-alkanes record
modern climate variability. Geochimica et Cosmochimica
Acta 68, 4877–4889.
Sauer, P.T., Eglinton, T.I., Hayes, J.M., Schimmelmann, A.,
Sessions, A.L., 2001. Compound-specific D/H ratios of lipid
biomarkers from sediments as a proxy for environmental and
climatic conditions. Geochimica et Cosmochimica Acta 65,
213–222.
Schimmelmann, A., Lewan, M.D., Wintsch, R.P., 1999. D/H
isotope ratios of kerogen, bitumen, oil, and water in hydrous
pyrolysis of source rocks containing kerogen types I, II, IIS,
and III. Geochimica et Cosmochimica Acta 63, 3751–3766.
Schimmelmann, A., Sessions, A.L., Boreham, C.J., Edwards,
D.S., Logan, G.A., Summons, R.E., 2004. D/H ratios in
176
J. Tuo et al. / Organic Geochemistry 37 (2006) 165–176
terrestrially sourced petroleum systems. Organic Geochemistry 35, 1169–1195.
Schouten, S., Rijpstra, W.I.C., Kok, M., Hopmans, E.C.,
Summons, R.E., Volkman, J.K., Sinninghe Damste´, J.S.,
2001a. Molecular organic tracers of biogeochemical processes
in a saline meromictic lake (Ace Lake). Geochimica et
Cosmochimica Acta 65, 1629–1640.
Schouten, S., Wakeham, S.G., Sinninghe Damste´, J.S., 2001b.
Evidence for anaerobic methane oxidation by archaea in
euxinic water of the Black Sea. Organic Geochemistry 32,
1277–1281.
Schouten, S., Wakeham, S.G., Hopmans, E.C., Sinninghe
Damste´, J.S., 2003. Biogeochemical evidence that thermophilic archaea mediate the anaerobic oxidation of methane.
Applied and Environmental Microbiology 69, 1680–1686.
Schulze, T., Michaelis, W., 1990. Structure and origin of
terpenoid hydrocarbons in some German coals. Organic
Geochemistry 16, 1051–1058.
Sessions, A.L., 2001. Hydrogen Isotope Ratios of Individual
Organic Compounds. PhD dissertation, Indiana University,
Bloomington, Indiana, USA, p. 149.
Sessions, A.L., Burgoyne, T.W., Schimmelmann, A., Hayes,
J.M., 1999. Fraction of hydrogen isotopes in lipids biosynthesis. Organic Geochemistry 30, 1193–1200.
Sessions, A.L., Sylva, S.P., Summons, R.E., Hayes, J.M., 2004.
Isotopic exchange of carbon-bound hydrogen over geologic
timescales. Geochimica et Cosmochimica Acta 68, 1545–1559.
Sheng, G.Y., Fu, J.M., Liu, D.H., Chen, X.J., Simoneit, B.R.T.,
Han, D.X., 1991. Tetracyclic diterpenoid enriched in Devonian cutinite coal. Oil and Gas Geology 12, 108–116.
Simoneit, B.R.T., 1986. Cyclic terpenoids of the geosphere. In:
Johns, R.B. (Ed.), Biological Markers in the Sedimentary
Record. Elsevier, Amsterdam, pp. 43–99.
Simoneit, B.R.T., Grimalt, J.O., Wang, T.G., Cox, R.E.,
Hatcher, P.G., Nissenbaum, A., 1986. Cyclic terpenoids of
contemporary resinous plant detritus and of fossil woods,
ambers and coal. Organic Geochemistry 10, 877–889.
Sinninghe Damste´, J.S., Rijpstra, W.I.C., Hopmans, E.C., Prahl,
F.G., Wakeham, S.G., Schouten, S., 2002. Distribution of
membrane lipids of planktonic Crenarchaeota in the Arabian
Sea. Applied and Environmental Microbiology 68, 2997–
3002.
Sternberg, L., 1988. D/H ratios of environmental water recorded
by D/H ratios of plant lipids. Nature 333, 59–61.
Dev, S., 1989. Terpenoids. In: Rowe, J.W. (Ed.), Natural
Products of Woody plants, vol. 1. Springer, Berlin, pp. 691–
807.
Sun, Y., Chen, Z., Xu, S., Cai, P., 2005. Stable carbon and
hydrogen isotopic fractionation of individual n-alkanes
accompanying biodegradation: evidence from a group of
progressively biodegraded oils. Organic Geochemistry 36,
225–238.
Tuo, J., Philp, R.P., 2005. Saturated and aromatic diterpenoids
and triterpenoids in Eocene coals and mudstones from China.
Applied Geochemistry 20, 367–381.
Tuo, J., Wang, X., Chen, J., Simoneit, B.R.T., 2003. Aliphatic
and diterpenoid hydrocarbons and their individual carbon
isotope compositions in coals from the Liaohe Basin, China.
Organic Geochemistry 34, 1615–1625.
Van der Meer, M.T.J., Schouten, S., Sinninghe Damste´, J.S., de
Leeuw, J.W., Ward, D.M., 2003. Compound-specific isotopic
fractionation patterns suggest different carbon metabolisms
among Chloroflexus-like bacteria in hot-spring microbial mats.
Applied and Environmental Microbiology 69, 6000–6006.
Volkman, J.K., Barrett, S.M., Blackburn, S.I., Mansour, M.P.,
Sikes, E.L., Gelin, F., 1998. Microalgal biomarkers: a review
of recent research developments. Organic Geochemistry 29,
1163–1179.
Wakeham, S.G., Schaffner, C., Giger, W., 1980. Polycyclic
aromatic hydrocarbons in recent lake sediments-II. Compounds derived from biogenic precursors during early diagenesis. Geochimica et Cosmochimica Acta 44, 415–442.
Xie, X., Nott, C.J., Avsejs, L.A., Volders, F., Maddy, D.,
Chambers, F.M., Gledhill, A., Carter, J.F., Evershed, R.P.,
2000. Paleoclimate records in compound-specific dD values of
lipid biomarker in ombrotrophic peat. Organic Geochemistry
31, 1053–1057.
Xiong, Y., Geng, A., Pan, C., Liu, D., Peng, P., 2005.
Characterization of hydrogen isotopic composition of individual n-alkanes in terrestrial source rocks. Applied Geochemistry 20, 455–463.
Yakir, D., DeNiro, M.J., 1990. Oxygen and hydrogen isotope
fractionation during cellulose metabolism in lemna gibba L.
Plant Physiology 93, 325–332.
Yang, H., Huang, Y.S., 2003. Preservation of lipid hydrogen
isotope ratios in Miocene lacustrine sediments and plant
fossils at Clarkia, north Idaho, USA. Organic Geochemistry
34, 413–423.
Yapp, C.J., Epstein, S., 1982. Climatic significance of the
hydrogen isotope ratios in tree cellulose. Nature 297, 636–639.
Zhang, Z., Zhao, M., Yang, X., Wang, S., Jiang, X., Oldfield, F.,
Englinton, G., 2004. A hydrocarbon biomarker record for the
last 40 kyr of plant input to Lake Heqing, southwestern
China. Organic Geochemistry 35, 595–613.