You are on page 1of 11

Organic

Geochemistry
Organic Geochemistry 37 (2006) 787–797
www.elsevier.com/locate/orggeochem

25-Norhopanes: Formation during biodegradation of


petroleum in the subsurface
a,*
Barry Bennett , Milovan Fustic a, Paul Farrimond b, Haiping Huang a,
Stephen R. Larter a
a
Petroleum Reservoir Group (PRG), Department of Geology and Geophysics, University of Calgary, 2500 University Drive NW,
Calgary, Alberta, Canada T2N 1N4
b
Integrated Geochemical Interpretation Limited, Hallsannery, Bideford, Devon EX39 5HE, UK

Received 3 January 2006; received in revised form 1 March 2006; accepted 13 March 2006
Available online 6 June 2006

Abstract

Quantitative data from petroleum systems in China (Liaohe basin) and Canada (Athabasca tar sands) support the the-
ory that 25-norhopanes are produced during biodegradation of petroleum in the subsurface. Within a single oil column,
both case histories showed increasing severity of degradation, indicated by destruction of hopanes and production of 25-
norhopanes downward to the oil–water contact. In the Athabasca samples between the [Peters, K.E., Moldowan, J.M.,
1993. The Biomarker Guide: Interpreting Molecular Fossils in Petroleum and Ancient Sediments. Prentice Hall, Engle-
wood Cliffs, New Jersey, p. 363] scale of biodegradation levels 5–9, concentrations of C28 20S triaromatic steroids and
other biodegradation-resistant compounds increased by 35%, reflecting a concentration effect as a consequence of removal
of more degradable compounds. Over the same interval, the concentrations of C28 17a 25-norhopane and C29 17a 25-
norhopane increased by an order of magnitude, thus requiring that the balance be met by their net production during
degradation.
A detailed molecular investigation of the Athabasca bitumen revealed that C30 17a hopane degrades faster than C29 17a
hopane, whilst the rate of formation of both C29 17a 25-norhopane and C28 17a 25-norhopane are similar, complicating a
straightforward interpretation of demethylation of hopanes to form 25-norhopanes. Hopane degradation in the Athabasca
tar sand may also occur without the production of 25-norhopanes. The results show that even within a single petroleum accu-
mulation, a number of mechanisms control changes in the abundance and composition of hopanes and 25-norhopanes.
 2006 Elsevier Ltd. All rights reserved.

1. Introduction degradation. However, despite this useful character-


istic, the origin of 25-norhopanes remains a
The presence of 25-norhopanes in crude oils is controversial subject. Three hypotheses prevail for
commonly recognised as an indicator of severe bio- the origin of 25-norhopanes in biodegraded oils:
(i) microbial demethylation of hopanes to 25-norho-
* panes through the removal of the methyl group at
Corresponding author. Tel.: +1 403 210 3916; fax: +1 403
2840074. C-10 in the hopane nucleus (Seifert and Moldowan,
E-mail address: bennettb@ucalgary.ca (B. Bennett). 1979; Peters and Moldowan, 1991; Moldowan and

0146-6380/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2006.03.003
788 B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797

McCaffrey, 1995), (ii) a relative enrichment of 25- taining n-alkanes, isoprenoid alkanes, and 25-
norhopanes derived from source rocks and already norhopanes is a mixture of a heavily degraded oil
present in the nonbiodegraded oil as a result of their (palaeobiodegraded oil containing 25-norhopanes)
greater resistance to degradation (Chosson et al., and ‘‘fresh oil’’ containing n-alkanes (Volkman
1992) and (iii) the microbes responsible for severe et al., 1983; Philp, 1983; Howell et al., 1984; Blanc
petroleum degradation may produce 25-norhopanes and Connan, 1992; Peters and Moldowan, 1993;
without any direct relationship to degradation of Mason et al., 1995; Campos et al., 1996; Rooney
hopane (Peters et al., 2005). et al., 1998; López et al., 1998; George et al.,
The formation of 25-norhopanes through the 1998, 2002; Scotchman et al., 1998; Dzou et al.,
microbial removal of the methyl group at C-10 in 1999; Nascimento et al., 1999; Grimalt et al.,
the hopane nucleus during the biodegradation of 2002; Pan et al., 2003). Considering the number of
petroleum is well documented in the literature (Seif- studies that use 25-norhopanes as indicators of pal-
ert and Moldowan, 1979; Rullkötter and Wendisch, aeodegraded oil, it is important to define the true
1982; Volkman et al., 1983; Requejo and Halpern, significance and limitations of 25-norhopanes when
1989; Peters and Moldowan, 1991; Moldowan and identified in petroleum accumulations. In this paper,
McCaffrey, 1995; Peters et al., 1996; Nascimento we investigate the inter-relationship of 25-norho-
et al., 1999; Alberdi et al., 2001; Tocco and Alberdi, panes and hopanes in two petroleum systems using
2002), and evidence supporting microbially induced a quantitative mass balance approach. We show
demethylation of hopanes to the 25-norhopanes is that although hopanes decrease while 25-norho-
clear. In a continuous oil column in a diatomite res- panes increase during severe biodegradation, mass
ervoir in California, Moldowan and McCaffrey balance arguments show that the relationship of
(1995) were able to show a concomitant and inverse hopanes to 25-norhopanes does not appear to be a
change with depth in concentrations of C35 hopane straightforward conversion.
and C29 hopane and their C34 25-norhopane and
C28 25-norhopane counterparts. A mass balance 2. Methods
approach using nondegraded diasteranes as ‘‘inter-
nal standards’’ also demonstrated that the concen- 2.1. Samples
trations of 25-norhopanes in heavily degraded oils
were similar to the concentrations of hopanes in The sample suite consisted of oil samples from
related nonbiodegraded oils (Volkman et al., 1983; the Liaohe petroleum system of China and two
Peters and Moldowan, 1991). sets of previously frozen core samples from the
Evidence supporting the existence of 25-norho- Athabasca tar sand deposit of Alberta, Canada.
panes in source rocks is also in no doubt, although A description of the Liaohe petroleum system
the C28 and C29 demethylated hopanes are minor in and effects of biodegradation on the petroleum
comparison to C29 17a hopane (Chosson et al., accumulation is in Huang et al. (2003). Briefly,
1992) and their distributions are usually only repre- the Liaohe suite consists of eight oil samples from
sented by the C29 17a 25-norhopane with a few 1490.8 to 1645.5 m within a 160 m oil column in
other members of the series, rather than the full the heavily biodegraded petroleum accumulation
C29-C35 pseudo-homologous series typically found in the youngest member of the Eocene Shahejie
in degraded oils. They have been reported as free Formation (ES1). For the Athabasca study, core
hydrocarbons in source rocks from Western Austra- material was obtained from the Lower Cretaceous
lia, although their origin was attributed to a McMurray Formation sandstone reservoir in the
reworked component of organic matter deriving Athabasca tar sands from two different locations
from biodegraded seeps (Noble et al., 1985). 25- in a north–south direction. The main investigation
Norhopanes have also been found in supposedly was based on seven bitumen-rich core samples
nondegraded oils (Blanc and Connan, 1992), sup- (Athabasca-1 well) representing the depth interval
porting the potential explanation that their high rel- from 4.4 m to a water saturated sand below
ative abundance in some degraded oils is due to 90.4 m. A second suite of samples (Athabasca-2
removal of more labile components (Chosson well) consisted of three bitumen-rich cores col-
et al., 1992). lected from 486 to 488 m just above an oil–water
Despite conflicting opinions on the origin of 25- contact and above the regional unconformity with
norhopanes, it is widely accepted that crude oil con- Devonian carbonates.
B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797 789

2.2. Recovery of aliphatic hydrocarbons from oils and hydrocarbons versus the internal standard 5b(H)-
core extracts cholane was assumed to be 1. Corrections for mass
spectral responses were introduced during the quan-
Saturated hydrocarbons were isolated from oil tification of C28 17a 25-norhopane (m/z 177) and
and core extracts following the procedure described C30 17a hopane (m/z 191) to account for the fact
by Bennett and Larter (2000). The Athabasca cores that both the AB and DE ring fragments contribute
were tar sands with the bitumen readily extracted by to the peak areas in their respective mass chromato-
repeatedly washing the core with solvent (dichloro- grams. The ratios of the AB/DE ring fragments
methane; CH2Cl2) until the extracts were colourless. were determined from mass spectra of neighbouring
The core extracts were de-asphaltened to obtain a carbon number homologues. The relative abun-
hexane-soluble fraction that was amenable to the dance of the AB and DE ring fragments from the
following solid-phase extraction (SPE) procedure. C29 17a 25-norhopane (m/z 177/m/z 191), C30 17a
Typically, 100 mg of oil or 50 mg de-asphaltened 22S and 22R 25-norhopanes (m/z 177/m/z 205)
core extract was absorbed onto a pre-weighed C18 are similar, giving a ratio of 0.47. For the correction
nonendcapped (NEC) SPE cartridge. The hydrocar- of C30 17a hopane the AB/DE ratio of 0.7 was
bon-containing fraction was collected in n-hexane obtained from the relative response in the C29 17a
(5 ml). The solvent volume was reduced to 1/2 ml norhopane (m/z 191/m/z 177) and C31 17a 22S
under a gentle stream of nitrogen gas. The saturated and 22R homohopanes (m/z 191/m/z 205). These
hydrocarbons were isolated from the hydrocarbon response factors were determined in oil samples rel-
fraction using a silver nitrate impregnated silica atively enriched in the 25-norhopanes and hopanes,
(Ag + silica) SPE cartridge (Bennett and Larter, respectively, to avoid complications arising from co-
2000). An aliquot (50 ll from 1/2 ml) of the n-hex- elution (see Fig. 1b, c).
ane (hydrocarbon-containing) eluent recovered dur-
ing C18 NEC SPE was added dropwise onto the 3. Results and discussion
Ag + silica, and the saturated hydrocarbons were
eluted in 2 ml of n-hexane. Aromatic hydrocar- 3.1. Determining the levels of petroleum
bons were subsequently recovered in 4 ml of biodegradation and identification of 25-norhopanes
CH2Cl2. The aliphatic and aromatic hydrocarbons
were analysed by GC-MS in selected ion monitoring We focus our discussion on case studies from two
(SIM) mode. petroleum systems from China (Liaohe) and Wes-
tern Canada (Athabasca) where a range of biodeg-
2.3. Gas chromatography-mass spectrometry radation levels was been encountered within each
petroleum accumulation (e.g., Huang et al., 2004).
Mass spectral characterisation of compounds in The levels of petroleum degradation are usually
the hydrocarbon fractions was carried out using ranked according to the 10 point scale of Peters
gas chromatography-mass spectrometry (GC-MS) and Moldowan (1993), with Peters and Moldowan
on a Hewlett-Packard 5890 GC (using splitless level 1 (abbreviated to PM 1) representing slightly
injection) interfaced to a HP 5970B quadrupole biodegraded oil and Peters and Moldowan level 10
mass selective detector (electron input energy (abbreviated to PM 10) severely degraded oil
70 eV, source temperature 250 C). (hopanes and diasteranes absent, C26–C29 aromatic
The saturated hydrocarbons were analysed using steroids attacked). The presence of 25-norhopanes
a DB-5 fused silica capillary column (30 m · in biodegraded oils is usually an indication that an
0.32 mm i.d. · 0.25 lm film thickness; J&W Scien- accumulation has been heavily biodegraded (PM 6
tific). The oven temperature program was: 40 C or greater). In one of the most severely biodegraded
(initial time, 2 min) to 175 C at 10 C/min to oils we analysed, the hopanes were degraded so
225 C at 6 C/min to 300 C at 4 C/min and held extensively (equivalent to PM 8) that the dominant
at 300 C for 20 min. compound in the mass chromatogram typically
The internal standards 5b(H)-cholane and squa- employed to identify hopanes (m/z 191) is C29 17a
lane were added to the core sample or oil prior to 25-norhopane (Fig. 1a). The distribution of 25-
C18 NEC SPE. Peak area integration during GC- norhopanes in the same oil consists of a fully devel-
MS analysis was by MASS LAB. The relative oped regular series of C26–C34 25-norhopanes
response factor (RRF) for individual saturated including the demethylated counterparts of Ts and
790 B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797

29N
m/z 191 (a)

29H
30H

28N
m/z 177 29N (b)

30N
26N 26N 31N
(Ts) (Tm) 32N
33N 34N
Intensity

30H
m/z 191 (c)

29H

31H
Tm 29N 32H
Ts 33H
34H 35H

Retention time (min)

Fig. 1. Mass chromatograms (a) m/z 191, displaying degraded hopanes and (b) m/z 177, demethylated hopanes in a severely (PM 8)
biodegraded oil, and (c) m/z 191, from a moderately biodegraded oil (PM 4). Homologous series of 25-norhopanes and hopanes having
17a stereochemistry are denoted by the carbon number followed by the letter N and H, respectively. The two commonly occurring
trisnorhopanes are designated Ts and Tm, while the demethylated compounds are labelled as 26N.

Tm (Fig. 1b). Overall, the distribution of the California, USA (Peters and Moldowan, 1991;
25-norhopanes bears a strong resemblance to the Moldowan and McCaffrey, 1995) and during petro-
regular hopane series in a ‘‘less altered’’ oil sample leum degradation under laboratory aerobic condi-
(compare Fig. 1b, c; c.f., Volkman et al., 1983), if tions (Bost et al., 2001). In contrast, the extended
a correction is applied for the enhanced responses hopanes may be degraded prior to the regular
in the m/z 177 and m/z 191, respectively, for C28 hopanes during laboratory experiments (Goodwin
17a 25-norhopane and C30 17a hopane (Fig. 2). It et al., 1983; Chosson et al., 1992) and in some bio-
has been reported in some cases that demethylated degraded crude oils (Seifert and Moldowan, 1979;
Ts and Tm are not observed and that the extended Rullkötter and Wendisch, 1982; Seifert et al.,
C35-hopanes also show resistance during petroleum 1984). These differences in hopane and 25-norho-
degradation under surface weathering conditions pane characteristics imply that biodegradation
(Requejo and Halpern, 1989). These observations alteration of hopanes can proceed in numerous
have also been reported in biodegraded oils from ways.
B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797 791

2500
(a)
2000

1500
µg g-1 oil
1000

500

0
Tm

C35 22R
C29 17α

C33 22R
C31 22R

C32 22R

C34 22R

C35 22S
C31 22S

C32 22S

C33 22S

C34 22S
C30 17 α
Ts

2000
(b)

1500
µg g-1 oil

1000

500

C34 22R
C30 22R

C32 22R

C33 22R
Tm - C26

C30 22S

C32 22S

C33 22S

C34 22S
Ts - C26

C28 17α

C29 17α

C3122R
C3122S

Fig. 2. Histogram plot of the concentrations of (a) C27– C35 17a hopanes in a moderately degraded oil (PM 4, Fig. 1c) and (b) C26–C34 17a
25-norhopanes in a severely degraded oil (PM 8, Fig. 1b), following correction for the enhanced mass spectral contribution (see Section
2.3) from C28 17a 25-norhopane and C30 17a hopane.

3.2. A quantitative investigation into the relationship noids have been completely removed, and traces
of hopanes and 25-norhopanes of 25-norhopanes are present, indicating that the
oil is moderately biodegraded (PM 5), whereas at
3.2.1. Liaohe case study the base of the oil column, hopanes are being
We first describe the quantitative changes in the destroyed, indicating a petroleum degradation
hopanes and 25-norhopanes in a biodegraded level of PM 8. The behaviour of C30 17a hopane
petroleum accumulation that originated from and the presumed product C29 17a 25-norhopane
lacustrine source rocks in China. The sample suite are shown in Fig. 3a. Toward the base of the oil
was acquired from the Liaohe Basin and was part column C29 17a 25-norhopane gradually increases
of a detailed investigation of the changes in the in abundance, coinciding with the loss of C30
hydrocarbons and nitrogen compounds resulting 17a hopane; this is also depicted by the ratio
from varying degrees of biodegradation (Huang C29 17a 25-norhopane/C30 17a hopane versus
et al., 2003, 2004). The petroleum was generated depth (Fig. 3b). Cases where 25-norhopanes show
from deeply buried Eocene ES3 Formation source an inverse relationship with hopanes have been
rocks and accumulated in freshwater lacustrine recognised in other heavily degraded oilfields,
sediments of the Eocene ES3 and ES1 formations. including the Wabasca Field in Alberta, Canada
The bulk petroleum composition and molecular (Brooks et al., 1988), the Tia Juana Field in the
data show systematic biodegradation gradients Maracaibo Basin, Venezuela (Tocco and Alberdi,
within the oil column (160 m). At the top of the 2002) and the Cymric Field in California (Moldo-
ES1 oil column, the n-alkanes and acyclic isopre- wan and McCaffrey, 1995).
792 B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797

(a) (b)
Concentrations - µg/g oil C29 17α 25-Norhopane
C30 17α Hopane
0 5000 10000 0 500 1000 0 0.10 0.20 0.30
1450
C30 17α Hopane C29 17α 25-Norhopane

Depth (m) 1500

1550

1600

1650

Fig. 3. (a) Concentrations of C30 17a hopane and C29 17a 25-norhopane and (b) ratio of C29 17a 25-norhopane/C30 17a hopane versus
depth in the ES1 oil column. The oil–water contact is located at 1650 m.

A number of lines of evidence from the present would expect to find the inverse response of 25-
case history support the formation of 25-norho- norhopane and hopane abundance; i.e., where
panes in the reservoir. In terms of mass balance, extended hopanes are preserved, a lack of extended
the concentrations of C29 17a 25-norhopane range 25-norhopanes would be expected. The hopane and
from 30 lg/g oil within the top of the oil column 25-norhopane composition data from Athabasca
to 1400 lg/g oil toward the base. A concentration affords such a comparative investigation.
increase from 30 to 1400 lg C29 17a 25-norhopane Here we present quantitative data for hopanes
per gram oil without formation of the compound and 25-norhopanes in two wells from the Athabasca
would require that more than 90% of the oil was tar sands deposit to further investigate the inter-
consumed between biodegradation levels PM 5–8, relationship of hopanes and 25-norhopanes.
which seems unreasonable. Furthermore, in geneti- Bitumens extracted from the Athabasca-1 well
cally related petroleum from the deeper but less bio- samples show a range of biodegradation from PM
degraded (PM 1–4; Huang et al., 2003) ES3 5 in the main oil column increasing downward to
reservoir, no 25-norhopanes were detected, implying PM 9 at the oil–water contact, where even diaster-
that their presence in the more degraded ES1 reser- anes are attacked. Compositional changes have also
voir is most likely linked to their neoformation dur- affected the hopane distributions, where C30 17a
ing higher levels of degradation. hopane decreases more than C29 17a hopane
The decrease in hopane and corresponding (Fig. 4). This feature has also been observed in oil
increase in 25-norhopane abundance down hole degradation experiments conducted under aerobic
(Fig. 3a) is not a quantitative conversion (c.f., conditions in the laboratory (Watson et al., 2002)
Moldowan and McCaffrey, 1995); C30 17a hopane and in severely biodegraded oils from the South Bel-
loss represents 8,700 lg/g oil, whilst C29 17a 25- ridge Field, California (Walters, 1993). Based on the
norhopane increases by only 1060 lg/g oil over the changing hopane distributions observed in the m/z
depth interval 1561 to 1645 m. Thus, it appears that 191 mass chromatogram, one would expect to find
demethylation of hopane may represent only one of higher amounts of C29 17a 25-norhopane compared
a number of degradation reactions contributing to to C28 17a 25-norhopane in the m/z 177 mass chro-
hopane loss. matogram. For comparative purposes, the distribu-
tions of the 25-norhopanes as a function of depth
3.2.2. Athabasca case study are shown in Fig. 4. At 27.7 m, trace contributions
Requejo and Halpern (1989) showed that in some from 25-norhopanes may be discernable, but with
cases during biodegradation the extended (C31+) increasing depth toward the oil–water contact, the
hopanes may be preserved compared to the C29 25-norhopanes become the major components in
and C30 hopanes. The converse also applies in some the m/z 177 mass chromatogram (Fig. 4). Over the
cases (Seifert and Moldowan, 1979). If 25-norho- depth interval 79–90.4 m, the relative abundances
panes are formed directly from hopanes, then one of C28 and C29 17a 25-norhopanes are similar, with
B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797 793

m/z 191 30H m/z 177


27.7 m 27.7 m 29H (a)
29H
23T

Tm

35H 28N

m/z 191 m/z 177 29H


79 m 23T 29H 79 m (b)

29N

30N
28N s
* R+30H
Intensity

35H

m/z 191 m/z 177


90.4 m 90.4 m
23T 29H (c)

29N

28N

30N
s
R+30H
* 35H

Time

Fig. 4. Representative m/z 191 and m/z 177 mass fragmentograms of the saturated hydrocarbon fractions isolated from Athabasca-1 tar
sand samples (a) 27.7 m, (b) 79 m and (c) 90.4 m depth. Key to 191 mass fragmentograms: 23T, C23 tricyclic terpane; Tm, C27 17a-
22,29,30-trisnorhopane; 29H, C29 17a hopane; 30H, C30 17a hopane; 35H, C35 17 a (H) 22S and 22R homohopane, * = contribution due
to C29 17a (H) 25-norhopane. Key to m/z 177 mass fragmentograms: 28N, C28 17a 25-norhopane; 29N, C29 17a 25-norhopane; 30N, C30
17a 25-norhopane.

no apparent enhancement of the C29 17a 25-norho- the regional unconformity. The degradation of
pane to match the more rapid decrease in the C30 hopane suggested a biodegradation level of PM 8,
17a hopane; thus a direct relationship between although unlike the Athabasca-1 well samples, the
hopane and 25-norhopane appears ambiguous. steranes and diasteranes were intact. However, like
The quantitative data in Figs. 5 and 6 confirm that the Athabasca-1 well samples, greater removal of
the concentrations of 25-norhopanes increase while C30 17a hopane compared to C29 17a hopane was
the hopanes decrease. Conversion of hopanes to 25- observed toward the oil–water contact. The concen-
norhopanes would be an appropriate explanation of tration of C30 17a hopane fell from 2390 to 1040 lg/
these down-hole concentration trends. However, g extract compared to a slight increase from 2600 to
considering the data in more detail, the more rapid 2690 lg/g extract for C29 17a hopane. However,
removal of C30 17a hopane compared to C29 17a there were no detectable 25-norhopanes in Athaba-
hopane is not matched by greater production of sca-2 well. The subsurface environment represented
C29 17a 25-norhopane compared with C28 17a 25- by the Athabasca-2 well does not appear to support
norhopane. 25-norhopane formation, whereas in the Athabasca-
In a second Athabasca case study, three core 1 well 25-norhopanes are produced. The oil column
samples from the Athabasca-2 well were sampled in the Athabasca-1 well overlies a water leg that is
from just above the oil–water contact, lying above within the McMurray Formation sandstones,
794 B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797

µg /g extract whereas the oil column in Athabasca-2 well directly


overlies the Devonian the carbonates. Therefore,
0 500 1000 1500 2000 2500 differences in the local water chemistry may affect
10
the development of the microbial environment. Fur-
ther work is required in the Athabasca region with
supporting aquifer data to understand what envi-
20 ronmental factors control 25-norhopane formation.
In Eocene Wilcox oils from South Texas, the onset
of hopane degradation was indicated in the C29 and
C30 hopanes, but no 25-norhopanes were identified
40
(Williams et al., 1986). In studies of seep oils from
Depth (m)

Greece (Seifert et al., 1984), the Wessex Basin (Bigge


and Farrimond, 1998), and heavily degraded seep
oils from the Central Adriatic Basin (Moldowan
60
et al., 1992), no 25-norhopanes were identified even
though hopanes were degraded. Active seeps from
offshore West Africa and North West Scotland
80 (Wenger and Isaksen, 2002) showed extensive petro-
leum degradation without the formation of 25-
norhopanes. A suite of tar balls collected from
along the California coast was also devoid of 25-
100
norhopanes even though hopanes were degraded
Fig. 5. Plot of concentrations of C28 17a 25-norhopane (h), C29 (Hostettler et al., 2004).
17a 25-norhopane (D), C29 17a norhopane (j), and C30 17a Although 25-norhopanes have been reported in
hopane (m) versus depth (m) in an Athabasca-1 tar sand core.
source rocks and nondegraded oils (Blanc and Con-
nan, 1992) we present mass balance evidence that
requires the formation of 25-norhopanes to explain
the geochemical trends observed in the Athabasca
µg /g extract cores. We have quantified a suite of compounds that
0 50 100 150 200 250
are known to be highly resistant to petroleum degra-
0 dation: monoaromatic C30 8,14-secohopanes, C28
20S tri-aromatic steroids and C21aaa pregnane.
The concentrations of these ‘‘resistant’’ compounds
20
increase towards the oil–water contact (Fig. 6) con-
sistent with their relative enrichment as other more
easily degradable components are selectively
removed. On average, the concentrations of these
Depth (m)

40 resistant markers increase by 35% over the well


interval and behave in a consistent manner, suggest-
ing that they are responding to the same process. In
60 contrast, the concentrations of 25-norhopanes
increase by an order of magnitude, indicating that
25-norhopanes are being produced during degrada-
tion in the Athabasca petroleum system.
80
3.2.3. Hopanes and 25-norhopanes: not a
straightforward relationship!
100 Brooks et al. (1988) suggested that at least two
mechanisms of biodegradation affect the biomarkers
Fig. 6. Plot of concentrations of C28 17a 25-norhopane (h), C29
17a 25-norhopane (D), C28 triaromatic steroid (·), C21 aaa
of the Cretaceous oil sands/heavy oils of Alberta:
pregnane (), and mono-aromatic C30 8,14-secohopane (d) firstly in oils where regular steranes have been
versus depth (m) in Athabasca-1 tar sand core. removed, no 25-norhopanes were identified
B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797 795

(Athabasca tar sands); secondly, when regular ster- amongst the hopane and 25-norhopane data sets
anes were present, 25-norhopanes were identified presented herein and in the literature. In the Ath-
(Wabasca tar sands). Contrary to the findings of abasca-1 well, C30 17a hopane degrades faster than
Brooks et al. (1988) we have identified 25-norho- C29 17a hopane, while the rate of C28 and C29 17a
panes in some Athabasca samples coinciding with 25-norhopane generation are similar, one would
the degradation of steranes and diasteranes. And expect faster production of C29 17a 25-norhopane
in other Athabasca samples where the steranes if they originated by hopane degradation. In the
appear to be intact and C30 17a hopane was signif- Athabasca-2 well, hopane degradation occurs with-
icantly removed; no 25-norhopanes were identified. out formation of 25-norhopanes. The observed
In summary, detailed quantitative molecular variations require that hopane degradation and
investigations into the origin of 25-norhopanes 25-norhopane formation proceed along different
reveal more complexity in the apparent inter-rela- pathways that may operate in association with the
tionship of hopanes and 25-norhopanes. In the lit- many different combinations of environmental con-
erature, strong evidence for the conversion of ditions and microbial processes. The geochemical
hopanes to 25-norhopanes is shown by the inverse data presented here indicate that a number of mech-
relationship between hopane and 25-norhopane anisms of hopane degradation are probably operat-
abundance (Moldowan and McCaffrey, 1995). ing even within a single petroleum deposit.
Our study of the Liaohe and Athabasca petroleum
deposits confirms this relationship, but shows that 4. Conclusions
conversion is not quantitative. Furthermore, data
from the Athabasca case study suggest that the The presence of 25-norhopanes is commonly
degree of conversion may be different for individual used to indicate that a petroleum accumulation
carbon number species. The quantitative depth-cor- has experienced at least heavy (PM 6) biodegrada-
related relationship between hopanes and 25- tion. Quantitative geochemical data from case stud-
norhopanes observed by Moldowan and McCaffrey ies for the Liaohe Basin and Athabasca tar sands
(1995) in the Cymric Field (California, USA) led suggest the operation of more than one mechanism
those authors to suggest that an oxidised intermedi- for hopane degradation and 25-norhopane forma-
ate was unlikely to accumulate. In the case histories tion. The formation of 25-norhopanes in both of
presented here, a 1:1 quantitative relationship was these biodegraded petroleum systems has been
not observed, suggesting that the balance in the shown using quantitative biomarker data, but mass
quantitative relationship may have to be met by balance arguments show that only a proportion of
the formation of as yet unknown products that the degraded hopanes give rise to 25-norhopanes.
may or may not be intermediates in 25-norhopane In some Athabasca samples, C30 17a hopane
formation. Candidate intermediates include hopa- degrades to a greater extent than C29 17a hopane,
noid structures bearing functionalities at the C-25 but there is no concomitant increase in the genera-
carbon, but so far only a tentative identification tion of C29 17a 25-norhopane compared to C28
of 28,30-bisnorhopan-25-oic acid has been pro- 17a 25-norhopane. An added complication is that
posed (de Lemos Scofield, 1990). The occurrence in some other Athabasca samples where hopanes
of hopanoic acids and 25-norhopanoic acids in bio- are degrading, no 25-norhopanes could be identi-
degraded oil may offer parallel information to that fied. Thus, even within a single oil deposit it appears
obtained from hydrocarbon biomarkers (Nasci- that a number of mechanisms control the biodegra-
mento et al., 1999) and help shed light on the dation of hopanes, with or without the formation of
behaviour of hopanes and 25-norhopanes. How- 25-norhopanes.
ever, the hopanoic acids and 25-norhopanoic acids,
when present in biodegraded petroleums, occur in Acknowledgements
concentrations orders of magnitude less than their
hydrocarbon counterparts (Bennett et al., unpub- The authors thank the Bacchus I sponsors (BP,
lished), suggesting they represent minor products Chevron, ConocoPhillips, ExxonMobil, ENI E&P
or short-lived intermediates. Division, Enterprise Oil, JNOC, Norsk Hydro,
Attempting to understand the inter-relationship Petrobras, Shell, TotalFinaElf) for providing
between hopanes and 25-norhopanes, remains chal- financial support. Encana are thanked for provid-
lenging since there appears to be no consistent trend ing sample material from the Athabasca-2 well.
796 B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797

The Alberta Ingenuity Centre for In-Situ Energy is Goodwin, N.S., Park, P.J.D., Rawlinson, T., 1983. Crude oil
also thanked for support (Milovan Fustic, PhD biodegradation under simulated and natural conditions. In:
Bjorøy, M. et al. (Eds.), Advances in Organic Geochemistry
student). The authors acknowledge that much of 1981. Wiley, Chichester, pp. 650–658.
the work was carried out at the NRG (Newcastle Grimalt, J.O., Campos, P.G., Berdie, L., Lopez-Quintero, J.O.,
University) prior to re-locating to Calgary (Barry Navarrete-Reyes, L.E., 2002. Organic geochemistry of the oils
Bennett, Steve Larter, Huang Haiping) and Bide- from the southern geological province of Cuba. Applied
ford (Paul Farrimond). We are also grateful to Geochemistry 17, 1–10.
Hostettler, F.D., Rosenbauer, R.J., Lorenson, T.D., Dougherty,
Drs. Mike Moldowan and Ken Peters for con- J., 2004. Geochemical characterization of tarballs on beaches
structive reviews. along the California coast. Part I – Shallow seepage impacting
the Santa Barbara Channel Islands, Santa Cruz, Santa Rosa
Associate Editor—Kenneth E. Peters and San Miguel. Organic Geochemistry 35, 725–746.
Howell, V.J., Connan, J., Aldridge, A.K., 1984. Tentative
identification of demethylated tricyclic terpanes in nonbiode-
References graded and slightly biodegraded crude oils from the Los
Llanos Basin, Colombia. Organic Geochemistry 6, 83–92.
Alberdi, M., Moldowan, J.M., Peters, K.E., Dahl, J.E., 2001. Huang, H., Bowler, B.F.J., Zhang, Z., Oldenburg, T.B.P., Larter,
Stereoselective biodegradation of tricyclic terpanes in heavy S.R., 2003. Influence of biodegradation on carbazole and
oils from the Bolivar Coastal Fields, Venezuela. Organic benzocarbazole distributions in oil columns from the Liaohe
Geochemistry 32, 181–191. basin, N.E. China. Organic Geochemistry 34, 951–969.
Bennett, B., Larter, S.R., 2000. Quantitative separation of Huang, H., Larter, S.R., Bowler, B.F.J., Oldenburg, T.B.P.,
aliphatic and aromatic hydrocarbons using silver ion-silica 2004. A dynamic biodegradation model suggested by petro-
solid-phase extraction. Analytical Chemistry 72, 1039–1044. leum compositional gradients within reservoir columns from
Bigge, M.A., Farrimond, P., 1998. Biodegradation of seep oils in the Liaohe basin, NE China. Organic Geochemistry 35, 299–
the Wessex Basin – a complication for correlation. In: 316.
Underhill, J.R., (Ed.), Development, Evolution and Petro- de Lemos Scofield, A., 1990. Nouveaux marquers biologiques de
leum Geology of the Wessex Basin, Geological Society, sediments et petroles riches en soufre: indentification et mode
London, Special Publication 133, pp. 373–386. de formation. PhD thesis, L’Universite Louis Pasteur de
Blanc, P., Connan, J., 1992. Origin and occurrence of 25- Strasbourg, Strasbourg, France.
norhopanes: a statistical study. Organic Geochemistry 18, López, L., Mónaco, S.L., Richardson, M., 1998. Use of molec-
813–828. ular parameters and trace elements in oil–oil correlation
Bost, F.D., Frontera-Suau, R., McDonald, T.J., Peters, K.E., studies, Barinas sub-basin, Venezuela. Organic Geochemistry
Morris, P.J., 2001. Aerobic biodegradation of hopanes and 29, 613–629.
norhopanes in Venezuelan crude oils. Organic Geochemistry Mason, P.C., Burwood, R., Mycke B., 1995. The reservoir
32, 105–114. geochemistry and petroleum charging histories of Palaeogene-
Brooks, P.W., Fowler, M.G., Macqueen, R.W., 1988. Biological reservoired fields in the Outer Witch Ground Graben. In:
marker and conventional organic geochemistry of oil sands/ Cubitt, J.M. and England, W.A., (Eds.), The Geochemistry of
heavy oils, Western Canada Basin. Organic Geochemistry 12, Reservoirs, Special Publication 86, The Geological Society of
519–538. London, pp. 281–301.
Campos, P.G., Grimalt, J.O., Berdie, L., Lopez-Quintero, J.O., Moldowan, J.M., Lee, C.Y., Sundararaman, P., Salvatori, T.,
Navarrete-Reyes, L.E., 1996. Organic geochemistry of Cuban Alajbeg, A., Gjukic, B., Demaison, G.J., Slougui, N.-E.,
oils - I. The northern geological province. Organic Geochem- Watt, D.S., 1992. Source correlation and maturity assessment
istry 25, 475–488. of select oils and rocks from the Central Adriatic basin (Italy
Chosson, P., Connan, J., Dessort, D., Lanau, C., 1992. In vitro and Yugoslavia). In: Moldowan, J.M., Albrecht, P., Philp,
biodegradation of steranes and terpanes: a clue to under- R.P. (Eds.), Biological Markers in Sediments and Petroleum.
standing geological situations. In: Moldowan, J.M., Albrecht, Prentice Hall, pp. 370–401.
P., Philp, R.P. (Eds.), Biological Markers in Sediments and Moldowan, J.M., McCaffrey, M.A., 1995. A novel microbial
Petroleum. Prentice Hall, pp. 320–349. hydrocarbon degradation pathway revealed by hopane
Dzou, L.I., Holba, A.G., Ramon, J.C., Moldowan, J.M., demethylation in a petroleum reservoir. Geochimica et
Zinniker, D., 1999. Application of new diterpane biomarkers Cosmochimica Acta 59, 1891–1894.
to source, biodegradation and mixing effects on Central Nascimento, L.R., Rebouças, L.M.C., Koike, L., Resi, F.de
Llanos Basin oils, Colombia. Organic Geochemistry 30, 515– A.M., Soldan, A.L., Cerqueira, J.R., Marsaioli, A.J., 1999.
534. Acidic biomarkers from Albacora oils, Campos basin, Brazil.
George, S.C., Lisk, M., Summons, R.E., Quezada, R.A., 1998. Organic Geochemistry 30, 1175–1191.
Constraining the oil charge history of the South Pepper Noble, R., Alexander, R., Kagi, R.I., 1985. The occurrence of
oilfield from the analysis of oil-bearing fluid inclusions. bisnorhopane, trisnorhopane and 25-norhopanes as free
Organic Geochemistry 29, 631–648. hydrocarbons in some Australian shales. Organic Geochem-
George, S.C., Boreham, C.J., Minifie, S.A., Teerman, S.C., 2002. istry 8, 171–176.
The effect of minor to moderate biodegradation on C-5 to C-9 Pan, C.C., Yang, J.Q., Fu, J.M., Sheng, G., 2003. Molecular
hydrocarbons in crude oils. Organic Geochemistry 33, 1293– correlation of free oil and inclusion oil of reservoir rocks in
1317. the Junggar Basin, China. Organic Geochemistry 34, 357–374.
B. Bennett et al. / Organic Geochemistry 37 (2006) 787–797 797

Peters, K.E., Moldowan, J.M., 1991. Effects of source, thermal geochemical oil-source correlation study. Organic Geochem-
maturity, and biodegradation on the distribution and isom- istry 29, 671–700.
erization of homohopanes in petroleum. Organic Geochem- Seifert, W.K., Moldowan, J.M., Demaison, G.J., 1984. Source
istry 17, 47–61. correlation of biodegraded oils. Organic Geochemistry 6,
Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide: 633–643.
Interpreting Molecular Fossils in Petroleum and Ancient Seifert, W.K., Moldowan, J.M., 1979. The effect of biodegrada-
Sediments. Prentice Hall, Englewood Cliffs, New Jersey, p. tion on steranes and terpanes in crude oils. Geochimica et
363. Cosmochimica Acta 43, 111–126.
Peters, K.E., Moldowan, J.M., McCaffrey, M.A., Fago, F.J., Tocco, R., Alberdi, M., 2002. Organic geochemistry of heavy/
1996. Selective biodegradation of extended hopanes to 25- extra heavy oils from sidewall cores, Lower Lagunillas
norhopanes in petroleum reservoirs. Insights from molecular Member, Tia Juana Field, Maracaibo Basin, Venezuela. Fuel
mechanics. Organic Geochemistry 24, 765–783. 81, 1971–1976.
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Bio- Volkman, J.K., Alexander, R., Kagi, R.I., Woodhouse, G.W.,
marker Guide. second ed. Wiley, New York. 1983. Demethylated hopanes in crude oils and their applica-
Philp, R.P., 1983. Correlation of crude oil from the San Jorges tions in petroleum geochemistry. Geochimica et Cosmochi-
Basin, Argentina. Geochimica et Cosmochimica Acta 47, mica Acta 47, 785–794.
267–275. Walters, C.C., 1993. Selective biodegradation of hopanes in oils
Requejo, A.G., Halpern, H.I., 1989. An unusual hopane biodeg- from the South Belridge Field. In: Øygard, K. (Ed.),
radation sequence in tar sands from the Pt. Arena (Monterey) Proceedings of the 16th International Meeting on Organic
formation. Nature 342, 670–673. Geochemistry. Falch Hurtigtrykk, Oslo, pp. 16–19.
Rooney, M.A., Vuletich, A.K., Griffith, C.E., 1998. Compound- Watson, J.S., Jones, D.M., Swannell, R.P.J., van Duin, A.C.T.,
specific isotope analysis as a tool for characterizing mixed oils: 2002. Formation of carboxylic acids during aerobic biodeg-
an example from the West of Shetlands area. Organic radation of crude oil and evidence of microbial oxidation of
Geochemistry 29, 241–254. hopanes. Organic Geochemistry 33, 1153–1169.
Rullkötter, J., Wendisch, D., 1982. Microbial alteration of Wenger, L.M., Isaksen, G.H., 2002. Control of hydrocarbon
17a(H)-hopanes in Madagascar asphalts: removal of C-10 seepage intensity on level of biodegradation in sea bottom
methyl group and ring opening. Geochimica et Cosmochimica sediments. Organic Geochemistry 33, 1277–1292.
Acta 46, 1545–1553. Williams, J.A., Bjorøy, M., Dolcater, D.L., Winters, J.C., 1986.
Scotchman, I.C., Griffith, C.E., Holmes, A.J., Jones, D.M., 1998. Biodegradation in South Texas eocene oils-effects on aromat-
The Jurassic petroleum system north and west of Britain: a ics and biomarkers. Organic Geochemistry 10, 451–461.

You might also like