Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Organic Geochemistry 32 (2001) 105±114

www.elsevier.nl/locate/orggeochem

Aerobic biodegradation of hopanes and norhopanes


in Venezuelan crude oils
F.D. Bost a, R. Frontera-Suau a,1, T.J. McDonald b,
K.E. Peters c,2, P.J. Morris a,d,*
a
Department of Microbiology and Immunology, Medical University of South Carolina, 221 Fort Johnson Rd.,
Charleston, SC 29412, USA
b
B & B Laboratories, 1902 Pinon Dr., College Station, TX 77845, USA
c
Mobil Technology Company, PO Box 650232, Dallas, TX 75265, USA
d
Marine Biomedicine and Environmental Sciences, Medical University of South Carolina, 221 Fort Johnson Rd.,
Charleston, SC 29412, USA
Received 12 January 2000; accepted 25 September 2000
(returned to author for revision 25 March 2000)

Abstract
The microbial degradation of two Venezuelan crude oils enriched in 25-norhopanes was examined after a 5-week
aerobic incubation using a microbial enrichment culture. Analysis of the oils using gas chromatography±mass spec-
trometry revealed degradation of the C28 tricyclic terpane, the C29±C34 17a(H),21b(H)-hopanes, and the C29
17a(H),21b(H)-25-norhopane. The C35 17a(H),21b(H)-hopane and 18a(H)-oleanane were conserved. Further, the C28±
C34 17a(H),21b(H)-25-norhopanes were degraded and no formation of 25-norhopanes was observed. Degradation
caused preferential removal of the 22R versus the 22S isomer in both the extended hopanes and 25-norhopanes,
implying that bacteria remove these compounds in aerobic environments. These data demonstrate 25-norhopane
degradation on a time scale similar to that for other biomarkers. # 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Biodegradation; 25-Norhopane; Hopane; Crude oil; Venezuela

1. Introduction Hopanes are a class of pentacyclic triterpane bio-


markers that originate from hopanoids in bacterial
Biomarkers are structurally complex components of membranes (Ourisson et al., 1984; Prince, 1987).
petroleum derived from biological molecular precursors, Numerous studies show that C30 17a,21b(H)-hopane
such as chlorophyll, sterols, and hopanoids (Peters and and its extended homologs (homohopanes) are biode-
Moldowan, 1993). Many biomarkers in crude oil are graded in the environment and laboratory (Goodwin et
resistant to biodegradation and are used by petroleum al., 1983; Peters and Moldowan, 1991; Chosson et al.,
geochemists to assess genetic relationships, thermal 1992; Parker and Acey, 1993; Peters and Moldowan,
maturity and biodegradation. The biomarker pro®le of 1993; Moldowan et al., 1995; Morris et al., 1995). Some
a crude oil is distinctive and diagnostic, often allowing reservoired crude oils contain 25-norhopanes, pre-
correlation of an oil to its source rock. sumably formed by demethylation of the A/B ring at the
C-10 position (RullkoÈtter and Wendisch, 1982; Volk-
man et al., 1983). The 25-norhopanes were ®rst
* Corresponding author. Tel.: +1-843-762-5533; fax: +1-
observed as ``degraded hopanes'' in an oil-impregnated
843-762-5535.
1 sandstone from the Uinta Basin in Utah (Reed, 1977).
Present address: Dept. of Environmental Sciences and
Engineering, 104 Rosenall Hall, University of North Carolina, Since that report, other investigators have observed 25-
Chapel Hill, NC 27599-7400, USA. norhopanes in heavily degraded reservoir oils (e.g. Sei-
2
Present address: Exxon Mobil Upstream Research Co., PO fert and Moldowan, 1979; Volkman et al., 1983; Seifert
Box 2189, Houston, TX 77252-2189, USA. et al., 1984; Requejo and Halpern, 1989; Moldowan et
E-mail address: morrisp@musc.edu (P.J. Morris). al., 1995). In most cases, a relative decrease in the
0146-6380/01/$ - see front matter # 2001 Elsevier Science Ltd. All rights reserved.
PII: S0146-6380(00)00147-9
106 F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114

abundance of the hopanes corresponds to increased 25- organisms were isolated from this culture, none of
norhopanes (Requejo and Halpern, 1989; Moldowan which have the ability to degrade C30 17a,21b-hopane in
and McCa€rey, 1995). In a later study of reservoir oils pure culture using crude oil as the sole carbon source
from the West Siberia and San Joaquin basins, Peters et (unpublished results). The data presented are from
al. (1996) concluded that the conversion of hopanes to experiments with one of two Venezuelan crude oils,
their corresponding 25-norhopanes was selective, with since the extent and pattern of degradation were com-
lower molecular-weight extended hopanes degraded parable for both. Our microbial culture demonstrated
before the high molecular-weight homologs. Further- simultaneous degradation of the hopanes and 25-nor-
more, they showed that C-25 demethylation favors the hopanes, suggesting that in many aerobic surface envir-
22S epimers of the C31 and C32 hopanes compared to onments these biomarkers have similar fates.
22R, while the opposite applies to the C34 and C35
hopanes. Molecular modeling showed that steric di€er-
ences between the two C-22 epimeric forms explain this 2. Experimental
stereoselective degradation.
The distribution of 25-norhopanes in petroleum 2.1. Venezuelan oils
reservoirs is not ubiquitous. Blanc and Connan (1992)
contended that 25-norhopanes are already present in The two oils used in this study are heavy production
petroleum and are ``unmasked'', or concentrated, oils collected from the Maturin-Temblador basin in
during biodegradation. In earlier experiments, though, Venezuela. These oils contain the C29±C35 17a,21b-
the 25-norhopanes were not a product of kerogen hopanes as well as the C28±C34 17a,21b-25-norhopanes.
cracking in laboratory pyrolysis studies of Western Table 1 lists various characteristics of the two oils.
Australian shales, suggesting that 25-norhopanes occur
in sediments as isolated compounds, not as complete 2.2. Microbial enrichment culture
series as in biodegraded oils (Noble et al., 1985). Addi-
tionally, Peters and Moldowan (1991) normalized the The microbial enrichment culture (LC culture) was
concentrations of various crude oil components in bio- originally enriched with soil from a creosote-con-
degraded and non-biodegraded oils to a conserved C27 taminated site in Fairhope, AL, using a Nigerian Bonny
diasterane to demonstrate that the concentration e€ect Light crude oil (Table 1) as the sole carbon source (2
of biodegradation could not account for the 25-norho- mg/ml). The LC culture was maintained with monthly
pane concentrations present in the biodegraded oil. transfers (4% inoculum) into fresh basal medium,
Biodegradation of hopanes commonly occurs without BMTM (Hareland et al., 1975), with Bonny Light crude
the formation of 25-norhopanes (Peters and Moldowan, oil. At the time of this experiment, the LC culture had
1993 and references therein). 25-Norhopanes have been been transferred for 27 consecutive months and had
observed in petroleum reservoirs where the hopanes are consistently maintained C30 17a,21b-hopane-degrading
demethylated prior to sterane alteration (Brooks et al., activity on Bonny Light crude oil.
1988). However, if the steranes are degraded prior to the
hopanes, then 25-norhopanes are not formed. In an
earlier study (Seifert and Moldowan, 1979), no 25-nor-
Table 1
hopane formation was observed in a Texas oil that had
Characteristics of Venezuelan and Bonny Light crude oils
been depleted of hopanes. This observation has been
duplicated in laboratory studies (Goodwin et al., 1983; Characteristics Venezuelan Venezuelan Bonny
Chosson et al., 1992; Morris et al., 1995). At a petro- oil (no. 14050) oil (no. 14103) Light crudea
leum re®nery landfarm, Moldowan et al. (1995) did not
API gravity 15 12 35
observe 25-norhopanes despite evidence of hopane
% Saturates 33 26 58
degradation in material that had been deposited at the (HPLC)
site almost 10 years earlier. This suggests that in aerobic % Aromatics 39 43 31
surface environments hopane degradation either yields (HPLC)
products other than the 25-norhopanes or that the 25- % Polars 28 31 11
norhopanes are subject to similar degradative mechan- (HPLC)
isms as the hopanes. % Asphaltenes 8 9 2
To determine the susceptibility of 25-norhopane to (isolated)
biodegradation, we investigated the fate of both the a
Oil used in enrichment of the LC culture for this study.
hopanes and 25-norhopanes in two Venezuelan crude Asphaltenes were removed prior to component analysis. Com-
oils (Table 1) using a microbial culture previously ponent analysis calculations for the saturate, aromatic, and
shown to degrade the C30 17a,21b-hopane in Bonny polar fractions were based on the normalized oil mass remain-
Light crude (Frontera-Suau et al., 1997). Nine micro- ing after asphaltene removal.
F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114 107

2.3. Time-course culture conditions 100% hexane to 100% dichloromethane. The saturate
fraction was treated with molecular sieve beads prior to
During the monthly transfer of the LC culture, a 4% analysis to remove the n-alkanes. The biomarkers were
inoculum (109 colony forming units/ml) was used to separated on an HP-1 fused silica capillary column (30
inoculate each 50 ml glass tube containing 10 ml of m0.25 mm i.d.). Gas chromatography/mass spectro-
BMTM supplemented with 2 mg/ml of either Venezue- metry analyses (m/z=177, m/z=191, m/z=217) were
lan oil as the sole carbon source. Triplicate inoculated performed using a Hewlett-Packard 5890 II gas chro-
samples and triplicate uninoculated control samples matograph interfaced to a Hewlett-Packard 5972 MSD
were set up for each of the following time points: 0, 1, 2, (operated at 70 eV in the selected ion mode, GC/MS/
3 and 5 weeks. Cultures were incubated at 30 C in the SIM).
dark and shaken at 200 rpm.
2.7. Biomarker quantitation
2.4. Crude oil extraction
Biomarker ratios were calculated using peak areas
At each time point triplicate uninoculated controls from the m/z=191 and m/z=217 chromatograms. For
and triplicate LC-inoculated samples were sacri®ced for quantitative analysis, the response factor for the surro-
each Venezuelan oil. Oil was extracted from the cultures gate standard was calculated by dividing the surrogate
by shaking three times with 10-ml aliquots of dichloro- concentration (5.0 mg/ml) by the respective peak area.
methane (Omnisolv HR-GC grade, 99.9%, EM Science, Concentrations for the C30±C35 17a,21b(H)-hopanes
Gibbstown, NJ). Extracts were combined, dried with (22S and 22R), 18a-oleanane, and C27 13b,17a-diaster-
anhydrous sodium sulfate (J. T. Baker, Phillipsburg, ane (20S) were determined by multiplying the respective
NJ), evaporated under vacuum to reduce volume, and peak areas (m/z=191, m/z=217) from the mass chro-
air-dried. Hexane (10-ml GC2 grade, 99.9%, Burdick & matograms by the response factor for the surrogate stan-
Jackson, Muskegon, MI) was then added to each dard. For each of the tricyclic terpanes (C28, C29, or C30),
sample to precipitate the asphaltenes prior to gas chro- the peak areas for the 22R and 22S epimers were added
matographic analysis. and divided by the combined value of the 22R and 22S
epimers for the conserved C35 17a,21b(H)-homohopane or
2.5. Gas chromatography (GC) by the peak area for the C27 13b,17a-diasterane (20S) from
the m/z=217 mass chromatograms. Also, the peak area
Deasphaltened samples in hexane were analyzed using for the C27 13b,17a-diasterane (20S) from the m/z=217
a Hewlett-Packard Model 5890 Series II Plus gas chro- mass chromatograms was divided by the combined peak
matograph equipped with a ¯ame ionization detector areas for the 22R and 22S epimers of the C35
(GC-FID) and an HP-5 column (25 m0.32 mm 17a,21b(H)-homohopane. The Ts to Tm ratios were
i.d.0.17 mm). The injector and detector temperatures calculated from the peak areas of the m/z=191
were 290 and 315 C, respectively. The carrier and com- chromatograms according to the formula Ts/(Ts+Tm).
bustion gases were helium and hydrogen, respectively. Peak areas for both C-22 epimers were used in the for-
The temperature program began at 50 C for 1 min and mula%C35 (22R+22S)/(C31±C35) (22R+22S) to com-
proceeded at 5 C/min to 310 C with a hold at 310 C for pute the homohopane index. The oleanane index was
20 min. calculated by dividing the peak areas of the m/z=191
chromatograms for 18a(H)-oleanane by those of C30
2.6. Gas chromatography±mass spectrometry (GC±MS) 17a,21b(H)-hopane. Calculations for the homohopane
and oleanane indices using the absolute concentration
The method of McDonald and Kennicutt (1992) was values were virtually identical to the values calculated
used to analyze the samples for biomarkers. In summary, using the peak areas. In the regular sterane to hopane
the hydrocarbon fraction of each sample (in dichloro- ratio, the C27, C28, and C29 aaa (20R+20S) and abb
methane) was added to the top of a chromatographic col- (20R+20S) regular sterane areas from the m/z=217
umn packed with 5 g of alumina. The surrogate standards, mass chromatograms were divided by the C29±C33
5b-cholane, d10-phenanthrene, and d12-chrysene (0.5 ml), 17a(H)-hopane (22R+22S) areas from the m/z=191
were added as a mixture to the chromatographic column mass chromatograms.
after addition of the sample. The sample/surrogate mixture
was eluted from the column with 15-ml dichloromethane,
collected in a 20-ml centrifuge tube, and gently evaporated 3. Results and discussion
to 500 ml with puri®ed nitrogen. The eluant was separated
into saturate fractions using high pressure liquid chroma- Fig. 1 shows the degradation of the n-alkanes and
tography (HPLC) with a Partisil 5 mm PAC Magnum acyclic isoprenoids in Venezuelan oil 14103 based on
HPLC preparative column and a solvent gradient from GC-FID analysis over 5 weeks. Uninoculated controls
108 F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114

Fig. 1. GC±FID chromatograms of Venezuelan oil 14103 extracted from microbial cultures at times 0, 1, 3 and 5 weeks.

for each time point did not di€er signi®cantly from the removal, generally at the same time as the diasteranes
time zero chromatograms shown. The n-alkanes, (Reed, 1977; Seifert and Moldowan, 1979). In Fig. 2,
pristane, and phytane degraded within the ®rst week, however, m/z=191 chromatograms reveal extensive
with few qualitative changes in the GC-FID pro®les of degradation of both the R and S epimers of the C28 tri-
culture extracts in the following weeks. Mass spectro- cyclic terpane after week 3. A recent study, based on
metric analysis of other biomarkers in these residues molecular volumes and surface areas, considers the sec-
revealed conservation of the steranes, and degradation of ond-eluting peak (inferred to be the 22R epimer) for the
the hopanes, 25-norhopanes, and the tricyclic terpanes. C26±C29 tricyclic terpanes more readily biodegraded by
As a class, the tricyclic terpanes are quite recalcitrant. a proposed C-10 demethylation process similar to that
Their degradation typically occurs well after hopane hypothesized for the hopanes (Peters et al., 1996; Peters,
F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114 109

Fig. 2. (A) GC±MS chromatograms (m/z=191) of Venezuelan oil 14103 crude oil extracts at times 0, 1, 3, and 5 weeks. (B) GC±MS
chromatograms (m/z=177) of Venezuelan oil 14103 crude oil extracts at times 0, 1, 3 and 5 weeks.

2000). A study of the tricyclic terpanes in heavy ratio over the same time decreased 82%. However,
Venezuelan reservoir oils showed preferential removal analysis of the C28 tricyclic terpane epimers revealed no
of the second-eluting tricyclic terpane peak (Alberdi et discernable preference (data not shown). The C29 and
al., 2000). In our laboratory study, however, no epimer C30 tricyclic terpanes in Venezuelan 14103, however,
speci®city was observed for the degradation of the C28 tri- remained relatively conserved over the course of the
cyclic terpane. After 5 weeks, the ratio of the C28 tricyclic experiment when compared to C35 17a,21b-homohopane
terpane to C35 17a,21b-homohopane had decreased 32% and C27 13b,17a-diasterane (20S) (Fig. 2, Table 2).
from time zero (Table 2). When the C28 tricyclic terpane 18a-Oleanane was more resistant to degradation than
is compared to the C27 13b,17a-diasterane (20S), the other compounds, yielding a 10-fold increase in the
110 F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114

Table 2
Biomarker ratios calculated for Venezuelan oil samples at time zero and after 5 weeks of biodegradation with the LC culturea

C28TT/ C29TT/ C30TT/ C35H/ C28TT/ C29TT/ C30TT/ C27D/ Ts/ Homohopane Regular Oleanane
C35Hb C35Hb C35Hb C30Hb C27Dc C27Dc C27Dc C35Hc Tmd index (%)e steranes/ indexg
hopanef

Time zero
Control 0.90.1 1.10.1 1.30.0 0.60.0 4.40.3 5.20.4 6.40.5 0.20.0 0.20.0 17.90.2 1.20.0 0.10.0
Sample 1.000.1 1.10.1 1.30.1 0.50.1 4.10.4 4.80.3 5.60.3 0.20.0 0.30.0 16.20.6 1.10.1 0.20.1
Week 1
Control 0.90.1 1.10.1 1.30.2 0.50.0 4.00.2 4.90.1 5.80.3 0.20.0 0.20.0 17.11.2 1.20.0 0.10.0
Sample 0.70.1 1.20.2 1.20.1 0.50.0 3.10.2 5.10.7 5.40.2 0.20.0 0.30.0 17.50.8 1.30.1 0.20.0
Week 2
Control 0.90.1 1.20.0 1.30.1 0.50.0 4.00.1 5.10.2 5.70.5 0.20.0 0.20.0 16.20.5 1.20.0 0.10.0
Sample 0.60.3 1.10.1 1.40.1 0.80.4 2.20.9 4.00.2 4.80.1 0.20.0 0.30.0 21.93.8 2.10.7 0.40.2
Week 3
Control 1.00.0 1.30.1 1.50.1 0.50.0 4.10.3 5.00.7 6.01.0 0.30.1 0.20.0 15.90.4 1.20.0 0.20.0
Sample 0.40.2 1.10.1 1.60.1 1.10.4 1.40.6 3.90.6 5.50.5 0.30.0 0.30.0 32.54.2 4.40.8 0.50.2
Week 5
Control 1.20.0 1.40.1 1.70.1 0.40.0 4.00.4 4.80.5 5.50.5 0.20.0 0.20.0 15.50.3 1.20.0 0.10.0
Sample 0.30.2 1.30.2 1.80.1 2.11.2 0.70.5 3.40.3 4.60.3 0.20.0 0.30.1 41.47.1 7.82.9 1.50.8
a
All values are the averages of triplicate samples with the standard deviations of those values. Controls are uninoculated cultures
maintained under the same conditions as the inoculated samples.
b
Calculated from m/z=191 mass chromatogram peak areas of the C28 through C30 tricyclic terpanes (TT) (22R+22S), C35
17a,21b(H)-homohopane (C35H) (22R+22S), and C30 17a,21b(H)-hopane (C30H).
c
Calculated from m/z=191 mass chromatogram peak areas of the C28 through C30 tricyclic terpanes (TT) (22R+22S), C35
17a,21b(H)-homohopane (C35H) (22R+22S), and C30 17a,21b(H)-hopane (C30H) along with the peak areas for the C27 13b,17a-
diasterane (20S) (C27D) from the m/z=217 mass chromatograms.
d
Calculated from m/z=191 mass chromatogram peak areas of the C27 17a(H)-22,29,30-trisnorhopane (Tm) and C27 18a(H)-
22,29,30-trisnorneohopane (Ts).
e
%C35 (22R+22S)/(C31±C35) (22R+22S) homohopanes as determined from peak areas from m/z=191 chromatograms. Calcula-
tions of the homohopane index using the absolute concentration data in Table 3 provided similar values.
f
In this ratio, the C27, C28, and C29 aaa (20R+20S) and abb (20R+20S) regular sterane areas from the m/z=217 mass chroma-
tograms and the C29±C33 17a(H)-hopanes (22R+22S ) from the m/z=191 mass chromatograms were used.
g
Calculated from m/z=191 mass chromatogram peak areas for 18a(H)-oleanane and C30 17a,21b(H)-hopane. Calculations of the
oleanane index using the absolute concentration data in Table 3 provided similar values.

oleanane index between the controls and inoculated sam- nonbiodegraded oils to indicate the redox potential of
ples at week 5 (Table 2). Degradation of three norhopane the source sediments, increased in the Venezuelan sam-
species, 17a(H)-28,30-bisnorhopane, 18a(H)-22,29,30- ples from 16% in controls at week 5 to 41% in week 5
trisnorneohopane (Ts), and 17a(H)-22,29,30-trisnorho- inoculated samples (Table 2). Additionally, the ratio of
pane (Tm), was observed on the m/z=191 chromatogram the recalcitrant C27 13b,17a-diasterane (20S) to the C35
(Fig. 2). Although both Ts and Tm were degraded in 17a,21b-homohopane did not signi®cantly vary over the
Venezuelan oil (Fig. 2), a slight increase in the Ts/Tm course of the experiment (Table 2). Similar trends in
ratio was observed (Table 2), implying that Tm is more hopane degradation were observed in the quantitative
readily biodegraded than Ts. data (Table 3). These data further illustrate the relative
By week 5, C30 17a,21b-hopane and C31±C34 17a,21b- conservation of C35 17a,21b-homohopane during the
homohopanes were reduced relative to C35 hopane at experiment.
time zero (Fig. 2). Signi®cant increases in both the C35 17a,21b-Homohopane conservation was observed
oleanane index and C35 17a,21b-homohopane:C30 in tar-sand samples from the Pt. Arena (Monterey)
17a,21b-hopane ratio reveal more extensive degradation Formation by Requejo and Halpern (1989). Moldowan
of C30 17a,21b-hopane compared to C35 homohopane et al. (1995) later demonstrated degradation of the C30
or oleanane (Table 2). Preferential degradation of the R through C34 hopanes with preservation of the C35
isomer over the S isomer was observed in the C31±C34 hopane in samples from a petroleum landfarm site.
17a,21b-homohopane degradation (Fig. 2). However, Peters et al. (1996) observed conservation of higher
the C35-homohopane index, commonly used for molecular-weight hopanes with preferential degradation
F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114 111

Table 3
Biomarker ratios and absolute concentrations of selected compounds in Venezuelan oil at time zero and week 5a

Time zero Week 5

Control Sample Control Sample

C30H 290.929.5 291.165.4 270.336.8 49.020.4


C31H, 22S 172.814.1 162.919.7 151.219.7 26.211.3
C31H, 22R 126.117.4 120.816.1 109.110.8 19.49.1
C32H, 22S 123.913.7 113.09.7 103.110.9 20.29.0
C32H, 22R 80.310.6 71.37.2 65.18.5 9.04.6
C33H, 22S 87.97.7 77.69.3 74.313.5 14.91.4
C33H, 22R 49.14.9 46.14.6 38.63.7 7.82.5
C34H, 22S 57.25.2 52.06.2 45.95.8 21.26.1
C34H, 22R 35.93.7 32.95.2 26.85.0 7.12.6
C35H, 22S 98.813.6 82.78.1 71.98.8 55.92.0
C35H, 22R 61.05.2 47.78.5 40.56.4 30.95.5
Oleanane 39.15.6 73.449.6 31.76.4 66.930.5
C27D, 20S 32.15.7 31.03.9 31.92.2 31.71.0
a
All values (ng/ml) are the averages of triplicate samples with the standard deviations of those values. Controls are uninoculated
cultures maintained under the same conditions as the inoculated samples. The response factor for the surrogate standard (see
Experimental) was calculated by dividing the surrogate concentration (5.0 mg/ml) by the respective peak area. Concentrations for the
C30±C35 17a,21b(H)-hopanes (22S and 22R), 18a-oleanane, and C27 13b,17a-diasterane (20S) were determined by multiplying the
respective peak areas (m/z=191, m/z=217) used to calculate the biomarker ratios in Table 2 by the response factor for the surrogate
standard.

of the 22R epimer. Subsequent molecular modeling of cyclic aromatic hydrocarbons (PAHs) by P. paucimobilis
the R and S epimers of the C34±C35 homohopanes EPA505. In a complex medium that included yeast
revealed that the S epimer conformation may sterically extract and glucose, resting P. paucimobilis EPA505
protect the C-25 methyl group from microbial attack cells were capable of growing on only 5 of the 17 PAHs
(Peters et al., 1996). Several microbially unique enrich- tested. However, ¯uoranthene-grown resting Pseudomo-
ment cultures in our laboratory from geographically nas paucimobilis EPA505 cells were capable of trans-
distinct soils show hopane-degrading activity with con- forming 11 of the 17 PAHS tested, suggesting that the
servation of the C35 hopane (unpublished data). presence of more labile carbon sources abrogates the
Conservation of the C35 hopane in the above studies transformation of more recalcitrant compounds.
contrasts to that of Goodwin et al. (1983), who reported The additional carbon sources included in the Chosson
degradation of the higher molecular-weight hopanes et al. (1992) study may have also stimulated the co-meta-
after 11.5 months (C35>C34>C33>C32>C31>C30). bolism of compounds in the West Rozel crude oil. A study
The Goodwin et al. (1983) study also observed degra- by Kachholz and Rehm (1978) described n-alkane co-
dation of the R epimer over the S epimer, observations metabolism, or incomplete mineralization, by ®ve species
similar to those in the previous examples. In their study, a of Bacillus during growth on glucose, peptone, and yeast
mineral salts medium was amended with crude oil as the extract. Ooyama and Foster (1965) described co-metabo-
source of inoculum and carbon. 25-Norhopanes were lism of cyclopentane, cyclohexane, cycloheptane, and
not detected despite hopane degradation. Chosson et al. cyclooctane to their corresponding cycloketones when
(1992) observed alteration of the steranes in the saturate grown in a medium with propane as the growth substrate.
fraction of West Rozel oil. Their hopane degradation by More complex saturated molecules, such as the hopanes,
pure cultures of Nocardia, Arthrobacter, and Myco- may also be transformed by a co-metabolic activity, as has
bacterium species, though not as pronounced as the been suggested by (1) demethylation to form norhopanes,
sterane degradation, showed the same C35>C34>C33> or (2) b-oxidation of the alkyl side chain. To support the
C32>C31>C30 hopane degradation sequence as latter, microbial conversion of 3a,5a-cyclosterols to 17-
observed by Goodwin et al. (1983). The mineral medium ketosteroids has been observed (Martin, 1977).
used by Chosson et al. (1992) incorporated glycerol and In addition to co-metabolic alteration, the saturated
yeast extract, potentially preferred carbon sources that ring structure of cycloalkanes may also be directly
could decrease the catabolic diversity of the micro- attacked. The general mechanism for the degradation of
organisms and might explain the altered hopane degra- cyclohexane involves the oxidative conversion to cyclo-
dation sequence in that study. To support this, Mueller hexanol, followed by conversion to cyclic ketones.
et al. (1990) described the biotransformation of poly- Hydrolysis leads to further enzymatic attack and
112 F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114

aromatization of the ring followed by ring cleavage by C35 hopanes in samples from the south end of the site.
dioxygenases (Perry, 1977; Trower et al., 1985, 1989). Despite the degradation of the hopanes, 25-norhopanes
Degradation of dehydroabietic acid by Flavobacterium were not detected. These observations are similar to the
resinovorum and Alcaligenes eutrophus is mediated degradation sequence observed for our Venezuelan oil,
through an attack on the saturated rings similar to that suggesting that conservation of C35 hopane and lack of
described above for cyclohexane (Biellman et al., 1973; 25-norhopane formation are a common degradation
Liss et al., 1997). Although potential pathways were not pathway in aerobic environments.
explored, the complete removal of isopimaric and Hopane degradation in petroleum reservoirs has been
dehydroabietic acids from the cultures of two Pseudo- observed under the following conditions: (1) the
monas-like isolates implies cleavage of the saturated hopanes are demethylated to 25-norhopanes prior to
rings in the molecule (Wilson et al., 1996). These sterane alteration, or (2) the steranes are degraded
mechanisms for degradation of smaller saturate mol- before the hopanes and the hopanes do not degrade to
ecules provide a framework for understanding microbial 25-norhopanes (Brooks et al., 1988). Interestingly, the
attack of more complex cyclic biomarkers. sterane pro®les in our biodegraded Venezuelan oils were
The LC enrichment culture was used to ascertain not altered during the experiment despite hopane and
whether a culture that is capable of degrading C30 25-norhopane degradation (data not shown, Table 2).
17a,21b-hopane is also capable of degrading the 25- In addition, no 25-norhopanes formed. The ratio of
norhopanes. The Venezuelan crude oils contain the C29 regular steranes to C29±C33 hopanes increased seven-
17a,21b-hopane, C30 17a,21b-hopane, and the entire fold by week 5 (Table 2). Hopane losses before sterane
suite of C31±C35 extended hopanes (Fig. 2). Of special degradation have been observed previously (Peters and
interest to this study, however, was the presence of the Moldowan, 1991). Further, a previous study using
C28±C34 17a,21b-25-norhopanes (Fig. 2), the putative microorganisms enriched from the same soil as the LC
demethylation products of the hopanes. Our data show culture showed degradation of the C27 diasteranes in a
that hopane and 25-norhopane display similar suscept- fossil fuel soil extract (Morris et al., 1995). In the pres-
ibility to degradation. Substantial changes in the m/z=177 ent study, however, microorganisms from the same soil
chromatograms were not observed until weeks 3 and 5 did not a€ect the steranes or diasteranes (data not
(Fig. 2). At week 5, the C28 and C29 17a, 21b-25-norho- shown, Table 2). These di€ering patterns of biomarker
panes (Fig. 2) were reduced, as were the R epimers of degradation in the Venezuelan oils and in the examples
the C30±C34 17a, 21b-25-norhopanes. At no time during cited above indicate that the composition of the carbon
this experiment was an increase in abundance of the source exerts control on microbial consortia.
17a,21b-25-norhopanes observed relative to C35 Oxygenated intermediates are expected as the result
hopane. In a separate experiment, we examined degra- of aerobic alteration of hopanes rather than a direct
ded Bonny Light crude oil residues daily for the possible progression to the saturated 25-norhopanes (Higgins and
formation of 25-norhopanes from hopane by the LC Gilbert, 1978). Aerobic biodegradation should produce
culture (data not shown). This daily analysis revealed no functionalized structures (hydroxyl moieties, carboxylic
formation of 25-norhopanes, despite C30±C34 hopane acid groups, ketones, etc.) at the position being altered
degradation. Accumulation of 17a,21b-25-norhopanes (Blanc and Connan, 1992). de Lemos Sco®eld (1990)
would be expected if 25-norhopanes were the demethyl- described a C-10 hopanoic acid as a possible inter-
ation products of the hopanes. Thus, given the similar mediate of hopane catabolism. The occurrence of the C-
susceptibility of hopane and 25-norhopane to degrada- 10 hopanoic acid was not monitored for this study;
tion, either no 17a,21b-25-norhopanes were formed however, future studies with the LC enrichment culture
from hopane transformation or the 25-norhopanes were will address the presence of hopanoic acids (Bost et al.,
transiently formed and rapidly degraded. 2000). To the best of our knowledge, though, these
Certain evidence suggests that hopane degradation is intermediate compounds have not been detected in
not always accompanied by the production of 25-nor- laboratory microbial degradation experiments.
hopanes. Goodwin et al. (1983) saw no 25-norhopanes Other compounds in oil, such as 18a(H)-oleanane,
in their laboratory-degraded Kent oil samples despite also contain a methyl group attached to C-10, yet are
hopane degradation. In laboratory studies of West conserved during degradation of the Venezuelan oil.
Rozel crude oil biodegradation by 73 aerobic bacteria, Peters et al. (1996) convincingly used molecular
25-norhopanes were not detected despite demonstrated mechanics to explain the conservation of C35 17a,21b-
hopane loss (Chosson et al., 1992). No 25-norhopane hopane, hypothesizing that a ``scorpion'' con®guration
formation was observed in samples of heavily bio- of the hopane molecule aliphatic side chain shields the
degraded oil seeps from Greece although the hopanes C-25 methyl group from microbial attack. Another
were degraded (Seifert et al., 1984). In a study of a mechanism must protect 18a(H)-oleanane, however,
landfarm, Moldowan et al. (1995) demonstrated degra- which lacks the side chain in the homohopanes. Simi-
dation of the C30 through C34 with conservation of the larly, the 17a(H),21b(H)-30-norhopane was conserved
F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114 113

in the Kent oil used in the Goodwin et al. (1983) study. Bost, F.D., Retzke, J.J., Frontera-Suau, R., Ross, B., McDonald,
This 17a(H),21b(H)-30-norhopane contains the same C-10 T.J., Morris, P.M., 2000. Insights into C30 17a,21b-hopane
methyl group present in the hopane and homohopane transformation by two crude oil-degrading enrichment cul-
molecules, yet remains relatively unchanged despite tures. American Society for Microbiology Abstracts, 21±25
May, Los Angeles, CA.
hopane degradation.
Brooks, P.W., Fowler, M.G., MacQueen, R.W., 1988. Biologi-
The data presented here suggest that hopane and 25-
cal marker and conventional organic geochemistry of oils
norhopane degradation share a common mechanism, sands/heavy oils, Western Canada Basin. Organic Geochem-
given the similar onset of degradation seen for both istry 12, 519±538.
classes of compounds. The data also dispute the notion Chosson, P., Connan, J., Dessort, D., Lanau, C., 1992. In vitro
that 25-norhopane is a recalcitrant endpoint to hopane biodegradation of steranes and terpanes: a clue to under-
degradation during our de®ned laboratory investiga- standing geological situations. In: Moldowan, J.M.,
tions. Given the similar pro®les of hopane degradation Albrecht, P., Philp, R.P. (Eds.), Biological Markers in Sedi-
in Venezuelan oil to previous observations (Requejo and ments and Petroleum. Prentice Hall, Englewood Cli€s, NJ,
Halpern, 1989; Moldowan et al., 1995), this degradation pp. 320±349.
pattern may be representative of what occurs in aerobic Frontera-Suau, R., McDonald, T.J., Morris, P.J., 1997. A
crude oil-degrading enrichment culture capable of hopane
surface environments. Tritz et al. (1999) determined that
biodegradation. American Society for Microbiology
the only product of tritium-labeled hopane oxidation by Abstracts, 4±8 May, Miami Beach, FL.
a cholesterol-induced Arthrobacter simplex was hop- Goodwin, N.S., Park, P.J.D., Rawlinson, A.P., 1983. Crude oil
17(21)-ene. Additional studies under aerobic and biodegradation under simulated and natural conditions. In:
anaerobic conditions are aimed at determining the Bjorùy, M. et al. (Ed.), Advances in Organic Geochemistry
mechanisms of degradation of a C30 17a(H),21b(H)- 1981. J. Wiley and Sons, New York, pp. 650±658.
hopane concentrate from Bonny Light crude oil by a Hareland, W., Crawford, R.L., Chapman, P.J., Dagley, S.,
complex microbial consortium (Bost et al., 2000). 1975. Metabolic function and properties of 4-hydroxy-phe-
Understanding the mechanism(s) and pathway(s) of nylacetic acid 1-hydroxylase from Pseudomonas acidovorans.
hopane and 25-norhopane transformation will provide a Journal of Bacteriology 121, 272±285.
Higgins, I.J., Gilbert, P.D., 1978. The biodegradation of
more universal understanding of the possible in-reser-
hydrocarbons. In: Chater, K.W.A., Somerville, H.J., (Eds.),
voir biodegradation phenomena that a€ect the The Oil Industry and Microbial Ecosystems. Heyden & Sons,
biomarker pro®les of crude oils. pp. 80±117.
Kachholz, T., Rehm, H.J., 1978. Degradation of long chain
alkanes by bacilli II. Metabolic pathways. European Journal
Acknowledgements of Applied Microbiology and Biotechnology 6, 39±54.
de Lemos Sco®eld, A., 1990. Nouveau marquers biologiques de
This work was supported by the Donors of the Pet- sediments et petroles riches en soufre: identi®cation et mode
roleum Research Fund, administered by the American de formation. PhD thesis, Universitie Louis Pasteur de
Chemical Society. This research was sponsored in part Strasbourg.
Liss, S.N., Bicho, P.A., Saddler, J.N., 1997. Microbiology and
by the U.S. Department of Energy's cooperative agree-
biodegradation of resin acids in pulp mill e‚uents: a mini
ment # DE-FCO2-98CH10902 to the Medical Uni- review. Canadian Journal of Microbiology 75, 599±611.
versity of South Carolina's Environmental Biosciences Martin, C.K.A., 1977. Microbial cleavage of sterol side chains.
Program. The authors would also like to recognize the Advances in Applied Microbiology 22, 29±58.
helpful reviews of the manuscript provided by Professor McDonald, T.J., Kennicutt, M.C., 1992. Fractionation of
Robert Alexander and an anonymous reviewer. crude oils by HPLC and quantitative determination of ali-
phatic and aromatic biological markers by GC±MS with
Associate EditorÐL. Ellis selected ion monitoring. LC±GC. Magazine of Separation
Science 10, 935±938.
Moldowan, J.M., McCa€rey, M.A., 1995. A novel microbial
hydrocarbon degradation pathway revealed by hopane
References
demethylation in a petroleum reservoir. Geochimica et Cos-
mochimica Acta 59, 1891±1894.
Alberdi, M., Moldowan, J.M., Peters, K.E., Dahl, J.E., in Moldowan, J.M., Dahl, J., McCa€rey, M.A., Smith, W.J.,
press. Stereoselective biodegradation of tricyclic terpanes in Fetzer, J.C., 1995. Application of biological marker technol-
heavy oils from Bolivar Coastal Fields, Venezuela. Organic ogy to bioremediation of re®nery by-products. Energy &
Geochemsitry. Fuels 9, 155±162.
Biellman, J.F., Branlant, G., Gero-Robert, M., Poiret, M., Morris, P.J., Shelton, M.E. and Chapman, P.J., 1995. Co-
1973. Degradation bacterienne de l'acide dehydroabietique contaminated sites: The biodegradation of fossil fuels in the
Flavobacterium resinovorum. Tetrahderon 29, 1227±1236. presence of polychlorinated biphenyls. In: Hinchee R.E.,
Blanc, Ph., Connan, J., 1992. Origin and occurrence of 25-norho- Hoeppel R.E., Anderson D.B. (Eds.), Bioremediation of
panes: a statistical study. Organic Geochemistry 18, 813±828. Recalcitrant Organics. Battelle Press, pp 123±130.
114 F.D. Bost et al. / Organic Geochemistry 32 (2001) 105±114

Mueller, J.G., Chapman, P.J., Blattman, B.O., Pritchard, P.H., Reed, W.E., 1977. Molecular compositions of weathered pet-
1990. Isolation and characterization of a ¯uoranthene-utiliz- roleum and comparison with its possible sources. Geochim-
ing strain of Pseudomonas paucimobilis. Applied and Envi- cia et Cosmochimica Acta 41, 237±247.
ronmental Microbiology 56, 1079±1086. Requejo, A.G., Halpern, H.I., 1989. An unusual hopane bio-
Noble, R., Alexander, R., Kagi, R.I., 1985. The occurrence of degradation sequence in tar sands from the Pt. Arena
bisnorhopane, trisnorhopane and 25-norhopanes as free (Monterey) Formation. Nature 342, 70±673.
hydrocarbons in some Australian shales. Organic Geochem- RullkoÈtter, J., Wendisch, D., 1982. Microbial alteration of
istry 8, 171±176. 17a(H)-hopanes in Madagascar asphalts: removal of C-10
Ooyama, J., Foster, J.W., 1965. Bacterial oxidation of cyclo- methyl group and ring opening. Geochimica et Cosmochi-
paranic hydrocarbons. Antonie van Leeuwenhoek Journal mica Acta 46, 1545±1553.
of Microbiology and Serology 31, 45±65. Seifert, W.K., Moldowan, J.M., 1979. The e€ect of biode-
Ourisson, G., Albrecht, P., Rohmer, M., 1984. The microbial gradation on steranes and terpanes in crude oils. Geochimica et
origin of fossil fuels. Scienti®c American 251, 44±51. Cosmochimica Acta 43, 111±126.
Parker, S.L. and Acey, R.A., 1993. Biodegradation of hopanes Seifert, W.K., Moldowan, J.M., Demaison, G.J., 1984. Source cor-
by Mycobacterium fortuitum. Journal of Cellular Biochem- relation of biodegraded oils. Organic Geochemistry 6, 633±643.
istry (Suppl. 17C), 196. Tritz, J.-P., Herrmann, D., Bisseret, P., Connan, J., Rohmer,
Perry, J.J., 1977. Microbial metabolism of cyclic hydrocarbons M., 1999. Abiotic and biological hopanoid transformation:
and related compounds. Critical Reviews in Microbiology 5, towards the formation of molecular fossils of the hopane
387±412. series. Organic Geochemistry 30, 499±514.
Peters, K.E., 2000. Petroleum tricyclic terpanes: Predicted phy- Trower, M.K., Buckland, R.M., Higgins, R., Grin, M., 1985.
siochemical behavior from molecular mechanics calculations. Isolation and characterization of a cyclohexane-metabolizing
Organic Geochemistry 31, 497±507. Zanthobacter sp. Applied and Environmental Microbiology
Peters, K.E., Moldowan, J.M., 1991. E€ects of source, thermal 49, 1282±1289.
maturity, and biodegradation on the distribution and iso- Trower, M.K., Buckland, R.M., Grin, M., 1989. Character-
merization of homohopanes in petroleum. Organic Geo- ization of an FMN-containing cyclohexanone mono-
chemistry 17, 47±61. oxygenase from a cyclohexane-grown Zanthobacter sp.
Peters, K.E., Moldowan, J.M., 1993. The biomarker guide: European Journal of Biochemistry 181, 199±206.
interpreting molecular fossils in petroleum and ancient sedi- Volkman, J.K., Alexander, B., Kagi, R.I., Woodhouse, G.W.,
ments. Prentice Hall, Englewood Cli€s, NJ. 1983. Demethylated hopanes in crude oils and their applica-
Peters, K.E., Moldowan, J.M., McCa€rey, M.A., Fago, F.J., tion in petroleum geochemistry. Geochimica et Cosmochi-
1996. Selective biodegradation of extended hopanes to 25- mica Acta 47, 785±794.
norhopanes in petroleum reservoirs. Insights from molecular Wilson, A.E.J., Moore, E.R.B., Mohn, W.W., 1996. Isolation
mechanics. Organic Geochemistry 24, 765±783. and characterization of isopimaric acid-degrading bacteria
Prince, R.C., 1987. Hopanoids: the world's most abundant from a sequencing batch reactor. Applied and Environ-
biomolecules? Trends in Biochemical Sciences 12, 455±456. mental Microbiology 62, 3146±3151.

You might also like