You are on page 1of 12

Journal of Petroleum Science and Engineering 191 (2020) 107308

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: http://www.elsevier.com/locate/petrol

Fractionation of alkylated carbazoles in petroleum during subsurface


migration: Evidence from molecular simulation and application in
sandstone reservoirs☆
Qiuya Han a, Meijun Li a, b, *, Xiaoqiang Liu a, c, Weidong Jiang c, **, Shengbao Shi a, Youjun Tang b,
Daxiang He b
a
State Key Laboratory of Petroleum Resources and Prospecting, College of Geosciences, China University of Petroleum, Beijing, 102249, China
b
Key Laboratory of Exploration Technologies for Oil and Gas Resources, Ministry of Education, College of Resources and Environment, Yangtze University, Wuhan,
430100, China
c
College of Chemistry and Environmental Engineering, Sichuan University of Science and Engineering, Zigong, 643000, China

A R T I C L E I N F O A B S T R A C T

Keywords: Carbazole (CA) and alkylated carbazoles are common nitrogen-containing heterocyclic aromatic compounds
Geo-chromatographic fractionation widely found in crude oils and sedimentary rocks extracts. Their geochemical significances in petroleum
Alkylcarbazole exploration have been widely investigated in previous studies. Using density functional theory (DFT) and mo­
Petroleum migration orientation
lecular dynamics (MD) simulation, we calculated the polarity of alkylated carbazole isomers/homologues and
Molecular dynamics simulation
Sandstone reservoir
their adsorption energy with water and α-quartz, respectively. The calculated results show that there are sig­
nificant differences in migration fractionation effects for alkylated carbazole isomers/homologues. This study
confirmed the validity of previously proposed carbazole migration indicators, i.e. 4-/1-MCA (the relative
abundance of 4-methylcarbazole to 1-methylcarbazole), 2,5-/1,8-DMCA (2,5-dimethylcarbazole to 1,8-dimethyl­
carbazole) and 2,4-/1,8-DMCA (2,4-dimethylcarbazole to 1,8-dimethylcarbazole) on the basis of quantum
chemistry and molecular simulation. All these three parameters have been successfully applied in a lacustrine
sandstone reservoir in the Fushan Depression, Beibuwan Basin (South China Sea). The calculations of Gibbs free
energy show that the differences in thermal stability between the isomers of alkylcarbazoles can be negligible.
The source input and depositional environment have no significant effect on these indicators for oils from a same
oil family. Hence, the values of these parameters are mainly controlled by subsurface petroleum migration
processes. Therefore, the molecular parameters relative to alkylated carbazoles can be used as practically
effective geochemical markers to trace the oil filling orientation and pathways for subsurface oil migration.

1. Introduction molecular markers tracing oil migration distances and filling pathways.
(Li et al., 1995, 1998b; Larter et al., 1996; Silliman et al., 2002; Wang
Organic nitrogen compounds with heterocyclic aromatic structures et al., 2004).
are representative nonhydrocarbon constituents of crude oils. (Tissot The biological origin of carbazole and its homologues/isomers in
and Welte, 1980). Heterocyclic aromatic nitrogen compounds in pe­ crude oils still remains controversial. According to the compositions of
troleum can be classified into two main categories: neutral pyrrolic and nitrogen compounds in Californian oils, Snyder (1965) proposed that
basic pyridinic compounds. As important neutral pyrrolic compounds, most alkylcarbazoles and alkylbenzocarbazoles in crude oils may be
carbazole (CA) and its derivatives have been widely applied in petro­ derived from alkaloids with an indole nucleus in higher plants and
leum geochemistry during past two decades because they are effective blue-green algae (cyanobacteria). However, Li et al. (1995) found that

The abstract was originally presented at the 35th Annual Meeting of The Society for Organic Petrology (TSOP) in Beijing, China.

* Corresponding author. State Key Laboratory of Petroleum Resources and Prospecting, College of Geosciences, China University of Petroleum, Beijing, 102249,
China.
** Corresponding author.
E-mail address: meijunli@cup.edu.cn (M. Li).

https://doi.org/10.1016/j.petrol.2020.107308
Received 26 December 2019; Received in revised form 12 February 2020; Accepted 14 April 2020
Available online 18 April 2020
0920-4105/© 2020 Elsevier B.V. All rights reserved.
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

the distribution and compositions of nitrogen compounds in crude oils indicators for the secondary petroleum migration.
from marine and lacustrine sources are significantly different from those Here we demonstrated some oil migration tracers relative to CAs
of higher wax oils derived from higher plants. It believes that alkaloids proposed in previous studies on the basis of dipole moment (μ) and
are not the main source of carbazoles in crude oils (Li et al., 1995). interaction energy (EIn) of alkylcarbazoles with α-quartz by MD simu­
Besides, a contribution from proteins and plant pigments within sedi­ lation and practical application in a typical sandstone reservoir.
mentary organic matter during diagenesis cannot be ruled out.
Carbazole and its C1 and C2 alkylated homologues are usually 2. Experiments and methods
detected and identified from crude oils and rock extracts by gas chro­
matography–mass spectrometry (GC–MS). In addition, gas chromatog­ 2.1. The identification and quantitation of alkylated carbazoles
raphy (GC)–nitrogen specific detection and direct insertion probe mass
spectrometry (PMS) are also common techniques (Li et al., 1992; Bowler The nitrogen fractions were isolated from crude oils samples using
et al., 1997). Isolation of the nitrogen fraction was mainly performed the method reported by Li et al. (1995). First, add 40 ml n-hexane to oil
using two-step liquid column (LC) chromatographic separation method samples and filter for deasphalting. Then, the n-hexane solutions were
reported by Later et al. (1981) and with slight modifications of adding a further separated into saturated hydrocarbons, aromatic hydrocarbons
deasphalting step prior to the LC steps (Li et al., 1995). and nonhydrocarbons using neutral alumina chromatography columns
CAs, important molecular markers in reservoir geochemistry, are eluting with 50 ml n-hexane, 50 ml methylbenzene and 70 ml chlor­
ubiquitous in petroleum and sedimentary organic matter although pre­ oform/methanol (98:2, v:v), respectively. Finally, the pyrrolic nitrogen
sent in such low abundance (Li et al., 1995). Previous studies indicated fraction were eluted with 50 ml n-hexane/toluene (1:1, v:v) from non­
that the occurrence and distributions of CAs are predominantly influ­ hydrocarbons using silicic acid chromatography columns.
enced by migration fractionation effect (Yamamoto, 1992; Li et al., GC–MS (gas chromatography–mass spectrometry) analysis of the
1995, 1998a; Larter et al., 1996; Wilhelms et al., 1997). In addition, the nitrogen fractions of oil samples was performed on an Agilent 5975i
effect of depositional facies and environment (Clegg et al., 1997; Bakr mass spectrometer coupled with an HP 6890 GC with 60 � 0.32 mm HP-
and Wilkes, 2002; Zhang et al., 2008), thermal maturity (Harrison et al., 5MS fused silica capillary column (0.25μ film thickness coating). Helium
1997; Li et al., 1997; Clegg et al., 1998a, 1998b; Horsfield et al., 1998; was used as a carrier gas at a constant flow of 1.0 ml/min. The GC oven
Zhang et al., 2011), and biodegradation (Zhang et al., 1999; Huang temperature was initially set to 80 � C for 1 min and then ramped to 150
et al., 2003) cannot be neglected. Therefore, these factors should be C at 15 C/min, then rose to 290 � C at 5 C/min. The injection tem­
� �

taken into account when utilizing CAs as migration markers. perature was set at 300 C. The MS was running in full-scan mode with a

CAs react with formation water and mineral surfaces via forming scanning range of 50–600 Da and electron ionization (EI) mode at 70 eV.
hydrogen bond between an H atom of its N–H functional group on the CA, MCA (methylcarbazole) and C2-CA (dimethylcarbazole and
pyrrolic ring and high electronegative atoms (like oxygen atom) in ethylcarbazole) can be identified in oils (Fig. 1) by comparison of rela­
carrier beds (Dorbon et al., 1984; Krooss et al., 1991). Thus, the absolute tive retention time and retention indices on m/z 167, 181, 195 mass
concentrations of CAs and the relative abundance for some alkylcarba­ chromatograms with those in published work (Li et al., 1992; Bowler
zole isomers will change during secondary migration processes. And a et al., 1997). The carbazoles were quantified using known amount of
series of geochemical parameters relative to carbazoles have been pro­ 9-phenylcarbazole as an internal standard which was added prior to
posed to trace the oil migration orientation. Previous studies indicated GC–MS analysis of the pyrrolic nitrogen fraction. The concentration of
that the formation of hydrogen bond could be influenced by alkylation individual carbazole compounds were obtained on basis of integrated
positions in terms of steric hindrance which leads to differences in peak areas compared with that of the internal standard.
fractionation effect during oil migration for alkylcarbazoles. The relative
enrichments of N–H-shielded to N–H-exposed isomers and the higher
homologues to the lower homologous species with increasing migration
distances have been observed in the laboratory experiments (Li and
Larter, 1993; Li et al., 1994; Larter et al., 2000) and successively applied
in case studies of a variety of fields (Li et al., 1995, 1998b).
In recent years, molecular simulations have been preliminarily
applied in petroleum geochemistry (Liu et al., 2016; Yan and Yuan,
2016; Li et al., 2018) based on quantum chemistry and statistical me­
chanics via computers. Duin and Larter (1998) used this technique to
simulate the affinity of benzacarbazole isomers for oil and water phases.
The simulation results indicate that benzo[a]carbazole is more easily
removed from the oil phase due to its higher affinity for the water phase
than benzo[c]carbazole, which explain partly why benzo[a]carbazo­
le/benzo[c]carbazole decreases with increasing migration distances.
However, less such MD simulations have been performed on alkylcar­
bazoles to reveal their microscopic mechanism of migration fraction­
ation during subsurface oil migration.
To a large extent, the interaction energy (EIn) of CAs adsorption on
silica can indirectly determinate the oil migration orientation in carrier
beds through molecular modeling. Li et al. (2018) have been proved
experimentally and theoretically that the relative abundances of 1-meth­
yldibenzofuran (1-MDBF) to 4-methyldibenzofuran (4-MDBF) could
serve as potential molecular indicators to trace oil filling pathways and
migration orientation. Ideally, as increasing migration distances, the
alkylcarbazole of a small interaction energy with silica is relatively
enriched comparing with their counterparts of a large interaction energy
with silica. Therefore, we may choose the alkylcarbazole isomer­
s/homologues with significant differences of interaction energy as Fig. 1. Identification of CA, C1-CA and C2-CA isomers in oils.

2
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

2.2. Density functional theory calculations 2008; Ma et al., 2012).


The Fushan Depression can be divided into four tectonic units from
Geometry optimizations were carried out at the B3LYP/6–311þþG south to north: the southern slope zone, the central structural zone, the
(d,p) level of theory by using the Gaussian 09 program (Fig. 2). Fre­ central fault trough zone and the northern step-fault zone. Moreover, a
quency outcomes were examined at the same level of theory to confirm series of sub-structural can be further identified mainly including
stationary points as zero imaginary frequencies. The thermodynamic Huangtong sag, Bailian sag, Yong’an anticline, etc.
properties of CA isomers were obtained including entropy, enthalpy, Four tectonic events occurring during the late Mesozoic to Neogene
electron energy, internal energy and Gibbs free energy. controlled the consequent tectonic evolution and sedimentary filling of
the Beibuwan Basin (Gong, 1997; Qiu and Gong, 1999; Liu et al., 2014).
2.3. MD simulations The Shenhu orogeny in the Late Cretaceous leaded to the formation of a
number of grabens and half-grabens. The Zhuqiong orogeny during the
All MD calculations were carried out with the consistent-valence Paleocene and Eocene have great effects on the formation of rifting,
force fields (cvff) (Dauber-Osguthorpe et al., 1988) by using Forcite governing the formation of the current configuration of the Beibuwan
code. Temperature was kept at 298 K for all simulations. The model of Basin. Fluvial and alluvial red pebbled sandstones and mudstones of the
H2O cell adopts the previous work by our research group (Li et al., Changliu Formation (E1c) were deposited during the early rifting period.
2018). In addition, α-quartz-(001) surface is investigated to probe the Subsequently, organic-rich mudstones of the Liushagang Formation
effect of small molecule adsorption. To achieve stable structures of (E2l) were deposited in a suite of well-developed lakes in the late rifting
adsorption systems were performed with a geometry optimization, fol­ stage, which formed the most significant source rock in the Beibuwan
lowed by MD simulations including (1) the simulation was run in an Basin. During the early Nanhai orogeny, the area was mainly filled with
NVT ensemble to ensure that the systems are at the correct temperature; lacustrine sediments and swamps, resulting in that the Weizhou For­
(2) after equilibration, the systems were run in the NVE ensemble. In all mation (E3w) was dominated by organic-rich mudstones with thin coal
MD simulations, 2 ns simulations were conducted to relax the systems interbeds. During Dongsha orogeny, block faulting controlled the basin
fully with 1 fs stepsize. In order to avoid interactions between adjacent evolution accompanied by extrusive volcanism (Liu et al., 2015).
structures for both H2O cell and α-quartz-(001) surface, vacuum thick­ The major discoveries in the Fusan Depression include condensate
ness set as 80 Å. The interaction energies between H2O/α-quartz-(001) and light oils from the Huachang uplift, the Bailian fault blocks and
surface and CA homologues (CAH) were calculated by using the Jinfeng faulted nose structure, and normal gravity oils in the Meitai fault
following equation, blocks in the first (E2l1) and third (E2l3) members of the Liushagang
Formation. Besides, heavy oils from the Yong’an and Bohou-Chaoyang
EIn ¼ (ETotal – ESurface – ECAH) (1) oil pools in the Weizhou Formation (E3w) were also discovered (Li
et al., 2008, 2009). Although the oils from different regions have a wide
Where ETotal is the energy of H2O/α-quartz-(001) surface and the CAH,
variety of physical property, the oils in this study sampled from the
ESurface is the energy of H2O/α-quartz-(001) surface without the CAH,
Huachang Oilfield are dominated by light oils and condensate with an
and ECAH is the energy of the CAH without the H2O/α-quartz-(001)
average gravity of 45 (oAPI). The density of these samples is commonly
surface.
around 0.8200 g/cm3 (Li et al., 2008).
In the Fushan Depression, the major source of oil and gas are the
3. Geological background black mudstone and shale of the Liushagang Formation occurring in
Huangtong and Bailian sags (Li et al., 2008; Liu et al., 2012). These
The Fushan Depression, located in the northern of the Hainan Island lacustrine fine-grained rocks contain on average 1.5% total organic
and southern of the Qiongzhou Strait, is a typical Mesozoic-Cenozoic carbon mainly consisting of type II and III kerogens (Ding et al., 2003),
“half-graben” rift in the southeastern of the Beibuwan Basin (Li et al., and the burial depth of which exceeds 3000 m, mainly within the hy­
2008). The Fushan Depression was developed on the basement of the drocarbon generative window.
Paleozoic strata and the Mesozoic intermediate acidic igneous rocks
(Chen et al., 1991). It is a NE-E trending depression bounded by the
Changliu Fault to the east, the Anding Fault to the south and the Lingao
Fault to the northwest, respectively, which is filled with over 9000 m of
Cenozoic sediments with an area of 2920 km2 (Fig. 3) (Luo and Pang,

Fig. 2. The optimized geometries of (a) 1-MCA, (b) 2-MCA, (c) 3-MCA, (d) 4-MCA, (e) 1,8-DMCA, (f) 1,4-DMCA, (g) 1,7-DMCA and (h) 2,5-DMCA at the B3LYP/
6–311þþG(d,p) level.

3
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Fig. 3. The schematic map showing the location and regional geological background (after Yang et al., 2016).

4. Results and discussion petroleum secondary migration are two primary potential
geo-chromatographic effects. The former is regarded to be of minor
4.1. Chemical structures and migration fractionation effects of alkylated importance with respect to primary migration owing to the low
CAs in petroleum permeability of source rocks and less water involved relatively in prin­
cipal phase of petroleum generation (Krooss et al., 1991). Previous
Adsorption, partition, ionic exchange and size exclusion are possible studies suggested that molecular fractionation mainly occurred in aro­
mechanisms resulting in geo-chromatography during subsurface matic compounds, especially nitrogen-containing compounds, whereas
migration of oils, of which adsorption is considered to be the prevailing that of saturated hydrocarbons is slight (Chakhmakhchev et al., 1983;
process (Carlson and Chamberlain, 1986; Krooss et al., 1991). Brother et al., 1991).
Liquid-liquid (resembling oil-pore water) and liquid-solid (resembling Migration fractionation effects occur for carbazole and their de­
oil-water free mineral) geo-chromatographic fractionation occurring in rivatives during the oil migration process. Because the nitrogen atom is

4
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

more electronegative than the carbon atom resulting in the electron Table 1
density on the CA molecule is shared unevenly between the N and C Connolly surface (Occupied Volume and Surface Area), interaction energies,
atoms. The N atom creates an active site on these molecules to enhance dipole moment and Gibbs free energies of carbazole, C1-CA and C2-CA isomers.
their polarity. Due to the polarity of CAs molecules, CAs tend to interact Carbazole Connolly Surface μ(Debye) Eln(kcal/mol) △G(kcal/
with other polar molecules in surrounding media via hydrogen bond Vo(Å3) Sa(Å2) H2O SiO2
mol)
during primary oil migration out of the source rocks and secondary
CA 163.01 181.71 1.6692 19.23 8.93
migration in the carrier beds (Dorbon et al., 1984; Li et al., 1995; Silli­
1-MCA 182.23 198.93 1.9995 19.78 6.08 0.0004
man et al., 2002). And the hydrogen bond mentioned above is formed by 3-MCA 182.04 200.50 1.5116 13.86 10.37 0.0010
the interaction between an H atom of the N–H functional group on 2-MCA 181.96 200.51 1.7942 17.59 8.77 0.0000
pyrrolic ring and atoms with higher electronegativity in organic matter 4-MCA 181.35 195.95 1.3244 14.91 10.50 0.0001
or on the surface of silica. Consequently, the amount of CAs theoretically 1,8- 200.69 215.93 2.2870 20.63 10.07 0.0006
DMCA
decreases along with oil migration. 1,3- 200.76 217.60 1.8846 17.62 10.82 0.0014
Alkylated carbazoles can be categorized into three groups, G1 (pyr­ DMCA
rolic N–H shielded), G2 (pyrrolic N–H partially shielded) and G3 (pyr­ 1,6- 200.90 217.59 1.7942 15.33 11.06 0.0012
rolic N–H exposed), respectively, based on the present or absence of DMCA
1,7- 200.66 217.58 2.0470 20.85 10.08 0.0006
alkyl substituent(s) at C-1 and C-8 positions which have different
DMCA
chemical properties due to the steric hindrance of the alkyl substituents 1,4- 198.30 213.76 1.6586 20.00 10.62 0.0020
(Li et al., 1995). The study by Yamamoto (1992) for migration frac­ DMCA
tionation of alkylbenzoquinolines suggested the preferential migration 1,5- 200.20 213.01 1.6534 17.79 10.34 0.0020
of nitrogen-masked isomers (NMIs) with the increasing of migration DMCA
2,6- 200.57 219.26 1.5207 18.79 11.23 0.0011
distances is attributed to NMIs’ weak adsorption on clay minerals. 1, DMCA
8-Dimethylcarbazole (1,8-DMCA) is an NMI and should have a weaker 2,7- 200.31 219.25 1.7613 16.91 10.67 0.0000
interaction with polar molecules in the carrier beds relative to NEIs, such DMCA
as 2,5-dimenthylcarbazole (2,5-DMCA). Therefore, the relative abun­ 1,2- 200.09 214.70 2.1554 16.69 10.89 0.0031
DMCA
dance of 2,5-DMCA to 1,8-DMCA could theoretically decrease with
2,4- 200.21 214.79 1.4841 19.96 11.17 0.0013
increasing migration distance. DMCA
2,5- 199.84 214.73 1.4732 18.36 11.64 0.0017
4.2. Molecular simulation of geo-chromatographic fractionation of DMCA
2,3- 199.56 216.07 1.7347 17.79 10.47 0.0042
alkylated CAs
DMCA
3,4- 199.22 211.70 1.1471 15.16 10.69 0.0065
The surface characteristics of molecules play significant roles in the DMCA
interaction between the molecules and the surrounding medium. Con­
nolly Surface (Connolly, 1983) were calculated with the optimization of
the equilibrium structure and by Atom Volumes & Surface (AVS) func­ (Yamamoto, 1992) and pyrrolic nitrogen compounds (Li et al., 1995)
tion in Materials Studio. The computed configuration followed: the grid with increasing migration distance. The authors contributed the mo­
spacing is set to 0.15 Å, van der Waals scale factor is set to1.00 Å, the lecular fractionation mainly to their stronger adsorption with clay
isosurface value is set to 0.05 Å, and the water probe radius is set to 1.5 mineral. To gain more insight into the interaction between CAs and
Å. (Liu et al., 2016). The calculated results and charge distributions are formation water/clay minerals (water molecules, α-quartz-(001) surface
shown in Table 1 and Fig. 4 (the negative partial charges are marked in as simulation models, respectively), molecular dynamics simulation
blue, while the positive partial charges are marked in yellow). As seen were performed in this study. Although there are also feldspar, clay
from the computed results, carbazole has the minimum area and volume minerals and rock fragments as well as the other matter in sandstone,
of Connolly Surface comparing to methlycarbazoles and dime­ quartz is the main component that causes the migration fractionation
thylcarbazoles, and it observed that slight increase in the value of the effect. Therefore, only the quartz is placed in the model considering the
area and volume of Connolly Surface with the increasing of substituents complexity of the simulation calculation and the occupied computing
numbers. However, there is no significant differences among carbazole resources. As shown in Table 1 and Fig. 5, for MCA isomers, the inter­
isomers with the same number of substituents. action energy (EIn) with α-quartz-(001) surface of 1-MCA is the mini­
The dipole moment can characterize the polarity of a molecule mum ( 6.08 kcal/mol), while that of 4-MCA is the maximum ( 10.50
linking the microcosmic and macroscopic performance properties. The kcal/mol). Therefore, the 1-MCA isomer, which has a relatively larger
calculated dipole moment of carbazole here is 1.67 Debye which is close dipole moment hasn’t gained a larger interaction energy on the
to an experimental value (1.7 Debye) (Shimamori and Sato, 1994). α-quartz-polarity surface comparing with the 4-MCA of a relatively
Therefore, the chosen calculation method in this study is quite reason­ smaller dipole moment. The methyl on 1-MCA can shield the hydrogen
able. The position of the substituents has significant effect on the dipole bond formation of N–H⋯O (O is the atom on α-quartz-surface), which
moment. Table 1 shows that 1-methylcarbazole (1-MCA) has the may lead to such results. This is why 4-MCA would be preferentially
maximum dipole moment value (2.00 Debye), whereas 4-methylcarba­ adsorbed on α-quartz-(001) surface during oil migration process than
zole (4-MCA) is the minimum one (1.32 Debye) among all methyl­ 1-MCA. Similarly, for DMCA isomers, the interaction energy (EIn) with
carbazole isomers. Judging from the dipole moment of α-quartz-(001) surface of 1,8-DMCA is the minimum ( 10.07 kcal/mol),
dimethylcarbazoles, 1,8-dimethylcarbazole (1,8-DMCA) yields the while that of 2,5-DMCA is the maximum ( 11.64 kcal/mol). Because
largest value (μ ¼ 2.29 Debye), whereas the smallest value (μ ¼ 1.47 two methyl groups at C-1 and C-8 positions of 1,8-DMCA totally shield
Debye) was presented in 2,5-dimethylcarbazole (2,5-DMCA). The large the N–H⋯O (O is the atom on α-quartz-surface) hydrogen bond forma­
differences in molecular polarity of alkylcarbazoles indicates there is tion, it suggests that 2,5-DMCA would be preferentially adsorbed on
significant differences in alkylcarbazoles’ ability to interact with the α-quartz-(001) surface during oil migration process than 1,8-DMCA
media of the carrier beds. Therefore, we can legitimately deduce that the though the dipole moment of 2,5-DMCA is smaller than that of 1,
relative abundance of alkylcarbazoles with different molecular polarity 8-DMCA.
in oils can be used to estimate oil migration distance. In addition, H2O cell (001) present adsorption as shown in Table 1
Previous studies elucidated a regular decline of nitrogen-exposed and Fig. 6. The EIn value of 1-MCA is 19.1824 kcal/mol while 4-MCA
relative to nitrogen-masked isomers in alkylbenzoquinolines presents a low interaction energy ( 14.9139 kcal/mol). 1,8-DMCA is

5
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Fig. 4. Connolly surface of CA and selective C1-and C2-CA isomers: (a) CA, (b) 1-MCA, (c) 1,8-DMCA, (d) 2,5-DMCA. The blue areas show the negative partial
charges, while yellow areas show the positive partial charges. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web
version of this article.)

Fig. 5. The adsorption of CA homologues on α-quartz-(001) surface (a) CA, (b) 4-MCA, (c) 1-MCA, (d) 1,8-DMCA, (e) 1,6-DMCA, (f) 2,5-DMCA.

adsorbed on the H2O cell (001) surface along with EIn of 20.6258 kcal/ filling pathways are theoretically supported.
mol, whereas 2,5-DMCA presents a lower EIn ( 18.3632 kcal/mol). It Furthermore, the dipole moment of 2,4-DMCA is the second lowest
can be possibly inferred that the larger interaction of 1-MCA and 1, 8- value after 2,5-DMCA, and the interaction energy (EIn) with α-quartz-
DMCA with water could make them easier to migrate along with for­ (001) surface of 2,4-DMCA is also close to that of 2,5-DMCA. Therefore,
mation water. it can be further proposed a potential molecular parameter, 2,4-/1,8-
Here we think that the two molecular geochemical parameters pro­ DMCA to trace the oil migration orientation. And the parameter shows
posed by previous studies, i.e. the ratio of 4-/1-MCA (4-methyl­ positive correlations with the most theoretical parameters i.e. 4-/1-MCA
carbazole/1-methylcarbazole: the relative abundance of compound and 2, 5-/1, 8-DMCA proposed above (Fig. 7). It will also be helpful for
peak area of 4-MCA to 1-MCA in m/z 181 mass chromatograms) and the tracing the oil migration orientation in the future.
ratio of 2,5-/1,8-DMCA (2,5-dimethylcarbazole/1,8-dimethylcarbazole:
the relative abundance of compound peak area of 2,5-DMCA to 1,8-
DMCA in m/z 195 mass chromatograms) to trace oil migration and

6
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Fig. 6. The adsorption of CA homologues on H2O cell (001) (a) CA, (b) 4-MCA, (c) 1-MCA, (d) 1,8-DMCA, (e) 1,6-DMCA, (f) 2,5-DMCA.

Fig. 7. Cross plots between (a) 2,4-/1,8-DMCA and 2,5-/1,8-DMCA, (b) 2,4-/1,8-DMCA and 4-/1-MCA in the Tertiary sandstone reservoir in the Fushan Depression,
Beibuwan Basin.

4.3. Application to sandstone reservoir in the Fushan oilfield of the high oleanane/C30hopane and C30diasteranes/C30hopane ratios and low
Beibuwan Basin gammacerane/C30hopane ratios, indicating an oxic depositional envi­
ronment with dominant terrigenous organic matter input (Li et al.,
4.3.1. Oil-to-oil correlation and oil families 2008). And these crude oils have a “V” shape distribution pattern of C27,
Previous studies on the oil-to-oil correlation of the Fushan oilfield C28 and C29 regular steranes, which further elucidate these samples
have shown that the crude oils from this region are characterized by clearly belong to the same oil family (Yang et al., 2016).

7
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

In present study, we also selected a total of 36 oils sampled from the


Eocene sandstone reservoir of the Liushagang Formation in the Fushan
Depression (Table 2). Based on the GC–MS analysis of these 36 oil
samples, various molecular markers such as normal alkanes, acyclic
isoprenoids and C27–C29 regular sterane could be obtained and applied
for oil-to-oil correlation and oil family classification. However, the crude
oils from the Fushan Depression are dominated by light oils and con­
densates with quite low concentrations of steroids and terpenoids. In
conventional GC–MS analysis, it is a little difficult to accurately identify
the compounds and quantify the relative abundance of individual bio­
markers due to the low signal-to-noise of the mass chromatograms.
Therefore, normal alkanes and acyclic isoprenoids were regarded as
useful geochemical parameters to classify oil family.
The cross plot of Ph/nC18 vs. Pr/nC17, can reflect a variety of
geochemical information, such as depositional environments, organic
matter type, thermal maturity and biodegradation, and can also be used
to oil-to-oil correlation analysis (Shanmugam, 1985; Yandoka et al.,
2015). In the cross plot of Ph/nC18 vs. Pr/nC17 ratios, all the oil samples
were clustered in one zone suggesting that the relative source rocks were Fig. 8. Cross plots of pristane to nC17 alkane (Pr/nC17) vs. phytane to nC18
deposited under an oxic environment with dominant terrigenous alkane (Ph/nC18) of crude oils from the Tertiary sandstone reservoir in the
organic matter input except for two samples, wells Jf6 (2734.0–2738.0 Fushan Depression Beibuwan Basin (after Shanmugam, 1985).
m) and Jfn1(2213.5–2216.0 m) with relatively higher values of Pr/nC17
and Ph/nC18, indicating they may be at lower maturity (Fig. 8). northeast to southwest in the Huachang oil/gas field and towards north
to south in the Bailian oilfield in general. Therefore, the overall migra­
4.3.2. Oil charging orientation and filling pathway tion orientation is generally identical and the major source kitchen is
We preliminarily applied the three proposed molecular parameters predicted to be located in the Bailian Sag. In addition, the Huangtong
relative to methylcarbazoles and dimethylcarbazoles to trace the di­ Sag may have minor contributions to oil wells in the west of the Hua­
rection of oil migration. The values of 4-/1-MCA, 2,5-/1,8-DMCA and chang oilfield. There are three main oil filling points around well H3 in
2,4-/1,8-DMCA ratios for wells in Fushan Depression are listed in Huachang oil/gas field, well L3 in the Bailian oilfield and well Hd1 in
Table 2. Fig. 9 shows the isopleth map of 4-/1-MCA ratios. The Huadong region, respectively.
decreasing direction of 4-/1-MCA values shows a tendency towards The isopleth map of 2,5-/1,8-DMCA ratios from the data of 36 oils

Table 2
Selected molecular geochemical parameters for oils from the Fushan Depression, Beibuwan Basin.
3
Well Depth (m) Formation Density (g cm ) 4-/1-MCA 2,5-/1,8-DMCA 2,4-/1,8-DMCA Pr/nC17 Ph/nC18 MPI-1

H1-1 3347.2–3401.1 E2l3 0.83 0.51 0.47 0.48 0.39 0.11 0.88
H2 2970.6–3014.0 E2l3 0.79 0.43 0.39 0.40 0.39 0.12 1.18
H2-1 2909.8–2954.6 E2l3 0.78 0.42 0.41 0.39 0.50 0.13 0.83
H2-2 2973.0–3021.0 E2l3 0.80 0.50 0.50 0.48 0.46 0.12 0.77
H2-3 2994.6–3084.0 E2l3 0.78 0.44 0.41 0.41 0.41 0.11 0.82
H3 3015.0–318206 E2l3 0.78 0.67 0.57 0.54 0.31 0.10 0.76
H3-1 3949.6–3286.2 E2l3 0.80 0.53 0.54 0.53 0.39 0.12 0.79
H3-2 3087.0–3144.6 E2l3 0.79 0.54 0.54 0.51 0.40 0.13 0.78
H3-3 3158.6–3221.0 E2l3 0.79 0.48 0.49 0.48 0.38 0.12 0.77
Hx4 3253.0–3259.6 E2l3 0.82 0.46 0.50 0.47 0.51 0.13 0.75
H5 2715.8–2731.0 E2l3 0.77 0.40 0.34 0.30 0.61 0.15 0.64
H7 3516.6–3521.0 E2l3 0.81 0.84 0.52 0.59 0.28 0.08 0.64
H4-3 3015.4–3060.6 E2l3 0.81 0.40 0.42 0.45 0.46 0.12 0.77
H10 2961.6–2981.8 E2l3 0.76 1.04 1.03 1.01 0.41 0.15 0.97
H2-18a 3151.6–3160.0 E2l3 n.d. 0.82 0.74 0.75 0.45 0.12 0.82
H3-12x 3194.4–3238.8 E2l3 0.79 0.68 0.47 0.48 0.35 0.11 0.78
H3-5 3109.1–3128.2 E2l3 0.76 0.83 1.23 1.06 0.46 0.16 0.84
H3-6 3114.4–3148.2 E2l3 0.77 0.94 1.10 1.03 0.45 0.14 0.84
H8-1 3585.2–3669.2 E2l3 0.79 0.61 0.57 0.65 0.38 0.12 1.16
H7-3 3141.4–3539.6 E2l3 0.80 0.57 0.49 0.57 0.37 0.13 1.16
Hd4-2 3818.7–3826.7 E2l3 n.d. 0.51 0.47 0.50 0.33 0.11 1.17
Hd5 3467.0–3473.4 E2l3 0.82 0.58 0.52 0.56 0.38 0.11 1.34
Hd2 2542.0–2563.0 E2l3 0.83 0.37 0.35 0.40 0.66 0.18 0.78
Hd1 3329.4–3372.8 E2l3 0.79 1.02 0.96 0.97 0.49 0.16 1.09
Hd1-1 3347.0–3384.0 E2l3 0.80 1.06 1.18 1.17 0.41 0.14 1.05
L1 2921.6–2934.8 E2l3 0.79 0.42 0.39 0.44 0.51 0.14 0.92
L11 2842.4–2846.0 E2l3 0.82 0.40 0.32 0.44 0.56 0.15 0.71
L1-2x 2795.0–2823.4 E2l3 0.79 0.47 0.32 0.36 0.50 0.15 1.26
L4 2678.4–2698.2 E2l3 0.75 0.39 0.40 0.41 0.40 0.09 1.35
L8x 2459.4–2474.4 E2l3 0.83 0.52 0.48 0.49 0.57 0.16 0.99
L9x 2140.0–2146.8 E2l3 0.82 0.51 0.39 0.48 0.63 0.18 0.93
L3 3390.6–3410.0 E2l3 n.d. 0.53 0.60 0.46 0.44 0.13 1.00
L12 2794.8–2801.0 E2l3 0.86 0.37 0.32 0.32 0.60 0.15 0.70
Jf4 1969.4–1986.8 E2l3 0.81 0.37 0.30 0.34 0.68 0.11 0.73
Jf6 2734.0–2738.0 E2l3 0.82 0.43 0.44 0.41 1.12 0.34 0.56
Jfn1 2213.5–2216.0 E2l3 0.84 0.36 0.31 0.35 0.85 0.22 0.79

8
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Fig. 9. The isopleth map of 4-/1-methylcarbazole ratios showing oil migration orientation and filling pathways of the Tertiary sandstone reservoir in the Fushan
Depression, Beibuwan Basin.

samples is illustrated in Fig. 10. The decreasing direction and the pro­ along well H3. Thus, it can be seen that three main oil filling points are
jecting locus of 2,5-/1,8-DMCA ratios indicates the oil migration direc­ also located around wells Hd1, L3 and H3, respectively.
tion and preferential filling pathways. The overall migration orientation From the results of molecule simulation, the properties of 2,4-DMCA
in the study area is from north to south: one filling pathway is charged to is close to 2,5-DMCA, and 2,4-/1,8-DMCA shows positive correlations
Huachang oil/gas field along wells Hd1-1, Hd1 and H2-18a; and the with the parameters i.e. 4-/1-MCA and 2,5-/1,8-DMCA. Therefore, we
other is charged to Balian oilfield along wells L3, L8x and L9x. Besides, proposed 2,4-/1,8-DMCA as a potential oil migration tracer and drawn
in the west area of Huachang oil/gas field, some oil wells may be the isopleth map of 2,4-/1,8-DMCA ratios in the oils (Fig. 11). The
charged partly from the Huangtong Sag from northwest to southeast isopleth trend and distribution of three molecular parameters relative to

Fig. 10. The isopleth map of 2,5-/1,8-dimethylcarbazole ratios showing oil migration orientation and filling pathways of the Tertiary sandstone reservoir in the
Fushan Depression, Beibuwan Basin.

9
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Fig. 11. The isopleth map of 2,4-/1,8-dimethylcarbazole ratios showing oil migration orientation and filling pathways of the Tertiary sandstone reservoir in the
Fushan Depression, Beibuwan Basin.

CAs are generally consistent, so the overall migration orientation and hydrocarbon maturity parameters, appearing to be as useful as vitrinite
filling pathways based on the 2,4-/1,8-DMCA ratios are similar to that of reflectance for maturity assessment (Radke et al., 1982; Radke, 1988).
the 4-/1-MCA and 2,5-/1,8-DMCA ratios. And the tracing results of the Cross-plots showing the relationship between 4-/1-MCA, 2,5-/1,
CAs are also in consistence with those of other molecular markers (Li 8-DMCA, 2,4-/1,8-DMCA and MPI-1 are illustrated in Fig. 12a, b and c,
et al., 2008, 2018; Yang et al., 2016). respectively. As shown in Fig. 12, there is no certain relationship be­
tween these three parameters and MPI-1, respectively, suggesting that
thermal maturity have no significant effect on these molecular
4.4. Effects of source input, depositional environment, biodegradation and
geochemical indicators.
maturity
Moreover, we further calculated Gibbs free energies of alkylated
carbazole isomers/homologues in order to illustrate the relationship
The source input and depositional environment have considerable
between the migration indicators and thermal maturity. Gibbs free en­
impact on the occurrence and distributions of CAs in oils and sedimen­
ergy (G ¼ H – TS) is a thermodynamic potential including the enthalpy
tary organic matter (Clegg et al., 1997; Bakr and Wilkes, 2002). The
effect and the entropy effect, which can be applied to describe the sta­
precondition of using molecular geochemical parameters to trace oil
bility of a molecule. Generally speaking, the lower potential energy of a
migration orientation is that all oils should belong to the same oil family.
system, the more stable it is. As can be seen from Table 1, the relative
In this study, all oils samples are derived from the same source kitch­
values of G (△G) could indicate a stable molecule. △G is defined as,
en/bed belonging to a single oil family as mentioned above in section
4.3.1. Therefore, the effects of source input and depositional environ­ △G ¼ Gi – Glowest (2)
ment for composition of CAs were negligible in this study.
Previous studies indicated that a slight biodegradation in reservoir Where Gi is a G of any of several isomers, Glowest is a G of the lowest
affects little to the carbazole and alkylcarbazoles in the crude oils. With molecule among isomers. For example, 2-MCA has a lowest (△G2-MCA ¼
the increasing of the biodegradation, the concentrations of carbazole G2-MCA – G2-MCA) ¼ 0.00 kcal/mol, whereas the △G for MCA isomers is
and alkylcarbazoles decrease for mid-degree biodegraded oils and relative to 2-MCA. The same situation is occurring in other isomers.
sharply decrease in the heavily biodegraded oils (Zhang et al., 1999; Gibbs free energies of MCA isomers and DMCA isomers here indicate no
Huang et al., 2003). However, the oils samples from Tertiary sandstone significant differences respectively (Table 1) which indirectly means
reservoirs in the Fushan Depression were not subjected to apparent they have little effect on thermal maturity.
biodegradation (Yang et al., 2016). In consequence, the influence of Therefore, the proposed migration orientation molecular parameters
biodegradation on the distribution of CAs can be neglected. in this study are principally controlled by migration fractionation effect.
In addition, the distribution of carbazoles are also related to maturity
differences. (Li et al., 1997; Clegg et al., 1998b; Zhang et al., 2011). The 5. Conclusions
study by Li et al. (1997) suggested that concentrations of various pyr­
rolic nitrogen compounds increase with increasing thermal maturity, By using density functional theory (DFT) and molecular dynamics
which can provide background information for calibrate such data in (MD) simulations, this study investigated the potential mechanisms for
migrated petroleum. The study by Zhang et al. (2011) showed the fractionation effect of alkylated carbazoles during oil migration process.
relative distributions of C1-, C2-carbazoles and benzocarbazoles dis­ Molecular dynamics simulations show that the differences of interaction
played “two-stage” variation in a wide maturity range indicating the energies between alkylated carbazoles and α-quartz/water of carrier
different geochemical controls between the immature and mature beds lead to their migration fractionation effect. For methylcarbazole
stages. The methylphenanthrene index (MPI-1) is the common aromatic (MCA) isomers, 4-methylcarbazole (4-MCA) has a much higher

10
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

2,5-/1,8-DMCA and 2,4-/1,8-DMCA are considered as effective migra­


tion indicators to trace oil migration orientation and filling pathways.
And three migration parameters were also successfully applied in
lacustrine sandstone reservoir in the Fushan Depression, Beibuwan
Basin.
The calculations of Gibbs free energy show that there are no signif­
icant differences of thermal dynamical stabilities among MCAs and
DMCAs, respectively, which means the proposed parameters are less
affected by thermal maturity. In this study, all oils samples are derived
from the same source kitchen/bed belonging to the same oil family. And
previous studies indicated that these oils samples were not subjected to
apparent biodegradation. Therefore, the variations of 4-/1-MCA, 2,5-/
1,8-DMCA and 2,4-/1,8-DMCA in petroleum during subsurface migra­
tion are mainly controlled by migration fractionation in this region.

Declaration of competing interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

CRediT authorship contribution statement

Qiuya Han: Data curation, Writing - original draft. Meijun Li:


Conceptualization, Writing - review & editing, Project administration.
Xiaoqiang Liu: Software, Methodology. Weidong Jiang: Visualization,
Resources, Investigation. Shengbao Shi: Methodology. Youjun Tang:
Funding acquisition. Daxiang He: Investigation.

Acknowledgements

This work was funded by the National Natural Science Foundation of


China [Grant No.41972148], the National Key R & D Program of China
(Grant No. 2017YFC0603102) and the Foundation of the State Key
Laboratory of Petroleum Resources and Prospecting, China University of
Petroleum-Beijing (No. PRP/open-1503). The authors are grateful for
the assistance of Lei Zhu and Jianfeng Zhang for the GC–MS analysis. All
calculated results were supported by Sichuan University of Science &
Engineering High Performance Computing Center for providing
computational.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.


org/10.1016/j.petrol.2020.107308.

References

Bakr, M.M.Y., Wilkes, H., 2002. The influence of facies and depositional environment on
the occurrence and distribution of carbazoles and benzocarbazoles in crude oils: a
case study from the Gulf of Suez, Egypt. Org. Geochem. 33, 561–580.
Bowler, B.F.J., Larter, S.R., Clegg, H., Wilkes, H., Horsfield, B., Li, M., Chem, A., 1997.
Dimethylcarbazoles in crude oils: comment on "liquid chromatographic separation
schemes for pyrrole and pyridine nitrogen aromatic heterocycle fractions from crude
oils suitable for rapid characterization of geochemical samples. Anal. Chem. 69,
3128–3129.
Fig. 12. Cross-plots showing the relationship between 4-/1-MCA, 2,5-/1,8- Brother, L., Engel, M.H., Krooss, B.M., 1991. The effects of fluid flow through porous
DMCA, 2,4-/1,8-DMCA and MPI-1 of crude oils from the Tertiary sandstone media on the distribution of organic compounds in a synthetic crude oil. Org.
reservoir in the Fushan Depression, Beibuwan Basin. Geochem. 17, 11–24.
Carlson, R., Chamberlain, D., 1986. Steroid biomarker-clay mineral adsorption free
energies: implications to petroleum migration indices. Org. Geochem. 10, 163–180.
adsorption energy with α-quartz-(001) surface than 1-methylcarbazole Chakhmakhchev, V., Punanova, S., Zharkov, N., 1983. Percolation of oil and changes in
its composition in porous media (based on experimental studies). Int. Geol. Rev. 25,
(1-MCA). And for dimethylcarbazole (DMCA) isomers, the interaction
1223–1228.
energy (EIn) with α-quartz-(001) surface of 2,5-dimethylcarbazole (2,5- Chen, M.-x., Xia, S.-g., Yang, S.-z., 1991. Local geothermal anomalies and their formation
DMCA) is the maximum, while that of 1,8-dimethylcarbazole (1,8- mechanisms on Leizhou Peninsula, South China. Sci. Geol. Sin. 4, 369–383.
DMCA) is the minimum. The interaction energy (EIn) with α-quartz- Clegg, H., Horsfield, B., Wilkes, H., Damst�e, J.S., Koopmans, M.P., 1998a. Effect of
artificial maturation on carbazole distributions, as revealed by the hydrous pyrolysis
(001) surface of 2,4-dimethylcarbazole (2,4-DMCA) is close to that of of an organic-sulphur-rich source rock (Ghareb Formation, Jordan). Org. Geochem.
2,5-dimethylcarbazole (2,5-DMCA). 29, 1953–1960.
Three theoretically proven migration parameters, i.e. 4-/1-MCA,

11
Q. Han et al. Journal of Petroleum Science and Engineering 191 (2020) 107308

Clegg, H., Wilkes, H., Horsfield, B., 1997. Carbazole distributions in carbonate and clastic Li, M., Yao, H., Stasiuk, L.D., Fowler, M.G., Larter, S.R., 1997. Effect of maturity and
source rocks. Geochem. Cosmochim. Acta 61, 5335–5345. petroleum expulsion on pyrrolic nitrogen compound yields and distributions in
Clegg, H., Wilkes, H., Oldenburg, T., Santamaría-Orozco, D., Horsfield, B., 1998b. Duvernay Formation petroleum source rocks in central Alberta, Canada ☆. Org.
Influence of maturity on carbazole and benzocarbazole distributions in crude oils Geochem. 26, 731–744.
and source rocks from the Sonda de Campeche, Gulf of Mexico. Org. Geochem. 29, Liu, E.-t., Wang, H., Lin, Z.-l., Li, Y., Ma, Q.-l., 2012. Characteristics and hydrocarbon
183–194. enrichment rules of transfer zone in Fushan Sag, Beibuwan Basin. J. Cent. S. Univ.
Connolly, M.L., 1983. Solvent-accessible surfaces of proteins and nucleic acids. Science 43, 3946–3953.
221, 709–713. Liu, E., Wang, H., Li, Y., Leonard, N.D., Feng, Y., Pan, S., Xia, C., 2015. Relative role of
Dauber-Osguthorpe, P., Roberts, V.A., Osguthorpe, D.J., Wolff, J., Genest, M., Hagler, A. accommodation zones in controlling stratal architectural variability and facies
T., 1988. Structure and energetics of ligand binding to proteins: Escherichia coli distribution: insights from the Fushan Depression, South China Sea. Mar. Petrol.
dihydrofolate reductase-trimethoprim, a drug-receptor system. Proteins-Struct. Geol. 68, 219–239.
Funct. Bioinform. 4, 31–47. Liu, E., Wang, H., Li, Y., Zhou, W., Leonard, N.D., Lin, Z., Ma, Q., 2014. Sedimentary
Ding, W., Wang, W., Ma, Y., 2003. Characteristics of Liushagang Formation petroleum characteristics and tectonic setting of sublacustrine fans in a half-graben rift
system in fushan depression of Beibuwan Basin. Offshore Oil 23, 1–6. depression, Beibuwan Basin, South China Sea. Mar. Petrol. Geol. 52, 9–21.
Dorbon, M., Schmitter, J.M., Garrigues, P., Ignatiadis, I., Ewald, M., Arpino, P., Liu, X.Q., He, X., Qiu, N.X., Yang, X., Tian, Z.Y., Li, M.J., Xue, Y., 2016. Molecular
Guiochon, G., 1984. Distribution of carbazole derivatives in petroleum. Org. simulation of CH4, CO2, H2O and N2 molecules adsorption on heterogeneous surface
Geochem. 7, 111–120. models of coal. Appl. Surf. Sci. 389, 894–905.
Duin, A.C.T.V., Larter, S.R., 1998. Application of molecular dynamics calculations in the Luo, Q., Pang, X., 2008. Reservoir controlling mechanism and petroleum accumulation
prediction of dynamical molecular properties. Org. Geochem. 29, 1043–1050. model for consequent fault and antithetic fault in Fushan Depression of Hainan area.
Gong, Z.S., 1997. Major Oil and Gas Fields of China Offshore. Petroleum Industrial Acta Pet. Sin. 29, 363–367.
Publisher. Ma, Q., Zhao, S., Liao, Y., Lin, Z.-l., 2012. Sequence architectures of paleogene
Harrison, E., Telnaes, N., Wilhelms, A., Horsfield, B., Van Duin, A., Bennett, B., Larter, S., Liushagang Formation and its significance in fushan sag of the Beibuwan Basin.
1997. Maturity Controls on Carbazole Distributions in Coals and Source Rocks. 18th Earth Sci. J. China Univ. Geosci. 37, 667–678.
International Meeting on Organic Geochemistry, Maastricht, pp. 235–236. Qiu, Z., Gong, Z., 1999. Offshore. In: Oil Exploration in China, vol. 4. Geological
Horsfield, B., Clegg, H., Wilkes, H., Santamaria-Orozco, D., 1998. Effect of maturity on Publishing House and Petroleum Industry Publishing House, Beijing.
carbazole distributions in petroleum systems: new insights from the Sonda de Radke, M., Welte, D.H., Willsch, H., 1982. Geochemical study on a well in the Western
Campeche, Mexico, and Hils Syncline, Germany. Naturwissenschaften 85, 233–237. Canada Basin: relation of the aromatic distribution pattern to maturity of organic
Huang, H., Bowler, B.F.J., Zhang, Z., Oldenburg, T.B.P., Larter, S.R., 2003. Influence of matter. Geochem. Cosmochim. Acta 46, 1–10.
biodegradation on carbazole and benzocarbazole distributions in oil columns from Radke, M., 1988. Application of aromatic compounds as maturity indicators in source
the Liaohe basin, NE China. Org. Geochem. 34, 951–969. rocks and crude oils. Mar. Petrol. Geol. 5, 224–236.
Krooss, B.M., Brothers, L., Engel, M.H., 1991. Geochromatography in petroleum Shanmugam, G., 1985. Significance of coniferous rain forests and related organic matter
migration: a review. Geol. Soc. Londn. Special Publ. 59, 149–163. in generating commercial quantities of oil, Gippsland Basin, Australia. AAPG (Am.
Larter, S., Bowler, B., Clarke, E., Wilson, C., Moffatt, B., Bennett, B., Yardley, G., Dan, C., Assoc. Pet. Geol.) Bull. 69, 1241–1254.
2000. An experimental investigation of geochromatography during secondary Shimamori, H., Sato, A., 1994. Dipole moments and lifetimes of excited triplet states of
migration of petroleum performed under subsurface conditions with a real rock. aniline and its derivatives in nonpolar solvents. J. Phys. Chem. 98, 13481–13485.
Geochem. Trans. 1, 54. Silliman, J.E., Li, M., Yao, H., Rong, H., 2002. Molecular distributions and geochemical
Larter, S.R., Bowler, B.F.J., Li, M., Chen, M., Brincat, D., Bennett, B., Noke, K., implications of pyrrolic nitrogen compounds in the Permian Phosphoria Formation
Donohoe, P., Simmons, D., Kohnen, M., 1996. Molecular indicators of secondary oil derived oils of Wyoming. Org. Geochem. 33, 527–544.
migration distances. Nature 383, 593–597. Snyder, L.R., 1965. Distribution of benzcarbazole isomers in petroleum as evidence for
Later, D.W., Lee, M.L., Bartle, K.D., Kong, R.C., Vassilaros, D.L., 1981. Chemical class their biogenic origin. Nature 205, 277 277.
separation and characterization of organic compounds in synthetic fuels. Anal. Tissot, B., Welte, D., 1980. Petroleum Formation and Occurrence, vol. 16. Springer-
Chem. 53 (11), 1612–1620. Verlag, pp. 372–373. Earth Sci. Rev.
Li, Meijun, Wang, Tieguan, Liu, Ju, Lu, Hong, Wu, Weiqiang, 2008. Occurrence and Wang, T.G., Li, S.M., Zhang, S.C., 2004. Oil migration in the Lunnan region, Tarim Basin,
origin of carbon dioxide in the fushan depression, Beibuwan Basin, south China sea. China based on the pyrrolic nitrogen compound distribution. J. Petrol. Sci. Eng. 41,
Mar. Petrol. Geol. 25, 500–513. 123–134.
Li, M., Larter, S., 1993. Interactions of organic nitrogen species with mineral water and Wilhelms, A., Telnaes, N., Larter, S., Steen, A., Bowler, B., Ronningsland, T., 1997.
organic networks: implications to petroleum geochemistry. In: 205th American Application of Migration Indicators (Benzocarbazoles) in Unraveling of the Filling
Chemical Society Meeting (Denver), Division of Geochemistry, American Chemical History of the Oseberg Field, North Sea. th International Meeting on Organic
Society (Abstract Paper). Geochemistry, Maastricht, pp. 22–26.
Li, M., Larter, S., Stoddart, D., Bjoroy, M., 1992. Practical liquid chromatographic Yamamoto, M., 1992. Fractionation of azaarenes during oil migration. Org. Geochem.
separation schemes for pyrrolic and pyridinic nitrogen aromatic heterocycle fraction 19, 389–402.
from crude oils suitable for rapid characterization of geochemical samples [J]. Anal. Yan, H., Yuan, S., 2016. Molecular dynamics simulation of the oil detachment process
Chem. 64, 1337–1344. within silica nanopores. J. Phys. Chem. C 120, 2667–2674.
Li, M., Larter, S.R., Frolov, Y.B., Bjoroy, M., 1994. Adsorptive interaction between Yandoka, B.M.S., Abdullah, W.H., Abubakar, M., Hakimi, M.H., Adegoke, A.K., 2015.
nitrogen compounds and organic and/or mineral phases in subsurface rocks. Models Geochemical characterisation of early cretaceous lacustrine sediments of bima
for compositional fractionation of pyrrolic nitrogen compounds in petroleum during formation, Yola sub-basin, northern benue trough, NE Nigeria: organic matter input,
petroleum migration. J. Separ. Sci. 17, 230–236. preservation, paleoenvironment and palaeoclimatic conditions. Mar. Petrol. Geol.
Li, M., Larter, S.R., Stoddart, D., Bjoroy, M., 1995. Fractionation of pyrrolic nitrogen 61, 82–94.
compounds in petroleum during migration: derivation of migration-related Yang, L., Li, M., Wang, T.G., Shi, Y., 2016. Dibenzothiophenes and
geochemical parameters. Geol. Soc. Londn. Special Publ. 86, 103–123. benzonaphthothiophenes in oils, and their application in identifying oil filling
Li, M., Liu, X., Wang, T.G., Jiang, W., Fang, R., Yang, L., Tang, Y., 2018. Fractionation of pathways in Eocene lacustrine clastic reservoirs in the Beibuwan Basin, South China
dibenzofurans during subsurface petroleum migration: based on molecular dynamics Sea. J. Petrol. Sci. Eng. 146, 1026–1036.
simulation and reservoir geochemistry. Org. Geochem. 115, 220–232. Zhang, C., Zhang, Y., Cai, C., 2011. Maturity effect on carbazole distributions in source
Li, M., Osadetz, K., Fowler, M., Snowdon, L., Stasiuk, L., Yao, H., Hwang, R., Jenden, P., rocks from the saline lacustrine settings, the western Qaidam Basin, NW China.
Grant, B., Idiz, E., 1998a. CASE STUDIES ON SECONDARY OIL MIGRATION IN THE J. Asian Earth Sci. 42, 1288–1296.
WILLISTON BASIN1. Williston Basin Symposium. Zhang, C., Zhang, Y., Zhang, M., Zhao, H., Cai, C., 2008. Carbazole distributions in rocks
Li, M., Wang, T., Liu, J., Zhang, M., Lu, H., Ma, Q., Gao, L., 2009. Biomarker 17α(H)- from non-marine depositional environments. Org. Geochem. 39, 868–878.
diahopane: a geochemical tool to study the petroleum system of a Tertiary lacustrine Zhang, C., Zhao, H., Mei, B., Chen, M., Xiao, Q., Wu, T., 1999. Effect of biodegradation
basin, Northern South China Sea. Appl. Geochem. 24, 172–183. on carbazole compounds in crude oils. Shiyou yu tianranqi dizhi (Oil Gas Geol.) 20,
Li, M., Yao, H., Fowler, M.G., Stasiuk, L.D., 1998b. Geochemical constraints on models 341–343.
for secondary petroleum migration along the Upper Devonian Rimbey-
Meadowbrook reef trend in central Alberta, Canada. Org. Geochem. 29, 163–182.

12

You might also like