You are on page 1of 13

Marine and Petroleum Geology 105 (2019) 32–44

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Research paper

Estimation of thermal maturity in the Bakken source rock from a T


combination of well logs, North Dakota, USA
Arash Abarghania,∗, Mehdi Ostadhassana, Bailey Bubacha, Peiqiang Zhaob,a
a
Department of Petroleum Engineering, University of North Dakota, Grand Forks, ND, 58202, United States
b
Institute of Geophysics and Geomatics, China University of Geosciences, Wuhan, 430074, China

ARTICLE INFO ABSTRACT

Keywords: Numerous attempts have been conducted to construct a reliable model to relate petrophysical and geomecha-
Bakken source rock nical characteristics of shale to geochemical properties. In this study, continuous logs were built to predict the
Bitumen reflectance thermal maturity of the Bakken Formation from a set of petrophysical logs. First, dynamic modulus (Ed) was
Thermal maturity created from shear and compressional sonic and, density logs and then converted to Ei (nanoindentation based
Wireline logs
Young's modulus) using the existing experimental relationship between Ed and Ei. Next, continuous TOC logs
were generated for the Bakken Shales based on an empirical relationship between well log derived Ei and TOC
contents that were measured in the lab. There was found an acceptable difference between estimated log-based
TOC and measured values in three separate wells. Solid bitumen reflectance as a reliable indicator of thermal
maturity was measured on all samples from the same wells in the scarcity/absence of vitrinite maceral. The
relationships between GR/NPHI/RHOZ wireline logs versus solid bitumen reflectance (BRO%) as the maturity
index was then investigated in twelve wells. Based on the strength or weight of the relationship between each log
and BRO%, a new parameter “Ʈ” was defined to represent thermal maturity from well logs. A good agreement
between parameter Ʈ and BRO% values was observed which provided an empirical equation to make the esti-
mation of BRO% values for the whole length of the shale members possible. Ultimately, the continuity of BRO%
values, would enable us to establish a relationship between BRO% and measured Tmax values to generate con-
tinuous Tmax logs for the Bakken Shale members. Using these newly developed well logs, it would become
possible to make a 3D property models for BRO% and Tmax (the thermal maturity property cube) for an accurate
petroleum system evaluation based on GR/NPHI/RHOZ logs.

1. Introduction (150 ft) in the central part of the basin (LeFever et al., 1991). The very
high gamma-ray response is the most recognizable petrophysical fea-
The Bakken Formation in the Williston Basin is one of the most ture and characteristic of the Bakken Formation (> 200 API units) from
important shale plays in North America. The Late Devonian - Early the Upper and Lower Shale members (Webster, 1984). Therefore, such
Mississippian Bakken overlies portions of North Dakota, Montana and features of the Bakken are widely used for the correlation purposes
the Canadian Manitoba and Saskatchewan (Fig. 1). It consists of four across the Williston Basin. Considering the fact that, there are lots of
members; Upper and Lower members which are known as the Bakken wells in the Bakken Formation with a complete set of petrophysical log
source rock and comprise of black shales, rich in organic matters, and data, we believe, if we can manage to establish a robust model to relate
the middle and Pronghorn members that are the reservoir rocks and petrophysical properties from the logs to the main geochemical char-
mainly composed of fine-grained carbonate-rich sandstone and siltstone acteristics of the source rock such as total organic carbon (TOC) or
with very low porosity and permeability. Bakken Formation is under- thermal maturity, it will be possible to construct continuous geo-
lain by Three Forks Formation and overlain by Lodgepole Formation chemical logs which lead to a 3D model for the source rock thermal
(Webster, 1984; Meissner, 1991; Khatibi et al., 2018; Abarghani et al., maturity and richness evaluations, basin wise. This model can help to
2018). Bakken Formation has no surface outcrops, and drill cuttings better understand source rock and carry out more accurate petroleum
and core samples are the only sources for the Bakken studies. In North system evaluation and finally reserve estimation.
Dakota, the Bakken Formation has a maximum thickness of 46 m


Corresponding authors.
E-mail addresses: arash.abarghani@und.edu (A. Abarghani), mehdi.ostadhassan@und.edu (M. Ostadhassan).

https://doi.org/10.1016/j.marpetgeo.2019.04.005
Received 3 August 2018; Received in revised form 1 April 2019; Accepted 5 April 2019
Available online 09 April 2019
0264-8172/ © 2019 Elsevier Ltd. All rights reserved.
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 1. Study area in North Dakota portion of the Williston Basin and wells location.

2. Methods and materials one to another, and create continuous geochemistry property models.
Furthermore, it may occur occasionally that there might not be ade-
Twelve wells were selected across North Dakota portion of the quate access to the potential source rock samples in a basin (frontiers
Williston Basin to improve spatial resolution. All wells contain a full set especially), or there isn't any published data to describe the source rock
of wireline logs including GR/NPHI/RHOZ (gamma ray, neutron por- with good areal extensions throughout the basin (Dembicki, 2016). In
osity, and density). Out of twelve wells, three also had shear and such scenarios, wireline logs are raised as the only source of data for
compressional sonic logs which were used to calculate geomechanical source rock interpretation. In this study, we tried to employ formation
parameters. A total of forty core samples of the Bakken Formation was density log (RHOZ), neutron porosity log (NPHI), GR log, shear and
collected from the North Dakota Geological Survey Wilson M. Laird compressional sonic log (DTSH3 and DTCO3, respectively) as well as
Core and Sample Library located on the University of North Dakota geochemical data and solid bitumen reflectance to construct a con-
campus. All samples were analyzed as the bulk powder using Basic/ tinuous geochemical log for the Bakken Formation. The main goal was
Bulk-Rock method by Rock-Eval 6 pyrolysis in order to obtain geo- to establish a method to investigate the thermal maturity of the Bakken
chemical parameters. In this technique, about 60 mg of ground bulk Formation based on wireline logs. Using acoustic velocity measure-
samples were heated for 3 min isothermally in an unreactive gas at- ments of studied wells with a complete set of wireline log data (wells
mosphere in the pyrolysis chamber and then, temperature was in- no. 5, 9, and 11) including compressional (DTCO3) and shear (DTSH3)
creased with a rate of 25 °C/min up to 650 °C. Residual carbon from the sonic and density log (RHOZ) and using empirical relations to estimate
pyrolysis stage were transferred into the oxygen oven and was burnt in dynamic elastic properties, “Eq. (1)” (Fjar et al., 2008), the dynamic
the oxidation stage. Using the IFP's (Institute France du Petrole) tra- elastic modulus was determined for each well:
demark Basic/Bulk-Rock method, geochemical parameters such as TOC
(total organic carbon) content, HI (hydrogen Index) and Tmax were 3 ts 2 4 tp 2
Ed = 2 2
calculated (Table 1). A detailed description of the programmed pyr- ts ts tp 2 (1)
olysis technique can be found in Behar et al. (2001). Polished epoxy-
mounted whole-rock blocks were prepared for all samples to carry out where Ed is the dynamic elastic modulus, ρ is the formation density, Δts
BRO% (bitumen reflectance) measurements in the scarcity/absence of and Δtp are the shear and compressional sonic travel time, respectively.
primary vitrinite maceral. BRO% was measured using a LEICA DM Shukla et al. (2013) developed a set of empirical relationships between
2500-P microscope equipped with a J&M photometer TIDAS MSP-200 geomechanical data (nanoindentation Young's modulus – Ei) and other
under TSOP/ICCP protocols and suggested procedures with a Sapphire properties such as porosity, mineralogy and total organic hydrocarbon
and a GGG (GadoliniumeGallium-Garnet) standard of 0.589% RO and (TOC) for various shale plays such as Barnett, Collingwood, Eagle Ford,
1.716% RO, respectively. Haynesville, Kimmeridge, Wolfcamp, and Woodford. They argued that
the nanoindentation method is an effective way to evaluate Young's
modulus in shales. They found a good agreement between na-
3. Results and discussion noindentation measured Young's modulus (Ei) and calculated dynamic
Young's modulus (Ed) using acoustic velocities and density in the above-
Numerous studies have been conducted by several researchers to mentioned shale plays. In order to compare the Bakken's wireline based
establish a robust relationship between geochemistry data and petro- Ed values with other shale plays, calculated Ed values were transformed
physical and geomechanical parameters (e.g., Schmoker and Hester, to Ei through the following empirical relations “Eq. (2)”presented by
1983; Passey et al., 1990; Aoudia et al., 2010; Kumar et al., 2012; Shukla et al. (2013):
Bocangel et al., 2013; Shukla et al., 2013; Dietrich, 2015). These cor- Ed = 0.97 × Ei + 1.91 (2)
relations are necessary to calibrate field data, predict properties from

33
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Table 1 Then, we established a new model between resulted Ei values and


Rock-Eval 6 pyrolysis Tmax, TOC and HI results for the studied samples, mea- Rock-Eval pyrolysis derived TOC (wt.%) for 18 core samples from the
sured BRO and Parameter “Ʈ” values. Bakken from wells no. 5, 9, and 11 (Fig. 2). Considering the residual
Well No. Member Depth Tmax TOC HI BRO Parameter Ʈ plots in Fig. 3 that are obtained from regression models of Fig. 2, it can
be seen that, although regression model is representing a low correla-
(ft.) °C wt.% mg HC/g % – tion coefficient but analysis of the various residual plots of the regres-
TOC
sion model confirm that the data distribution is normal and can be
1 L. Bakken 3774.0 420 3.53 611 0.38 2.59 utilized in this study for prediction of TOC amounts that are estimated
3785.0 410 14.93 526 0.37 6.75 with an acceptable margin of error. Therefore, it would be possible to
3 U. Bakken 5996.0 425 16.29 550 0.47 10.77 generate continuous TOC logs for the shale members of the Bakken
5999.0 426 7.33 491 0.53 2.69
based on wireline derived geomechanical data (Fig. 5). Similar to other
4 U. Bakken 8325.0 428 16.27 557 0.44 6.24
5 U. Bakken 7337.5 438 12.1 470 0.51 8.11 shale plays (Barnett, Collingwood, Eagle Ford, Haynesville, Kimmer-
7342.7 437 10.86 434 0.51 7.42 idge, Wolfcamp, and Woodford - Shukla et al., 2013) Young's modulus
L. Bakken 7419.0 435 17.13 428 0.52 7.95 values decrease with increase in TOC amounts which are confirmed in
7431.1 435 13.06 431 0.52 12.74 several studies for examining the effect of organic matter on Young's
6 U. Bakken 7625.5 436 17.82 465 0.47 1.91
modulus with nanoindentation. Additionally, the Bakken exhibits more
7631.0 431 21.22 481 0.43 4.65
L. Bakken 7707.2 437 14.65 392 0.44 6.14 amounts of TOC (wt.%), and smaller Young's modulus values (GPa)
7718.6 435 14.01 479 0.43 8.32 compared to other studied shale plays (Fig. 4).
7 U. Bakken 8203.2 431 18.07 569 0.47 8.01 Estimating the total organic carbon content of the source rock is the
8207.5 425 20.98 558 0.47 12.17
first step in source rock richness or petroleum system evaluation. The
L. Bakken 8279.0 431 13.17 494 0.44 7.76
8291.7 431 14.61 559 0.44 6.11
next step and the most important stage is to investigate the thermal
8 L. Bakken 8987.5 435 6.09 395 0.46 4.67 maturity of the source rock which is a governing factor to decide if the
8995.5 434 14.03 453 0.50 7.45 source rock has produced any hydrocarbons. In this regard, various
9 U. Bakken 10433.0 448 11.92 276 0.74 3.42 techniques are suggested by different researchers to quantify thermal
10435.7 447 14.93 316 0.77 4.19
maturity of the source rock such as the measurement of vitrinite, bi-
10436.4 447 12.69 260 0.76 4.72
10437.0 450 13.39 349 0.85 4.42 tumen or zooclasts reflectance, programmed pyrolysis (e.g., Rock-Eval
L. Bakken 10527.0 448 16.66 270 0.83 7.00 pyrolysis), thermal alteration index, the crystallinity of clay minerals,
10541.0 448 11.16 255 0.81 3.56 fluid inclusion analysis, liptinite maceral group fluorescence (e.g.,
10547.0 448 13.13 253 0.74 4.47
Kübler, 1967; Staplin, 1969; Alpern et al., 1978; Stach et al., 1982;
10555.0 449 13.26 257 0.72 5.38
10 L. Bakken 9886.0 432 15.76 531 0.49 5.07
Espitalie, 1986; Sweeney and Burnham, 1990; Mukhopadhyay and
11 U. Bakken 11158.0 452 12.31 177 0.99 3.28 Dow, 1994; Lowrie et al., 1996). However, all of these methods would
11159.0 453 9.34 171 0.94 2.97 provide us with discrete data points, are exhaustive and require good
11162.0 451 10.42 176 0.98 3.27 accessibility to the samples. Wireline logs are another important source
L. Bakken 11201.0 452 16.36 171 1.00 6.83
of data to get valuable information about the thermal maturity of
11203.0 451 9.08 151 0.95 3.81
11205.0 451 9.77 151 0.92 2.38 source rocks. They are important since unlike all above-mentioned
12 U. Bakken 9447.5 426 12.66 592 0.47 6.03 approaches which provide discrete data points, wireline logs could
9453.5 424 14.80 615 0.42 5.92 generate continuous data. This continuous data can later be used as an
L. Bakken 9503.5 425 17.32 563 0.44 11.53
input for establishing 3D property models for a better insight about the
9507.1 422 19.78 572 0.25 18.19
14 U. Bakken 9670.4 429 11.11 550 0.48 2.69
source rock in the basin scale.
9673.3 428 6.99 473 0.46 3.07 Resistivity and sonic wireline logs have frequently been used for
interpreting thermal maturity in source rocks. Meissner (1978) used

Fig. 2. The relationship between nanoindentation Ei and Rock-Eval pyrolysis derived TOC.

34
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 3. Residual plots and details of statistical analysis for the regressed linear trend between Ei and TOC.

Fig. 4. Indentation Young's modulus (Ei) versus TOC for the Bakken samples and other major shale plays (modified after Shukla et al., 2013). Note that, the Bakken
samples exhibit higher values of TOC and smaller Young's modulus compared to other shale plays.

resistivity logs from the Bakken to differentiate immature areas/inter- (1990) proposed a technique based on sonic transit time and resistivity
vals and generate mature zones by mapping the resistivity values based logs (Δlog R overlay technique) to quantitatively estimate of TOC (wt.
on the idea that resistivity would increase as a result of hydrocarbon %). Experiments demonstrated this approach is an effective technique
generation. Later as a follow up to this method, it was explained that for predicting TOC (wt.%) for maturity levels less than about LOM 10.5
expulsion of hydrocarbons out of kerogen can lead to replacement with (Level of Organic Metamorphism - a measure of thermal maturity).
formation salt water if porosity is good enough (Dembicki, 2016) which However, this would underestimate TOC (wt.%) at higher maturities
can influence the resistivity response of the source rock. Passey et al. (Passey et al., 2010; Dembicki, 2016). A combination of formation

35
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 5. Continuous TOC logs and core samples TOC values for the whole length of the Upper and Lower members for wells no. 5, 9 and 11.

density and neutron porosity logs with resistivity were used by Hinds vitrinite reflectance for the Cambay Basin in India. Another correlation
and Berg (1990) for predicting thermal maturity of the Upper Cretac- between vitrinite reflectance and acoustic transit time, formation den-
eous Austin Chalk Formation in Texas Gulf Coast. There was not any sity, resistivity, and neutron porosity log were presented by Ge and Li
estimation of thermal maturity in this study and only a qualitative (2003) for determining the maturity in coal beds. Same combination
description was presented. (resistivity/formation density/neutron porosity) were used by Zhao
Other researchers (e.g., Smagala et al., 1984; Schmoker and Hester, et al. (2007) for calculating vitrinite reflectance and interpreting
1990) tried to establish a correlation between vitrinite reflectance (as a thermal maturity for the Barnett Shale. However, this study was not
major indicator of thermal maturity - Mukhopadhyay and Dow, 1994) able to establish a continuous VR log for the whole thickness of the
and wireline resistivity logs (Dembicki, 2016). Similar attempts were formation. Zhao et al. (2015) established a relationship between for-
proposed by Lang (1994) and McTavish (1998) to estimate thermal mation density and neutron porosity logs to determine the coal ranks
maturity based on the relationship between vitrinite reflectance and (equal to thermal maturity in the source rock). While, this method is not
sonic wireline log. Lang (1994) argued that it would be possible to applicable to shales due to a significant difference in geochemical
predict source rock maturity based on sonic travel time difference characteristics between coal and shale (Zhao et al., 2019). Continuous
considering that both sonic transit time and thermal maturity should evaluation of vitrinite reflectance for shale gas reservoirs is presented
increase with depth. The same approach was employed by Tyagi et al. by Jin (2015) based on the relationship between total porosity and
(2011) for setting up a correlation between acoustic transit time and vitrinite reflectance in the Jiaoshiba area. Kadkhodaie and Rezaee

36
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

(2017) used sonic time difference and resistivity logs to define a new
index (RRS) which illustrates that, an increase in vitrinite reflectance
should exhibit an increase in acoustic transit time and oil saturation as
well. The vitrinite reflectance index (RRS) was then defined based on
the sonic transit time and the reciprocate of the resistivity index.
Though, considering the fact that organic-rich shale resistivity is not
defined only by oil saturation, but TOC contents too, some limitations
on its general application is expected. Collectively, as Dembicki (2016)
has discussed, independent models should be generated locally for each
source rock because of the complexity of parameters that can affect
thermal maturity and also various parameters which may influence
wireline logs recordings.
To make a reliable database as a reference for comparing log de-
rived data from the thermal maturity point of view, forty core samples
were selected from the Bakken Upper and Lower shales, and after
preparing crushed rock pellets, reflectance measurements were carried
out on them. In the scarcity/absence of vitrinite maceral in the samples,
bitumen particles reflectance was considered as thermal maturity in-
dicator. Vitrinite is derived from high plants, therefore vitrinite will be
scarce or absent especially in Lower Paleozoic and also in some marine
strata such as the Bakken. Many studies (e.g., Gentzis and Goodarzi,
1990; Khorasani and Michelsen, 1993; Mukhopadhyay, 1994; Landis
and Castaño, 1995; Schoenherr et al., 2007; Kelemen et al., 2010;
Mastalerz et al., 2018) argued that solid bitumen can be used as the Fig. 7. SGR log of the Well No. 14. Uranium clearly has a larger contribution in
the Bakken gamma-ray emission compare to Thorium.
indicator of maturity in the scarcity/absence of vitrinite. However, due
to some reasons such as the presence of multiple populations of solid
bitumen in the studied samples or variation in texture and morphology, reflected light microscopy (Fig. 6), at least 50 measurements were done
employing bitumen reflectance as thermal maturity should be exercised on each sample and results were reported as BRO% (Table 1).
with cautious (Mastalerz et al., 2018). Selecting an appropriate group of The cross-plots of measured BRO% with GR, NPHI, and RHOZ log
bitumen population in samples based on petrographic studies under the values from twelve wells are shown in Fig. 8 (A, B, and C). These figures

Fig. 6. Solid bitumen was used in the scarcity/absence of vitrinite maceral. Photomicrographs of the Bakken solid bitumen samples. (A) Elongated grain of solid
bitumen (BRO,ran = 0.45%), Well #14 depth 9670.4 ft (B) Solid bitumen (BRO,ran = 0.53%), Well #6 depth 7631 ft; (C) Blocky grain of solid bitumen
(BRO,ran = 0.62%), Well #5 depth 7419 ft; (D) Solid bitumen particle with relatively high reflectance (BRO,ran = 0.76%), Well #6 depth 7707.2 ft. The red square in
the middle of each image is a scale of 5 μm of each side. All photomicrographs were taken using a 50× oil immersion objective. (For interpretation of the references
to colour in this figure legend, the reader is referred to the Web version of this article.)

37
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 8. Cross-plots of BRO% versus NPHI (A), GR (B) and RHOZ (C). The relationship between BRO% and parameter “Ʈ” (D).

represent that measured BRO% values have a relatively good agreement Bakken. Kerogen type II has a moderately high H/C and moderate O/C
with NPHI and RHOZ logs but a weak relationship with GR log. How- atomic ratios and produces mainly naphthenic oil (Dembicki, 2016).
ever, statistical analysis of the residual plots (Fig. 9) is not exhibiting Increasing thermal maturity, aliphatic components (rich in hydrogen)
any specific patterns and therefore data distribution is normal. In- content start to decrease due to contrithe bution of hydrogen atoms in
creasing thermal maturity (high BRO% values), hydrocarbons are gen- the generated and expelled petroleum from the kerogen composition,
erated and expelled, and therefore TOC and hydrogen index while aromatic components (rich in oxygen) show an increasing trend
(HI = 2TOC ) amounts in source rocks decreases (Daly and Edman, (Khatibi et al., 2018). Considering the fact that hydrogen will decrease
S × 100

1987; Dembicki, 2016). As a result of this process, RHOZ values start to in kerogen chemical composition during thermal maturity advance-
increase with an increase in BRO% values. Considering organic matter ment, neutron log starts to read lower amounts of hydrogen index. This
lower density (∼0.9–1.1 g/cc) compared to shale's mineral matrix being said, the observed trend between BRO% as an indicator of thermal
(∼2.25 g/cc), formation density log (RHOZ) could be used for esti- maturity and neutron log value (Fig. 8, A) could be verified. GR values
mating organic matter content of the source rock zones (Meyer and demonstrate a weak relationship with BRO%, though, collectively, there
Nederlof, 1984; Dembicki, 2016). NPHI values also decrease due to is a decrease in BRO% with increasing GR values. The gamma-ray log
petroleum generation and expulsion from the source rock. In this pro- measures the natural gamma ray emitted by Uranium, Thorium, and
cess, the hydrogen contents will be decreased, and, as a result, smaller Potassium. Generally, organic-rich source rocks should high gamma-ray
values will be recorded by neutron porosity (NPHI) log. Neutron por- emissions due to their high uranium (U) contents. This is explained by
osity log hydrogen index (IH) which is an indicator of the formation the existence of suitable conditions for uranium concentrations in the
richness of hydrogen content is the main factor considered by neutron depositional environments of organic matters under anoxic/euxinic
log and defined as the hydrogen content of the formation (wt.%) to the circumstances (Fertl and Chilingar, 1988; Dembicki, 2016). The Spec-
weight (%) of hydrogen in water. Fresh water has a hydrogen index of tral Gamma-Ray log of well no. 14 indicates that uranium has a large
1.0. Organic matter has a high hydrogen index and an increase in contribution in the Bakken gamma-ray emission compared to Th
neutron log, especially when combined with decreasing in density log (Fig. 7). Therefore, the increasing values in the bulk gamma-ray could
(due to organic matter's low density compared to surrounding minerals) infer to an increase in the total organic matter. Considering, the higher
could be a very effective approach for recognizing organic rich intervals amounts of TOC in the Bakken immature samples (with lower BRO%
in the sedimentary sequences (Rider, 1986; Dembicki, 2016). Previous values), this decrease in BRO% values which takes place with an in-
studies (e.g., Webster, 1984; Osadetz et al., 1992; Osadetz and crease in gamma-ray values could be well justified.
Snowdon, 1995; Jin and Sonnenberg, 2014; Khatibi et al., 2018; The correlation coefficient (R2) represents the degree of the re-
Abarghani et al., 2018) indicate that, amorphous oil-prone marine type lationship between two parameters in cross-plot diagrams. A higher
II kerogen is the most abundant constituent organic matter in the correlation coefficient value refers to a stronger agreement between

38
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 9. Residual Plots of statistical analysis for the regressed line between BRO and NPHI, GR, RHOZ, and Ʈ presented in Fig. 8. No specific pattern is recognized in the
residual plot diagrams.

related parameters if the primary reason for the relationship is valid. In between Ʈ and BRO%. It is found that “Ʈ” values decrease with in-
our study, based on the above discussion this relationship can be ex- creasing BRO% values in the Bakken samples. Using obtained empirical
plained as GR and NPHI values both decrease with thermal advance relationship between measured BRO% and “Ʈ”, now continuous logs of
which is equivalent to the increase of RHOZ values. This happens when estimated BRO% for the entire section of the Upper and Lower shale
hydrocarbons expel from the OM which causes an increase in formation members of the Bakken formation with a relatively high degree of ac-
density. Therefore, based on equation (3), when the attempt is to relate curacy can be generated, as seen in Fig. 10 (A to G) for wells no. 1, 3, 5,
multiple log values such as GR/NPHI/RHOZ that are correlated to a 6, 7, 8 and 12 with the Bakken depth less than 10000 ft.
separate parameter (BRO%) individually, multiplication of R2 (corre- However, some errors in estimated values for wells with depths
lation between each log and BRO%) to each log value (GR/NPHI/ above than 10000 ft were identified. As it could be seen in Fig. 8-D,
RHOZ) would represent the weight (strength) of each correlation (GR- there is a steep slope in the regression line between parameter “Ʈ” and
BRO, NPHI-BRO and RHOZ-BRO) in overall estimation of the new measured BRO% > 0.7. This is inferred that as the maturity increases
parameter (Ʈ). Considering different observed correlations between and the source rock enters the oil window (Tmax about 435 °C) a rapid
BRO % with GR/NPHI/RHOZ logs separately, we decided to define a increase in BRO% values occurs (Fig. 11-A). The previous study by the
new empirical parameter based on the weight of the relationships be- authors (Abarghani et al., 2018) indicates that such BRO% values could
tween these three logs individually and BRO% value. In the next step, be detected in wells where the Bakken depths is more than 10000 ft.
depending on the type of the relationship between each log with BRO%, Furthermore, the rapid decrease in HI value amounts is also visible for
a new empirical parameter “Ʈ” can be defined as follows: samples with depth beyond 10000 ft (Fig. 11-B). To remedy this pro-
blem and decrease the error margin in the estimated BRO% values, a
constant should be added to the conversion equation. Using true mea-
2 2
(RGR BRo × GRValue) × (RNPHI BRo × NPHIValue )
= 2
(RRHOZ BRo × RHOZ Value ) (3) sured BRO% values, it was decided to add 0.2% per 1000 ft for the wells
where the Bakken depths becomes deeper than 10000 ft. Applying this
where, RGR2
BRo is the correlation coefficient between GR and measured conversion coefficient to the correlation equation for well 9 and 11
BRO% (equal to 0.0474 for this study), RNPHI2
BRo and RRHOZ BRo are the
2
where the Bakken is deeper than 10000 ft (Fig. 10H and I). has led to a
correlation coefficients between NPHI and RHOZ with measured BRO%, good estimation of BRO% in these two wells as well.
respectively (equal to 0.3183 and 0.2177, respectively for this study). Ʈ Cross-plot of measured BRO% and Tmax for 40 samples of the Bakken
is a parameter showing the overall relationship between GR, NPHI and cores (Fig. 12), confirms a good agreement between these two im-
RHOZ at each depth related to organic matters concentration. Ʈ is portant parameters for interpreting thermal maturity. Converting esti-
larger in intervals with the higher concentration of organic matter, mated BRO% values for the entire section of the Upper and Lower Shale
where higher GR and NPHI values and lower RHOZ values are re- members to the equivalent Tmax values enables us to generate con-
corded. tinuous Tmax logs for each shale members in all wells and evaluate
We then, calculated parameter “Ʈ” for forty samples and cross- thermal maturity in a continuous format. In the next step, using these
plotted it versus measured BRO% (Fig. 8, D) to find out the relationship continuous logs, we can make 3D property models for BRO% and Tmax

39
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 10. Continuous BRO% logs for the whole length of the Upper and Lower members of the Bakken Formation. Very good agreement between estimated and real
measured values for BRO% is seen for wells no. 6, 5, 8, 1, 7, 11, and 9. Acceptable values were estimated for wells no. 12 and 3.

to generate a cube of thermal maturity for source rock evaluation based could be calculated in a similar approach for other source rocks in
on GR/NPHI/RHOZ logs in basin scale. It is noteworthy to mention various basins.
that, due to the complexity of the various parameters which can affect
BRO% and Tmax in different basins, we cannot develop a universal es- 4. Conclusion
timation from this method and it exclusively should be used for the
Bakken Shale members. However, the newly defined parameter "Ʈ" In this study, we used to shear and compressional sonic logs and

40
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 10. (continued)

formation density log to calculate dynamic Young's modulus (Ed). Using relationship between our newly defined parameter and measured
experimental relationship developed for other major shale plays such as BRO%. Resulted logs are in good agreement with true measured BRO%
the Barnett, Collingwood, Eagle Ford, Haynesville, Kimmeridge, values and could be used directly for thermal maturity interpretation or
Wolfcamp and Woodford shales, calculated Ed values were converted to conversion to Tmax values. This was done through the correlation that
equivalent indentation Young's modulus (Ei). Continuous TOC logs was obtained between these two parameters for the Bakken core sam-
were generated based on the relationship between Ei and core samples ples. Finally, this enables us to perform thermal maturity interpretation
Rock-Eval 6 pyrolysis derived TOC values. In the scarcity/absence of based on estimated Tmax values. Using this new technique, it is possible
vitrinite maceral, solid bitumen reflectance measurement as a reliable to utilize continuous BRO% or Tmax data in generating thermal maturity
thermal maturity indicator was carried out on all core samples. 3D models and provide a better insight of the source rock in the basin
Relationships between BRO% and GR/NPHI/RHOZ wireline logs were scale. Although this method exclusively has developed for the Bakken
investigated and based on the strength in the relationship between each shale members, the procedure for developing parameter “Ʈ” could be
log and BRO% values; a new empirical parameter “Ʈ” was defined. used for any other source rock to create a parameter specific to that
Continuous BRO% logs for the whole length of the Upper and Lower basin.
members of the Bakken Formation were created based on the empirical

41
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 10. (continued)

42
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Fig. 11. Cross-plots of Tmax and Hydrogen Index vs depth. There is a clear boundary between immature and mature samples on ∼435 °C after a depth of 10000 ft (A).
A fast decrease in HI values could be seen after this depth in the samples (B).

Fig. 12. The observed relationship between real measured BRO% and Rock-Eval 6 pyrolysis derived Tmax.

43
A. Abarghani, et al. Marine and Petroleum Geology 105 (2019) 32–44

Acknowledgment hydrocarbons. Org. Geochem. 271, 137–149.


Lang, W.H., 1994. The determination of thermal maturity in potential source rocks using
interval transit time/interval velocity. Log. Anal. 35, 47–59.
The authors wish to thank North Dakota Geological Survey, Core LeFever, J.A., Martiniuk, C.D., Dancsok, E.F., Mahnic, P.A., 1991. Petroleum potential of the
Library, for giving us access to the Bakken core samples, particularly middle member, Bakken formation, Williston Basin. In: Williston Basin Symposium.
Jeffrey Bader, state geologist as well as Kent Hollands, lab technician. Lowrie, A., Hamiter, R., Fogarty, M.A., Orsi, T., Lerche, I., 1996. Thermal and time-
temperature index (TTI) patterns during geologic evolution of North and central Gulf
We'd also like to express our appreciation to Dr. Thomas Gentzis and Dr. of Mexico. In: AAPG Bulletin, 80(9), Abstracts of Meeting: AAPG Gulf Coast
Humberto Carvajal-Ortiz from Core Laboratories in Houston, TX, for Association of Geological Societies, San Antonio, Texas, United States, October 2–4,
providing us with the geochemistry data and fruitful discussions on 1996. American Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 1508.
Mastalerz, M., Drobniak, A., Stankiewicz, A.B., 2018. Origin, properties, and implications
solid bitumen reflectance. Sincere thanks to anonymous reviewers for of solid bitumen in source-rock reservoirs: a review. Int. J. Coal Geol. 195, 14–36.
constructive comments which improved the quality of the manuscript. McTavish, R.A., 1998. Applying wireline logs to estimate source rock maturity. Oil Gas J.
96, 76–79.
Meissner, F.F., 1978. Petroleum geology of the Bakken Formation, Williston Basin, North
Appendix A. Supplementary data
Dakota and Montana. In: Estelle, D., Miller, R. (Eds.), The Economic Geology of the
Williston Basin, 1978 Williston Basin Symposium. Montana Geological Society, pp.
Supplementary data to this article can be found online at https:// 207–230.
doi.org/10.1016/j.marpetgeo.2019.04.005. Meissner, F.F., 1991. Petroleum geology of the Bakken Formation Williston Basin, North
Dakota and Montana. Guidebook to Geology and Horizontal Drilling of the Bakken
Formation. American Association of Petroleum Geologists, pp. 19–42.
References Meyer, B.L., Nederlof, M.H., 1984. Identification of source rocks on wireline logs by
density/resistivity and sonic transit time/resistivity crossplots. AAPG (Am. Assoc.
Pet. Geol.) Bull. 68, 121–129.
Abarghani, A., Ostadhassan, M., Gentzis, T., Carvajal-Ortiz, H., Bubach, B., 2018. Mukhopadhyay, P.K., 1994. In: Vitrinite Reflectance as Maturity Parameter: Petrographic
Organofacies study of the Bakken source rock in North Dakota, USA, based on organic and Molecular Characterization and its Applications to Basin Modeling, vol 570
petrology and geochemistry. Int. J. Coal Geol. 188, 79–93. American Chemical Society, Symposium Series.
Alpern, B., Durand, B., Durand-Souron, C., 1978. Proprieties optique de residus de la Mukhopadhyay, P.K., Dow, W.G., 1994. Vitrinite reflectance as a maturity parameter.
pryrolyse des kerogenes. Rev. Inst. Fr. Petrol 33, 867–890. Vitrinite reflectance as a maturity parameter 570, 1–24.
Aoudia, K., Miskimins, J.L., Harris, N.B., Munich, C.A., 2010. Statistical analysis of the Osadetz, K.G., Snowdon, L.R., 1995. Significant Paleozoic petroleum source rocks, their
effects of mineralogy on rock mechanical properties of the Woodford shale and the distribution, richness and thermal maturity in Canadian Williston Basin (southeastern
associated impacts for hydraulic fracture treatment design. In: Paper ARMA 10-303 Saskatchewan and southwestern Manitoba). Geol. Surv. Can. Bull. 487, 60.
Presented at the 44th US Rock Mechanics Symposium and 5th U.S. - Canada Rock Osadetz, K.G., Brooks, P.W., Snowdon, L.R., 1992. Oil families and their sources in
Mechanics Symposium, Salt Lake City, Utah, USA, pp. 27–30 (June). Canadian Williston Basin, (southeastern Saskatchewan and southwestern Manitoba).
Behar, F., Beaumont, V., Penteado, H.D.B., 2001. Rock-Eval 6 technology: performances Bull. Can. Petrol. Geol. 40 (3), 254–273.
and developments. Oil Gas Sci. Technol. 56, 111–134. Passey, Q.R., Bohacs, K., Esch, W.L., Klimentidis, R., Sinha, S., 2010. From oil-prone
Bocangel, W., Sondergeld, C., Rai, C., 2013. September. Acoustic mapping and char- source rock to gas-producing shale reservoir - geologic and petrophysical char-
acterization of organic matter in shales. In: SPE Annual Technical Conference and acterization of unconventional shale gas reservoirs. In: Society of Petroleum
Exhibition. Society of Petroleum Engineers. Engineers Conference Paper131350, Presented at the CPS/SPE International Oil &
Daly, A.R., Edman, J.D., 1987. Loss of organic carbon from source rocks during thermal Gas Conference and Exhibition, Beijing, China, June 8–10, 2010 29 pp.
maturation. AAPG Bull. 71. Passey, Q.R., Creaney, S., Kulla, J.B., Moretti, F.J., Stroud, J.D., 1990. A practical model for
Dembicki, H., 2016. Practical Petroleum Geochemistry for Exploration and Production. organic richness from porosity and resistivity logs. AAPG Bull. 74 (12), 1777–1794.
Elsevier. Rider, M.H., 1986. The Geological Interpretation of Well Logs.
Dietrich, A.B., 2015. The Impact of Organic Matter on Geomechanical Properties and Schmoker, J.W., Hester, T.C., 1983. Organic carbon in Bakken formation, United States
Elastic Anisotropy in the Vaca Muerta Shale. master’s thesis. pp. 122. portion of the Williston Basin. AAPG (Am. Assoc. Pet. Geol.) Bull. 67, 2165–2174.
Espitalie, J., 1986. Use of Tmax as a maturation index for different types of organic matter: Schmoker, J.W., Hester, T.C., 1990. Formation resistivity as an indicator of oil generation –
comparison with vitrinite reflectance. In: In: Burrus, J. (Ed.), Proceedings of Meeting: 1st Bakken formation of North Dakota and Woodford shale of Oklahoma. Log. Anal. 31, 1–9.
IFP Exploration Research Conference, Carcans, France, June 3–7, 1985, vol 44. Technip, Schoenherr, J., Littke, R., Urai, J.L., Kukla, P.A., Rawahi, Z., 2007. Polyphase thermal
Paris, France, pp. 475–496 Thermal Modeling in Sedimentary Basins. evolution in the Infra-Cambrian Ara Group (South Oman Salt Basin) as deduced by
Fertl, W.H., Chilingar, G.V., 1988. Total organic carbon content determined from well maturity of solid reservoir bitumen. Org. Geochem. 38, 1293–1318.
logs. Soc. Petrol. Eng. Form. Eval. 3, 407–419. Shukla, P., Kumar, V., Curtis, M., Sondergeld, C.H., Rai, C.S., 2013. January.
Fjar, E., Holt, R.M., Raaen, A.M., Risnes, R., Horsrud, P., 2008. Petroleum Related Rock Nanoindentation studies on shales. In: 47th US Rock Mechanics/Geomechanics
Mechanics, vol 53 Elsevier. Symposium. American Rock Mechanics Association.
Gentzis, T., Goodarzi, F., 1990. A Review of the Use of Bitumen Reflectance in Smagala, T.M., Brown, C.A., Nydegger, G.L., 1984. Log-derived indicator of thermal
Hydrocarbon Exploration with Examples from Melville Island, Arctic Canada. Rocky maturity, Niobrara formation, denver basin, Colorado, Nebraska, Wyoming. In:
Mountain Section (SEPM). Woodward, J., Meissner, F.F., Clayton, J.L. (Eds.), Hydrocarbon Source Rocks of the
Hinds, G.S., Berg, R.R., 1990. Estimating Organic Maturity from Well Logs, Upper Greater Rocky Mountain Region. Rocky Mountain Association of Geologists, pp.
Cretaceous Austin Chalk. Texas Gulf Coast. 355–363.
Jin, J., 2015. Geophysical Modeling of Shale Gas: A Case Study of Longmaxi Formation in Stach, E., Mackowsky, M-Th, Teichmüller, M., Taylor, G.H., Chandra, D., Teichmüller, R.,
Jiaoshaiba Area. Doctoral thesis. Yangtze University, Wuhan (In Chinese with 1982. Stach's Textbook of Coal Petrology, second ed. Gebrüder Borntraeger, Berlin, pp.
English Abstract). 535.
Jin, H., Sonnenberg, S.A., 2014. Characterization for Source-Rock Potential of the Bakken Staplin, F.L., 1969. Sedimentary organic matter, organic metamorphism, and oil and gas
Shale in the Williston Basin, North Dakota and Montana: AAPG Search and Discovery occurrence. Bull. Can. Petrol. Geol. 17 (1), 47–66.
Article 80356. Sweeney, J.J., Burnham, A.K., 1990. Evaluation of a simple model of vitrinite reflectance
Kadkhodaie, A., Rezaee, R., 2017. Estimation of vitrinite reflectance from well log data. J. based on chemical kinetics. AAPG (Am. Assoc. Pet. Geol.) Bull. 74 (10), 1559–1570.
Pet. Sci. Eng. 148, 94–102. Tyagi, A., Abedeen, A., Dutta, T., 2011. Petrophysical evaluation of shale gas reservoir: a
Kelemen, S.R., Walters, C.C., Kwiatek, P.J., Freund, H., Afeworki, M., Sansone, M., case study from Cambay Basin, India. In: SPE/DGS Saudi Arabia Section Technical
Lamberti, W.A., Pottorf, R.J., Machel, H.G., Peters, K.E., Bolin, T., 2010. Symposium and Exhibition, Al-Khobar, Saudi Arabia, May 15-18.
Characterization of solid bitumens originating from thermal chemical alteration and Webster, R.L., 1984. Petroleum source rocks and stratigraphy of the Bakken formation in
thermochemical sulfate reduction. Geochem. Cosmochim. Acta 74 (18), 5305–5332. North Dakota. In: RMAG Guidebook, Williston Basin, Anatomy of a Cratonic Oil
Khatibi, S., Ostadhassan, M., Tuschel, D., Gentzis, T., Bubach, B., Carvajal-Ortiz, H., 2018. Province, pp. 268–285.
Raman spectroscopy to study thermal maturity and elastic modulus of kerogen. Int. J. Xiang, G., Tiao, L., 2003. On well logging evaluation method for physical property of
Coal Geol. 185, 103–118. coalbed. Well Logging Technol. 27 (2), 129–131.
Khorasani, G.K., Michelsen, J.K., 1993. The thermal evolution of solid bitumens, bitumen Zhao, S., Liu, L., Fu, D., 2007. On early evaluation of hydrocarbon source rocks based on
reflectance, and kinetic modeling of reflectance: application in petroleum and ore seismic data. J. Oil Gas Technol. (J. Jianghan Petroleum Inst.) 29 (5), 76–79 (In
prospecting. Energy Sources 15 (2), 181–204. Chinese with English Abstract).
Kübler, B., 1967. La crystallinité de l'illite et les zones tout à fait supérieures du index of Zhao, P., Mao, Z., Zhao, Jin D., Sun, P.B., Sun, W., Pang, X., 2015. Investigation on log
illite métamorphisme. In: Etages Tectoniques (Colloque de Neuchâtel, 18-21 Avril responses of bulk density and thermal neutron in coal bed with different ranks. J.
1966). Neuchâtel University Institute Geology, Neuchâtel, pp. 105–122. Geophys. Eng. 12 (3), 477–484.
Kumar, V., Sondergeld, C.H., Rai, C.S., 2012. Nano to macro mechanical characterization Zhao, P., Ostadhassan, M., Shen, B., Liu, W., Abarghani, A., Liu, K., Luo, M., Cai, J., 2019.
of shale. In: Paper SPE 159804 Presented at the Annual Technical Conference and Estimating thermal maturity of organic-rich shale from well logs: case studies of two
Exhibition, San Antonio, Texas, USA, pp. 8–10 (October.). shale plays. Fuel 235, 1195–1206.
Landis, Ch, Castaño, J., 1995. Maturation and bulk chemical properties of a suite of solid

44

You might also like