Вы находитесь на странице: 1из 25

SPE-198129-MS

Optimization of Development of Heavy Oil Reservoirs through Geochemical


Characterization

Celal Hakan Canbaz, Ege University; Melek Deniz-Paker and Fatma Bahar Hosgor, Petroleum Experts LLC; Dike
Putra, Rafflesia Energy; Raul Moreno, Smart Recovery; Cenk Temizel and Ahmad Alkouh, College of Technical
Studies

Copyright 2019, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Kuwait Oil & Gas Conference and Show held in Mishref, Kuwait, 13 - 16 October 2019.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Geochemistry is not only a well-known tool in providing a better understanding of the distribution of fluids
in the reservoir rock but also an efficient kit in developing reservoir by decreasing the uncertainty throughout
the characterization process. Utilizing geochemistry, not only efficiently identify the fluids and type of
oil alteration drastically laterally and vertically over short distances in heavy oil reservoirs where such
differences are of significant importance in production of heavy oils in these already challenging reservoirs,
but also outline the value of geochemistry to justify the value of information in the process of more robust
reservoir characterization and management of heavy oil reservoirs.
A conceptual model representative heavy oil reservoir recovery is utilized to compare the recoveries
between a case where geochemistry is applied to characterize the reservoir and another case where
geochemical methods are not employed by using a full-physics commercial reservoir simulator. A sensitivity
and optimization software is coupled with the reservoir simulator to outline the relative significance of the
important parameters in the recovery process.
Geochemical characterization, not only, provides information on gas content and its likely behavior where
it can also lead to better decisions on completion strategies to avoid zones of different viscosity, but also
the essential correlation between the geochemistry and the thermodynamics of heavy oil. Comprehensive
reservoir characterization leads to a more robust identification of reservoir fluids where such knowledge
will greatly enhance the efficiency thus the economics of the process that is especially important in low oil
price environments. There is lack of studies recently on the application of geochemical characterization on
the recovery of the process analyzing the relative significance of components, key drivers and the value
of the information throughout the process, even though some authors have been published their research
on geochemistry and its use in the characterization of the reservoirs. Our study outlines a comprehensive
background including latest developments, investigates the key factors, and the value of information on
comparative cases considering the relevant components of the process.
2 SPE-198129-MS

Introduction
Today, with the rapid declining of conventional hydrocarbon resources together with the remote locations
and hostile environments of the world's remaining resources, expensive and less successful exploration
drilling, petroleum geochemistry is increasingly becoming vital in the oil industry with its wide modern
geochemical techniques. To improve the exploration success and production efficiency, all the components
of prospective petroleum system are needed to be characterized thoroughly and petroleum-generating
capacity of the source rock must be evaluated. Petroleum geochemistry assists reservoir characterization and
management by providing useful input for basin and petroleum system modeling with the characterization
of the elements and processes that control the richness and distribution of petroleum source rock (England,
2005; McCarthy et al., 2018; Katithi and Opar, 2016; Brooks and Welte, 1984).
Geochemistry is the study of the earth chemistry, chemical processes, cycles, reactions that govern the
composition of rocks, distribution and movement of chemical elements in both time and space (Hannigan,
2007). The inception point of petroleum geochemistry can be considered as the discovery of chlorophyll-
like organic compounds (porphyrins) in crude oil by Albert Treibs (1934). After that, major oil companies
had started to performe researches about the formation and movement of oil and gas in the subsurface and
try to prospect petroleum using source rocks (Hunt and Meinert, 1958).

Petroleum Geochemistry
Petroleum geochemistry is used at many different stages of oil field, from basin study to investigations of
petroleum properties and production evaluations. Some applications of reservoir geochemistry can be listed
as follows; determination of source rock type and maturity, identification of reservoir fluid type from rock
extracts, assessment of basin palaeohydrology/oil alteration controls and seal efficiency, determination of
fluid contact, evaluation of hydrocarbon continuity, analysis of commingled oils for production allocation,
Sw calculation, characterization of reservoir bitumens and tar mats and determination of tar sands location,
identification of potential production problems associated with deasphaltation, production monitoring to
assess production plan and EOR, evaluation of workovers and assessment of tubing leakage - mixed
production problems (Kaufman et al., 2002; Larter and Aplin, 1995).
Fig. 1 demonstrates the value of petroleum geochemistry in exploration. The study shows that exploration
efficiency was increased to 63% when the geophysical and geochemical techniques are used together
(Sluijk and Parker, 1986). From 2000 to the present, petroleum geochemistry continues to advance in
every aspect. Geochemistry can be classified as inorganic, physical and organic geochemistry. Petroleum
geochemistry is a section of organic geochemistry which focuses on carbon bearing compound found in
geologic environments. The science of petroleum geochemistry is the application of chemical principles to
the study of the origin, generation, migration, accumulation and alteration of petroleum, and the use of this
knowledge in exploration and recovery of oil and gas (Fig. 2).
SPE-198129-MS 3

Figure 1—An assessment of approaches to exploration, by comparing random drilling, exploration based on trap size from
geophysics, and the use of geophysics in conjunction with petroleum geochemistry. (Sluijk and Parker, 1986; Dembicki, 2017)

Figure 2—Major Divisions of Petroleum geochemistry (modifed from Brooks and Welte, 1984).

Source Rock
The main element of petroleum geochemistry is the source rocks. Source rocks are the sedimentary rocks
from which petroleum has been generated or contains sufficient organic matter to generate petroleum.
Sedimentary rocks contain soluble organic matter bitumen (solid vein-filling material, pitch, tar, and asphalt)
and pyrobitumen which is solidified bitumen residue that is insoluble in organic solvents, and kerogen that
represents the major organic carbon reservoir in the world. Kerogen is the particulate fraction of organic
formed from the breakdown and diagenesis of the macerals that are the components of plants, animals, and
bacteria deposited in the rock. It is economically desirable to differentiate petroleum exploration areas as
oil-prone or gas-prone. This can be achieved by evaluating the type of kerogen. Generally, hydrogen-rich
4 SPE-198129-MS

kerogens (Types I and II) are considered oil-prone, and hydrogen-poor kerogen (Type III) is considered
mainly gas-prone (Horsfield et al., 1988).
As illustrated in modified Van Krevelen diagram (Fig. 3a), there are three types of hydrocarbon-
generating kerogens are found in the environment. (Tissot et al., 1974; McCarthy et al., 2011). Type I
kerogen has a high initial H/C and low initial O/C atomic ratios and produces mainly waxy oil. Type II
kerogen has a moderately high H/C and moderate O/C atomic ratios in its initial state and produces mainly
naphthenic oil. Type III kerogen is named as humic and has a low initial H/C, high initial O/C atomic ratios
and produces mainly gas. Additionally, Type IV kerogen is inert with no hydrocarbon-generating potential.
(Tissot and Welte, 1984). The quality and type of kerogen determine petroleum potential and depends on
the burial depth and the thermal history of the rocks. Source rocks, which are found at the oil window are
thermally mature and able to generate petroleum (Fig. 3b).

Figure 3—a) Kerogen type and maturation. b) Thermal Transformation of Kerogen (McCarthy et al., 2011).

A proper evaluation of the source rocks and its contents are crucial to reach the hydrocarbons reservoirs
because the knowledge of richness or amount of organic matter, type and thermal maturity of source rock
help to locate active source rocks and predict the amount of oil and gas available in the basin. The source
rock maps are very important for the oil companies since they reduce exploration risk. (Demaison, 1984;
Peters and Cassa, 1994). The geochemical analysis is the most powerful tool for providing analytical data
to identify and map source rocks.

Source Rock Analysis


Various techniques are employed for the evaluation of the hydrocarbon generating capacity of source rock.
The primary key to an accomplished source rock study is the good quality sample taken from the sediment.
Geochemical testing of drill cuttings, canned cuttings samples, drilling mud samples, core, and sidewall core
samples is useful in the determination of the organic matter in the rock and can aid an exploration program.
After proper sampling, screening procedures such as total organic carbon (TOC), Rock-Eval pyrolysis,
visual kerogen/vitrinite reflectance analysis or gas chromatography (GC) analysis can be performed.
TOC is the amount of organic carbon present in the rock and it is the first screening parameter for
quantifying organic richness. Since oil and gas potential is related to its carbon content, it is used to assess
SPE-198129-MS 5

the quantity of organic matter. But it is not a clear indicator of petroleum potential (Dembicki, 2017). The
initial carbon assessment is generally followed by pyrolysis and vitrinite reflectance analysis. Table 1a and
1b, tabulate the geochemical parameters describing the kerogen type and the petroleum potential of source
rock as per TOC and Pyrolysis analysis.

Table 1—(a) Geochemical Parameters Describing (a) Kerogen Type (Quality) and the
Character of Expelled Products (b) the Petroleum Potential of an Immature Source Rock.

The Rock-Eval Pyrolysis analysis is considered to be the most valuable geochemical exploration tool
used to determine the kerogen type matter, thermal maturity, and the generation capability of source rocks
(Al-Areeq, 2018). It is used to establish the amount of kerogen transformed into petroleum and how much
can be transformed at a higher temperature. In this method, the rock sample is heated gradually while the
amount of hydrocarbons generated is measured. At a certain temperature, oil and gas which has already been
generated in the source rock are expelled and measured as the S1 peak. On further heating, new petroleum
is formed from the kerogen by pyrolysis and measured as S2 peak. This represents the amount of oil and
gas that could have been generated if the source rock had been buried deeper. The quantity of CO2 formed
during pyrolsis is S3 peak. S2/S3 is a gross indicator of kerogen quality dependent upon degree of maturation
and S1/S1+S2 the productivity index or the amount of free hydrocarbon in the rock sample as a function of
the total hydrocarbon possible (Bjørlykke, 2015; Clementz at al.,1979). Fig. 4 shows well site geochemical
logs obtained on sidewall cores and drill cuttings for a sample well.

Figure 4—Geochemical Well Site Logs on Sidewall Cores and Drill Cuttings (Clementz at al., 1979).
6 SPE-198129-MS

Two important parameters, hydrogen index (HI = S2/ TOC wt%, equivalent to H/C atomic ratio in van
Krevelen diagram) and oxygen index (OI = S3/TOC wt%, equivalent to O/C atomic ratio in the kerogen)
are also obtained with rock-eval pyrolysis data.
S2 parameter is generally used as the principle source richness indicator. TOC-S2 and TOC-PY cross-
plots can be used to determine hydrocarbon-generating potential. The population with the higher S2 and
PY (poyential yield) values show greater hydrocarbon-generating potential (Fig. 5). Vitrinite is a type of
kerogen particle, or maceral, formed from the thermal alteration of lignin and cellulose in plant cell walls.
It is used for assessing maturation of kerogen. The vitrinite reflectance in rock systems allows prediction
of the relative locations and depths at which undiscovered accumulations of oil versus gas are reservoired
(Fig. 6). Oil typically is generated from buried organic matter in the range of approximately 0.6–1.2% Ro
whereas gas generation generally commences at 1.0% Ro and greater (Tissot and Welte, 1984; Hunt, 1996;
Hackley and Cardott, 2016).

Figure 5—a) Pyrolysis S2 versus TOC plot showing generative source rock potential. b) Source rock rating and hydrocarbon
generative potential based on the plot of petroleum (PY) versus TOC for the analyzed rock samples (Al-Areeq, 2018).
SPE-198129-MS 7

Figure 6—Optical thermal maturity parameters and zones of hydrocarbon generation (Hartkopf-Fröder et al., 2015).

As the mentioned analysis deals with the bulk properties of the organic matter in sediments, some tools
provide a means of determining the distribution of different compound types found in the organic matter.
Gas chromatography (GC) is one of the commonly used separation techniques that provides a higher degree
of separation to resolve more detail about the composition of the bitumen instead of the coarse separation
obtained by the liquid column chromatography. The gas chromatograms are evaluated qualitatively to
determine the potential of a rock to generate oil versus gas, as well as identify kerogen type. As can be
seen in Fig. 7a, oil-prone rocks exhibit prominent n-paraffin peaks with the bulk of the material in the
greater than C15 carbon number range. Moreover, GC can also be used for the contamination recognition.
Contamination of cutting and core samples, from oil-based drilling mud and drilling mud additive can be
recognized from gas chromatograms. An example, Fig. 7b shows the distinctive pattern for an olefin-based
synthetic drilling mud base oil in a source rock extract.
8 SPE-198129-MS

Figure 7—(a) Rock extract gas chromatograms from source rock with kerogen type II, (b) Source extract
gas chromatogram showing contamination of a synthetic oil-based drilling mud (Dembicki, 2017).

Additionally, there are combined analysis method such as PGC. This method integrates Rock-Eval
pyrolysis and GC to reveal the composition of the hydrocarbons and other compounds that make up the S2
peak. This information provides a direct indicator of the kerogen type in the source rock and the types of
hydrocarbons that can be generated by the kerogen (Dembicki, 2017).
Geochemical logs of Rock-Eval pyrolysis and TOC, vitrinite reflectance, lithology, mud log gas, and
related data are indispensable tools for the petroleum industry. These logs identify petroleum source rocks,
the thermal maturation gradient and in situ and migrated petroleum shows. A comparison between fluid
distributions from a section of the Eldfisk chalk reservoir estimated from a geochemical log (Iatroscan)
and MSFL (induction log) (Stoddart et al. 1995) is given in Fig. 8. In general, the logs are similar, but
the geochemical logs give a greater resolution. Applications of geochemical logs can be used to locate
perforations in low-quality reservoirs or assess fluid composition prior to testing (Larter and Aplin, 1995).

Figure 8—Determination of petroleum saturations by geochemical and electric logs (Larter and Aplin, 1995).

Pay Zone Detection


Geochemical techniques for pay zone detection are good supplementary tools for finding or corroborating
the presence of reservoirs. Generally, Mud Gas Analysis, Mud Gas Data Interpretation, Fluorescence and
Cut, Isotubes, Rock-Eval Pyrolysis, Solvent Extraction/Gas Chromatography and Thermal Extraction–Gas
Chromatography (TEGC) techniques are used in various pay zones such as bypassed or overlooked pay
zones in low resistivity sands, fractured shales, oil-based muds, fresh water basins (Maende and Jarvie,
SPE-198129-MS 9

2008). The S1 peak from Rock-Eval pyrolysis can detect potential pay zones through the analysis of cuttings
collected over the potential reservoir internal.
Solvent Extraction/Gas Chromatography technique utilizes a simple solvent extraction and gas
chromatographic analysis of cuttings or sidewall cores from reservoir intervals (Baskin et al., 1995;
Bjørlykke, 2015) to compare the composition of multiple zones to look for diagnostic characteristics that
indicate gas, oil, and/or water in the reservoir rock (Fig. 9). Assement of the type of hydrocarbon present in
the pay zones can also be achieved through TEGC (thermal extraction gas chromatographic) fingerprinting.
(Maende and Jarvie, 2008).

Figure 9—An example of using whole oil gas chromatogram “fingerprinting” to confirm wireline log
interpretations of the contents of selected reservoir and the locations of fluid contacts (Baskin et.al, 1995).

In this technique, the cuttings are thermally extracted in a gas chromatograph and the subsequent gas
chromatographic analysis provides a “fingerprint” of the hydrocarbon distribution in the rock (Fig. 10).
Pay zones are classified as condensate, normal oil, biodegraded oil or waxy oil as per hydrocarbon types.
Some post-accumulation processes may cause compositional changes and modify the original properties
of petroleum. These processes can be occurred due to temperature and pressure (in-reservoir maturation),
moving waters (water washing and biodegradation) or natural gas (deasphalting by precipitation of solid
bitumen within the reservoir) (Brooks and Welte, 1984). Fig. 11 illustrates the alteration parameters of crude
oil.

Figure 10—Example of TEGC fingerprints of rocks containing various quantities of oil (Maende and Jarvie, 2008).
10 SPE-198129-MS

Figure 11—Alteration of Crude oils in reservoirs (adapted from Milner et al., 1977).

The heavy oils and tar sand bitumens generated almost entirely by the process of biodegradation, with
relatively minor accumulations due to immature oil being expelled directly from source rocks (Larter et al.,
2007). Biodegradation is an important alteration process which removes light components from oil to form
heavy oil and bituminous tar sands. It is simply a hydrocarbon oxidation process and causes a reduction in
the API gravity and the content of saturated and aromatic hydrocarbons relative to polar compounds, while
increasing oil viscosity, acidity and sulphur content, and the concentration of certain metals in the oil.
By knowing the degree of biodegradation in the reservoir can reduce the exploration risk and optimize
the production (Huc, 2003). Fig. 12 shows the ternary diagram of varying degree of biodegradation and
total ion current (TIC) fragmentograms of representative samples (Bata et al., 2016). The TICs showed the
presence of unresolved complex mixture (UCM) humps, which is a common characteristic of biodegradated
oils (Gouch et al., 1992; Ventura et al., 2008).

Figure 12—Ternary plot showing bulk properties of oils extracted from


Cretaceous oil sand samples and inserted TIC fragmentograms (Bata et al., 2016).

Oil-to-Oil and Oil-to-Source Rock Correlations


Oil-to-oil and oil-to-source rock correlations are used to check the geochemical relationships between
oils and between oils and source rocks. While the oil-to-oil correlation provides identification of active
source rocks and oil families, oil-to-source rock correlations help to determine pathway and time of the oil
SPE-198129-MS 11

migration. With the improved mass spectrometer technology, biomarker distributions became one of the
key elements of these correlations’ studies.
Biomarkers are complex organic compounds in sediments and crude oils that have the same basic
structure as compounds found in living organisms. (Dembicki, 2017; Peters and Cassa, 1994). The detailed
study of geochemical characteristics of oils provides a better understanding of the origin of oils. In the
example given above (Fig. 13), the key biomarkers of steranes and terpanes seem to be unaffected by the
degradation process, these biomarkers are used to assess the characteristics of the source rock that generates
these oils. (Kaixi, et al., 2015)

Figure 13—Typical biomarker characteristic of saturated hydrocarbon for extracted


oil from the Liuhua11-1 Oilfield (a case from LH11-1-4) (Kaixi, et al., 2015).

Geochemical Characterization in Different Type of Reservoirs, Examples of


Applications Throughout World and the Challenges
The demand for a well-understood reservoir model has been increasing for both conventional and
unconventional reservoirs due to increasing energy demands. Many working groups including engineers,
geologists, and geophysicists are collecting and analyzing the data obtained from logs, well tests, core
analyses, etc. to understand the complex nature of a reservoir, to determine the parameters (permeability,
porosity, maturity, etc.) affecting its performance, and to make operational decisions to increase recoveries.
Geochemistry is one of the tools reveals many information regarding reservoir fluid properties, source
rock potential, thermal maturity, reservoir compartmentalization, biodegradation, hydrocarbon seeps, etc.
(Robertson, 2016; Ortega-Lucach et al., 2018). Compositional variations in reservoirs can be due to the
kerogen type of source rock, thermal maturity and several alteration effects including the nature of fluid
flow in the reservoir. Petroleum compositions also depend on the type of reservoirs, and each formation
may have unique variations due to the rock and fluid properties, heterogeneities, and any operations within
the formations as well as reservoir temperature and pressure (Robertson, 2016; Jarvie et al., 2015). Thus,
geochemical characterization is applicable for both conventional and unconventional reservoirs including
oil, gas, or both as well as geothermal reservoirs and different tools for geochemical analyses can help
engineers/scientists to understand the reservoir at various stage of the field and can assist them to increase
recovery factor with decreasing operational costs by integrating these data with petrophysical, well log, and
well test data.
In their study, Khisamov et al. (2018) explained how a rock-forming organic matter transforms into a
hydrocarbon reservoir as shown in Fig.1 and they remarked that oil generation is not only a function of the
conversions of kerogen into hydrocarbon but also it depends on the stages of conversion rates of organic
matter to kerogen itself. The conversion rates of each stage determine the fluid type at a particular time in
the oil and gas system like early gas generation during maturation of organic matter, so the hydrocarbon
at the surface and in the reservoir can be estimated from the degree of matrix heterogeneity, the type and
nature of the organic and inorganic components (Khisamov et al., 2018). Thus, biodegradation can occur in
12 SPE-198129-MS

a reservoir has low reservoir temperature (20 – 60 °C), an oil-water contact, microorganisms, and nutrients
such as nitrate and phosphate (Wei and Yang, 2015).
Several geochemical methods help to determine the properties of organic matter and the usage of
each method can be chosen according to rock system, hydrocarbon type, and economic evaluations.
These methods are Rock-Eval and LECO/TOC tests, Vitrinite reflectance (Vr) and high-resolution solution
methods such as Pemex, gas chromatography (GC), gas chromatography- mass spectrometry (GCMS), X-
ray Photoelectron Spectroscopy (XPS), and saturates-aromatic-resin-asphaltene (SARA) analyses. Also,
geochemical logging tools help engineers to determine the minerology, chemistry, and lithology of the
formation (Herron et al., 1992; Pemper et al., 2018). Geochemical analyses for unconventional oil and gas
lead to the origin of the hydrocarbon, it is thermal maturity, as well as the location of oil and gas windows for
reservoir development. In unconventional reservoirs and hybrid systems (conventional and unconventional),
the production depends on the fluid composition since petroleum is fractionated during expelling from
the source rock and is became rich in saturates and aromatics rather than heavier components and the
fractionation process continues during production leaving heavier components in the reservoir (Jarvie et al.,
2015). Rokosh et al. (2010) gathered over 270 core samples from Duvernay Muskwa formations in Alberta
and combined geochemical and geological analyses to determine the shale gas potential of these formations.
The Rock-Eval/TOC results showed that the highest TOC value obtained from a carbonate mudstone in the
east Duvernay Basin.

Figure 14—Schematic representation of transformation mineral-organic polymer into hydrocarbons (Khisamov et al., 2018).

After industry’s focus on tight sand and organic shales of the Bone Springs and Wolfcamp formations
in the Delaware Basin, Schwartz et al. (2014) conducted geochemical analysis for headcap and mud
gases and produced liquids; origin of the gas and its thermal maturity levels measured with Rock-Eval
(molecular and stable isotope composition). They indicated that the mud gases are lighter than the headcap
gas since they mixed with biogenic gas and source rock maturity level is in the oil window. Zumberge et
al. (2016) combined the data of TOC and pyrolysis of oil with produced oil and gas compositions obtained
by biomarkers and stable carbon isotope to understand how transformed hydrocarbons from thermally
maturing source rock migrate to the more interconnected porous reservoirs. Their study for organic-rich Late
Devonian Woodford formation of Anadarko Basin showed that more mature resources allow high maturity
fluid flow and they made a comparison of maturity; gas is the highest whereas source rock has the lowest
SPE-198129-MS 13

maturity for organically-lean Late Devonian Bakken formation of Williston Basin. On the other hand, their
study for and Permian Wolfcamp and Spraberry formations of Midland Basin showed high maturity oil
migrating from immature sediments. Rittenhouse et al. (2017) analyzed highly complex Leonard Reservoir
of Delaware Basin in New Mexico by focusing light oil to condensates. They used several core samples from
four wells and conducted LECO/TOC, Rock-Eval as well as vitrinite reflectance to determine the origin
and maturity of the fluids. The results showed that the thick organic-rich mudstones interbedded with leaner
carbonates and fluid type is predominantly Type II marine kerogen. Jagadisan and Heidari (2018) studied the
wettability of organic-rich mudrocks and kerogen at various levels of thermal maturity by XPS and sessile
drop methods to provide a better understanding of their effects on fluid flow and their experimental results
show that determining the effect of geochemical parameters improves the understanding flow reservoir fluid
flow and may help to design secondary recovery methods. Ortega-Lucach et al. (2018) used two continues
analogous coring from Eagle Ford formation in Northeast Mexico to perform geochemical analyses by
using Rock-Eval, Vitrinite Reflectance, Pemex and to determine the source rock origin and thermal maturity
for the optimum oil and gas windows. The results showed the geochemistry of the formation is like the
Eagle Ford in the USA and geochemical analyses help to identify the Lower Eagle Ford formation as the
most promising for production. Iriarte et al. (2018) conducted several experiments with crushed Niobrara
formation core and proppant samples to investigate the chemical interactions between the formation,
fracturing fluid, and the proppant with combining geomechanical changes. These samples are subjected to
the different low salinity water and they observed the chemical interactions affected the rock stiffness and
strength. Makarychev et al. (2018) characterized the well-known conventional Shilaif formation in UAE as
an unconventional play by integrating geochemical analyses and well logging technologies. In their study,
they performed LECO/TOC, Rock-Eval, GC, GCMS, SARA analysis and bulk kinetics on core samples
to map the maturity of the synclines. The results led maximum 1.15 Vr maturity in the synclines whereas
increasing height decrease maturity level for anticlines and, they found heavy bitumen in the source rock
for the syncline B.
In conventional oil and gas reservoir, geochemical characterization has been focused on the
origin, biodegradation, and spatial distribution of hydrocarbons in the reservoir as well as reservoir
compartmentalization or channeling, especially for heavy oil reservoir since these properties are crucial
for a better well completion, well-optimized field development and secondary/tertiary recovery operations
(Henshaw et al., 1996; El Gezeeri et al., 2013; Sereda and James, 2014; Carpenter, 2015; Wei and Yang,
2015; Robertson, 2016; Landaeta and Valencia, 2017). Perkins et al.’s (1992) study on the steam injection
for South Midway Sunset heavy oil reservoir showed that by using inert and reactive aqueous-based tracer,
one can simply evaluate the in-situ formation flow pattern, reservoir temperature and reactivity of the
mineral along with the formation. Tobey et al. (1993) conducted several pyrolytic tests for samples from
Uthmaniyah reservoir in Ghawar Field to determine the origin of the tar, its properties and effects on the
production, and determine the future well locations. Their results showed that tar samples have soluble
and insoluble components, also some bitumen which leads variations in spatial distribution that may affect
on well locations. Henshaw et al. (1998) studied geochemical analysis by gas chromatography as a tool
for evaluating the degree of biodegradation and combined the data with viscosity test results to map the
viscosity of the San Joaquin Valley in California. The heavy oil in the valley is known as immature and
is expelled from organic-rich shales, so Henshaw et al. observed highly variated viscosity distribution
along the valley that can influence the field production performance and project economics. Another study
regarding heavy oils made by Sardinas et al. (2005) in a Carbonate Cuban Field and they used six oil samples
to characterize the heavy oil field. Their gas chromatography test results showed that the oil properties
are similar which leading wells are interconnected. Jones et al. (2007) studied the real-time tar assessment
in Saudi Arabia by pyrolytic analysis from drilling cuts to confirm well placement, evaluate the intervals
contain tar, and geosteering the horizontal wells out of tar mats. For their specific problems with tar, they
found real-time pyrolysis tests in-site as a successful tool to field development. El Gezeery et al. (2013)
14 SPE-198129-MS

also used geochemical analysis in order to drill multilateral wells to Burgan reservoir in Kuwait which
is classified as Lower and Upper Burgan by the properties of the sand; blocky and channel sands. It also
contains vertically stacked channel sands along with a fault network connected to the aquifer. Firstly, they
conducted a high-resolution Chemostratigraphy (XRF) analysis as a pilot study to determine the mineral
distribution along with the three wells. Then, they drilled the multilateral wells as geochemical steering
based on the evaluation of formation from the pilot study with petrophysical data. Sereda and James (2014)
studied drill-cut and core samples from two fields in McMurray formation to perform geochemical analysis
for predicting sealing and baffling shale layers with potential steam chambers that may cause slow bitumen
production. The formations in the Athabasca oil sands have biodegraded oil such as heavy oil and bitumen
and degradation cause the lighter bitumen migrating to the top the reservoir whereas heaviest bitumen
residing at the bottom. The baffle and barrier predictions are crucial for the success of any steam-assisted
gravity drainage (SAGD). They detected even thin shale layers by using geochemical analysis and their
results got along production and geological data and they concluded the baffles and barriers caused the
steam chamber growth. They stated that GCMS analysis may not be applicable for the heterogeneous
bitumen. Wei and Yang (2015) also conducted geochemical analysis to characterize the bitumen in a highly
fractured Grosmont Carbonate reservoir for Saleski steam pilot. They used several vertical cores to extract
the bitumen and analyzed molecular composition to determine the biomarkers. Their results showed that
the units C and D degraded from same source rock and they are separated by the Marl acting as a baffle.
Ronaldson’s (2016) study also highlights that the geochemical analysis can be a great tool to characterize the
baffle and barriers, also, reservoir compartmentalization. His study summarized the GCMS analysis of core
samples obtained from McMurray Formation of Alberta. He stated that the oil chemical composition can be
variated along the reservoir due to biodegradation, gravitational settling, the flow pattern in the reservoir and
geochemical analysis can be a useful tool for well placement and recovery. In their study, Chakhmakhchev
et al. (2017) various geochemical test to evaluate the thermal maturity of the Precambrian light oil and
condensate reservoirs of the East Siberia in Russia. These reservoirs are known as the oldest petroleum
and gas systems located in a harsh environment. The data obtained from whole oil GC showed that the
Precambrian petroleum system has more mid-chain methylated alkanes whereas these alkanes are lesser
in the condensates of the Baykit and Nepa-Botuoba basins. They observed high variations in the degree
of maturity obtained by GCMS for the overmatured condensates. The biomarkers of Nepa-Botuoba Basin
showed that the oils and condensates are expelled from the same source rock. Landaeta and Valencia (2017)
combined the PVT and geochemical analyses to obtain a viscosity map of Huyapari Field at Orinoca Heavy
Oil Belt in Venezuela. They conducted crude fingerprint analyses to determine the degree of biodegradation
and they also used biomarkers to estimate oil viscosities across the field and using biomarkers along other
test results minimized the error in viscosity data. They observed vertical biodegradation levels due to
the presence of a water leg and microorganisms in the system and integrating the PVT and geochemical
analyses to estimate oil viscosity enhanced the understanding of the sweet spots across the field. Also,
they found geochemical analysis as an economical option for field development. In their study, Johansen
et al. (2018) combined asphaltene gradient analyses with geochemical analysis (by GC) to investigate the
connectivity of the Ivar Aasen oilfield in the Norwegian North Sea that is very complex in nature and located
in Triassic channel sands. After 1-year production, they observed that the asphaltenes confirm the reservoir
connectivity mostly of the field. Also, they observed mild biodegradation oil across the field and the entered
oil in the field had been already degraded.
The geochemical analyses are very effective tools to characterize the reservoir and field development,
especially if they are combined with petrophysical, geological, and production data. But engineers and
scientists might be faced with several challenges when they need this useful tool. Some of the challenges
can be summarized as:
SPE-198129-MS 15

- The number of the representative core, fluid and/or drill-cuttings samples are generally limited, and
to determine the rock and fluid properties such as reservoir compartmentalization, barriers, fluid
compositions, the degree of maturity with low uncertainty require gathering many samples that may
increase the operational costs. Also, these valuable samples should be handled very carefully before
and during the tests.
- Thermal maturity tests are limited for the condensate and light oil reservoirs since these reservoirs
are lack of longer chain biomarkers and polycyclic biomarkers.
- The usage of vitrinite reflectance can be limited for the samples contain very little vitrinite.
- The high-resolution methods are expensive and require more run times.
- The usage of GCMS is limited to estimate the steam chamber height for bitumen samples having
heterogeneity in composition.

Modeling and Simulation


A full-physics commercial simulator is used to identify the effects of the key factors of geochemical
characterization, as well as the value of information on comparative cases. 250 gridblocks (25×10×1) were
used in the model. Surface pressure is set to 14.7 psi with 20 oC surface temperature where the wellhead
pressure is 145 psi and wellhead temperature is 75 oC. Reservoir permeability distribution is described
between 150mD to 3000 mD. Reservoir porosity is described as 0.343 with the pore pressure of 1465 psi
and rock compressibility of 2×10-6 psi-1.
Two different fluids as Aqueous phase and Oleic phase used in the model. Aqueous phase consists of
water, Chloride, Sodium, and Calcium with the viscosity of 1cp and density of 991 kg/m3. Compressibility
of aqueous phase described as 4.94×10-7 psi-1. In Oleic phase, Acid surfactant (1 cP, 991 kg/m3, 4.94 × 10-7
psi-1), Oil (120 cP, 812.54 kg/m3, 9.29 × 10-7 psi-1), and Solution Gas (0.889554 cP, 363.939 kg/m3, 9.29 ×
10-7 psi-1) described with their input parameters.
10 specific lateral distances between 0m and 225m is used in the model. Effect of injecting different
brines with different concentrations of Ca++ and Na+, and also effect of these injections in the cumulative
oil recovery and oil saturation is simulated for these distances. The lateral distribution of solid phase
concentration which also indicates the concentration of clay content is also modeled.
Effect of Calcium concentration in aqueous fluid modeled for three different salinity injections to a
sandstone reservoir. The distribution of Ca++ concentration with the lateral distance from the well location
(0m) is given in Figure 15. Three different plots represent the low concentration injection, high concentration
injection, and a case that includes both high and low concentration injection respectively.
16 SPE-198129-MS

Figure 15—Ca++ concentration in distance of the horizontal section for sandstone reservoir

Concentration of Ca++ in the Aqueous fluid is also shown in Table 2. Herein, all the results are close
to each other with some slight differences that can be ignored for each specific lateral distances. The
maximum value reaches up to 1.72×10-4 mole/L in 50-75 m, 125-150 m, and 200-225 m, when the lowest
value is 1.23×10-4 mole/L for the specific depth of 25m. These slight differences indicates that the Ca++
ion exchange between the injection fluid and the reservoir rock (sandstone) is not affected by the Ca++
concentration of the injection fluid. Table 2 shows the detailed data of results for the scenarios of different
Ca++ concentrations.

Table 2—Ca++ Concentrations of different salinity injections


SPE-198129-MS 17

Na+ concentration in the injection fluid also investigated by applying the same scenario that the injection
fluid includes Ca++. Similarly, three different concentrations as, low, high, and High-Low investigated
separately.

Figure 16—Na+ concentration in distance of the horizontal section for sandstone reservoir

Change of aqueous fluid Na+ concentrations with depth are almost same for all three scenario. Although
the fluctuations can be neglected, the plots showed a reverse behavior of Ca++. Its worth to point out that,
Na+ concentrations get the values which are so close to zero within the depths 50-75m, 125-150m, and
200-225m. Table 3 shows the detailed results of the injection scenarios of different Na++ concentrations.

Table 3—Na+ Concentrations of different salinity injections


18 SPE-198129-MS

In a system that Ca++ is the dominant ion for reactions, cumulative oil recovery for three different
scenarios is given in Figure 17. According to the plots, the highest recovery value reaches up to 925.9
m3 (5823.7 barrels) with the low concentration case, and the lowest value belongs to the low and high
concentration case with 725.8 m3 (4565.1 barrels). When the values compared with the base case scenario
which only includes production, there is a significant improvement by modifying the salinity of the injection
fluid.

Figure 17—Cumulative Oil production comparison of four different scenarios

Cumulative oil production for base case and the scenarios of injection with three different concentrations
is also given in annual basis for 9 years period of time (2007-2016) in Table 4.

Table 4—Cumulative Oil Production for four different scenarios


SPE-198129-MS 19

Similar behavior can also be monitored in average oil production values. It is observed that the injection
with low salinity concentration reaches the highest average oil production between 2007 and 2016. Herein,
base case scenario shows a significant decrease from 57.6 m3 to 6.4 m3 (around 8 times less) within 9 years
period). However, injection with different salinity concentrations reduces the decline ratio (around 3 times
less for all concentrations in 9 years period).

Figure 18—Average Oil production comparison of four different scenarios (2007-2016)

Additionally, brine injection resulted in significant increase of oil production for all cases such as; low
concentration (7.68 times in 2016), high concentration (7.4 times in 2016), and low & high concentration
(6.26 times in 2016).

Table 5—Average Oil Production values for four different scenariosin annual basis.

Oil saturation results before the Losal starts and at the end of the losal shown in 3D by using STARS
fluid model in Figure 19.
20 SPE-198129-MS

Figure 19—3D Model of Oil Saturation Distribution before and after the Losal

Simulation ending time is given as 1st of Jan 2017 and the oil saturation distribution along the the
horizontal section at the end of running time is given in Figure 20. The figure clearly shows that the oil
saturation have some significant local increases in all injection cases compared with the base case scenario.
It sourced by the attraction of brine which enables a multi-ion exchange and mobilize more oil phase that
left in the ruck surface. It alters the divalent cation environment and enable it to become a single cation one.
Hence, the oil saturation signature shows a noteworthy changes in the distance of the horizontal section.
SPE-198129-MS 21

Figure 20—Model of Oil Saturation Distribution with Lateral Distance

All injection cases showed close results with slight differences. However, maximum Oil saturation
value reaches up to 51.9% in base case scenario, when it reaches up to 71% with the injection scenarios.
Additionally, solid-phase concentration which indicates the clay content is also modeled for the injection
cases with three different concentration (Figure 21). Results showed some peaks in solid phase concentration
in specific lateral distances such as; 50m, 125m and 200m. These distances are same locations where the
oil saturation gets the lowest and Ca++ concentration gets the first highest results for all scenarios in a
Ca++ reaction dominated system.

Figure 21—Model of Solid-Phase (Clay Content) Distribution with Lateral Distance

Conclusions and Recommendations


Geochemical Characterization is one of the methods that eases taking decisions such as; selecting best
completion tools and design, avoiding unproductive reservoir zones, and describing best salinity level
for flooding operations. The study gave a comprehensive background about the latest developments, key
22 SPE-198129-MS

factors, with the worldwide examples. It also successfully demonstrated a description of optimal salinity
level of injection fluid. The distributions of Ca++ and Na+ concentrations, oil saturation and solid-phase
concentration in a lateral distance were analyzed for a Ca++ reactions dominated system. Results were
given below:

• In a Ca++ Reaction dominated system, injection of low salinity brine shows more effective results
compared with high salinity brine injection as well as the case that low and high salinity brine
injected together in a row.
• As a result of having a Ca++ reaction dominated system, Na+ shows almost zero value in lateral
distances that Ca++ concentration gets the highest values.
• Low concentration brine gets the highest cumulative oil recovery compared with other cases and
its almost 6.6 times more than the base case scenario which does not have any injection (at the
ending time).
• Brine injection significantly affect the mobility of residual oil saturation and oil saturation values
get close results for all brine injection cases when its up to 60% more than the base case scenario
in some specific lateral distances.
• Solid-phase concentrations get highest values in lateral distances that the oil saturation get the
lowest and Ca++ concentration gets the highest results for all cases. The amount of clay content
prevents the residual oil phase to mobilize at these specific distances and increasing amount of
Ca++ ions in the environment decreases the oil saturation at these distances.
As all runs performed by considering that the system is a Ca++ reactions dominated, it is recommended
to model Na+ reactions dominated system in future works to compare the results of both mechanisms.

References
1. England, W., E. (2007). Reservoir geochemistry — A reservoir engineering perspective. Journal
of Petroleum Science and Engineering 58, 344–354.
2. McCarthy, K., Rojas, K., Niemann, M., Palrnowski, D., Peters, K., Stankiewicz, A. (2011). Basic
petroleum geochemistry for source rock evaluation. Oilfield Review. 23. 32-43.
3. Katithi, D. M., & Opar, D. O. (2016, December 5). Petroleum Geochemistry of the Loperot-1
Well in Lokichar Basin, Kenya. SPE-AFRC-2557372-MS SPE/AAPG Africa Energy and
Technology Conference, Nairobi City, Kenya.
4. Brooks, J. and Welte, D. (1984). Advances in Petroleum Geochemistry, Volume 1, Academic
Press.
5. Kaufman, R. L., Dashti, H., Kabir, C. S., Pederson, J. M., Moon, M. S., Quttainah, R., & Al-
Wael, H. (2002, June 1). Characterizing the Greater Burgan Field: Use of Geochemistry and Oil
Fingerprinting. SPE-78129-PA, SPE Reservoir Evaluation & Engineering.
6. Larter S. R. and Aplin A. C. (1995). Reservoir geochemistry: methods, applications and
opportunities, Geological Society, London, Special Publications, 86, 5-32.
7. Hannigan, R. (2007). Chapter 1 What goes around comes around: Today’s environmental
geochemistry. Developments in Environmental Science. 5. 10.1016/S1474-8177(07)05001-2.
8. Treibs, A., 1934. The occurrence of chlorophyll derivatives in an oil shale of the upper Triassic.
Annalen 517, 103–114
9. Hunt, J.M., Meinert, R., (1958). Petroleum prospecting. U.S. Patent 2,854,396.
10. Sluijk, D., Parker, J.R., (1986). Comparison of predrilling predictions with postdrilling outcomes,
using Shell’s prospect appraisal system. In: Association of Petroleum Geologists Studies in
Geology 21: Oil and Gas Assessment: Methods and Applications American, pp. 55–58
SPE-198129-MS 23

11. Dembicki, H., Jr. (2017). Practical Petroleum Geochemistry for Exploration and Production,
Elsevier.
12. Horsfield, B., Yordy, K.L., Crelling, J.C. (1988). Determining the petroleum-generating potential
of coal using organic geochemistry and organic petrology. Organic Geochemistry, Volume 13,
Issues 1–3, Pages 121-129.
13. Tissot, B., Durand, B., Espitalié, J. and Combaz, A. (1971, March). Influence of Nature and
Dagenesis of Organic Matter in Formation of Petroleum. AAPG Bulletin 58, no.3:1809-1851.
14. Tissot, B.P. and Welte, D.H. (1984) Petroleum Formation and Occurrence. 2nd Edition, Springer-
Verlag, Berlin, 699 p.
15. Demaison, G. J., (1984). The generative basin concept, in Demaison G.J. and Munis R. J., eds.,
Petroleum Geochemistry and Basin Evaluation: AAPG Memoir 35, p. 1-14.
16. Peters, K., Cassa, M., R. (1994). Applied Source Rock Geochemistry. AAPG Memoir, Volume
60.
17. Al-Areeq, N., M. (2018, February). Petroleum Source Rocks Characterization and Hydrocarbon
Generation. Recent Insights in Petroleum Science and Engineering, Mansoor Zoveidavianpoor,
IntechOpen.
18. Clementz, D., M., Demaison, G., J., Daly, A., R. (1979, January). Well Site Geochemistry By
Programmed Pyrolysis. Proc., 11th Proceedings of the Annual Offshore Technology Conference
1(1).
19. Bjørlykke, K. (2015). Petroleum Geoscience from Sedimentary Environments to Rock Physics.
pringer-Verlag Berlin Heidelberg.
20. Hunt, J.M. (1996). Petroleum Geochemistry and Geology. second ed. W.H. Freeman and
Company, New York (743 pp).
21. Hackley, P. and Cardott, B. (2016). Application of organic petrography in North American shale
petroleum systems: A review. International Journal of Coal Geology, Volume:163.
22. Hartkopf-Fröder, C., Königshof, P., Littke, R., Schwarzbauer, J. (2015). Optical thermal maturity
parameters and organic geochemical alteration at low grade diagenesis to anchimetamorphism: A
review. Int. J. Coal Geol.
23. Stoddart, D., P., Hall, P., B., Larter, S., R., Brasher, J., Li, M., Bjoroy, M. (1995). The reservoir
geochemistry of the Eldfisk Field, Norwegian North Sea.
24. Maende, A. and Jarvie, D. (2008, December). Finding Bypassed or Overlooked Pay zones using
Geochemistry Techniques. International Petroleum Technology Conference, IPTC 12918, Kuala
Lumpur, Malaysia.
25. Baskin, D.K., Hwang, R.J., Purdy, R.K., 1995. Predicting gas, oil and water intervals in Niger
Delta reservoirs using gas chromatography. American Association of Petroleum Geologists
Bulletin 79, 337–350.
26. Milner, C., W., D., Rogers, M., A., Evans, C., R. (1977). Geochem. Explor. 7, 101-153.
27. Huc, A. (2003). Petroleum Geochemistry at the Dawn of the 21st Century. Oil & Gas Science
and Technology-Rev. IFP-volume: 58, 233-241.
28. Bata, T., Parnell, J., Bowden, S., Boyce, A., Leckie, D. (2016). Origin of heavy oil in Cretaceous
petroleum reservoirs. Bulletin of Canadian Petroleum Geology. 64. 106-118.
29. Gouch, M.A., Rhead, M.M. and Rowland, S.J. 1992. Biodegradation studies of unresolved
complex mixtures of hydrocarbons: model UCM hydrocarbons and the aliphatic UCM. Organic
Geochemistry, v. 18, p. 17–22.
30. Ventura, G.T., Kenig, F., Reddy, C.M., Frysinger, G.S., Nelson, R.K., Van Mooy, B. and Gaines,
R.B. 2008. Analysis of unresolved complex mixtures of hydrocarbons extracted from Late
24 SPE-198129-MS

Archean sediments by comprehensive two-dimensional gas chromatography (GC× GC). Organic


Geochemistry, v. 39, p. 846–867.
31. Larter, S., Adams, J., Gates, I. (2007, May). The Impact of Oil Viscosity Heterogeneity on
Production Characteristics of Heavy Oil and Tar Sand (HOTS) Reservoirs Part III: The Origin of
Highly Non Linear Oil Viscosity Gradients and the Design of Geotailored Recovery Processes
Suitable for such Compositionally Graded Reservoirs, CSPG/CSPE GeoConvention 2007,
Calgary, Alberta, Canada, May 14-17.
32. Kaixi, J., Wenxiang, H., Nian, X., Peng Li, Changchun, H., Qingzheng, G., (2015, September).
Chinese Journal of Geochemistry, Volume 34, Issue 3, pp 320–329.
33. Robertson, David R. K. (2016) Geochemistry as a State of Art Reservoir Investigative Tool. At
the GeoConvention 2016: Optimizing Resources, Calgary, Canada, 7 – 11 March.
34. Ortega-Lucach, S., Gutierrez-Caminero, L., Torres-Vargas, R., and Murillo-Muneton, G.
(2016). Geochemical Characterization of the Eagle Ford Formation in Northeast Mexico. At the
Unconventional Resources Technology Conference, Houston, Texas, USA, 23 – 25 July.
35. Jarvie, D. M., Jarvie, B. M., Weldon, W. D., and Maende, A. (2015). Geochemical Assessment
of In Situ Petroleum in Unconventional Resource Systems. At the Unconventional Resources
Technology Conference, San Antonio, Texas, USA, 20 – 22 July.
36. Khisamov, R., Skibitskaya, N., Kovalenko, K. Bazarevskaya, V., Samokhvalov, N., Bolshakov,
M., Kuzmin, V., Marutyan, O., Navrotskiy, O., Pugo, T. (2018). Well Logging Data Interpretation
in Oil and Gas Source Rock Sections Based on Complex Petrophysical and Geochemical
Analysis Results. At the SPE Russian Petroleum Technology Conference, Moscow, Russia, 15 –
17 October.
37. Wei, W. and Yang, D. (2015). Geochemical Bitumen Characterization for the Saleski Steam Pilot
in the Grosmont Carbonate Reservoir. At the SPE Canada Heavy Oil Technical Conference,
Calgary, Canada, 9 – 11 June.
38. Herron, S. L., Petricola, M. J. C. and Dove, R. E. (1992). Geochemical Logging of a Middle East
Carbonate Reservoir. Journal of Petroleum Technology, 44 (11), 79 – 82.
39. Pemper, R., Pereira, A., Hou, G., Dolliver, D., Tudge, J., Kharrazi, J., Chok, H., Schmid, G.,
Mekic, N., Blankinship, T., Epstein, R., Cave, T., MacPherson, A. (2018). A New Geochemical
Logging Tool for Determination of Formation Chemistry and Minerology in both Conventional
and Unconventional Reservoirs. At the SPE Annual Technical Conference and Exhibition, Dallas,
Texas, USA, 24 – 26 September.
40. Rokosh, C. D., Anderson, S. D. A., Beaton, A. P., Berhane, M., Pawlowicz, J. (2010).
Geochemical and Geological Characterization of the Duvernay and Muskwa Formations in
Alberta. At the Canadian Unconventional Resources & International Petroleum Conference,
Calgary, Canada, 19 – 21 October.
41. Schwartz, K. Muscio, G., Nester, P., Easow, I, Javalagi, M. (2014). Petrophysical and
Geochemical Evaluation of an Avalon Shale Horizontal Well in the Devonian Basin. At the
Unconventional Resources Technology Conference, Denver, Colorado, USA, 25 – 27 August.
42. Zumberge, J. E., Curtis, J. B., Reed, J. D., Brown, S. W. (2017). Migration Happens:
Geochemical Evidence for Movement of Hydrocarbons in Unconventional Petroleum Systems.
At the Unconventional Resources Technology Conference, San Antonio, Texas, 1 – 3 August.
43. Rittenhouse, S., Li, Y., Hughston-Kennedy, K. Fritz, J., Pritchard, J., Cassel, L., Baum, V., Liem,
S., Mooney, T. (2017). Delaware Basin Leonard Reservoir Characterization, New Mexico and
Texas. At the Unconventional Resources Technology Conference, Austin, Texas, USA, 24 – 26
July.
SPE-198129-MS 25

44. Jagadisan, A. and Heidari, Z. (2018). Impacts of Geochemical Properties on Wettability


of Kerogen and Organich-rich Mudrocks. At the Unconventional Resources Technology
Conference, Houston, Texas, USA, 23 – 25 July.
45. Iriarte, J., Katsuki, D., Tutuncu, A. N. (2018). Geochemical and Geomechanical Alterations
Related to Rock-Fluid-Proppant Interactions in the Niobrara Formation. At the SPE International
Conference and Exhibition on Formation Damage, Lafayette, Louisiana, USA, 7 – 9 February.
46. Makarychev, G., Leyrer, K., Baig, M. Z., Laer, P. V. (2018). Integration of Geochemical
and Petrophysical Analyses for the Shilaif Formation Characterization. At the Abu Dhabi
International Petroleum Exhibition & Conference, Abu Dhabi, UAE, 12 – 15 November.
47. Henshaw, P. C., Carison, R. M. K., Pena, M. M., Boduszynski, C. E., Rechsteiner, C. E., and
Shafizadeh, A. S. G. (1998). Evaluation of Geochemical Approaches to Heavy Oil Viscosity
Mapping in San Joaquin Valley, California. At the SPE Western Regional Meeting, Bakersfield,
California, 10 – 13 May.
48. El Gezeeri, T., Ismail, A. A., Al-Anezi, Kh., Al Jeaan, M., Silambuchlvan, J. K., Padhy, G.
S., Atul, W. ad Al Shoeibi, A. (2013). Drilling of Multilateral Wells Aided with Geochemical
Analysis. At the International Petroleum Technology Conference, Bijing, China, 26 – 28 March.
49. Sereda, J. N. and James, B. R. (2014). A Case Study in the Application of Bitumen Geochemistry
for Reservoir Characterization in SAGD Development. At the SPE Canada Heavy Oil
Conference, Alberta, Canada, 10 – 12 June.
50. Carpenter, C. (2015). Drilling of Multilateral Wells in Kuwait Aided with Geochemical Analysis.
Journal of Petroleum Technology, 67 (02), 1176 – 1183.
51. Landaeta, G. S. and Valencia, L. E. (2017). Extra-Heavy Oil Viscosity Estimation Using PVT
and Geochemical Analyses: Applications at Huyapari Field, Orinoco Heavy Oil Belt, Venezuela.
At the SPE Latin America and Caribbean Petroleum Engineering Conference, Buenos Aires,
Argentina, 18 – 19 May.
52. Tobey, M. H., Halpern, H. I., Cole, G. A., Lynn, J. D., Al-Dubaisi, J. M., and Sese, P. C. (1993).
Geochemical Study of Tar in the Uthmaniyah Reservoir. At the SPE Middle East Oil Technical
Conference and Exhibition, Bahrain, 3 – 5 April.
53. Sardinas, Z. D., Fernandez, O. P., Quintero, J. O. L., Rivera, C. L., and Tapanes, N. (2005).
Geochemistry Characterization and Continuity studies in a Carbonate Cuban Field. At the SPE
Latin American and Caribbean Petroleum Engineering Conference, Rio de Janeiro, Brazil, 20 –
23 June.
54. Jones, P., Halpern, H., Dahan, M., Bellaci, I., Neuman, P., Akkurt, R., Al-Qathami, S., Al-
Amoudi, M, Malki, K., Dix, M., and Zerek, R. (2007). Implementation of Geochemical
Technology for “Real-Time” Tar Assessment and Geosteering: Saudi Arabia. At the Offshore
Mediterranean Conference and Exhibition, Ravenna, Italy, 28 – 30 March.
55. Chakhmakhchev, A., Shiganova, O., Andrus, V., and Tchistiakov, A. (2017). Geochemical Tools
to Correlate Light Petroleum and Evaluate Thermal Maturity: A Case Study of Precambrian
Condansates from East Siberia, Russia. At the SPE Russian Petroleum Technology Conference,
Moscow, Russia, 16 – 18 October.
56. Johansen, Y. B., Rinna, J., Betancourt, S. S., Forsythe, J. C., Achourov, V., Cnas, J. A., Chen,
Li, Zu, J. Y., and Mullins, O. C. (2018). Asphaltene Gradient Anlayses by DFA Coupled with
Geochmeical Analysis by GC and GCxGC Indicate Connectivity in Agreement with One Year
of Production in Norwegian Oilfield. At the SPE Annual Technical Conference and Exhibition,
Dallas, Texas, USA, 24 – 26 September.

Вам также может понравиться