You are on page 1of 20

SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Mapping and Modeling Large Viscosity and Asphaltene Variations in a
Reservoir Undergoing Active Biodegradation
Richard R. Jackson, Julian Y. Zuo, and Ankit Agarwal, Schlumberger; Bernd Herold and Sanjay Kumar, Cairn
India Ltd; Ilaria De Santo, Hadrien Dumont, Cosan Ayan, and Oliver C. Mullins, Schlumberger

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Amsterdam, The Netherlands, 27–29 October 2014.

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
Viscosity is one of the key reservoir fluid properties. It plays a central role in well productivity and
displacement efficiency and has a significant impact on completion strategies. Accurately assessing areal
and vertical variations of viscosity will lead to more realistic reservoir simulation and optimal field
development planning. Downhole fluid analysis (DFA) has successfully been used to measure the
properties of reservoir fluids downhole in real time. DFA has excellent accuracy in measuring fluid
gradients which in turn enable accurate thermodynamic modeling. Integration of DFA measurements with
the thermodynamic modeling has increasingly been employed for evaluating important reservoir proper-
ties such as connectivity, fluid compositional and property gradients. The thermodynamic model is the
only one that has been shown to treat gradients of heavy ends in all types of crude oils and at equilibrium
and disequilibrium conditions. In addition, fluid viscosity depends on concentration of heavy ends that are
associated with optical density measured by DFA. Therefore, mapping viscosity and optical density
(heavy end content) is a new important application of DFA technology for use as assessment of reservoir
architectures and a mutual consistency check of DFA measurements. In this case study, a very large
monotonic variation of heavy end content and viscosity is measured. Several different stacked sands
exhibit the same profiles. The crude oil at the top of the column exhibits an equilibrium distribution of
heavy ends, SARA and viscosity, while the oil at the base of the oil column exhibits a gradient that is far
larger than expected for equilibrium. The fluid properties including SARA contents, viscosity and optical
density vary sharply with depth towards the base of the column. The origin of this variation is shown to
be due to biodegradation. GC-chromatographs of the crude oils towards the top of the column appear to
be rather unaltered, while the crude oils at the base of the column are missing all n-alkanes. A new model
is developed that accounts for these observations that assumes biodegradation at the oil-water contact
(OWC) coupled with diffusion of alkanes to the OWC. Diffusion is a slow process in a geologic time
sense accounting for the lack of impact of biodegradation at the top of the column. An overall
understanding of charging timing into this reservoir and expected rates of biodegradation are consistent
with this model. The overall objective or providing a 1st-principles viscosity map in these stacked sand
reservoirs is achieved by this modeling. Linking DFA with thermodynamic modeling along with precepts
from petroleum systems modeling provides a compelling understanding of the reservoir.
2 SPE-170794-MS

Introduction
Downhole fluid analysis (DFA) has successfully been used to measure the properties of reservoir fluids
downhole at in-situ reservoir conditions, in real time during downhole formation testing and sampling
operations (Mullins 2008). DFA also has excellent accuracy in measuring fluid gradients which in turn
provide valuable inputs to thermodynamic and reservoir fluid modeling. While the cubic equation of state

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


(EoS), (Peng & Robinson 1976) is well now established to model gas-liquid compositional gradients and
saturation pressure gradients of reservoir fluids, there had been no thermodynamic treatment to model
gradients of dissolved (or colloidally suspended) asphaltenes in reservoir fluids. This deficiency is largely
due to the lack of understanding of the molecular and colloidal sizes of asphaltenes in crude oil, and even
in laboratory solvents. In recent years, the molecular and nanocolloidal structures of asphaltenes has been
clarified and codified in the Yen-Mullins model (Mullins 2010, 2011, Mullins et al. 2012). Recently,
important confirmation of this model has been obtained from various sources (Majumdar et al. 2013, Korb
et al. 2013, Goual et al. 2011, 2014).
With the size of asphaltene nanostructures resolved in the Yen-Mullins model, the effect of gravity on
asphaltenes can be now expressed explicitly. As such, the gravity term has been added to the Flory-
Huggins polymer solution theory to obtain the Flory-Huggins-Zuo equation of state (FHZ EoS); and a
number of technical publications by several worked have described its practical application (Freed et al.
2010, 2014, Zuo et al. 2010, 2010, 2013) to treat gradients in many reservoir studies. This approach is the
only one which has been demonstrated to treat asphaltene concentration gradients in all types of crude oils
and hydrocarbons; spanning the range from retrograde gas-condensates (with high gas-oil-ratio, GOR) to
heavy oils (with low gas-oil-ratio and high asphaltene content (Zuo et al. 2013).
With the equilibrium model of reservoir fluids determined with the FHZ EoS, it is now possibly to treat
dynamic process from a thermodynamic perspective, where a transient is occurring in the reservoir. One
transient process that is of great significance in oilfields is biodegradation. Biodegradation of oil columns
occurs at the oil-water contact; as the microbes live in the water but consume oil chemical components
(Bennett et al. 2013). The microbes preferentially consume specific chemical species in the crude oils. For
example, n-alkanes are the first chemical species consumed by microbes in the biodegradation process as
indicated in the Peters-Moldowan rank of biodegradation (Peters et al. 2005). Large fluid gradients have
been observed in cases of biodegradation (Bennett et al. 2013, Peters et al. 2005, Larter et al. 2003).
Diffusion is well known to be a critical component of controlling rates of biodegradation (Head et al.
2003). Nevertheless, the inability to model asphaltene concentration gradients largely precluded the ability
to develop specific thermodynamic models of the effect of biodegradation on asphaltene and viscosity
gradients in oil reservoirs.
Biodegradation in oil reservoirs often creates a large compositional gradient and makes composition
distributions in a reservoir away far from equilibration, yielding a large viscosity variation (Head et al.
2003). Under certain conditions, living microorganisms (primarily bacteria, but also yeasts, molds, and
filamentous fungi) can alter and/or metabolize various classes of compounds present in oil, a set of
processes collectively called oil biodegradation. Biodegradation affects oil spills and surface seeps.
Furthermore, as noted more than 30 years ago, biodegradation also alters subsurface oil accumulations.
Shallow oil accumulations (⬍ 80 degC reservoir temperature) are commonly found to be biodegraded and
altered to some degree. In fact, many of the world’s petroleum are often severely biodegraded, with oil
in the shallow, super-giant Orinoco and Athabasca tar sands in Venezuela and Canada as well known
examples. Smaller, but still giant, accumulations of biodegraded oil occur elsewhere throughout the world.
Oil biodegradation by bacteria can occur under both oxic and anoxic conditions, albeit by the action
of different consortia of organisms. In the subsurface, oil biodegradation occurs primarily under anoxic
conditions, mediated by sulphate reducing bacteria in cases where dissolved sulfate is present, or
methanogenic bacteria in cases where dissolved sulphate is low. Although subsurface oil biodegradation
SPE-170794-MS 3

does NOT require oxygen, it does require certain essential nutrients (e.g., nitrogen, phosphorus, potas-
sium), which can be provided by dissolution/alteration of minerals in the water leg. Empirically, it has
been noted that biodegraded oil accumulations occur in reservoirs that are at temperatures less than 80
degC. At higher temperatures, it appears that many of the microorganisms involved in subsurface oil
biodegradation cannot exist. However, not all oil accumulations at temperatures less than 80 degC are
biodegraded. Because subsurface oil biodegradation does NOT require oxygen, and can occur at tem-

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


peratures up to 80 degC, in-reservoir biodegradation can occur even at many thousands of feet below the
surface (e.g. Head et al. 2003), as long as the geothermal gradient is sufficiently low, and nutrients (e.g.,
nitrogen, phosphorus, potassium) are available through the water leg (e.g. Larter et al. 2006). Gas
chromatograph-mass spectroscopy is often used for oil biodegradation investigation for dead oil samples.
Until now there have been no technical publications about reservoir oil biodegradation investigation using
downhole fluid analysis (DFA) and PVT sample analysis.
The study described in this technical contribution, which uses a case study of data from various wells
from an oilfield in Rajasthan, India is probably one of the first studies of this type. We examine large
asphaltene concentration and viscosity gradients, along with large gradients in other fluid properties in an
oil reservoir with five stacked reservoir sequences which are undergoing active and in-situ biodegradation.
Extensive fluid property measurements, both downhole and laboratory, are performed in this study. Fluid
data from four wells across the field is acquired showing consistency across the reservoir. In particular,
accurate asphaltene gradient measurements are performed utilizing downhole fluid analysis. The combi-
nation of the FHZ EoS together with a diffusive model and with asphaltene nanoaggregates is used to
perform the thermodynamic modeling of the asphaltene gradients. Diffusion of alkanes to the oil-water
contact is presumed to be the rate limiting step in this process. The predominant effect on asphaltene
concentration is consistent with consumption of alkanes and some alkylaromatics. The thermodynamic
model is shown to account for both the upper oil column, which is thermodynamic equilibrium, and the
lower section of the oil column – which is mostly dominated by the dynamic processes of diffusion and
biodegradation. Each of the five stacked siliclastic reservoir sequences is shown to exhibit the same fluid
properties, establishing the robustness of the measurements and the analysis. The ability to provide an
explicit thermodynamic model of asphaltene gradients – even in cases where the reservoir is undergoing
dynamic processes advances reservoir engineering studies, and provide information of reservoir modeling
and field development planning for efficient oil production.
Case Study – Overview of the Field and Reservoir Fluid Properties
The oil field and the four wells described in this case study are located in the Barmer Basin in the west
of Rajasthan, northwest India (Figure 1). The field contains hydrocarbons within fluvial dominated
siliclastic reservoir units or sands within a formation of mostly Paleocene age. The reservoir is a
multilayer sequence consisting of about 250 m thickness of medium to thickly bedded, fine to coarse-
grained sandstones interbedded with iron rich mudstones. The reservoir sequences comprise of good
quality reservoir sands with porosity ranging 18-33% (average 25%) and permeability’s up to 20 D
(average 5D). The net-to-gross (N/G) for each sand varies from 45% in the upper reservoir formations to
90% in the lower reservoir formation sequences (Compton 2009, Zittel et al. 2008). The initial formation
pressures are low as the reservoir sequences are at relatively shallow depths (Figure. 1). A more
comprehensive review of the petroleum geology and petrophysics of the Barmer Basin and our oil field
is provided by Compton et al. (2009) and Zittel et al. 2008.
A 3D view of the formation structures penetrated by the four wells in this study is shown in Figure 1
(left) and a schematic diagram is shown in Figure 1 (right). The reservoir was discovered in 2005. Wells
1 and 2 were drilled during initial exploration and field appraisal. The reservoir started production from
January 2012. Wells 3 and 4 were drilled in 2012 approximately six months after field production was
established. The reservoir has several different units or members which are described here as FB1, FB2,
4 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 1—(Left) 3D view of a multilayer reservoir structure penetrated by the four wells included in this study. (Right) Schematic diagram of the
reservoir with five-stacked reservoir sequences and a common initial oil-water-contact.

Figure 2—Pressure-depth plots showing pressures and gradients with the excess pressure analysis values (dark blue circles).

FB3, FB4, and FB5. The reservoir in this study is a shallow reservoir with depths ranging from
approximately 290 to 455 m true vertical depth subsea (TVDSS) and measured well depth (MD) is 479
to 666 m. The reservoir temperature varies from 55 to 61 °C and the formation pressure ranges from 620
to 890 psi. At the top of the oil column, the bubble point pressures of the reservoir fluids are close to the
formation pressures. Stock tank oil (STO) API gravity varies from 19 to 32 °API. Gas-oil ratio (GOR) is
low and ranges from 9 to 16 m3/m3. Oil viscosity varies significantly and ranges from 20 to 150 cP at
reservoir conditions (Zittel et al. 2008).
Formation Pressures and Formation Pressure Analysis
Distributed pressure measurements acquired with wireline formation tester tools, prior to production
during the exploration and appraisal phase, and then later during field production have helped in
understanding the vertical connectivity of the reservoir intervals within and across the field (Figure 2).
Pressure gradient data from the initial exploration wells prior to field production have been analyzed
using conventional gradient analysis techniques and also excess pressure (Brown 2003) or residual
SPE-170794-MS 5

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 3—Schematic of the wireline modular formation sampling tool with downhole fluid analysis (DFA) sensors and focused sampling probe
modules. An advanced fluid analyzer module is shown at top right, which measures optical density and hydrocarbon compositions, gas-oil-ratio
(GOR), fluid density, viscosity, and fluorescence intensity among other fluid properties. A focused sampling probe (bottom right) uses a concentric
sampling arrangement and two synchronized pumps which accelerates and enables acquisition of fluid samples with very low contamination of filtrate
from oil-based-mud (OBM) drilling fluid compared to conventional sampling tools.

pressure methods to help evaluate subtle variations in fluid density (Figure 2). Pressure gradient analysis
of the wells of indicate subtle vertical variations in fluid density and non-linear pressure gradients.
Reservoir Fluid Sampling and Downhole Fluid Analysis
Downhole fluid analysis (DFA) measurements were conveyed in the study wells by an advanced and
modular wireline formation sampling tool (Mullins 2008). Figure 3 shows a schematic diagram of the
formation testing tool with the DFA sensors, and with a focused sampling probe module (Akurt et al.
2007). One of the basic methods employed for DFA is optical spectroscopy as shown in Figure 3 (top
right). Optical spectrometers measure the visible-near-infrared (vis-NIR) spectrum of the formation fluids
passing through the flowline of the MDT tool as a function of time. The vis-NIR absorption spectrum
provides the magnitude of electronic absorption which is dependent on the asphaltene content. In addition,
the CH vibrational overtones and combination bands in the NIR spectral range are sensitive to the alkane
composition. In the case of crude oil, the methane CH4 , the –CH3 and –CH2 groups have characteristic
peaks in the NIR region that can be treated to give information on fluid composition (Mullins et al. 2001,
Fujisawa et al. 2002).
Figure 4 depicts the visible to near-infrared (NIR) absorption spectra of several common reservoir
fluids. Large variations in the absorption response are observed for different types of hydrocarbons, with
a graded transition from the ‘colorless’ gases and condensates to much darker heavy oils (Mullins 2008).
At any wavelength (␭), the absorption (A␭) of a sample is defined as the logarithm of the transmittance,
i.e. the log of the intensity ratio of the incident light (I0) and transmitted light (I). The optical density
(OD␭) is expressed in Equation 1 as:
(1)
6 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 4 —Visible to NIR absorption spectra of different types of reservoir and well fluids. Visible light is best suited for distinguishing relative
asphaltene content (‘color’). The NIR spectrum is useful for water detection, distinguishing water from oil and identifying the type of oil and oil
composition.

Figure 5—An example line of section for wells from the case study, showing gamma-ray, formation pressures, excess pressure (Pexcess) analysis, DFA
optical density (OD) measurements together with FHZ EoS model for the various reservoir units across the field.

For the wells in our study where formation testing, fluid sampling and DFA measurements have been
acquired, we have analyzed this data in the context of the 3D reservoir model – by visualization and
comparison of lines of well sections across the field. This approach enables us to compare potential
barriers between wells in the reservoir, such as faults, and to compare measurements in reservoir units
between wells to see how they relate to each other. This helps assessment of any differences in reservoir
fluids and their distribution; and enables us to map and establish fluid regions in a 3D reservoir model.
Figure 5 provides an example panel of a line of section and with well section displays from our field.
SPE-170794-MS 7

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 6 —The Yen-Mullins model [refs.] depicting the predominant molecular and colloidal structures of asphaltenes in crude oils and in laboratory
solvents. At low concentrations such as in light oil, asphaltenes are dispersed as a true molecular solution (left). At higher concentrations such as in
black oils, asphaltenes are dispersed as nanoaggregates (center). At yet higher concentrations such as in mobile heavy oils, asphaltenes are dispersed
as clusters of nanoaggregates (right).

Flory-Huggins Zuo EoS and Yen-Mullins Model of Asphaltenes


The Flory-Huggins-Zuo equation of state was developed by Freed et al. (2010, 2014) and Zuo et al. (2010,
2011) to account for asphaltene gradients in oilfield reservoirs and in laboratory centrifugation experi-
ments. It is the first EoS for predicting asphaltene concentration gradients with depth in oil reservoirs; and
is the only EoS that has successfully modeled condensates through heavy oils (Zuo et al. 2013). The FHZ
EoS relies on the colloidal characterization of the crude oils. For example, for heavy oils, large asphaltene
gradients are obtained due to the relatively large cluster size. If an EoS does not incorporate asphaltene
clusters from the Yen-Mullins model, it will be unable to obtain the observed large gradients in heavy oil
columns (Zuo et al. 2013, Mullins et al. 2013). Because of the simple form of the FHZ EoS, it can use
various measurements of reservoir fluid properties in particular by employing asphaltene concentration
ratios or optical density ratios at two different depths in the reservoir. The FHZ EoS adds a gravity term
to the Flory-Huggins theory and accounts for asphaltene gradients, a significant object for EoS.
The FHZ EoS is given by:
(2)

where OD, R, ␾, ␯, ␦, T, g, ␳, and h are the optical density, universal gas constant, volume fraction,
molar volume, solubility parameter, temperature, gravitational acceleration, density and depth, respec-
tively. Subscript a denotes the properties of asphaltenes; subscripts h1 and h2 stand for the properties at
depths h1 and h2, respectively. It should be pointed out that the solubility parameter, molar volume, and
density of bulk fluids, temperature, pressure and compositions are dependent on depth.
For low-GOR fluids, the solubility and entropy terms can approximately be canceled out due to the
opposite influence on the asphaltene concentration gradient. Thus Eq. (2) can be rewritten as
(3)

The first application of Eq. (3) to an oilfield column was for a highly under saturated low-GOR black
oil, with incorporation of nanoaggregates from the Yen-Mullins model (Mullins et al. 2007). In Eq. (2),
OD, T, ␳ and h are available from DFA and other measurements. The five unknowns are the asphaltene
density as well as the molar volumes (v) and solubility parameters (␦) of both oil and asphaltenes,
respectively. The asphaltene density and solubility parameter are estimated by the correlations. The molar
volume and solubility parameter of live fluids are estimated by a cubic EoS. The remaining unknown is
the size of asphaltenes which is tightly constrained by the Yen-Mullins model asphaltene of asphaltenes
(Mullins 2010).
8 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 7—Example of DFA optical density measurements from our India case study versus the laboratory analysis and measurement of asphaltene
content from the corresponding reservoir fluid samples where DFA was performed. Asphaltene wt.% versus optical density displays a linear
relationship (dashed line).

The Yen-Mullins model depicts the predominant molecular and colloidal structures of asphaltenes in
crude oils and in laboratory solvents as shown in Figure 6. At low concentrations such as in light oil,
asphaltenes are dispersed as a true molecular solution (left, ~1.5 nm in diameter). At higher concentrations
such as in black oils, asphaltenes are dispersed as nanoaggregates (center, ~2 nm in diameter). At yet
higher concentrations such as in mobile heavy oils, asphaltenes are dispersed as clusters of nanoaggre-
gates(right, ~5 nm in diameter).
The FHZ EoS is the only one that has been shown to treat gradients of asphaltenes in all types of crude
oils and at equilibrium and disequilibrium conditions. Details about recent advances in the FHZ EoS for
asphaltene gradients and formation evaluation are provided in the reference by Zuo et al. 2013.
Optical density is linearly related to asphaltene content. The DFA measurements and PVT analysis
measurements demonstrate that the oil column in our study from India, also follows a linear relationship
between n-C7 asphaltene content and optical density at 1290 nm as shown in Figure 7 below. The data for
this case is consistent with many previous determinations of crude oil color and asphaltene content which
have been described in some reservoirs (Mullins et al. 2001, Fujisawa et al., 2002).

Case Study – Oil Composition, Properties of the Reservoir Fluids and


Spatial Distribution
For the wells described in this case study, the formation testing tool configured with DFA sensors and a
focused sampling probe (Figure 2, see Akurt et al. 2007) was used to measure formation pressures, fluid
compositions and properties; and to acquire fluid samples with low OBM filtrate contamination downhole
at multiple depths within the reservoir. The downhole fluid samples which were acquired in the study
wells, were then degased at surface with subsequent SARA (saturates, aromatics, resins and asphaltenes)
analysis and fluid property measurements performed. Gas chromatography (GC) was performed for the
degased samples.
Figure 8(a) shows the variations of saturates, aromatics and resins ⫹ asphaltenes for the degased oil
samples from Well 1. The curves shown in these graphics are simple polynomial curve fits just to illustrate
the data trends. The gas chromatograms of the degased oil samples are illustrated in Figure 8(b). A
relatively large compositional gradient is observed towards the base of the oil column. There is evidence
for increasing biodegradation towards the base of the oil column – which results in decreasing saturated
hydrocarbons and increasing resins ⫹ asphaltenes.
SPE-170794-MS 9

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 8 —Diagram illustrating the variations of saturates, aromatics and resins ⴙ asphaltenes (SARA) from analysis of oil samples acquired during
wireline formation testing in one of the study wells. The dashed curves denote the best curve fitting. Increasing effects of oil biodegradation are
apparent towards the base of the oil column; which results in decreasing saturated hydrocarbons and increasing resins ⴙ asphaltenes.

In the upper section of the reservoir sequence, the crude oil samples contain large and comparable
quantities of n-alkanes which indicates they are not altered by biodegradation. However, in the lower
sequence of the oil column, n-alkanes are seen to be decreasing, and are almost all missing at the base of
the oil column. The OBM filtrate contamination of the oil sample with its narrow n-alkane distribution is
observed in these samples. The gas chromatogram of the oil based mud (OBM) filtrate is also shown in
Figure 8 for reference. The GC trace of the OBM filtrate is readily distinguishable from that of the crude
oil.
Figure 8 displays the GC chromatograms of the degased crude oils versus depth across the entire
hydrocarbon column for our reservoir. The GC chromatogram of the OBM filtrate is also shown (Figure
8, top-right), which does contaminate some of the crude oil samples. The GC chromatograms for the crude
oil samples in the upper section of the oil column are very similar with dominant n-alkane composition.
In contrast, in the lower half of the oil column, there is a strong trend towards reduced n-alkane
contribution. This is seen in comparison of the sharp n-alkane peaks versus the elevated, broad undif-
ferentiated complex mixture (UCM). At the very base of the column there is almost no n-alkane
contribution (except for the OBM filtrate contamination, which is circled). This GC chromatography data
is consistent with extensive biodegradation at the base of the column, with decreasing biodegradation at
10 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 9 —GC chromatograms for the entire oil column. The GC of the OBM filtrate appears in the upper right. The crude oils from the upper half
of the hydrocarbon column in the reservoir have very similar chromatograms with plentiful n-alkanes. The GC chromatograms in the lower half of
the oil column show decreasing n-alkane composition with depth. OBM contamination becomes more pronounced at the base of the oil column due
to sampling problems associated with sampling fluids with high viscosity values. Blue circles are the DFA measurements of optical density acquired
during the sampling operations; and the solid blue curve is the best curve fitting.

points above the base of the oil column. Virtually no biodegradation is detected in the upper section of
the oil column as the n-alkanes are uniformly prominent. OBM filtrate contamination is generally worse
at the base of the column where the increased viscosity and higher viscosity values of the oils makes
sample acquisition more challenging.
Figure 10 depicts a comparison of the variations of the optical density measured by the DFA sensor,
with the live fluid viscosity and degased oil API gravity measured by the PVT laboratory for the four wells
in our study. The dashed curves denote the best curve fitting of the data. The fluid properties slightly
change at the top of the oil column whereas the fluids significantly become heavier toward the base of the
oil column owing to oil biodegradation at the oil-water contact.
Based on the reference fluid, and isothermal equilibrium assumption, compositional gradients with
depth were calculated by the method of Zuo et al. (2010, 2013). The model predicted results indicate there
is almost no compositional grading in the oil column. Therefore, the cubic EoS failed to predict the
asphaltene and viscosity gradients, in particular, in the lower half of the oil column. The oils in our
reservoir have low GOR values, ranging from about 9 to 16 m3/m3; and with a relatively high asphaltene
content. As would be expected and previously discussed, conventional cubic EoS methods cannot treat
asphaltene concentration gradients successfully (Freed et al. 2010, 2014, Zuo et al. 2010, 2013). The FHZ
SPE-170794-MS 11

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 10 —(a) Optical density variation; (b) Live fluid viscosity variation; (c) API gravity variation. The dashed blue curves are the best polynomial
fit. Fluid properties slightly change at the top of the oil column whereas fluids significantly become heavier toward the base of the oil column owing
to severe biodegradation at the oil-water contact.

EoS was then employed to predict the asphaltene (or optical density) gradient using 2 nm in diameter of
asphaltene nanoaggregates in terms of the Yen-Mullins model (Mullins 2010, Mullins et al. 2012). The
FHZ EoS with the gravity term only can be used to estimate the asphaltene gradient due to low- GOR oil.
That is, the solubility term of the FHZ EoS is largely invariant for a relatively homogenous low-GOR
crude oil at low reservoir pressure as predicted by the cubic EoS. The predictions by the FHZ EoS are
illustrated in Figure 11. The FHZ EoS is working well at the top of the oil column. The lower sands have
much bigger optical density (asphaltene wt.%) than the equilibrium distribution predicted by the FHZ
EoS. The optical density is higher than the equilibrium value by a factor of 3 at the base of the oil column.
The implication is that thermodynamic equilibrium of asphaltenes applies over the upper half of the oil
column, while a significant disequilibrium in asphaltene concentration applies in the lower half of the oil
column.
Simple Diffusion Model for Evaluation of Reservoir Oil Biodegradation
Although asphaltene nanoscience models, such as the Flory-Huggins-Zuo Equation of State (FHZ EoS)
utilizing the Yen-Mullins model of asphaltenes (Mullins 2010, 2011, Zuo et al. 2013) are well established
– the impact of oil biodegradation now needs to be taken into account. In this section we discuss
observations from our case study and demonstrate how we may incorporate biodegradation in our FHZ
EoS models by incorporating a term for diffusion.
The schematic diagram shown in Figure 11 depicts a one dimensional model of the reservoir and oil
biodegradation system and process. For our sequence pf reservoir units with a common initial oil-water
contact as shown in Figure 1 (left), its dip angle (␣) can be used to convert the real distance [h/sin(␣)] to
the true vertical depth subsea (TVDSS, h) as shown in Figure 12.
At the oil-water contact, depth h ⫽ 0. There is a common water saturated zone below the oil-water
contact and a multilayered reservoir sequence which is oil bearing above the oil-water contact. Following
the same approach as the development of the FHZ EoS, whole oil components are divided into two
groups: maltenes (m) and asphaltenes (a). Microbes can consume components of alkanes (and some
12 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 11—Comparison of the DFA optical density measurements with FHZ EoS predictions. The upper section of the oil column within the reservoir
can be accounted for by applying the FHZ EoS model with asphaltene nanoaggregates. However, the oils in the lower part of this reservoir sequence
have much larger optical density values (asphaltene wt.%) than the equilibrium distribution predicted by the FHZ EoS for the upper section of this
reservoir. From these measurements and the model results we conclude the lower sequence of the oil column within this reservoir is not equilibrated.

Figure 12—Schematic diagram to illustrate the oil biodegradation system and process. (a) Maltene (mainly alkane) concentration distribution; (b)
Asphaltene concentration distribution. C and h are the mole concentration (it can be replaced by mass concentration) and TVDSS. At the oil-water
contact, h ⴝ 0, h increases upwards.

alkylaromatics) in the group of maltenes. In the following, for simplicity, the components consumed by
microbes are referred to as alkanes.
To simplify the complex processes involved in oil biodegradation, and establish a simple diffusive
model for biodegradation, we make the following assumptions:
1) Microbes consume the alkanes rapidly in comparison with diffusion times of the components to the
oil-water contact. The rate-limiting step is presumed to be alkane diffusion down to the oil-water
contact.
2) Initial oil compositions of the reservoir are equilibrated.
3) Biodegradation occurs at the oil-water contact where microbes live (Peters et al. 2005, Head et al.
2003, Larter et al. 2003).
SPE-170794-MS 13

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 13—Asphaltene mass percentage with time and depth. The green solid curve is the equilibrium distribution calculated by the FHZ EoS.
Biodegradation tends to lead to an increase in asphaltene content.

4) Diffusion of asphaltenes upwards the oil column is ignored because biodegradation makes solvents
(maltenes) better for dissolving asphaltenes that may compensate upward diffusion of asphaltenes.
This diffusive model, and the initial and boundary conditions and parameters, together with solutions
and supporting equations to couple diffusion with the FHZ EoS and Yen-Mullins model are presented in
detail by Zuo et al. (Manuscript in prep.).
Application of the FHZ EoS with Diffusive Model to the Case Study:
Results and Discussion
The diffusive model was applied to the oil reservoir with the biodegraded reservoir sequences which has
been previously described. For this study it is assumed the effective diffusion coefficient, D ⫽ 5.0⫻10⫺7
cm2/s for maltenes (actually alkanes) and the dip angle is assumed to be 15°. Following our approach we
calculate the maltene concentration distributions with time and depth. To apply this method, two partial
densities of maltenes at initial conditions (or at the top of the oil column) and at the base of the oil column
are required. At the top of the oil column, the FHZ EoS is applied to compute the equilibrium asphaltene
gradient (then obtaining asphaltene weight fraction) and fluid mass density (␳0) was measured. Therefore,
the initial mass concentration of the maltenes can be determined. Similarly, at the base of the oil column,
mass density and mass fraction of maltenes were also measured. Thus, the change in asphaltene
concentration is known as well. Therefore, the only undetermined parameter is time, which can be
obtained by fitting the measured asphaltene gradient data of the lower half of the oil column.
The initial fluid of the oil column is assumed to be equilibrated. At the beginning of biodegradation,
microbes consume all alkanes at the oil-water contact, thus yielding an alkane (maltene) concentration
contrast. The alkane concentration contrast makes the alkanes diffuse downwards. Then the microbes
consume the alkanes diffused down at or near the oil-water contact. Because biodegradation processes are
more rapid than diffusion, the limiting step is alkane diffusion. This process continues slowly. The alkanes
gradually disappear upward and the concentration of maltenes decreases upward to shallower depths with
time. Reservoir fluid density and fluid density gradients with depth can be measured at reservoir
conditions by downhole fluid analysis techniques (O’Keefe et al. 2007), and/or determined by measure-
ments from fluid samples in the PVT laboratory. The oil density is then able to be populated based on
these measurements. With the density gradient distribution, the mass concentration of asphaltenes is then
calculated in our approach, and the mass fraction of asphaltenes is also computed. Details of our approach
14 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 14 —The optical density gradient predicted by the FHZ EoS and the diffusion model at t ⴝ 7.5ⴛ106 years and ␣ ⴝ 15°. The FHZ EoS matches
the DFA measurements OD very well across the upper section of the reservoir where no oil biodegradation effects are observed. The FHZ EoS with
the diffusive model terms is in good agreement with the measured OD gradient across the lower reservoir section of the oil column.

are provided in Zuo et al. (Manuscript in preparation). The calculated asphaltene mass percentage with
depth and time for our India reservoir example is illustrated below in Figure 13.
Then time (t) is adjusted to match the DFA measured OD data where asphaltene weight fraction is
converted to OD using a correlation developed from analysis of the phase behaviour from laboratory
measurements and experimental data. The optimized time t ⫽ 7.5⫻106 years if D ⫽ 5⫻10⫺7 cm2/s and
␣ ⫽ 15°. As previously described, the reservoir comprises of formations of Palaeocene age. The formation
temperature varies from 55 to 61 °C. The oils within the reservoir in our case study are likely still
undergoing active biodegradation; with 7.5 million-year time frame for biodegradation a reasonable value
and consistent with many published.
The OD fitting result from optimized time (t ⫽ 7.5⫻106 years) is depicted in Figure 14. The FHZ EoS
gives the equilibrium distribution of optical density, which matches the DFA sensor measured OD values
very well in the upper section of the oil column, where no biodegradation effects are apparent. The
diffusive model matches the OD gradient nicely in the lower half of the oil column. Because the
biodegradation time is not long enough, the fluids within the upper reservoir units are still containing
n-alkanes. This is consistent with the gas chromatograms as shown in Figures 8(b) and 9 and it also
explains why the FHZ EoS approach can match the upper reservoir section.
As described in the literature by Head et al. (2003) and Larter et al. (2003), many oil reservoirs with
high levels of alteration by biodegradation have typically lost up to 50% of their mass of C6⫹ components.
This observation is in agreement with the factor of 3 increase of asphaltenes from equilibrium distribution
predicted by the FHZ EoS with 2 nm in diameter at the base of the oil column in our India study. The loss
of alkane and some alkyl-aromatic components decreases the oil volume yielding an increase in
asphaltene concentration.
Oil viscosity plays a central role in well productivity and displacement efficiency, and impacts
completion strategies and field development plans. Accurately assessing areal and vertical variations of
viscosity will lead to more realistic reservoir simulation and field development models. Since it has been
demonstrated that oil viscosity varies exponentially with optical density (OD) (Zuo et al. 2012), then
viscosity can be related to the OD calculated by the diffusive model.
Figure 15 compares the viscosity calculated by the equation above, with the measured viscosity from
analysis of the PVT samples acquired during formation testing and sampling. The modeled viscosity is in
SPE-170794-MS 15

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 15—Viscosity gradient predicted by the viscosity model compared with viscosity measurements from the reservoir fluid samples. Oil viscosity
increases by a factor of ~8 due to biodegradation near the base of the oil column.

very good agreement with the laboratory measurements of the oil viscosity performed on the fluid samples
acquired during formation testing.
Discussion of Case Study and Biodegraded Oil Systems and Reservoir
Charge Histories
For some reservoirs involving oil biodegradation, multiple oil charges and the reservoir charge history
must be considered. For example, in a case study of an oil reservoir in Brazil, the reservoir crude oil was
shown to be enriched in n- alkanes indicating little biodegradation, and in 25-norhopanes indicating
extensive biodegradation. The explanation in this case was that the first reservoir charge occurred when
the reservoir temperature was less than 80 degC, and extensive biodegradation of this oil occurred (leading
to formation of 25-horhopanes) (Pomerantz et al. 2010). Subsequently the reservoir underwent subsidence
and heated beyond 80 degC, thereby causing paleopasteurization of the reservoir. Subsequent to this
event, a further charge of oil into the reservoir occurred accounting for the presence of n- alkanes in the
reservoir crude oil (Pomerantz et al. 2010). Moreover, after all these processes, the asphaltenes then
equilibrated at least locally within the reservoir. More recently in an unpublished technical study from the
North Sea, a series of petroleum “spill-charge” reservoirs have been observed showing an increase in
asphaltene concentration gradients where the Peters-Moldowan rank ranges from 1 (little biodegradation)
to 4 (significant biodegradation).
For the shallow and low pressure reservoir described in our India case study, the reservoir temperature
remains quite low, so biodegradation is current. A multiple charge scenario of the sort previously
discussed does not apply for our India case study. Nevertheless, we must consider whether current
reservoir charging is playing a significant role in establishing the hydrocarbon distributions in this
reservoir. Lets presume the Stainforth charge mechanism applies where a charge of late, light hydrocar-
bons undergoes piston-like displacement of an earlier charge of heavier hydrocarbon (Stainforth et al.
2004). Such a sequence typically applies during basin development and subsidence processes, where the
later reservoir charge is more mature. Active biodegradation would only increase the density of the earlier
charge strengthening the gravity segregation which causes the layering in the Stainforth charge mecha-
nism. If significant charging were current, then there would not be equilibrated asphaltenes in the upper
half of the reservoir. Instead, the asphaltene gradient would be in accord with that of the charge sequence.
The length of time to equilibrate asphaltenes via diffusion in the upper half of the column is directly
comparable to the length of time required to diffuse alkanes to the oil-water contact in the lower half of
16 SPE-170794-MS

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Figure 16 —(Left) A schematic diagram to illustrate the diffusive model. (Right) Diagram of the case study optical density data shown with FHZ Eos
model results with diffusion term for evaluation of reservoir oil biodegradation. From the study and the results this suggests biological alteration and
biodegradation of these oils has occurred within the reservoir, i.e. in situ.

Figure 17—Diagram displaying the main reservoir fluid regions and assessment in 3D with a fluid model. This example shows two different fluid
regions in the reservoir which are connected in this hydrocarbon zone; with water below the hydrocarbon column.

the reservoir. Because the asphaltenes are determined to be equilibrated in the upper half of the reservoir,
we conclude that current charging is not significant, and is not determining the distribution of hydrocarbon
types in this reservoir. this analysis is enabled by employing the industry’s first EoS for asphaltene
gradients, it is now possible to know what an equilibrated asphaltene gradient is. Consequently, for the
analysis of our India reservoir in question as a diffusive gradient is reinforced, the contribution of current
reservoir charging is not significant.
Conclusions
In this case study of an oilfield in Rajasthan, India, we have described monotonic variations of asphaltene
concentration and a viscosity gradient. Extensive fluid property measurements, both downhole and
laboratory, were conducted during this study. Fluid data from four wells across the field were acquired
SPE-170794-MS 17

showing consistency across the formation and reservoir sequences. In particular, accurate asphaltene
concentration gradient measurements were performed utilizing downhole fluid analysis. Combined with
the FHZ EoS, a simple diffusive model was developed for the entire observed asphaltene distribution in
each of the reservoir units within the reservoir. This model and the model results are illustrated in Figure
16.
In Figure 16, the crude oil in the upper half of the oil column exhibits an equilibrium distribution of

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


asphaltenes matching predictions of the Flory-Huggins-Zuo Equation of state (FHZ EoS) using asphaltene
nanoaggregates of the Yen-Mullins model. However, the bottom half of the reservoir reveals a large
asphaltene gradient approximately three times larger than the equilibrium predictions from the FHZ EoS.
This increase in asphaltenes creates a very large (8x) viscosity gradient and is a major production concern.
In addition, these shallow reservoirs are undergoing active biodegradation at temperatures of 55 °C to 61
°C.
A simple diffusive model coupled with the FHZ EoS is shown to account for the entire observed
asphaltene distribution in each of the five sand layers. Alkanes (and some aromatics) are rapidly consumed
at the oil-water contact at the base of the oil column. The rate limiting step is the diffusion of these
compounds to the oil-water contact. The loss of these oil components decreases the oil volume yielding
an increase in asphaltene concentration and a significant increase in viscosity. The ability to account for
asphaltene concentration and viscosity variations in this stacked reservoir sequence by applying a simple
diffusive model coupled with the FHZ EoS and the Yen-Mullins model provides a robust model for input
to reservoir engineering studies, and application to field development planning for well placement, water
flood design and to help improve production efficiency.
Figure 17 illustrates how our the results of our analysis and fluid models can be incorporated in a 3D
reservoir model – in this particular example for we can display the water saturated zones and two different
fluid regions (upper and lower reservoir intervals) which are connected in the hydrocarbon zone of our
reservoir.

Acknowledgements
The authors thank Cairn India Ltd and Schlumberger management for giving permission to present this
paper. Martyn Beardsell is also thanked for his help with prototyping and developing new interpretation
software workflows for visualization and analysis of formation testing and DFA measurements within a
reservoir model.

References
Akkurt, R.; Bowcock, M.; Davies, J.; Del Campo, C.; Hill, B.; Joshi, S.; Kundu, D.; Kumar, S.;
O’Keefe, M.; Samir, M.; Tarvin, J.; Weinheber, P.; Williams, S.; Zeybek, M. 2007. Focusing on
Downhole Fluid Sampling and Analysis. Oilfield Review Winter 2006/2007, 18(4), 4 –19
Bennett, B.; Adams, J.J.; Gray, N.D.; Sherry, A.; Oldenburg, T.B.P.; Huang, H.; Larter S.R.; Head,
I.M. 2013. The controls on the composition of biodegraded oils in the deep subsurface – Part 3. The
impact of microorganism distribution on petroleum geochemical gradients in biodegraded petroleum
reservoirs, Org. Geochem. 2013, 56, 94 –105.
Brown, A. 2003. Improved Interpretation of Wireline Pressure Data. AAPG Bulletin, 87, February
2003, 295–311.
Compton, P. M. 2009. The geology of the Barmer basin, Rajasthan, India, and the origins of its major
oil reservoir, the Fatehgarh Formation, Petroleum Geoscience, 15, 117–130.
Freed, D. E., Mullins, O. C., Zuo, J. Y. 2010. Theoretical Treatment of Asphaltenes in the Presence
of GOR Gradients. Energy & Fuels, 24(7), 3942–3949.
Freed, D. E.; Mullins, O. C.; Zuo, J. Y. 2014. Heuristics for Equilibrium Distributions of Asphaltenes
in the Presence of GOR Gradients, In Press – Submitted. Energy & Fuels, 2014.
18 SPE-170794-MS

Fujisawa, G.; Van Agthoven, M.A.; Rabbito, P.; Mullins, O.C. 2002. Near-Infrared Compositional
Analysis of Gas and Condensate Reservoir Fluids at Elevated Pressures and Temperatures. Applied Spec.
56, 1615–1620.
Goual, L.; Sedghi, M.; Zeng, H.; Mostowfi, F.; McFarlane, R.; Mullins, O.C. 2011. On the Formation
and Properties of Asphaltene Nanoaggregates and Cluster by DC-Conductivity and Centrifugation. Fuel,
90, 2480 –2490.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Goual, L.; Mostowfi; F.; Pomerantz, A.E.; McFarlane, R.; O.C. Mullins, O.C. Cluster of Asphaltene
Nanoaggregates by DC-Conductivity and Centrifugation, 2014 (Submitted to Energy & Fuels).
Head, I. M., Martin Jones, D., and Larter, S. J. 2003. Biological Activity in the Deep Subsurface and
the Origin of Heavy Oil. Nature, 426, November 2003, 344 –352.
Korb, J.P.; Louis-Joseph, A.; Benamsili, L. 2013. Probing Structure and Dynamics of Bulk and
Confined Crude Oils by Multiscale NMR Spectroscopy, Diffusometry, and Relaxometry. J. Phys. Chem.
B 2013, 117, 7002–7014.
Larter, S.R.; Wilhelms, A.; Koopmans, M.; Aplin, A.; Di Primio, R.; Zwach, C., Erdmann, M.;
Telnaes, N. 2003. The Controls on the Composition of Biodegraded Oils in the Deep Subsurface—part
1: Biodegradation Rates in Petroleum Reservoirs. Org. Geochem. 2003, 34, 601–613.
Larter, S. J., Huang, H., Adams, J., Bennett, B., Jokanola, O., Oldenburg, T., Martin Jones, J., Head,
I., Riediger, C., Fowler, M. 2006. The Controls on the Composition of Biodegraded Oils in the Deep
Subsurface: Part II – Geological Controls on Subsurface Biodegradation Fluxes and Constraints on
Reservoir-Fluid Property Prediction. AAPG Bulletin, 90, 921–938.
Majumdar, R.D.; Gerken, M.; Mikula, R.; Hazendonk, P. Validation of the Yen–Mullins Model of
Athabasca Oil-Sands Asphaltenes using Solution-State 1H NMR Relaxation and 2D HSQC Spectroscopy.
2013. Energy & Fuels, 27 (11), 6528 –6537.
Mullins, O.C.; Daigle, T.; Crowell, C.; Groenzin, H.; Joshi, N.B. 2001. Gas-Oil Ratio of Live Crude
Oils Determined by Near-Infrared Spectroscopy, Applied Spectros. 55, 197–201.
Mullins, O. C. 2008. The Physics of Reservoir Fluids: Discovery through Downhole Fluid Analysis.
Schlumberger Press, Houston, TX, USA.
Mullins O. C. 2010. The Modified Yen model. Energy Fuels, 24(4), 2179 –207.
Mullins, O. C. 2011. The Asphaltenes. Annual Review of Analytical Chemistry, 4, 393–418.
Mullins, O. C.; Sabbah, H.; Eyssautier, J.; Pomerantz, A.; Barre, L.; Andrews, A.; Ruiz-Morales, Y.;
Mostowfi, F.; McFarlane, R.; Goual, L.; Lepkowicz, R.; Cooper, T.; Orbulescu, J.; Leblanc, R.; Edwards,
J.; Zare, R.N., 2012. Advances in Asphaltene Science and the Yen-Mullins Model, Energy & Fuels, 26,
3986 –4003.
O’Keefe, M., Godefroy, S., Vasques, R., Agenes, A., Weinheber, P., Jackson, R. R., Ardila, M.,
Wichers, W., Daungkaew, S., and De Santo, I. 2007. In-Situ Density and Viscosity Measured by Wireline
Formation Testers, Paper SPE110364 presented at the SPE Asia Pacific Oil and Gas Conference and
Exhibition, Jakarta, Indonesia, October 30 –November 1, 2007.
Peng, D.-Y.; Robinson, D. B. 1976. A New Two-Constant Equation of State. Ind. Eng. Chem.
Fundam., 15, 59 –64.
Peters, K.E.; Walters, C. C.; Moldowan, J. M. 2005. The Biomarker Guide: Biomarkers and Isotopes
in Petroleum Systems and Earth History, Volumes 1 & 2, Cambridge University Press.
Pomerantz, A.E.; Ventura, G.T.; A.M McKenna, J.A. Cañas, J. Auman, K. Koerner, D. Curry, Nelson,
R.L.; Reddy, C.M.;Rodgers, R.P.; Marshall, A.G.; K.E. Peters, Mullins, O.C. 2010. Combining Biomarker
and Bulk Compositional Gradient Analysis to Assess Reservoir Connectivity, Org. Geochem. 41 (8),
812–821.
SPE-170794-MS 19

Seifert, D. J., Zeybek, M., Dong, C., Zuo, J. Y., Mullins, O. C. 2012. Black Oil, Heavy Oil and Tar
in One Oil Column Understood by Simple Asphaltene Nanoscience. Paper SPE 161144 presented at the
Abu Dhabi International Petroleum Exhibition and Conference held in Abu Dhabi, UAE, 11-14 November
2012.
Stainforth, J. G. 2004. New Insights into Reservoir Filling and Mixing Processes, in Cubit, J.M.,
England, W.A., Larter, S. (Eds.), Understanding Petroleum Reservoirs: Toward an Integrated Reservoir

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023


Engineering and Geochemical Approach, Geological Society, London, Special Publication, 2004.
Zittel, R. J., Beliveau, D., O’Sullivan, T., Mohanty, R., and Miles, J. 2008. Reservoir Crude Oil
Viscosity Estimation from Wireline NMR Measurements – Rajasthan, India. SPE Reservoir Evaluation &
Engineering, June 2008, 545–553.
Zuo, J. Y.; Freed, D.; Mullins, O. C.; Zhang, D.; Gisolf, A. 2010. Interpretation of DFA Color
Gradients in Oil Columns Using the Flory-Huggins Solubility Model. Paper SPE 130305. Presented at the
CPS/SPE International Oil & Gas Conference and Exhibition in China, Beijing, China, 8 –10 June 2010.
Zuo, J. Y.; Freed, D.; Mullins, O. C.; Zhang, D. 2010. DFA Profiling of Oil Columns with Asphaltene
Gradients. Paper SPE 133656. Presented at the SPE Annual Technical Conference and Exhibition held in
Florence, Italy, 19 –22 September 2010.
Zuo, J. Y.; Mullins, O. C.; Freed, D.; Zhang, D.; Dong, C.; Zeng, H. 2011. Analysis of Downhole
Asphaltene Gradients in Oil Reservoirs with a New Bimodal Asphaltene Distribution Function. J. Chem.
Eng. Data, 2011, 56(4), 1047–1058.
Zuo, J. Y.; Mullins, O.C.; Freed, D.; Elshahawi, H.; Dong, C.; Seifert, D.J. 2013. Advances in the
Flory–Huggins–Zuo Equation of State for Asphaltene Gradients and Formation Evaluation. Energy &
Fuels, 2013, 27, 1322–1335.
Zuo, J. Y., Mullins, O. C., Elshahawi, H., Ramaswami, S., Dong, C., Dumont, H., Zhang, D., and
Ruiz-Morales, Y. 2013. Advanced Formation Evaluation Using DFA and Asphaltene Flory-Huggins-Zuo
EOS. Paper SPE 164596 presented at the North Africa Technical Conference and Exhibition held in Cairo,
Egypt, 15-17 April 2013.
Zuo, J. Y., Jackson, R. R., Agarwal, A., Herold, B., Kumar, S., De Santo, I., Dumont, H., Ayan, C.,
Beardsell, M., and Mullins, O. C. Manuscript In Preparation. A Diffusion Model Coupled with the
Flory-Huggins-Zuo Equation of State and Yen-Mullins Model Accounts for Large Viscosity and As-
phaltene Variations in a Reservoir Undergoing Active Biodegradation. Manuscript in Preparation, Energy
& Fuels.
Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/14ATCE/All-14ATCE/SPE-170794-MS/1496036/spe-170794-ms.pdf/1 by Schlumberger Oilfield UK Plc user on 14 February 2023
SPE-170794-MS
20

You might also like