You are on page 1of 18

SPWLA 57th Annual Logging Symposium, June 25-29, 2016

ROCK CLASSIFICATION IN THE EAGLE FORD SHALE THROUGH


INTEGRATION OF PETROPHYSICAL, GEOLOGICAL, AND
GEOCHEMICAL CHARACTERIZATION
Shahin Amin, and Matthew Wehner, Texas A&M University, Zoya Heidari, The University of Texas at Austin, and
Michael Tice, Texas A&M University

Copyright 2016, held jointly by the Society of Petrophysicists and Well Log Ratio between the two wells. For the interbedded
Analysts (SPWLA) and the submitting authors.
This paper was prepared for presentation at the SPWLA 57th Annual Logging wackestone-limestone facies, YM average estimate in
Symposium held in Reykjavik, Iceland June 25-29, 2016. well no. 1 was approximately 10% higher than well no.
2, which can be the reason for the difference in their
ABSTRACT production.

Formation evaluation and production design is often INTRODUCTION


challenging in organic-rich mudrocks due to
complexities in petrophysical and compositional Reliable rock classification in organic-rich mudrocks
properties, and post-depositional hydrocarbon should take into account petrophysical, compositional
generating mechanisms such as thermal maturation over and geomechanical properties of the formation.
time. Petrophysical parameters such as porosity, Properties such as storage capacity, fluid saturations,
permeability and fluid saturations are important, but not mineralogy, along with quantity and type of organic
sufficient to fully characterize organic-rich mudrocks. matter are important factors in formation evaluation of
Integration of petrophysical, geochemical and organic-rich mudrocks. Furthermore, inclusion of elastic
geomechanical data is therefore required for a reliable properties in rock classification is essential for the
rock classification in source rocks. This paper focuses on assessment of hydraulic fracturing performance.
integrated rock classification in the Eagle Ford Shale in However, due to the complexity in matrix composition
South Texas, consisting of organic-rich fossiliferous and pore structure of organic-rich mudrocks, assessment
marine shale deposited in Late Cretaceous. of these parameters is challenging, and requires proper
integration of core and well-log data.
We first performed joint inversion of triple-combo,
spectral gamma ray and elemental capture spectroscopy Geochemical data have been substantially acquired and
(ECS) logs to estimate depth-by-depth volumetric interpreted in organic-rich mudrock formations.
concentration of minerals, porosity, and fluid Integration of geochemical data (i.e. X-ray fluorescence
saturations. In the absence of acoustic measurements, or ECS logs) with core mineral concentration
concentrations and shape (i.e., aspect ratio) of minerals measurements such as X-ray diffraction (XRD) enables
were used as inputs to the Self-consistent Approximation accurate quantitative mineral modeling, leading to a
(SCA) model, to estimate depth-by-depth effective reliable estimation of petrophysical and compositional
elastic properties such as Young's Modulus (YM) and properties. Improvement of the results is achieved with
Poisson's Ratio (PR). We then classified the rocks based the use of statistical element-mineral or mineral-mineral
on geologic texture and geochemical properties, as well relationships in the petrophysical model (Quirein et al.,
as well-log based estimates of petrophysical and 2010). Interpretation of inorganic geochemical data is
geomechanical parameters. also used in determination of clay types, and modeling
total organic carbon. Furthermore, analysis of trace
We successfully applied a well-log based rock element concentrations in the rocks provides insights on
classification to two wells located in the oil window of paleoproductivity and paleoredox conditions at the site
Eagle Ford formation. Well no. 1 produced an additional of deposition of organic-rich mudrocks (Wright and
20% of hydrocarbons in the first 90-day of its Ratcliffe, 2010). Redox sensitive trace metals, which are
production. Through the analysis of the results, we less soluble under reducing conditions are enriched in
observed similar petrophysical properties and organic oxygen depleted sedimentary facies. However, some
content of the reservoir quality classes in both wells. trace elements require both reducing conditions and
However, we noticed differences in estimates of elastic organic matter to accumulate appreciably in shales.
parameters such as Young’s Modulus and Poisson’s Thus, elements such as Mo, V, and U provide
1
SPWLA 54th Annual Logging Symposium, June 22-26, 2013

information about redox state at the water column during identified and grouped along the zone of interest.
deposition and also quality of organic Petrophysical and geochemical measurements were
matter/hydrocarbons (Tribovillard et al, 2006; conducted on more than 200 core samples, where gas-
Brumsack, 2006; Lewan, 1984). filled porosity, pressure decay permeability, and TOC
were the discriminant factors between rock classes in the
In the past, geoscientists have introduced different Haynesville shale.
techniques to classify sedimentary rocks based on
similarities in particular properties. Among previous Production of hydrocarbons from organic-rich mudrocks
rock classification studies, Archie classified carbonates is not only dependent on the petrophysical properties. To
based on their pore structure (Archie, 1952). Archie’s maximize wellbore-reservoir contact, and create high
porosity-based classification was among the earliest to permeability pathways for hydrocarbon flow, practices
differentiate limestones based on petrophysical such as horizontal drilling, and hydraulic fracturing are
properties, for selection of reservoir quality rocks widely used in the industry. Hence, geomechanical
(Archie, 1952). Increase in availability and use of thin- evaluation in organic-rich mudrocks is required to assess
section images, particularly for interpretation of the effectiveness of these completion operations. In a
depositional environments, shaped the texture-based well-log-based rock classification approach that
rock classification schemes. Classification based on accounted for elastic properties, Saneifar et al. (2013)
depositional texture found to be helpful adjunct to other included a brittleness index as an additional factor of
classifications such as mineralogical classifications. Folk rock quality to organic-richness, kerogen porosity, and
(1962) introduced a detailed texturally-based quartz and illite volumes. This integrated study used a
classification that incorporated grain size, roundness, predetermined number of geological facies in
sorting and packing as well as grain composition using Haynesville Shale as an input to the classification. The
microscopic images. In a similar approach, Dunham results of the classification was then cross-validated with
(1962) classified carbonate rocks based on the ratio of the thin-section petrography images for each class.
grain to matrix, into two main groups: grain-supported, Furthermore, Aderibigbe et al. (2016) incorporated stress
and mud-supported. As of today, the original and profiles in a production-oriented rock classification,
modified versions of Dunham’s carbonate classification applied to the Wolfcamp formation. The completion
is still being used by geologists in interpretation of intervals were selected based on the final results of the
depositional environments. integrated rock classification.

Variety of methods and algorithms have been proposed Previous publications have demonstrated the significant
in the past, to characterize and classify organic-rich impact of reliable rock classification on enhancement of
mudrocks. Kale et al. (2010) studied 800 core plugs from formation evaluation, and subsequently an improved
four wells to identify and group similar facies in a selection of completion intervals (Suarez-Rivera et al.,
quantitative core-based classification. Statistical 2011; Petriello et al., 2013). Reliability of rock
methods such as principal component analysis (PCA), classification can be assessed depending on the input
and clustering algorithms were used to group similar parameters and its application (e.g., reserve- or
rocks based on porosity, mineralogy, total organic production-oriented classification). In this paper we
carbon (TOC) and capillary pressure measurements. conduct a geochemical and geological rock facies
Although, quantitative core-based classifications are classification in the lower Eagle Ford (LEF) formation,
valuable, the data acquisition requires substantial and determine an optimum number of petrophysical rock
amount of laboratory time at an additional cost. classes. Then, we develop a well-log-based algorithm
Therefore, well-log-based classifications are favored to that assimilates petrophysical, compositional and
the industry providing complete vertical coverage, geomechanical properties of the rocks. Furthermore, we
availability of data at all wells, and lower costs. In a well- apply the introduced petrophysical rock classification to
log-based study, Popielski et al. (2012) conducted a wells located in the field where core data is not available.
classification using “k-means” clustering and factor Finally, we compare the results of the petrophysical rock
analysis (i.e., used in reduction of the dimension of input classification in wells with different hydrocarbon
data), to group rocks with similar well-log responses, and production history. The main contribution of this paper
well-log derived estimates. Furthermore, Petriello et al. is introduction of an integrated rock classification
(2013) introduced a well-log-based multivariate scheme that honors geochemical, geological, and
classification that employed Heterogeneous Rock petrophysical rock properties. Such rock classification
Analysis (HRA) algorithm. Similar rock patterns were approach is required in organic-rich mudrocks for
2
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

improving decisions made for well completions and element-to-mineral and mineral-to-mineral as additional
production planning. constraints in the multi-mineral analysis to decrease non-
uniqueness of the inversion results and to improve the
The following sections include the methods employed reliability of the estimated properties. Finally, we
for geochemical, geological and petrophysical compare the well-log-based estimates of mineral
classifications, as well as the results obtained in the concentrations against core XRD measurements to
lower Eagle Ford formation. validate the reliability of the multi-mineral model. We
estimate water saturation in the lower Eagle Ford using
METHOD the modified Simandoux model (Bardon and Pied,
1969). The constant parameters in the model are obtained
This section includes the step-by-step procedures through calibration against core measurements. The
conducted to obtain geochemical, geological, and same constant parameters are used for water saturation
petrophysical classifications in the lower Eagle Ford assessment in wells where core data is not available.
formation. We start with well-log-based petrophysical
evaluation of the formation including the assessment of Depth-by-depth assessment of elastic moduli. In the
elastic properties. Then we describe the methods for absence of acoustic logs, we apply the self-consistent
geochemical, geological, and petrophysical approximation (SCA) model to calculate depth-by-depth
classification techniques. Finally, we integrate the results elastic properties such as bulk and shear moduli. SCA
of the petrophysical, geochemical, and geological theorem assumes that the constituents of the mixture are
classifications to select the reservoir quality rock classes, immiscible, in an isotropic elastic media. The inputs to
and identify the candidate intervals for well completions. the SCA equations are (a) volumetric concentration of
Finally, we apply the well-log based rock classification each constituent, (a) elastic moduli specific to each
to a nearby well with a different hydrocarbon production, component, and (c) the geometric shape factor associates
to investigate possible factors affecting the well with that component (Berryman, 1995). SCA equations
performance. to solve for composite bulk and shear moduli are
described by
Well-log-based petrophysical evaluation. We determine
Lower Eagle Ford boundaries based on gamma ray and N
uranium concentration from spectral gamma ray logs. A x K K * P*i 0 (1)
i i SC
significant increase in total gamma ray and uranium i 1
content determines the boundary, when moving from
upper to lower Eagle Ford. Furthermore, a unique pyrite- and
rich marker bed indicates the transition to lower Eagle
Ford from the top (Lock et al, 2011). In the bottom of the N
x µ µ* Q*i 0, (2)
lower Eagle Ford, the boundary with Buda limestone is i i SC
i 1
well-known due to the changes in basic well-log
responses such as gamma-ray, neutron porosity, and where xi is the volumetric concentration of component i,
density logs. Total organic carbon content (TOC) is N is the total number of constituents, K*SC and SC are
estimated using a correlation obtained between TOC the bulk and shear moduli of the host rock, Ki and i are
(wt%) from core measurements and bulk density from the bulk and shear moduli of the component i. P*i and Q*i
well logs. Assuming a constant kerogen density, depth- are factors dependent of the shape assigned to each
by-depth volumetric concentration of kerogen is then constituent. We assign shape factors for each component
estimated. The estimates for volumetric concentration of based on the analysis of two-dimensional (2D) X-ray
kerogen are inputs to the inversion-based multi-mineral elemental distribution maps. Finally, Young’s Modulus
analysis workflow. (YM or E) and Poisson’s Ratio (PR or v) are calculated
via
We jointly interpret triple-combo and ECS logs to
estimate depth-by-depth mineral concentrations and
petrophysical properties such as porosity and fluid 9 K * µ*
E SC SC (3)
saturations. In this process, determination of the mineral *
3K µ*
types is based on X-ray diffraction (XRD) measurements SC SC
(average of one sample per 15 feet) in the lower Eagle
Ford interval. Furthermore, we acquire correlations of and

3
SPWLA 54th Annual Logging Symposium, June 22-26, 2013

3K * - 2 µ * oil equivalent per day (MMBOE/d) accompanied by 5


SC SC . (4) billions of cubic feet per day (BCF/d) of natural gas
v
2 3K * µ* production in 2015 (U.S. Energy Information
SC SC
Administration (EIA) , 2015). The Eagle Ford discovery
well was drilled in LaSalle County by Petrohawk in
Geological rock classification. Variation in depositional 2008, with initial flow of 7.6 MMCF/d of natural gas.
environment moving upward from the base of Eagle Since then, over 20,000 Eagle Ford drilling permits were
Ford is noticeable in the texture of the extracted cores. issued in Texas, with substantial increase in production
We identify these differences in texture based on full every year (Texas Railroad Commission, 2015). During
core photos in addition to thin-section images. The the Jurassic and Cretaceous periods in the northwest part
names assigned to the lithofacies are based on (a) the of the Gulf of Mexico, the paleotopography was related
ratio of grain to matrix (percentages of fossils in the to the evolution of the Comanche Reef Platform as it
matrix) according to Dunham’s carbonate classification grew and eroded based on relative sea level. The
(b) degree of lamination, and (c) type of the existing deposition of the Eagle Ford sediments was on a
fossils. carbonate platform that had uneven topography due to
erosion (Faust, 1990; Donovan and Staerker, 2010).
Geochemical rock classification. To classify rocks based Thus, when the platform began to be flooded, sometime
on inorganic geochemical data, we collect depth-by- from early to middle Cenomanian at the beginning of the
depth core X-ray fluorescence (XRF) measurements in Late Cretaceous, the basins experienced restriction and
the lower Eagle Ford formation. Through the analysis of episodic anoxia. There is evidence that the event
major and trace element concentrations, we classify the depositions in the Eagle Ford Group not only included
formation into chemostratigraphic units. The ash bed falls but also storm deposits. The Eagle Ford
geochemical classification explains vertical variations in sediments primarily consists of alternating
paleo productivity and redox conditions in depositional mudstone/marlstones with limestones along with
environments, and post depositional processes such as varying levels of organic content (up to 10 wt% TOC),
diagenesis, organic maturation, and hydrocarbon and occasional bentonitic ash beds. The Lower Eagle
generation. Finally, we conduct a class-by-class analysis Ford Formation (which in the subsurface contains the
of parameters such as hydrogen index, oxygen index, and unconventional reservoir) appears to have been
thermal maturity, based on Rock-Eval Pyrolysis core deposited during basin restriction, storm deposition and
measurements. episodic anoxia; these factors likely contributed to the
ultimate incorporation of organic matter (Wehner et al.,
Petrophysical rock classification. After depth-by-depth 2015). These conditions make the Eagle Ford formation
estimation of petrophysical, compositional, and highly heterogeneous in the vertical dimension. Some of
geomechanical properties, we conduct a well-log-based this vertical heterogeneity can be predicted by sequence
rock classification, adopting a hierarchical clustering stratigraphy (Donovan and Staerker, 2010; Donovan et
algorithm to group similar rocks by minimizing the total al., 2012; Donovan et al., 2013). Gardner et al., (2013)
intra-class variance in each step (Ward, 1963). We assign investigated the lateral variation of Eagle Ford Group
the input number of rock classes based on the total sediments and noted a difference depending on whether
number of geofacies obtained from the combination of upper or lower part of the Eagle Ford Group was
core and geochemical studies. considered. However, the variation is very minor at
scales up to 5km, unlike the vertical variation. Factors
FIELD EXAMPLE: EAGLE FORD FORMATION that influence vertical variation include sea-level at time
of deposition, sediment source, organic productivity,
We applied the described method to the Lower Eagle storm frequency, and dominant diagenetic mechanisms.
Ford formation. This section includes an overview on the
geology of the Eagle Ford formation and the results of The Eagle Ford sequence overlies Buda limestone, and
the described rock classification schemes in two wells is overlain by Austin Chalk. Depositional sequences of
located in the oil window of the Eagle Ford formation. Eagle Ford are a transgressive lower member, and a
highstand upper member. The unconventional reservoir
Geologic Background of Eagle Ford formation. Eagle associated with organic-rich mudstones is in the lower
Ford formation, located in Texas, is the biggest North part of Eagle Ford. Moving from Maverick Basin in
American shale play in terms of liquid hydrocarbon South Texas to San Marcos Arch in Central Texas, the
production, peaking approximately 1.6 million barrels of upper section of Eagle Ford is truncated beneath the
4
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

unconformity at the top of Eagle Ford (Donovan and Rock-Eval pyrolysis measurements show high vertical
Staerker, 2010). Lower Eagle Ford is characterized by variation in organic properties such as TOC, hydrogen
high organic content, and can be distinguished by index (HI), and thermal maturity. Type II oil-prone
increase in Uranium content from spectral gamma ray kerogen is present in well no. 1, with Tmax values in the
logs (ranging from 5- 15 ppm in LEF section). Thin- range of 430-450 °C. In terms of mineralogy, calcite,
section image analysis is conducted in most of the quartz, illite mixed layers, and kaolinite are dominant
literature to describe the depositional texture of the rocks minerals in the depth intervals of interest. Furthermore,
based on the observed features. Dunham’s carbonate XRD measurements show pyrite weight concentrations
classification scheme is commonly used as a basis for of up to 5%. The breakdown of clay minerals indicates
textural analysis. Proportions of grains to matrix, presence of illite-smectite and illite-mica layers at the top
carbonate to clays, and degree of lamination are included interval, whereas the base of LEF contain a mixture of
in the naming of the facies. The order of facies from base illite and kaolinite.
of Eagle Ford toward the upper section, indicates a
succession from mudstones to laminated
wackestone/packstone facies. Thus, the geologic facies
are not independent of stratigraphy and depositional
environment (Harbor, 2011; Fairbanks, 2012; and
Workman, 2013). Furthermore, these authors describe
facies in several wells located in different parts of the
basin, showing a high degree of variation in the field.
Present ash beds (usually less than 2 in thick) in the Eagle
Ford are classified as massive bioturbated or bentonitic
claystones with very high clay content. The packstone-
grainstone facies are limestone beds (up to 2 ft thick),
composed of over 75% calcite. Minerals found in the
Eagle Ford consists mainly of calcite, quartz, mixed
layer clays, kerogen, and pyrite. Additionally, minor Fig. 1 The location of the two wells evaluated in this
amounts of dolomite, plagioclase, feldspars and apatite paper. They are located in the Maverick Basin at the
are reported. The breakdown of clay minerals show high Eagle Ford Shale play, north of Edwards Shelf Margin
percentages of illite mixed layers, kaolinite and minor (Modified from EIA, 2014).
chlorite content.
Petrophysical evaluation: We first estimated
The available wells for this study are located in the oil petrophysical and compositional properties in the LEF
window of Eagle Ford formation in Maverick Basin, just section of well no. 1 and well no. 2. Tracks 2-8 of Fig. 2
north of Edwards Reef Margin. Based on previously and Fig. 3 show the well-logs and well-log based
published literature, and interpretation of well-logs, the estimates of petrophysical and compositional properties
upper Eagle Ford (UEF) section appears to be truncated in well no. 1 and well no. 2, respectively. We then
beneath the unconformity of the top of Eagle Ford, applied the SCA technique to estimate depth-by-depth
having a thickness of lower than 40 ft in this area elastic moduli such as bulk and shear modulus, taking
(Donovan and Staerker, 2010; Workman, 2013; Hentz into account the mineral volumetric concentrations and
and Ruppel, 2010). The total thickness of the lower Eagle their associated grain shape factors (α). We determined
ford interval is approximately 145 feet in well no. 1 and the possible ranges for individual elastic moduli of the
125 feet in well no.2, consisting of organic-rich constituents from published literature (Mavko et al.,
marlstone-limestone laminae. Fig. 1 shows the 2009; Prasad et al., 2002; Sone et al., 2013; Vanorio et
geographic locations of well no. 1 and well no. 2. al., 2003), while the mineral shape factors were obtained
Available rock data for these wells cover the lower Eagle using X-ray elemental maps. Fig. 4 shows examples of
Ford, which is organic-rich and known to be the most core X-ray elemental distribution images in core samples
productive interval in the formation. Triple-combo, from LEF. Based on the analysis of X-ray images, quartz
spectral gamma ray, and ECS logs in addition to core and pyrite appeared to be more spherical shaped. Aspect
measurements such as XRD and Rock-Eval Pyrolysis, ratio (α) of 1 is assigned to these minerals. On the other
thin-section images, and depth-by-depth XRF hand, for clay minerals where the structure consists of
measurements were acquired for this study. nanometer scale layers, we assigned an aspect ratio of
0.01. Fig. 4 shows the identified calcite grains to be
5
SPWLA 54th Annual Logging Symposium, June 22-26, 2013

compacted and interbedded with clays in an elliptical estimated acoustic-wave slowness and those from
form at specific zones, and in another place more acoustic well logs in the wells where acoustic
spherically shaped. Therefore, we assigned two different measurements are available (e.g., well no. 2). The same
shape factors for calcite constituent (0.1 for elliptical, input parameters are used in the wells where acoustic
and 1 for spherical grains). Table 1 shows the input measurements are not recorded to estimate elastic
constants that minimized the difference between properties.

Fig. 2 Well no. 1, Lower Eagle Ford Field Example: conventional well-logs, estimates of petrophysical, compositional
and elastic properties, and the results of geochemical, geological, and petrophysical classifications. Tracks from left
to right include, Track 1: depth; Track 2: high resolution GR (HGR), caliper (HCAL); Track 3: array induction
resistivity logs (AT10-AT90); Track 4: neutron porosity (in water-filled limestone units, NPHI) and bulk density
(RHOB); Track 5: estimates of volumetric concentrations of minerals; Track 6: estimates of total porosity (POR);
Track 7: estimates of water saturation (Sw); Track 8: well-log-based estimates estimates of TOC (derived from bulk
density log) compared to core measurements; Track 9: estimates of Young’s modulus (YM); Track 10: estimates of
Poisson’s ratio (PR); Track 11: outcome the of geochemical rock classification; Track 12: outcome of the geological
rock classification; Track 13: outcome of the well-log based petrophysical rock classification (PRC)

6
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Fig. 3 Well no. 2, Lower Eagle Ford Field Example: conventional well-logs, estimates of petrophysical, compositional
and elastic properties, and the results of the petrophysical classification. Tracks from left to right include, Track 1:
depth; Track 2: high resolution GR (HGR); Track 3: array induction resistivity logs (AT10-AT90); Track 4: neutron
porosity (in water-filled limestone units, NPHI) and bulk density (RHOZ); Track 5: estimates of volumetric
concentrations of minerals; Track 6: estimates of total porosity (POR); Track 7: well-log-based estimates of water
saturation compared to core measurements (Sw); Track 8: well-log-based estimates of TOC (derived from bulk density
log) compared to core measurements; Track 9: estimates of Young’s modulus (YM); Track 10: estimates of Poisson’s
ratio (PR); Track 11: outcome of the well-log based petrophysical rock classification (PRC) based on estimates of
total porosity, water saturation, total organic carbon, total clay volume, Young’s modulus, and Poisson’s ratio.

7
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Table 1: Input parameters to the SCA model

SCA Calcite Calcite Illite Mixed


Quartz Pyrite Kaolinite Kerogen
Parameters (spherical) (elliptical) Layers

K (GPa) 75 75 38 147 8 6 5

µ (GPa) 40 40 44 132 6 2 3
α 1 0.1 1 1 0.01 0.01 0.1

Fig. 4 Mineral grains identified using 2D X-ray


elemental distribution maps from the lower Eagle Ford
core samples.

Fig. 5 shows the estimates of petrophysical and


compositional properties, as well as comparison between
the estimates of compressional- and shear-wave
slowness from SCA and acoustic well logs (DTCO and
DTSH) in well no. 2. The SCA results provide a higher
resolution assessment of acoustic compressional- and
shear-slowness. Although, we observe a higher errors in
estimates of shear-wave slowness, the approximation
remain reliable for majority of the depth intervals. Due
to limited number of X-ray images we used a single
model for the LEF interval. However, availability of
large number of high resolution core images can Fig. 5 Well no. 2, Lower Eagle Ford Field Example:
potentially enable the assignment of zonal shape factors conventional well-logs, estimates of volumetric
based on rock facies classification, resulting in a more concentration of minerals, and acoustic well-logs. Tracks
robust and reliable model. After application of the from left to right include, Track 1: depth; Track 2: high
calibrated SCA model to well no.1 and well no.2, we resolution GR, caliper; Track 3: array induction
estimated depth-by-depth values of YM and PR. Tracks resistivity logs; Track 4: neutron porosity (in water-filled
9-10 of Fig. 2 and Fig. 3 show the final estimates of YM limestone units), bulk density; Track 5: estimates of
and PR in well no. 1 and well no. 2, respectively. volumetric concentrations of minerals; Track 6: SCA-
based estimates of compressional-wave slowness
compared to the acoustic well-log measurement
(DTCO); Track 7: SCA-based estimates of shear-wave
slowness compared to the acoustic well-log
measurement (DTSH).

8
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Chemostratigraphy of lower Eagle Ford. We divided the distinguished by its low TOC content (2-4 wt%). LEF3
lower interval of Eagle Ford formation into five is similar to LEF4, except it contains higher TOC (> 4
geochemical classes (LEF1 to LEF5 in descending order wt%), and lower calcium (Ca) content. Transitioning to
from the bottom upward), based on significant LEF2, a significant increase in concentration of trace
stratigraphic variations in major elements (Ca, Si, Al, K, elements such as vanadium (V), nickel (Ni),
Fe, S), trace elements (U, V, Ni, Mo), and total organic molybdenum (Mo), and uranium (U) is apparent. At the
carbon concentrations. Fig. 6 shows depth-by-depth core very top of the lower Eagle Ford, LEF1 has a linearly
XRF measurements obtained in the lower Eagle Ford, in decreasing Ca content, and likewise, increasing Al and
well no.1 as well as the geochemical rock classification Si content. Vanadium concentration remains similar to
results. Through analysis of the results, we highlight LEF2, and a significant decrease in nickel concentration
distinguishing geochemical characteristics among the is observed. shows the Table 2 lists class-by-class
classes: the lowest unit, LEF5 is at the base of Lower distribution of the elemental concentrations (i.e., S, U, V,
Eagle Ford, on top of Buda Limestone, and is Ni), and total organic content (TOC), that were used as
chemostratigraphically distinct. Uranium content is the the basis of discrimination between the geochemical
highest in this interval, and the Si/Al ratio is very low classes. Track 11 of Fig. 2 shows the chemostratigraphic
(~1.5-3). In terms of organic geochemistry, TOC is units in well no. 1.
moderate to high (4-5 wt%). The next unit, LEF4 is

Fig. 6. Well no. 1, Lower Eagle Ford Field Example: conventional well-logs, and vertical distribution of
concentrations of major and trace elements in the lower Eagle Ford used to divide the section into five
chemostratigraphic units. Tracks from left to right include, Track 1: depth; Track 2: Chemostratigraphy of lower Eagle
Ford; Track 3: high resolution GR; Track 4: array induction resistivity logs; Track 5: neutron porosity (in water-filled
limestone units), bulk density; Tracks 6-15: calcium (Ca), silicon (Si), aluminum (Al), potassium (K), iron (Fe), sulfur
(S), uranium (U), molybdenum (Mo), vanadium (V), and nickel (Ni) weight concentrations obtained from core XRF
measurements; Track 16: total organic carbon weight concentration (LECO™ TOC measurements)
9
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Table 2: Distribution of core XRF measurements of elemental weight concentrations separated by geochemical
rock class in well no. 1.
Geochemical RC S (wt%) U (ppm) V (ppm) Ni (ppm) TOC (wt%)
LEF1 1.4 ± 0.2 7.0 ± 1.3 404 ± 102 64 ± 24 4.3 ± 0.6
LEF2 2.3 ± 0.6 7.5 ± 1.3 391 ± 112 126 ± 65 5.0 ± 0.8
LEF3 2.5 ± 0.7 6.2 ± 1.7 176 ± 61 81 ± 37 4.4 ± 0.9
LEF4 2.2 ± 0.5 6.9 ± 1.5 207 ± 97 70 ± 33 3.0 ± 0.5
LEF5 2.6 ± 0.6 12 ± 2.0 192 ± 62 75 ± 34 4.1 ± 0.4

The key petrophysically relevant observations from the


aforementioned chemostratigraphic analysis are
summarized as follows.

While both the top and the bottom of Lower Eagle Ford
contain clays, the clay type distribution is not the same.
Kaolinite is significant in the lowermost unit, LEF5, as
presented by the very low Si/Al ratio (<2), and low
potassium (0.2- 0.3 wt%). These characteristics indicate
that illite is diluted by the presence of a clay that is not
potassic, with a lower silicon content.
(a)
The best reservoir rock classes, LEF1 and LEF2 appear
to be associated with intervals of extreme enrichment of
specific trace metals including V and Ni. These two
elements, along with U and Mo, are considered to be
useful indicators for productivity and preservation of
organic matter, as well as for different types of anoxia
(Tribovillard et al., 2006; Wilde et al., 2004). Uranium
and molybdenum can easily accumulate in marine
sediments at high concentrations. When the anoxia is
strong enough, it will reduce U and Mo, allowing their
precipitation into solids that can be incorporated in
sediments. On the other hand, while V and Ni are
reduced under anoxic conditions in the ocean, they are (b)
more efficiently incorporated into the sediments by
bonding with specific organic molecules under anoxia.
According to Lewan (1984), one implication is that V
and Ni concentrations, and their relative proportions can
convey information about kerogen type and presence of
trapped hydrocarbons. Fig. 7 shows a class-by-class
distribution of the Rock-Eval parameters including S1,
HI, and Tmax. The S1 and S2 values from Rock-Eval
measurements confirm that hydrocarbon content is high
when V and Ni are high.

The concentrations of trace elements appeared to (c)


correspond more closely with hydrocarbon content,
Fig. 7 Summary of the Rock-Eval pyrolysis
compared to TOC. While the TOC varies between
measurements (a) S1, (b) HI, and (c) Tmax. The results are
geological rock classes, it does not directly correlate with
presented separately in the five identified geochemical
the amount of in-situ hydrocarbons.
classes.
10
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Lithofacies classification. Through the analysis of thin- scale inter-lamination of mudstones with limestone
sections and core images obtained from 145 feet of laminae, composed of foraminifera. Frequently, the
conventional, we identified four lithofacies in the lower proportions of mudstone to foraminifera-rich limestone
Eagle Ford interval. This classification takes into change in these facies. Reported mineralogy indicates
account the texture, grain-to-matrix ratio, fossil types, carbonate content as low as 45%, and total clay content
and sedimentary structures observed in the core. Track up to 30% in weight, consisting of illite and kaolinite.
12 of Fig. 2 show the lithofacies classification in well no.
1. 3. Bedded foraminiferal mudstone/nodular limestone:
Dark-gray to black illite-rich mudstones with faint
Moving upward from the base of the Eagle Ford, we laminations, where the limestone beds are nodular, or
observed a transition from a massive and weakly have a thickening-thinning (carbonates nodules
laminated texture, to a repetitive alternating pair of described by Dawson, 2000). Often internal structures of
lithofacies that dominates the majority of the Lower the nodular limestones are obliterated by pore-filling
Eagle Ford section. The pair consists of two main sub- cementation and recystrallization. The transitions
facies: foraminiferal mudstone and packstones. These between mudstones and limestones is sharp in these
facies alternate at different thickness scales from sub- facies. Moreover, since the bounding surfaces of the
millimeter up to decimeter. Furthermore, mixtures of nodular laminations are often curving, it is possible in
these facies are common depending on the scale of that in some cases, the nodular limestones is laterally
observation. Typically, this pair of facies has minimal or discontinuous.
no bioturbation and transitions between facies is sharp.
The foraminiferal mudstones consists of bioclastic grains 4. Bedded foraminiferal wackestone/limestone: The
(10-100 microns), carbonate mud and clays. The wackestone is dark grey to black due to high organic
bioclastic grains are mostly foraminiferal tests but can content ranging from 4-6 wt%. Wackestones contain thin
include other skeletal remains of ostracodes, diatoms, foraminiferal lags and isolated bivalves. The limestone
and bivalves. Perhaps with the exception of bivalves, the beds with thickness of few inches to a feet, are
bioclastic grains in the Lower Eagle Ford are planktonic foraminiferal packstone and grainstones transitioning at
with no benthic ones reported to date (Lowery et al., sharp boundaries with the wackestones. The abundance
2014). The packstone-grainstones are essentially of bioclastic grains in this lithotype is high in the lower
composed of planktonic foraminifera but without Eagle Ford unlike the deeper mudstone facies.
significant clays or other siliciclastics. They tend to be in
very light grey color. Diagenetic packstone-grainstones
look much like the regular packstone-grainstones but
contain many diagenetic features including
recrystallization, calcite cement, concretions, and
dolomitization. Fig. 8 shows core photographs
representing different lithofacies.

The four identified lithofacies in the Lower Eagle Ford


include massive argillaceous mudstone, laminated
argillaceous foraminiferal mudstone, bedded
foraminiferal mudstone/nodular limestone, bedded
foraminiferal wackestone/limestone, which are
described as follows.

1. Massive argillaceous mudstone: the mudstone facies


with little or no discernable bedding structures or internal
structures, are present in the base of Eagle Ford. Low
counts of foraminifera is identified in these mudstones,
and clay content is high compared to other lithofacies,
consisting of illite mixed layers and kaolinite.
Fig. 8 Examples of core images associated with the
2. Laminated argillaceous foraminiferal mudstone: these lithofacies in the lower Eagle Ford, from well no. 1.
facies have a medium grey color due to the millimeter-
11
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Fig. 9 through Fig. 13 show different examples of the


thin-section images associated with foraminiferal
packstone/grainstone, wackestone and mudstone facies.
These images show a variation in proportions of
bioclastic grains to mud. Foraminifers are the most
abundant fossils in the lower Eagle Ford formation, and
are composed of calcite. In some cases, they are affected
by dissolution. Due to the organic-richness of the rocks,
the color of the matrix mud is dark grey to black. Based
on Dunham’s carbonate classification, the grain-to-
matrix ratio is over 90 % for packstones and grainstones.
This ratio in wackestones is between 10-90 %, and less
than 10 % for mudstones. Fig. 11 Thin-section example of foraminiferal
wackestone facies in LEF, calcite-filled forams are
noticeable in the image

Fig. 9 Thin-section example of foraminiferal Fig. 12 Thin-section example of foraminiferal mudstone


packstone/grainstone facies, associated with a limestone facies in the lower Eagle Ford. Dissolved forams are
nodule at the top of LEF noticeable in this image

Fig. 10 Thin section example of skeletal packstone facies Fig. 13 Thin-section example of laminated foraminiferal
observed in the LEF, an inoceramid shell is evident in mudstone facies. Sub-mm foraminiferal band is evident
this image in the image

12
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Petrophysical rock classification (PRC). Inputs to this


well-log-based classification include porosity, water
saturation, TOC, total clay volume, Young’s modulus,
and Poisson’s ratio. Track 13 of Fig. 2, and track 11 of
Fig. 3, show the petrophysical classification results in
well no. 1 and well no. 2, respectively. Furthermore, Fig.
14 shows cross-plots of the estimates of petrophysical
and compositional properties for both wells. Tables 3
and 4 include class-by-class statistics of the well-log-
based classification results in well no. 1 and well no. 2,
respectively.

We ranked the petrophysical classes through an


integrated rock characterization approach. To fully
characterize the petrophysical classes, we interpreted the (a)
results of the previously described geological and
geochemical classifications, with the results of
petrophysical classification, presented in this section.
The characteristics of petrophysical rock classes are
described as follows.

PRC1 occurs interbedded throughout the rock classes


except the massive argillaceous mudstones.
Geologically, this class is a carbonate with low porosity
and occurs interbedded with other PRCs, with a very low
hydrocarbon content (i.e., less than 0.3 wt%).

PRC2 corresponds to the massive argillaceous


mudstones. This class is notable for kaolinite content up
to 15% (vol%). TOC is high and the rock contains small (b)
volumes of free hydrocarbon.

PRC3 is associated with the laminated argillaceous


foraminiferal mudstones. It appears to be a mudstone
with low porosity and low TOC (i.e., less than 3 wt%).

PRC4 is interbedded with PRC1 in the bedded


foraminiferal mudstone/nodular limestone. It is a porous
(approximately 8.5% porosity) mudstone with moderate
to high TOC content (3-5 wt%), and the lowest estimated
Young’s modulus among the other rock classes. The
water saturation is the highest in this rock class.

PRC5 represents organic-rich foraminiferal wackestone


facies. Petrophysically, it is similar to PRC4, with
(c)
slightly higher hydrocarbon content, and lower water
saturation estimates compared to PRC4. Fig. 14 Well-log based estimates of (a) TOC versus
porosity, (b) TOC versus volumetric concentration of
clay, and (c) water saturation versus porosity in well no.
1 and well no. 2. Different colors represent different
petrophysical rock class.

13
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Table 3: Well no. 1: class-by-class statistics of the well-log-based rock classification results in the LEF

Petrophysical RC Porosity (%) TOC (wt%) Vclay (%) Sw (%) YM PR


PRC1 3.5 ± 0.6 3.8 ± 0.8 12 ± 4 17 ± 7 39 ± 6 0.27 ± 0.01
PRC2 5.5 ± 0.8 2.8 ± 1.1 29 ± 4 20 ± 3 26 ± 4 0.29 ± 0.01
PRC3 3.9 ± 0.3 2.7 ± 0.5 24 ± 4 16 ± 3 34 ± 2 0.27 ± 0.01
PRC4 8.5 ± 1.1 4.8 ± 0.5 14 ± 2 28 ± 4 22 ± 2 0.33 ± 0.01
PRC5 7.0 ± 1.3 5.0 ± 0.7 12 ± 4 16 ± 3 28 ± 3 0.30 ± 0.01

Table 4: Well no. 2: class-by-class statistics of the well-log based rock classification results in the LEF

Petrophysical RC Porosity (%) TOC (wt%) Vclay (%) Sw (%) YM PR


PRC1 4.6 ± 0.8 2.1 ± 0.8 9±3 28 ± 7 36 ± 6 0.28 ± 0.01
PRC2 6.6 ± 1.1 2.7 ± 0.7 20 ± 3 28 ± 6 26 ± 4 0.30 ± 0.01
PRC3 4.8 ± 0.7 2.4 ± 0.6 17 ± 2 30 ± 3 28 ± 5 0.30 ± 0.01
PRC4 8.1 ± 0.9 3.5 ± 0.4 20 ± 3 39 ± 6 23 ± 3 0.32 ± 0.01
PRC5 7.0 ± 1.0 4.5 ± 0.7 17 ± 3 18 ± 4 25 ± 3 0.30 ± 0.01

In determination of reservoir quality rock classes, we reservoir quality rock classes in the LEF. PRC5 and
took into account factors such as storage capacity, PRC4 are organic-rich, oil-prone, and porous mudrocks
organic-richness, thermal maturity of kerogen, and free that contain substantial quantities of in-situ hydrocarbon.
hydrocarbon content in the rocks. Fig. 15 shows Rock-
Eval measurements of S1 versus TOC in the LEF, After detection of the highest reservoir quality rock
separated by petrophysical class. It is evident from the classes in the formations, it is essential to assess the
data that high TOC is not a direct indicator of high free ability to create effective fracture networks as pathways
hydrocarbon content in the rocks. For instance, at TOC for hydrocarbon transport. Although PRC4 is considered
concentrations of over 4 wt%, PRC5 and PRC4 to be a high reservoir quality class, it is not recommended
generated significantly higher free hydrocarbons, as a good candidate for selection of completion zones
comparing to PRC3 and PRC2. due to low YM and high PR estimates. Based on the
stated criteria, and integration with geochemical, and
geological classification results, we selected XX430-
XX470 ft. as the best candidate interval for further
completions in well no. 1.

Through comparison of the well-log based classification


results in well no.1 and well no.2, we observed similar
petrophysical characteristics in the reservoir quality rock
classes between the two wells. However, the estimates of
YM in PRC5 was approximately 10 % higher in well no.
1. This can be the reason for the 20 % lower hydrocarbon
production in well no. 2.

In terms of production, well no. 1 produced a cumulative


Fig. 15 Well no. 1: Rock-Eval pyrolysis measurements volume of 54,000 BOE hydrocarbons, and a water
of S1 versus TOC in the LEF. Different colors represent production of 1,000 barrels (BBL) in the first 90 days
different petrophysical rock classes after completions. While well no. 2 produced 43,000
BOE hydrocarbons, and 10,000 BBL of water at the
Through integration of petrophysical and geochemical same time interval. Fig. 16 shows the cumulative 90-day
data, we determined PRC5 and PRC4 as the best hydrocarbon, and water production for well no. 1, and
well no. 2.
14
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Geologically, these classes represent organic-rich


foraminiferal wackestone and mudstone facies.
Geochemical classification results showed significant
enrichment of trace elements such as vanadium and
nickel, and subsequently higher hydrocarbon contents in
the reservoir quality classes. In well no. 1, where
abundant core sample were available, through
combination of geologic, geochemical, and
petrophysical classification results, we selected depth
interval of XX430- XX470 ft. as the best candidate
interval for completion purposes.

We applied the petrophysical classification to a nearby


well, where core samples were not available for
Fig. 16. Cumulative 90-day hydrocarbon and water lithofacies and chemofacies characterization. Production
production in well no.1 and well no.2. Green and blue history shows a 20% lower hydrocarbon production in
bars represent hydrocarbon and water production, the first 90-days in well no. 2 compared to that of well
respectively. no. 1. The results of the well-log based rock
classification in well no. 2 showed similar petrophysical
CONCLUSIONS estimates as well no. 1. However, the estimated elastic
properties between the wells were slightly different,
In this paper we introduced an integrated well-log based which could be the reason for the difference in
rock classification method that takes into account production.
geochemical, petrophysical, and geomechanical
properties of the formation rocks. This paper documents ACKNOWLEDGMENTS
the application of this method on two wells located in the
lower Eagle Ford formation. We jointly interpreted Research reported in this paper was funded by the
conventional and geochemical well logs to estimate the Joint Industry Research Program on “Multi-Scale
petrophysical, compositional, and elastic properties to Formation Evaluation of Unconventional and
characterize the organic-rich mudrocks in the lower Carbonate Reservoirs,” jointly sponsored by
Eagle Ford. Aramco Services Company, BHP Billiton, BP,
Chevron, ConocoPhillips, and Devon Energy. We
We identified four lithofacies based on the analysis of thank the management and technicians at the W. D.
rock texture, using core and thin-section images. Von Gonten Laboratories for their support and for
Furthermore, we analyzed the elemental distribution making their facility available for the experimental
throughout the formation, and identified five work reported in this paper. Special gratitude goes
chemostratigraphic units based on the significant to Mr. Andrew Russell for his technical support
changes in the XRF-derived elemental concentrations of related to XRD measurements.
sulfur, uranium, vanadium, nickel, and organic carbon,
measured depth-by-depth on the core samples. We ACRONYMS
determined the number of petrophysical rock classes
based on the lithofacies classification results. Finally, in 2D Two-dimensional
a well-log based classification method, we employed BBL Barrel
hierarchical clustering algorithm, to classify the LEF into BCF Billions of Cubic Feet
five different petrophysical rock classes. BOE Barrel of Oil Equivalent
ECS Elemental Capture Spectroscopy
The outcome of the introduced rock classification EIA Energy Information Administration
technique resembled a stratigraphic distribution of the HI Hydrogen Index
petrophysical rock classes that is similar to the geologic LEF Lower Eagle Ford
facies classification in well no. 1. Through integration of PCA Principal Component Analysis
the results of the geological, geochemical, and PR Poisson’s Ratio
petrophysical rock classifications, we selected two PRC Petrophysical Rock Classification/Class
reservoir quality rock classes: PRC5 and PRC4. RC Rock Class
15
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

SCA Self-consistent Approximation Recent Prganic Carbon-Rich Sediments:


TOC Total Organic Carbon Implications for Cretaceous Black Shale
UEF Upper Eagle Ford Formation, Palaeogeography,
XRD X-ray Diffraction Palaeoclimatology, Palaeoecology, 232(2),
XRF X-ray Fluorescence 344-361.
YM Young’s Modulus
Dawson, W. C., 2000, Shale Microfacies: Eagle Ford
NOMENCLATURE Group (Cenomanian-Turonian) North-
Central Texas Outcrops and Subsurface
E Young’s modulus, GPa Equivalents, Gulf Coast Association of
K Bulk modulus, GPa Geological Societies Transactions, 50, 607-
Q*i Shape coefficient of an inclusion in self- 621.
consistent approximation model
S1 Free hydrocarbon content measured in rock Donovan, A.D. and Staerker, T.S., 2010. Sequence
pyrolysis test, weight fraction Stratigraphy of the Eagle Ford (Boquillas)
Sw Water saturation, fraction Formation in the Subsurface of South Texas
N Number of components in self-consistent and Outcrops of West Texas, Gulf Coast
approximation model Association of Geological Societies
P*i Shape coefficient of an inclusion in self- Transactions, 60, 861-899.
consistent approximation model
Tmax Temperature at S2 peak in pyrolysis, °C Dunham, R. J., 1962, Classification of Carbonate
Rocks According to Depositional Textures,
Vclay Total clay volume, fraction
Classification of Carbonate Rocks--A
xi Volumetric concentration of component i in Symposium, 108-121.
self-consistent approximation model, fraction
Geometric grain shape factor Fairbanks, M. D., 2012, High Resolution Stratigraphy
Shear modulus, GPa and Facies Architecture of the Upper
v Poisson’s ratio, fraction Cretaceous (Cenomanian-Turonian) Eagle
Ford Group, Central Texas, University of
REFERENCES Texas, Austin, Texas.

Archie, G. E., 1952, Classification of Carbonate Faust, M.J., 1990, Seismic Stratigraphy of the Mid-
Reservoir Rocks and Petrophysical Cretaceous Unconformity (MCU) in the
Considerations, AAPG Bulletin, 36(2), 278- Central Gulf of Mexico
298. Basin, Geophysics, 55(7), 868-884.

Aderibigbe, A., Chen Valdes, C., and Heidari, Z., Folk, R. L., 1962, Spectral Subdivision of Limestone
2016, Integrated Rock Classification in the Types. Classification of Carbonate Rocks--A
Wolfcamp Shale based on Reservoir Quality Symposium, 62-84.
and Anisotropic Stress Profile Estimated
from Well Logs, Interpretation, 4(2), SF1- Harbor, R. L., 2011, Facies Characterization and
SF18. Stratigraphic Architecture of Organic-Rich
Mudrocks, Upper Cretaceous Eagle Ford
Bardon, C. and Pied, B., 1969, Formation Water Formation, South Texas, University of
Saturation in Shaly Sands, SPWLA 10th Texas, Austin.
Annual Logging Symposium, Houston,
Texas, USA, 25-28 May. Hentz, T. F., and Ruppel, S. C., 2010, Regional
Lithostratigraphy of the Eagle Ford Shale:
Berryman, J.G., 1995, Mixture Theories for Rock Maverick Basin to East Texas Basin, Gulf
Properties. Rock Physics & Phase Relations: A Coast Association of Geological Societies
Handbook of Physical Constants, 205-228. Transactions, 60, 325-337

Brumsack, H. J., 2006, The Trace Metal Content of Kale, S., Rai, C. and Sondergeld, C., 2010, Rock
16
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

typing in gas shales, SPE Annual Technical Integrating Core Data and Wireline
Conference and Exhibition, Florence, Italy, Geochemical Data for Formation Evaluation
19-22 September and Characterization of Shale-Gas
Reservoirs, SPE Annual Technical
Lewan, M.D., 1984, Factors Controlling the Conference and Exhibition, Florence, Italy,
Proportionality of Vanadium to Nickel in 19-22 September.
Crude Oils, Geochimica et Cosmochimica
Acta, 48(11), 2231-2238. Saneifar, M., Aranibar, A., and Heidari, Z., 2013,
Rock Classification in the Haynesville Shale-
Lock, B. E., Peschier, L., and Whitcomb, N., 2010, Gas Formation Based on Petrophysical and
The Eagle Ford (Boquillas Formation) of Val Elastic Rock Properties Estimated From Well
Verde County, Texas-A Window on the Logs. SPE Annual Technical Conference and
South Texas Play, Gulf Coast Association of Exhibition, New Orleans, Louisiana, USA,
Geological Societies Transactions, 60, 419- 30 September-2 October.
434.
Sone, H. and Zoback, M.D., 2013, Mechanical
Lowery, C.M., Corbett, M.J., Leckie, R.M., Watkins, D., Properties of Shale-gas Reservoir Rocks—
Romero, A.M. and Pramudito, A., 2014, Part 1: Static and Dynamic Elastic Properties
Foraminiferal and Nannofossil Paleoecology and Anisotropy, Geophysics, 78(5), D381-
and Paleoceanography of the Cenomanian– D392.
Turonian Eagle Ford Shale of Southern
Texas, Palaeogeography, Palaeoclimatology, Suarez-Rivera, R., Deenadayalu, C., Chertov, M.,
Palaeoecology, 413, 49-65. Hartanto, R.N., Gathogo, P. and Kunjir, R.,
2011, Improving Horizontal Completions on
Mavko, G., Mukerji, T. and Dvorkin, J., 2009, The Rock Heterogeneous Tight-Shales, Canadian
Physics Handbook: Tools for Seismic Analysis Unconventional Resources Conference,
of Porous Media. Cambridge university press. Calgary, Alberta, Canada, 15-17 November.

Petriello, J., Marino, S., Suarez-Rivera, R., Tribovillard, N., Algeo, T. J., Lyons, T., and
Handwerger, D. A., Herring, S., Woodruff, Riboulleau, A., 2006, Trace Metals as
W., & Stevens, K., 2013, Integration of Paleoredox and Paleoproductivity Proxies:
Quantitative Rock Classification with Core- an Update, Chemical geology, 232(1), 12-32.
Based Geologic Studies: Improved Regional-
Scale Modeling and Efficient Exploration of Vanorio, T., Prasad, M., and Nur, A., 2003, Elastic
Tight Shale Plays, SPE Unconventional Properties of Dry Clay Mineral Aggregates,
Resources Conference and Exhibition-Asia Suspensions and Sandstones, Geophysical
Pacific, Brisbane, Australia, 11-13 Journal International, 155(1), 319-326.
November.
Ward Jr, J. H., 1963, Hierarchical Grouping to
Popielski, A. C., Heidari, Z., and Torres-Verdin, C., Optimize an Objective Function. Journal of
2012, Logs in Hydrocarbon-Bearing Shale, the American Statistical Association,
SPE Annual Technical Conference and 58(301), 236-244.
Exhibition, San Antonio, Texas, USA, 8-10
October. Wehner, M., Tice, M. M., Pope, M. C., Gardner, R.,
Donovan, A. D., and Staerker, T. S., 2015,
Prasad, M., Kopycinska, M., Rabe, U., and Arnold, Anoxic, Storm Dominated Inner Carbonate
W., 2002, Measurement of Young's Modulus Ramp Deposition of Lower Eagle Ford
of Clay Minerals using Atomic Force Formation, West Texas, Unconventional
Acoustic Microscopy. Geophysical Research Resources Technology Conference, San
Letters, 29(8), 13-1–13-4. Antonio, Texas, USA, 20-22 July.

Quirein, J., Witkowsky, J., Truax, J. A., Galford, J. E., Wilde, P., Lyons, T. W., and Quinby-Hunt, M. S.,
Spain, D. R., and Odumosu, T., 2010, 2004, Organic Carbon Proxies in Black
17
SPWLA 57th Annual Logging Symposium, June 25-29, 2016

Shales: Molybdenum, Chemical Geology, Research Program on “Multi-Scale Rock Physics”


206(3), 167-176. starting 2016. She received a Ph.D. (2011) in petroleum
engineering from The University of Texas at Austin.
Workman, S. J., 2013, Integrating Depositional Zoya is one of the recipients of the 2016 SPE (Society of
Facies and Sequence Stratigraphy in Petroleum Engineers) regional Formation Evaluation
Characterizing Unconventional Reservoirs: award, the 2014 TEES (Texas A&M Engineering
Eagle Ford Shale, South Texas: Master’s Experiment Station) Select Young Faculty Fellows
Thesis, Western Michigan University, award from the College of Engineering at Texas A&M
Kalamazoo, Michigan. University, the 2012 SPE Petroleum Engineering Junior
Faculty Research Initiation Award, and the 2015 SPE
Wright, A. M., and Ratcliffe, K., 2010, Application of Innovative Teaching Award. Her research interests
Inorganic Whole Rock Geochemistry to include petrophysics, borehole geophysics, rock physics,
Shale Resource Plays, Canadian inverse problems, completion petrophysics, and
Unconventional Resources and International reservoir characterization of unconventional and
Petroleum Conference, Calgary, Alberta, carbonate reservoirs.
Canada, 19-21 October.
Michael Tice is an Associate Professor in the
ABOUT THE AUTHORS Department of Geology & Geophysics at Texas A&M
University. He earned his Ph.D. in Geology from
Shahin Amin is a M.Sc. student in the Harold Vance Stanford University in 2006, and worked as a post-
Department of Petroleum Engineering at Texas A&M doctoral researcher at the California Institute of
University. He received his B.Sc. degree in Petroleum Technology from 2005-2007. Michael’s research
Engineering from Texas A&M University. During the interests focus on sedimentary geology and
summer of 2015, he worked as a Petrophysicist Intern at geochemistry, geobiology, and water-rock interactions.
Occidental Petroleum Corporation. Shahin is currently a
graduate research assistant at Texas A&M University,
under supervision of Dr. Zoya Heidari. His research
interests include petrophysics, formation evaluation and
rock classification in unconventional reservoirs.

Matthew Wehner is a Ph.D. student in the Department


of Geology and Geophysics at Texas A&M University.
He received his B.Sc. degree in Geology from Texas
A&M University. He was nominated as ConocoPhillips
Research Fellow in 2009 and 2010. Matthew is currently
working as a part-time Geologist for Von Gonten
Laboratories in College Station, Texas. His research
interests includes, X-ray fluorescence spectrometry,
quantitative stratigraphy and their application in basin
analysis and unconventional reservoir characterization.

Zoya Heidari is an assistant professor in the Department


of Petroleum and Geosystems Engineering at The
University of Texas at Austin. Before joining The
University of Texas at Austin, she was an assistant
professor at Texas A&M University in College Station
and the Chevron Corporation faculty fellow in Petroleum
Engineering from September 2011 to August
2015. Zoya has been the founder and the director of the
Texas A&M Joint Industry Research Program on “Multi-
Scale Formation Evaluation of Unconventional and
Carbonate Reservoirs” from 2012 to 2015 and the
University of Texas at Austin Industry Affiliated
18

You might also like