Sunteți pe pagina 1din 20

URTeC: 2669624

Rock Typing in Eagle Ford, Barnett, and Woodford formations


Ishank Gupta*, Chandra Rai, Carl Sondergeld, Deepak Devegowda, Mewborne School
of Geological and Petroleum Engineering, The University of Oklahoma.
Copyright 2017, Unconventional Resources Technology Conference (URTeC) DOI 10.15530/urtec-2017-2669624

This paper was prepared for presentation at the Unconventional Resources Technology Conference held in Austin, Texas, USA, 24-26 July 2017.

The URTeC Technical Program Committee accepted this presentation on the basis of information contained in an abstract submitted by the author(s). The contents of this paper
have not been reviewed by URTeC and URTeC does not warrant the accuracy, reliability, or timeliness of any information herein. All information is the responsibility of, and, is
subject to corrections by the author(s). Any person or entity that relies on any information obtained from this paper does so at their own risk. The information herein does not
necessarily reflect any position of URTeC. Any reproduction, distribution, or storage of any part of this paper without the written consent of URTeC is prohibited.

Abstract

Shales are the most commonly found sedimentary rocks on earth. Most US shale plays are massive with different
maturity regions and varying prospects. There is a paradigm shift in the understanding of shale anisotropy and
micro-structure in the last decade. The focus has now shifted on identifying the sweet spots and optimum zones for
completion. Rock Typing is one of the sought-after techniques to achieve this objective and it has become an
integrated part of the unconventional characterization workflow.

In this work, rock typing was done using an integrated workflow utilizing laboratory petrophysical measurements.
The rock types were derived using machine learning clustering algorithms namely K-means and Self Organizing
Maps (SOM). The integrated workflow was applied in three different shale plays namely Eagle Ford, Barnett, and
Woodford.

Three different rock types were identified. In general, Rock Type 1 had the highest porosity and Total Organic
Carbon (TOC) indicative of highest storage and source rock potential, respectively. Rock Type 1 was also the key
rock type controlling the production. Rock Type 2 had intermediate porosity and TOC while Rock Type 3 had the
lowest porosity and TOC.

Next, core derived rock types had to be upscaled to logs. Support Vector Machines (SVM), a classification
algorithm was used for upscaling. It was trained using a dataset consisting of depths at which both core and log data
were available. Different logs like gamma ray, resistivity, neutron and density were used for upscaling. Finally, a
Rock Type Ratio (RTR) was defined from rock type logs based on fraction of Rock Type 1 over gross thickness.
The ratio so developed was found to have a strong correlation with normalized oil equivalent production rate.

A total of 22 wells with core data were considered for rock typing in the three shale plays. The rock types were
upscaled to 95 wells at a cumulative over 20,000 ft. depth interval. The workflow shown in this paper can easily be
extended to other datasets in other plays. The manual approach on the other hand can be prohibitively time-
consuming.

Introduction

Over the years, a large number of methods have been developed for rock typing in conventional reservoirs. Most of
these methods were based on core or log derived permeability-porosity cross-plots. These methods were particularly
successful in sandstones and carbonates owing to a large dynamic range of permeability and porosity values in these
formations. Although well logs do not provide permeability estimates directly, several correlations were developed
to obtain it indirectly from other logs (Timur 1968, Coates and Dumanoir 1974, Thomeer 1983). If Nuclear
Magnetic Resonance (NMR) logs are available, Timur-Coates and SDR models can also be used to get permeability
estimates. Pittman (1992) suggested using r35 cutoffs based on mercury injection experiments to determine different
rock types. Amaefule et al. (1993) introduced Rock Quality Index (RQI) by modifying Kozeny-Carman equation.
He also proposed Flow Zone Indicator (FZI) cut-offs to define the different rock types.
URTeC 2669624 2

Walls (1982), Randolph et al. (1984), Davies et al. (1993), and Rushing et al. (2008) found commonly used rock
typing methods inadequate for tight reservoirs. They believed that other attributes such as rock texture and
composition, core-based descriptions, and clay mineralogy are also important for rock typing in tight reservoirs.

Kale et al. (2010) carried out rock typing for unconventional play namely Barnett and he included several
petrophysical parameters such as TOC, mineralogy, helium porosity and mercury injection capillary pressure
(MICP). Sondhi (2011), Gupta (2012), Aranibar et al. (2013), and Li et al. (2015) have presented similar approaches
to rock typing in unconventional reservoirs.

This study focusses on Barnett, Woodford and Eagle Ford shales. The study is unique as it uses two independent
techniques like K-Means and SOM to predict rock types. The upscaling of core based rock types to log based rock
types using the Support Vector Machine algorithm is also unique to this workflow. Finally, fidelity of the rock
typing was verified against production data acquired from several wells that also had well logs.

In this paper, as a first step, core-based rock typing was carried out using several clustering algorithms namely SOM
(Kohonen and Honkela 2007) and K-means (Macqueen 1967). In order to upscale rock types from cores to logs, a
supervised classification technique called Support Vector Machines (SVM) (Cortes and Vapnik 1995) was
employed. Finally, the upscaled rock type logs were correlated with production data to test the robustness of the
rock typing exercise.

Laboratory Core Measurements

The lab petrophysical measurements commonly used for rock typing are helium porosity, mineralogy, TOC, SRA
(S1 and S2 values) and mercury injection capillary pressure (MICP). The laboratory measurements were done for
263 samples (in Eagle Ford), 211 samples (in Barnett) and 411 samples (in Woodford).

Porosity
The helium porosity measurement procedure is outlined in Karastathis (2007). The methodology is based on use of
Boyles law to determine grain volume. Bulk volume is determined by mercury immersion. Finally, porosity is
calculated from the knowledge of grain and bulk volumes.

Mineralogy
Fourier Transform Infrared Spectroscopy (FTIR, Sondergeld and Rai 1993, Ballard 2007) technique is used to
determine mineralogy of the samples. The technique uses an in-house algorithm to invert the absorbance spectrum
for 16 different minerals namely quartz, calcite, dolomite, illite, smectite, kaolinite, chlorite, pyrite, orthoclase,
oglioclase, mixed clays, albite, anhydrite, siderite, apatite and aragonite. The composition is reported in terms of
weight percent.

Total Organic Carbon (TOC)


TOC was measured using dry pyrolysis technique (Law 1999). Three factors that determine the source rock potential
are source rock richness, source rock quality and source rock maturity. TOC is indicative of richness of the source
rock. A higher TOC may indicate higher source rock potential.

Source Rock Analysis (SRA)


SRA analysis is done by a pyrolysis flame-ionization detection (PFID) technique. At the first stage, the sample is
heated at the rate of 25 C per minute up to 300 C. During this stage, volatile hydrocarbons are releases which are
measured as a S1 peak. In the next stage, sample is heated to a higher temperature of 550 C. During this stage,
kerogen in the sample is pyrolyzed and recorded as S2 peak. The CO2 released during this process is recorded as S3
peak. The residual carbon is recorded as S4 peak. Generally, S1 peak corresponds to movable hydrocarbons and S2
represents immovable hydrocarbons. More details are available in Tissot and Welte (1984).

Mercury Injection Capillary Pressure (MICP)


Mercury capillary pressure is measured by a Micromeritics AutoPore IV machine. The pressure steps are varied
from 5 to 60,000 psia. The mercury intrusion and extrusion from the sample is measured as a function of pressure.
The incremental and cumulative intrusion and extrusion curves are recorded and interpreted to assess the
connectivity of the sample along with other data like helium porosity.
URTeC 2669624 3

Geological and Petrophysical Review

This section describes the geological and petrophysical background for different shale plays namely Eagle Ford,
Barnett, and Woodford. This information is critical as it provides constraints on the rock typing and helps
understand different rock types in terms of depositional processes.

Eagle Ford
Eagle Ford is one of the most prominent shale plays in US. It is late cretaceous in age with more than 1.5 billion
barrels oil, and 4.2 Tcf gas produced so far. It varies in thickness between 150 to 450 ft (Callantine 2010). The
depths of Eagle Ford formation vary from 7,000 ft. to 12,000 ft (CLR 2010). Eagle Ford is slightly over-pressured
with pressure gradient varying from 0.4 to 0.7 psi/ft. (CLR 2010). It has 4 different basins namely Maverick,
Hawkville, San Marcos and East Texas. Traditionally, some experts do not consider East Texas basin as a part of
Eagle Ford play due to its very high clay content compared to rest of the Eagle Ford.

Breyer et al. (2013) discusses the depositional environment for Eagle Ford formation. They suggested that the
sediment influx happened from North-east direction. East Texas basin situated in the north-eastern part of the Eagle
Ford shale play mainly had deltaic deposits. This is the primary reason for higher clay fraction in this basin. The
central, western, and southern parts of the play were dominated by marine shelf and slope deposits. The marine
deposits are very rich in carbonates. The slope deposits (in the south) are deeper compared to shelf deposits (in the
north). The slope deposits have a higher maturity compared to the shelf deposits. Thus, as one goes north, maturity
decreases and there is a transition from gas to oil (Tuttle 2010).

EOG 2010 reported an isopach thickness map for Eagle Ford formation. They reported that western and southern
parts of the play have a higher thickness (going up to 350 ft.) while northern and eastern parts of the play have lower
thickness (up to 60 ft.). The thickness increases as one goes seaward. TOC is another important factor governing the
prospectivity of a shale play. Generally, lower Eagle Ford is richer in TOC. The TOC varies from 1 to 6 % in upper
Eagle Ford and from 2 to 12 % in lower Eagle Ford (Tian 2014). The thicker parts of the Eagle Ford play are
associated with higher TOC.

Fig. 1: Wells with core and log data for rock typing in Eagle Ford. 12 wells had core data (shown as red bubbles). They spread throughout the
Eagle Ford play but were mainly limited to the condensate window. Core data were available for 263 depth points. Additional wells (shown as
black bubbles) were taken for correlation of rock types with production data. They did not have core data but had the required logs.

The wells having the core and log data and which were used for rock typing are shown in Fig. 1. Twelve wells were
available which had the core data (shown as red bubbles). They spread throughout the Eagle Ford play but were
mainly limited to the condensate window. Core data were available for 263 depth points. Out of these 12 wells,
URTeC 2669624 4

triple combo logs were available in only 3 wells while 3 more had gamma ray and resistivity. Thus, gamma ray and
resistivity logs were used for upscaling from core to log level. Additionally, seventeen wells (shown as black
bubbles) were taken which had no core data but had the required logs namely gamma ray and resistivity. The rock
type logs were created in these wells to get large sample size for correlation of production data with rock type logs.

Barnett
Barnett is one of the prominent shale gas plays in US. It is a Mississippian shale located in the Delaware and Fort
Worth basins in North Texas (Pollastro et al. 2007). It has produced more than 69 MMbbl oil, and 19.2 Tcf gas so
far. It varies in thickness between 100 to 700 ft. (Kinley 2008). The depths of Barnett formation vary between
7,000-18,000 ft. (Kinley 2008). Barnett is slightly over-pressured with average pressure gradient of 0.52 psi/ft. (Slatt
2008).

Barnett shale play lies in Fort Worth basin which was formed during Paleozoic. Barnett was deposited in a back-
arch setting (Bruner and Smosna 2011). Barnett play consists of upper Barnett, lower Barnett and Forestburg
limestone. Lower Barnett lies directly over a regional angular unconformity. The Forestburg limestone separates the
upper and lower Barnett shale members. It is quite thick in the north (~200 ft.) and thins as one goes south (few
feet). Towards the south, it becomes difficult to distinguish upper Barnett from lower Barnett. Bruner and Smosna
2011 also reported an isopach map for Barnett. The north-eastern and eastern portions of the play have the highest
thickness.

Sarmiento et al. 2013 discussed TOC and vitrinite reflectance data for Barnett. They showed that majority of the
play (including the thick eastern and north-eastern parts) have low TOC. Southern part of the play has the highest
TOC. But the maturity data (vitrinite reflectance, % Ro) showed that majority of the high TOC region lies in
immature window. The thick eastern and north-eastern parts of the basin having lower TOC lie in the gas maturity
window. This explains the high gas potential of the Barnett play. Singh (2008) identified 10 different lithofacies in
Barnett. They were namely siliceous non-calcareous mudstone, siliceous calcareous mudstone low calcite, siliceous
calcareous mudstone high calcite, silty-shale, phosphatic deposits, limy mudstone, dolomitic mudstone, calcareous
laminae, concretions and fossiliferous deposits. Out of these, lithofacies 1, 2, 3 and 6 are dominant and are
responsible for majority of the petrophysical variation. Majority of the cored interval also consists of these four
lithofacies. Lithofacies 1 and 2 are associated with high TOC and high porosity. On the other extreme, Lithofacies 6
is very tight with little porosity and very low TOC.

a) b)

Fig. 2: (a) Wells with core and log data for rock typing in Barnett. 3 wells (shown as red bubbles) were available. Core data were available for
211 depth points. Additional 44 wells (shown as black bubbles) were taken for correlation of rock types with production data. (b) Wells with core
and log data for rock typing in Woodford. 7 wells (shown as red bubbles) had core data. Core data were available for 411 depth points. Additional
12 wells (shown as black bubbles) had triple combo logs but no core data. Rock type logs were populated in these 12 wells for correlation with
production data.
URTeC 2669624 5

The wells having the core and log data and which were used for rock typing are shown in Fig. 2 (a). Three wells
were available which had core data. They lie in over mature, gas rich, stratigraphically thick part of the Barnett shale
play (shown as red bubbles). Core data were available for 211 depth points. Out of these 3 wells, only gamma ray
and resistivity logs were available for 2 wells. Thus, gamma ray and resistivity logs were used for upscaling from
core to log level. Additionally, forty-four wells (shown as black bubbles) were taken which had no core data but had
the required logs namely gamma ray and resistivity. The rock type logs were created in these wells for correlation
with production data.

Woodford
The Woodford shale located in the Anadarko, Arkoma and Ardmore basins in Oklahoma and Texas, is a Devonian-
Mississippian shale. The shale has produced more than 87 MMbbl oil and 4.6 Tcf gas so far. The thickness of the
shale varies between 150 to 400 ft. (CLR 2010). The depths of Woodford formation vary between 4,800 to 10,000
ft. (CLR 2010). Woodford formation is over-pressured with pressure gradient varying between 0.60-0.65 psi/ft.
(CLR 2010).

Woodford shale was deposited in Devonian age, around 360 million years ago (CLR 2010). It was deposited in the
marine environment as an organic rich shale. The low oxygen environment facilitated preservation of oil prone
organic matter. During the early Pennsylvanian, plate collision resulted in the formation of Anadarko, Ardmore and
Arkoma basins, followed by rapid subsidence and sedimentation, during the late Pennsylvanian. It was during this
period that the majority of the overlying sandstone reservoirs were deposited. By the early Permian period, oil
generation and migration into overlying conventional reservoirs had started.

The Anadarko basin has the thickest Woodford formation among the three basins (Caldwell and Johnson 2013).
Comer (2005) reported TOC and Vitrinite reflectance data for the Woodford shale. The data show that the Anadarko
basin has the highest TOC and it lies in the oil maturity window. Thus, Anadarko basin is likely to have the best oil
production in Woodford shale play. South Central Oklahoma Oil Province (SCOOP) and STACK (Sooner Trend
Anadarko Basin Canadian and Kingfisher Counties), the most prolific Woodford shale regions lie in the Anadarko
basin.

The common lithologies found in Woodford formation are black shale, chert, sandstone, siltstone and dolostone. The
most productive lithologies are siliceous and include cherts and cherty black shales. Siliceous formations in
Woodford are highly brittle and natural fractures are also seen. Chert and Quartz in Woodford have different sources
and distributions. Quartz is detrital in origin while chert is biogenic and represents siliceous micro-organisms
namely Radiolaria. Chert deposits are organic rich and where they are thermally mature, they form optimum
exploration targets.

The wells with both core and log data and which were used for rock typing are shown in Fig. 2 (b). Seven wells
were available which had the core data. The majority of these well were in Anadarko basin (shown as red bubbles).
Core data were available for 411 depth points. Triple combo logs were available in all 7 wells. Thus, gamma ray,
resistivity, neutron and density logs were used for upscaling from core to log level. Additionally, twelve wells
(shown as black bubbles) were taken which had no core data but had triple combo logs. The rock type logs were
created in these wells for correlation with production data.

Rock Typing Methodology

The workflow used for rock typing is given in Fig. 3. The key petrophysical parameters used for rock typing are
porosity, TOC, and mineralogical compositions. These data were selected as these represent the most commonly
done measurements in the lab. These measurements have high accuracy and lower associated errors. Finally, they
explain the maximum variance in the data and are sufficient to distinguish different rock types.

Generating Clusters using PCA, K-means and SOM


Rock typing involves grouping rock samples that share certain characteristics. While this may be easily done if the
inputs comprise two or three measurements, it can be challenging to accomplish when the input data have high
dimensionality. Principal component analysis (PCA) (Hotelling 1933) can be used to reduce data dimensionality.
The idea behind PCA is to identify directions in multidimensional space that contain most of the variations observed
in the data. Principal components are linear combination of different input parameters (TOC, porosity, mineralogy in
URTeC 2669624 6

this study). Mathematically, the principal components are the eigenvector of the covariance matrix of the original
attributes. In general, the first few principal components explain most of the variance in the data. Fig. 4 a) shows a
plot of principal components versus variance in the Eagle Ford data. In this dataset, the first three principal
components explain 90 % of the variance in the data. Thus, instead of using 5 variables (like TOC, porosity, quartz,
clays, and carbonates) for clustering, it is sufficient if the first three principal components are used.

Core Data Porosity, Mineralogy, Total


Organic Carbon

Clustering Algorithms K-Means, Self


Organizing Maps

Test Clusters Source Rock Analysis,


Nano-indentation data

Upscale rock types Support Vector


Machines

Correlating Clusters to Production data

Fig. 3: This figure illustrates the workflow for rock typing. In Step 1, petrophysical data from the lab is collected. In Step 2, rock types are
defined by clustering. Next, the clusters are quality checked. Then, rock types are upscaled to logs. Finally, rock type logs are analyzed with
production data to ascertain the key rock type controlling the production.

Clustering algorithms such as K-Means and Self Organizing Maps (SOM, Chon and Park, 2008) are then used with
selected principal components to define rock types. Prior to clustering, determination of optimum number of clusters
is done. The optimum number of clusters was chosen to be three in this study because of the analysis presented in
Fig. 4 b which plots intra-cluster variance (red curve) and inter-cluster variance (green curve) as a function of the
number of clusters. Intra-cluster variance (Sum of Squares Within, SSW) refers to the average variance within a
cluster. Inter-cluster variance (Sum of Squares between, SSB) refers to the distinctions between the clusters. The
point at which they start flattening out define the optimum number of clusters. Increasing the number of clusters
beyond this point does not significantly improve the definition of rock types.

Fig. 4: K-means clustering schematic. a) Principal Component Analysis. It is done to reduce the dimensionality of the clustering problem. The
key Principal components are taken for rock typing which explain the maximum variance in the data b) Determination of optimum number of
clusters requires analysis of the intra-cluster variance (red curve) and inter-cluster variance (green curve). Intra-cluster variance refers to the
average variance within a cluster. Inter-cluster variance refers to the distinctions between the clusters. The point at which the curves start
flattening out defines the optimum number of clusters (in this case rock types). c) 3 Clusters obtained using K-means plotted in 3D. Each axis is
one Principal Component.
URTeC 2669624 7

K-means clustering to identify rock types is a multi-step process. In step 1, the number of clusters are determined.
This number can be chosen by the user based on experience or decided using the SSW-SSB method explained
earlier. In step 2, the cluster centroids are randomly assigned. In step 3, points closest to a centroid are assigned to
that group. This process is basically equivalent to reducing the sum-of-squares-within (SSW) variance. In step 4,
centroids are reassigned as mean of all the data in a particular group. This process is repeated until convergence is
achieved. Fig. 4a shows the original data while Fig. 4c shows the results of the clustering procedure for 3 clusters.

SOM is another form of clustering technique which is similar to neural net analysis (Chon and Park, 2008). In SOM,
it is not required to predefine the number of clusters. A 2D training map is created from the input data. The training
map has nodes and each node has the same number of vectors as the input data. In this particular case, each data
point or training map node has three vectors representing the three principal vectors. Next, a node is randomly
selected from training map and compared with input data grid. The data point on the grid which is closest to the
training node is defined Best Matching Unit (BMU). The neighborhood of the BMU is multiplied by a weight factor
to transform that area into more like the training node. This process is repeated a large number of times until
convergence is achieved. The convergence is achieved once the clusters develop on the data grid and do not change
with subsequent iterations.

The three clusters identified from SOM are shown in Fig. 5 a. The characteristics of the three clusters are shown in
the rose diagram (or pie diagram) in Fig. 5 b. Fig. 5 shows the results from Eagle Ford. From SOM clustering, Rock
Type 1 (RT1) has the highest porosity and highest TOC. It also shows that Rock Type 3 (RT3) is associated with
high clays, high porosity but low TOC. On comparing, both SOM and K-Means clustering provided very similar
results. Table 1 shows the characteristics of different rock types obtained from K-means and SOM clustering
techniques for all the three shale plays namely Eagle Ford, Barnett, and Woodford.

Fig. 5: a). Clusters created on a SOM map b) Rose diagram (or pie diagram) from SOM. Rose diagram shows the characteristics of different rock
types/clusters. On comparing, the three clusters obtained from SOM are very similar to those obtained from K-means.

Upscaling Rock Types using classification algorithms


Rock types were upscaled from petrophysical measurements on plugs to logs using Support Vector Machines (SVM,
Steinwart and Christmann 2008). SVM technique identifies boundaries between different clusters to separate them.
These boundaries are called hyperplanes. For the rock typing application, the SVM model was trained using a test
dataset. Different log suites were used for different plays depending on the data availability (gamma Ray and
resistivity in Eagle Ford, Barnett, and triple combo in Woodford). A test dataset was prepared for each play which
included depth points where both log and core data were available. A large part of the test dataset was used to train
the model and then a prediction was made on a small portion of the test dataset. The efficiency or accuracy of the
classifier was gauged by the predictions of the rock type from the SVM classifier against those obtained by using
core data in the test dataset. Radial SVM model gave the best results and therefore, it was used for prediction. The
trained model was then used to predict rock types for the remaining depths and well logs for which core data are
unavailable. The distribution of rock types along the wellbore were then correlated with the production data to
identify the key rock type controlling the production.
URTeC 2669624 8

Table 1: Characteristics of different rock types (data represents 25-75 percentile)


Eagle Ford TOC Porosity Carbonates Clays Quartz S1 (mg/gm
(wt %) (vol %) (wt %) (wt %) (wt %) rock)
Rock Type 1 3.5-5.5 7.0-9.0 55-70 10-30 0.0-6.0 2.0-6.0
Rock Type 2 1.5-3.0 4.5-6.5 70-85 5-15 0.0-4.0 1.5-3.5
Rock Type 3 1.5-3.0 7.0-9.0 15-35 45-70 0.0-4.0 1.0-2.0
Woodford TOC Porosity Carbonates Clays Quartz S1 (mg/gm
(wt %) (vol %) (wt %) (wt %) (wt %) rock)
Rock Type 1 4.0-9.0 5.5-9.0 10-20 18-30 35-50 1.0-5.0
Rock Type 2 3.0-5.0 5.5-7.0 5-15 42-55 15-30 0.0-4.0
Rock Type 3 0.5-4.5 2.0-4.0 65-90 0-10 0-10 0.0-2.0
Barnett TOC Porosity Carbonates Clays Quartz S1 (mg/gm
(wt %) (vol %) (wt %) (wt %) (wt %) rock)
Rock Type 1 3.5-5.5 6.0-8.0 5-15 32-40 28-36 0.7-1.1
Rock Type 2 2.5-4.5 6.0-8.0 7-18 42-52 12-20 0.3-0.8
Rock Type 3 1.5-3.5 3.5-6.0 30-55 15-30 11-20 0.3-0.7

Results and Discussion

Core-Derived Rock Typing


The core based rock types derived using clustering techniques were quality checked using apriori petrophysical
knowledge. Higher values of TOC, S1 and porosity indicate higher storage and source rock potential.

Brittleness is another key petrophysical property governing fracture propagation and well productivity in
unconventional reservoirs. Mineralogy data from FTIR can be used as a proxy for brittleness. Quartz mineral is very
brittle while clay minerals are ductile. Equation 1 or equation 2, given by Jarvie et al. (2007) and Wang and Gale
(2009), respectively, can be used to calculate brittleness index from mineralogy.

. (2007) = .......(1)
++
+
(2009) = .(2)
+++
Eagle Ford
The parameters governing storage and source potential for different rock types in Eagle Ford are shown in Fig. 6.
Clearly, Rock Type 1 has the highest porosity, TOC and S1 values. Thus, Rock Type 1 has the highest storage and
source potential. The average mineral content for the different rock types in Eagle Ford are shown in Fig. 7. All rock
types have little quartz and thus cannot be differentiated based on quartz content for brittleness. However, Rock
Type 3 has the highest clay percentage. Wells 10, 11 and 12 are rich in Rock Type 3. These wells lie in detrital
deltaic deposits. This explains why these wells are different and clay rich compared to other wells in carbonate rich
marine shore/shelf deposits. The Rock Type 3 has high porosity but very little TOC/S1. Thus, Rock Type 3 is
ductile and has poor source rock potential.

Fig. 6: Parameters governing storage and source potential in Eagle Ford. Clearly, Rock Type 1 has the highest porosity, TOC and S1 values.
Porosity is a direct indicator of storage potential. High amount of TOC and S1 values generally suggest higher source rock potential. Thus, Rock
Type 1 has the highest storage and source potential.
URTeC 2669624 9

Fig. 7: Average mineral content for different rock types in Eagle Ford. Rock Type 3 has high clay content. Wells rich in Rock Type 3 lie in East
Texas basin. This rock type has high porosity but poor source rock potential.

Mercury injection capillary pressure data identified three distinct capillary pressure curves for three different rock
types. The incremental and cumulative Hg intrusion plots were normalized by helium pore volume. The plots for the
three rock types are shown in Fig. 8. Normalization of the mercury pore volume with helium pore volume results in
normalized intrusion scale between 0 to 1 and helps to determine connectivity of the sample.

Fig. 8: Representative normalized incremental and cumulative mercury intrusion plots for the three rock types in EF. It shows that the connected
pore volume decreases as we go from Rock Type 1 to Rock Type 3 and Rock Type 2. Rock types 1 and 3 had similar range of helium porosity
but Rock Type 3 has lower connectivity then Rock Type 1. This is likely due to its clay rich nature where clays deposited at pore lining and pore
throats reduce the connectivity.

The mercury to helium pore volume ratio for Rock Type 1, varies between 0.6 to 0.75 for different samples. The
ratio varies between 0.5 to 0.65 for Rock Type 3 samples and between 0.3 to 0.4 for Rock Type 2 samples. Thus, the
connectivity decreases as we go from Rock Type 1 to Rock Type 3 and Rock Type 2. It is interesting to note that
both rock types 1 and 3 had similar range of helium porosity but Rock Type 3 has lower connectivity then Rock
URTeC 2669624 10

Type 1. This is likely due to dispersed clay type which deposits at pore lining and pore throats and thereby, degrades
connectivity.

The shape of the capillary pressure curve also clearly distinguishes the three rock types. In Rock Type 2 samples, the
normalized incremental intrusion curve increases monotonously without reaching a plateau or an inflection point
even at 60,000 psia. At 60,000 psia, equivalent pore size which the mercury could pass through is 3 nm. This shape
is characteristic of very tight rocks where the dominant pore size may be smaller than 3 nm.

In Rock Type 1 and Rock Type 3 samples, the capillary pressure curve exhibits a distinct maximum before 60,000
psia. The inflection point refers to the dominant pore throat. A larger value of the dominant pore throat means higher
permeability. For Rock Type 1 samples, the average dominant pore throat size was 13 nm while for Rock Type 3
samples, it was 5 nm. Thus, Rock Type 1 samples had the highest permeability. As discussed before, Rock Type 3
samples may be effected by presence of clay.

Barnett
The parameters governing storage and source potential for different rock types in Barnett are shown in Fig. 9. Rock
Type 1 has the highest storage and source potential. The average mineral content for different rock types in the
Barnett are shown in Fig. 10. Rock Type 1 has high quartz content and is likely the most brittle of the three rock
types. Coupled with high storage and source rock potential, this rock type is expected to have the largest impact on
production. Rock Type 2 is the most ductile of the three rock types due to highest clay content.

Fig. 9: Parameters governing storage and source potential in Barnett. Clearly, Rock Type 1 has the highest storage and source rock potential.

Fig. 10: Average mineral content for different rock types in Barnett. Rock Type 1 has high quartz content. It is more brittle compared to the other
two rock types. Rock Type 2 has the highest clay percentage and is the most ductile of the three rock types.

MICP data is shown in Fig. 11. In Rock Type 1 samples, the cumulative intrusion plot show that the ratio of
mercury to helium volume varies between 0.65 to 0.80. In Rock Type 2 samples, this ratio varies between 0.40 to
0.60 and it varies between 0.30 to 0.45 for Rock Type 3 samples. Thus, it shows the connected pore volume
decreases as we go from Rock Type 1 to Rock Type 2 and Rock Type 3. It is interesting to note that both rock types
1 and 2 had similar range of helium porosity but Rock Type 2 has much lower connectivity then Rock Type 1.
Again, this is likely due to dispersed clay type which deposits at pore lining and pore throats, and thereby degrades
connectivity.

The shape of the capillary pressure curve in Rock Type 3 samples, is characteristic of very tight rocks where the
dominant pore size may be smaller than 3 nm. Considerable hysteresis between saturating and desaturating curves is
evident in both Rock Type 1 and Rock Type 2, implying there is real Hg intrusion into the samples. However, in
URTeC 2669624 11

Rock Type 3, cumulative intrusion curves show that saturating and desaturating curves almost overlie. Lack of
hysteresis is a sign of false intrusion/blank effect due to sample compression at high pressures. Samples exhibiting
this type curve have high calcite content and very low porosity.

For Rock Type 1 samples, the average dominant pore throat size was 8 nm while for Rock Type 2 samples, it was 4
nm. Thus, Rock Type 1 samples had the highest permeability. As discussed before, Rock Type 2 samples may be
effected by presence of clay.

Fig. 11: Representative normalized incremental and cumulative mercury intrusion plots for the three rock types in Barnett. The connected pore
volume decreases as we go from Rock Type 1 to Rock Type2 and Rock Type 3. Rock types 1 and 2 had similar range of helium porosity but
Rock Type 2 has lower connectivity then Rock Type 1. This is likely due to its clay rich nature where clays deposited at pore lining and pore
throats reduce the connectivity.

Woodford
The parameters governing storage and source potential for different rock types in Woodford are shown in Fig. 12.
Again, Rock Type 1 has the highest porosity, TOC and S1 values. MICP data and average mineral content for
different rock types in the Woodford are shown in Fig. 13 and Fig. 14, respectively. Rock Type 1 has high quartz
content and is likely the most brittle of the three rock types. Again, it is expected to have the largest impact on
production.

Fig. 12: Parameters governing storage and source potential in Woodford. Again, Rock Type 1 has the highest storage and source rock potential.
URTeC 2669624 12

Fig. 13: MICP data for Woodford. The connected pore volume decreases as we go from Rock Type 1 to Rock Type 2 and Rock Type 3. Lower
connectivity in Rock Type 2 in spite of high helium porosity is likely due to high clay content.

Fig. 14: Average mineral content for different rock types in Woodford. Rock Type 1 has high quartz content. It is more brittle compared to other
two rock types. Rock Type 2 has the highest clay percentage and is the most ductile of the three rock types.

In Rock Type 1 samples, the cumulative intrusion plot show that the ratio of mercury to helium volume varies
between 0.65 to 0.80. In Rock Type 2 samples, this ratio varies between 0.50 to 0.65 and it varies between 0.40 to
0.55 for Rock Type 3 samples. Thus, the connected pore volume decreases as we go from Rock Type 1 to Rock
Type 2 and Rock Type 3.

The shape of the capillary pressure curve shows that Rock Type 3 samples are characterized by high carbonate
percentage and have a higher grain density. For Rock Type 1 samples, the average dominant pore throat size was 6
nm while for Rock Type 2 samples, it was 4 nm. Thus, Rock Type 1 samples had the highest permeability. As
discussed before, Rock Type 2 samples may be effected by presence of clay.
URTeC 2669624 13

Extending Core-Based Classification to Well Logs


Next, the core-based classification was extended to the well log data. This is necessary to determine rock type logs
for uncored wells and uncored intervals in cored wells. Rock types were upscaled using SVM.

Eagle Ford
In Eagle Ford, gamma ray and resistivity logs were available for six out of the twelve wells which had the core data.
Thus, these two logs were used for upscaling. The distribution of gamma ray and resistivity for different rock types
are shown in Fig. 15. Rock Type 1 and Rock Type 3 both show high gamma ray possibly due to high TOC and high
clays, respectively. Also, Rock Type 1 shows a high resistivity due to high oil saturation and Rock Type 3 shows a
lower resistivity due to high water saturation and high clays. Thus, core and log data are consistent with each other.
The rock type logs for two sample wells (W10 on left, W6 on right) in the Eagle Ford are shown in Fig. 16. W10 is
in East Texas basin and is thus, very rich in Rock Type 3. W6 is in San Marcos basin and is rich in Rock Type 1.
W6 is expected to have a better production rate than W10.

Fig. 15: Gamma ray and resistivity distribution for different rock types in the Eagle Ford from the depth points at which both core and log data
were available. Rock Type 1 and Rock Type 3 both show high gamma ray possibly due to high TOC and high clays, respectively. Also, Rock
Type 1 shows a high resistivity due to high oil saturation but Rock Type 3 shows a lower resistivity due to high water saturation and high clays.
Thus, core and log data are consistent with each other and logs can be used for upscaling.

Fig. 16: Rock type logs for two sample wells (W10 on left, W6 on right) in Eagle Ford. W10 is located in East Texas basin and is very rich in
Rock Type 3. W6 is located in San Marcos basin and is rich in Rock Type 1.
URTeC 2669624 14

Barnett
In Barnett, again gamma ray and resistivity were the only logs available in two out the three wells which had the
core data. Thus, these two logs were used for upscaling. Fig. 17 shows the distribution of gamma ray and resistivity
for different rock types. Rock Type 1 has low density, high gamma ray and high neutron porosity consistent with
high TOC and high porosity measured in the lab. Rock Type 3, on the other hand, has highest density and lowest
neutron porosity consistent with high carbonates in lab measured mineralogy. The rock type logs for two sample
wells (W13 on left, W14 on right) are shown in Fig. 18. Out of lower and upper Barnett, lower Barnett is richer in
Rock Type 1. Thus, lower Barnett has higher TOC and quartz compared to upper Barnett.

Fig. 17: Gamma ray and resistivity distribution for different rock types in Barnett. Rock Type 1 shows high gamma ray and resistivity due to high
TOC.

Fig. 18: Rock type logs for two sample wells (W13 on left, W14 on right) in Barnett. Out of lower and upper Barnett, lower Barnett is richer in
Rock Type 1. Thus, lower Barnett has higher TOC and quartz compared to upper Barnett.
URTeC 2669624 15

Woodford
In Woodford, triple combo logs were available for all seven wells that had core data. Thus, gamma ray, neutron,
density, and resistivity logs were used for upscaling. Fig. 19 shows the distribution of gamma ray, neutron porosity
and density logs for different rock types. The logs were consistent with the core data and were found adequate for
upscaling the rock types. Fig. 20 shows the rock type logs for two sample wells (W16 on left, W18 on right).

Relating rock types to Production Data


In the three shale plays, Rock Type 1 has the highest storage and source rock potential. In Woodford and Barnett, it
is also the most brittle rock type. Thus, Rock Type 1 is expected to be the key driver of the production. A Rock Type
1 ratio (RTR) was created by dividing the Rock Type 1 thickness with the gross thickness (i.e. RT1+RT2+RT3) for
all the wells. This was then correlated with normalized production. A positive correlation between the two was seen
in all the three shale plays which bolstered the confidence in the validity of the rock typing exercise.

Fig. 19: Gamma ray, density and neutron logs distribution for different rock types in Woodford. Rock Type 1 has low density, high gamma ray
and high neutron porosity consistent with high TOC and high lab measured porosity. Rock Type 3, on the other hand, has highest density and
lowest neutron porosity consistent with high carbonates in lab measured mineralogy.

Fig. 20: Rock type logs for two sample wells (W16 on left, W18 on right) in Woodford.
URTeC 2669624 16

Eagle Ford
The spatial location of the wells for which rock types were upscaled are shown in Fig. 1. Red bubbles represent
wells with cores and black bubbles represents additional wells which did not have the core data. All the wells were
horizontal wells and their lateral length varied from 600 to 6000 ft. The production was normalized by the lateral
lengths. The comparison of RTR with normalized production is shown in Fig. 21.
1000
BOE/per lateral ft

100
(bbl/ft)

10

1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Rock Type Ratio (RT1/RT1+RT2+RT3)

Fig. 21: Normalized production correlated with the Rock Type Ratio (RTR) in Eagle Ford. A strong positive correlation is evident suggesting
Rock Type 1 is the key rock type controlling the production.

Normalized production here refers to first 24 months cumulative barrel of oil equivalent (BOE) normalized by the
lateral lengths. A strong positive correlation was seen suggesting that Rock Type 1 is the key rock type controlling
the production. There are some outliers which are expected due different completion treatments, etc.

Barnett
The spatial location of wells for which rock types were upscaled are shown in Fig. 2 (a). All the wells were vertical
wells. Fig. 22 shows the comparison of RTR with normalized production. Normalized production here refers to first
24 months cumulative gas normalized by the zone thickness. A strong positive correlation was seen suggesting that
Rock Type 1 is the key rock type controlling the production.

There are multiple trends evident in Fig. 22. The set of wells lying along the green trend line show high productivity
while wells lying along red trend line show relatively lower productivity. The wells having high productivity are
from different counties namely Denton, Wise and Parker. The commonality among these high productivity wells is
that they were all completed by one operator. The wells lying along the red trend line were completed by other
operators. Thus, it appears that the reason for multiple trends in Fig. 22 can be attributed to different completion
practices used by various operators.

1000
Months) per ft (bbl/ft)
Cumulative Gas (24

800
600
400
200
0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Rock Type Ratio (RT1/RT1+RT2+RT3)

Fig. 22: Normalized production correlated with the Rock Type Ratio (RTR) in Barnett. A strong positive correlation is evident suggesting Rock
Type 1 is the key rock type controlling the production. The set of wells lying along the green trend line show high productivity while wells lying
along red trend line show relatively lower productivity. The commonality among these high productivity wells is that they were all completed by
one operator. Thus, it appears that the reason for multiple trends in the figure can be attributed to different completion practices used by various
operators.

Woodford
Same exercise was carried in Woodford wells. Fig. 2 (b) shows the location of the wells. Some of the wells were
horizontal, some were vertical. Fig. 23 shows the comparison of RTR with normalized production. Vertical well
URTeC 2669624 17

production was normalized by zone thickness and horizontal well production by lateral length. A positive correlation
between normalized production and RTR is evident.

100
BOE/per lateral ft (bbl/ft) 80
a)
60
40
20
0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Rock Type Ratio (RT1/RT1+RT2+RT3)

1000
b)
BOE/per ft (bbl/ft)

800

600

400

200

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Rock Type Ratio (RT1/RT1+RT2+RT3)

Fig. 23: Normalized production correlated with the Rock Type Ratio (RTR) in Woodford. a) Correlation plot for horizontal wells. b) Correlation
plot for vertical wells. A positive correlation is evident on both the plots suggesting that Rock Type 1 is the key rock type controlling the
production.

Conclusions

The current paper develops an integrated work flow for rock typing using lab measurements and well logs. The
workflow correlates the core-based rock types with available logs and generates rock type logs. These logs show a
strong correlation with the production from the well. Data mining algorithms such as K-Means, SOM, SVM, etc. are
very powerful in handling large amount of data and finding meaningful associations between different data types.
The rock types were upscaled to 95 wells at over 20,000 ft depth interval. The additional benefit is that the workflow
is largely automated making the rock typing exercise rapid. Table 1 shows the characteristics of different rock types
for different plays.

The rock type logs can be used by engineers to optimize perforations and decide on number of fracture stages. Rock
Type 3 is poor source rock and may not warrant any perforation or fracturing. Rock Type 1 is the best reservoir
rock. It not only has the highest storage and source rock potential, but is also the most brittle. Therefore, Rock Type
1 can be selectively perforated and fractured to save cost and maximize production from a well.

Some other applications of the rock typing which are routinely carried out in conventional reservoirs and will most
certainly benefit unconventional reservoirs are - 3D reservoir modeling, sweet spot identification along with seismic
data, new well locations, and improved volumetric estimates.

Nomenclature
BOE Barrel of Oil Equivalent
FTIR Fourier Transform Infrared Spectroscopy
LOM Level of Organic Maturity
MICP Mercury Injection Capillary Pressure
PCA Principal Component Analysis
URTeC 2669624 18

PFID Pyrolysis Flame-Ionization Detection


RQI Reservoir Quality Index
RTR Rock Type Ratio
SOM Self-Organizing Maps
SRA Source Rock Analysis
SSB Sum of squares between
SSW Sum of squares within
SVM Support Vector Machines
TOC Total Organic Carbon
Vp P-wave from Ultrasonic data
Vs S-wave from Ultrasonic data

Acknowledgement
We would like to thank Devon, Cimarex, Shell, Conoco Phillips, BHP, Pioneer, Apache and Drilling Info for
providing the cores and logs to carry out the study. Their generous contribution has made this study possible.

References
Amaefule, J. O., Altunbay, M., Tiab, D., Kersey, D. G., & Keelan, D. K. 1993. Enhanced Reservoir Description:
Using Core and Log Data to Identify Hydraulic (Flow) Units and Predict Permeability in Uncored Intervals/Wells.
Presented at SPE Annual Technical Conference and Exhibition, 3-6 October, Houston, Texas. SPE-26436-MS.
doi:10.2118/26436-MS.

Aranibar, A., Saneifar, M., & Heidari, Z. 2013. Petrophysical Rock Typing in Organic-Rich Source Rocks Using
Well Logs. Presented at Unconventional Resources Technology Conference, 12-14 August, Denver, Colorado. SPE-
168913. doi: 10.1190/URTEC2013-117.

Ballard, B. D. 2007. Quantitative Mineralogy of Reservoir Rocks Using Fourier Transform Infrared Spectroscopy.
Presented at SPE international Student Paper Contest, 11-14 November, Anaheim, California. SPE-113023-STU.
doi:10.2118/113023-STU.

Breyer, J.A., Denne, R., Funk, J., Kosanke, T., and Spaw, J. 2013. Stratigraphy and Sedimentary Facies of the Eagle
Ford Shale (Cretaceous) between the Maverick Basin and the San Marcos Arch, Texas, USA. Presented at AAPG
Annual Convention and Exhibition, Pittsburgh, Pennsylvania, May 19-22, 2013.

Bruner, K., and Smosna, R. 2011. A Comparative Study of the Mississippian Barnett Shale, Fort Worth Basin, and
Devonian Marcellus Shale, Appalachian Basin. DOE/NETL-2011/1478, US DOE, April, 2011.

Caldwell, C.D. and Johnson, P.G. 2013. Anadarko Woodford Shale: Improving Production by Understanding
Lithologies/Mechanical Stratigraphy and Optimizing Completion Design. Oral presentation at AAPG Education
Directorate Woodford Shale Forum, Oklahoma City, Oklahoma, April 11, 2013.

Callantine, S. 2010. Eagle Ford Shale Play. Investor Report, Chesapeake Energy, Oklahoma City, Oklahoma,
August 2010.

Chon, T.S., and Park, Y.S. 2008. Self-Organizing Map, In Encyclopedia of Ecology, edited by Sven Erik Jorgensen
and Brian D. Fath, Academic Press, Oxford, Pages 3203-3210, ISBN 9780080454054,
http://dx.doi.org/10.1016/B978-008045405-4.00907-1.

CLR. 2010. Anadarko Woodford: The SCOOP, Internal Report, Continental Resources Limited, Oklahoma City,
Oklahoma.

Coates, G.R. and Dumanoir, J.L. 1974. A New Approach to Improved Log-Derived Permeability. The Log Analyst,
(January-February 1974), pp. 17.
URTeC 2669624 19

Comer, J. B. 2005. Facies distribution and hydrocarbon production potential of Woodford Shale in the southern
Midcontinent. Presented at Unconventional Energy Resources in the Southern Midcontinent Symposium, Norman,
Oklahoma, Oklahoma Geological Survey, Circular 110, p. 51-62.

Cortes, C. and Vapnik, V. 1995. Support-vector networks. Machine Learning. 20 (3): Book Pg. 273297. doi:
10.1007/BF00994018.

Davies, D. K., Williams, B. P. J., & Vessell, R. K. 1993. Reservoir Geometry and Internal Permeability Distribution
in Fluvial, Tight, Gas Sandstones, Travis Peak Formation, Texas. SPE Reservoir Engineering.
http://dx.doi:10.2118/21850-PA.

EOG Resources. 2010. South Texas Eagle Ford. Internal Report, EOG Resources, Houston, Texas.

Gupta, N. 2012. Multi-scale characterization of the Woodford shale in West-central Oklahoma from Scanning
electron microscope to 3D seismic, PhD Thesis, Oklahoma U., Norman, Oklahoma.

Hotelling, H. 1933. Analysis of a complex of statistical variables into principal components. Journal of Educational
Psychology, 24, 417441, and 498520.

Jarvie, D. M., Hill, R., J., Ruble, T., E., and Pollastro, R., M. 2007. Unconventional shale-gas systems: The
Mississippian Barnett Shale of North-Central Texas as one model for thermogenic shale-gas assessment: AAPG
Bulletin, V. 91, P. 475499, http://dx.doi: 10.1306/12190606068.

Kale, S. 2009. Petrophysical Characterization of Barnett Shale Play, Master Thesis, Oklahoma U., Norman,
Oklahoma.

Kale, S., Rai, C., & Sondergeld, C. 2010. Rock Typing in Gas Shales. Presented at SPE Annual Technical
Conference, Florence, Italy, 19-22 September. SPE 134539. http://dx.doi:10.2118/134539-MS

Karastathis A. 2007. Petrophysical Measurements on Tight Gas Shale. Master Thesis, University of Oklahoma,
Norman, Oklahoma.

Kinley, T. 2008. Hydrocarbon potential of the Barnett Shale (Mississippian), Delaware Basin, west Texas, and
southeastern New Mexico. AAPG Vol (92): 967-991 http://dx.doi.org/10.1306/03240807121.

Kohonen, T. and Honkela, T. 2007. Kohonen network. Scholarpedia, doi:10.4249/scholarpedia.1568.


Law, C. 1999. Evaluating Source Rocks, Chapter 6, AAPG special volumes, Petroleum Geology/Handbook of
Petroleum Geology: Exploring for Oil and Gas Traps, Pages 6-1 - 6-41, Edited by Edward A. Beaumont and
Norman H. Foster.

Li, H., Hart, B., Dawson, M., & Radjef, E. 2015. Characterizing the Middle Bakken: Laboratory Measurement and
Rock Typing of the Middle Bakken Formation. Presented at Unconventional Technology Conference,20-22 July,
San Antonio, Texas. SPE-178676-MS. doi: 10.2118/178676-MS.

MacQueen, J. B. 1967. Some Methods for classification and Analysis of Multivariate Observations. Proceedings of
5th Berkeley Symposium on Mathematical Statistics and Probability. University of California Press. pp. 281297.

Pittman E.D. 1992. Relationship of Porosity and Permeability to Various Parameters Derived from Mercury
Injection Capillary Pressure Curves for Sandstones. AAPG Bulletin Vol. 76(2): 191-198, ISSN 0149-1423.

Pollastro, R.M., Jarvie, D.M., Hill, R.J., and Adams, C.W. 2007. Geologic framework of the Mississippian Barnett
Shale, Barnett-Paleozoic total petroleum system, Bend archFort Worth Basin, Texas, AAPG Bulletin 91 (4): 405-
436.

Randolph, P. L., Soeder, D. J., & Chowdiah, P. 1984. Porosity and Permeability of Tight Sands. Presented at
Unconventional Gas Recovery Symposium, 13-15 May, Pittsburg, PA. SPE-12836. doi:10.2118/12836-MS.
URTeC 2669624 20

Rushing, J. A., Newsham, K. E., & Blasingame, T. A. 2008. Rock Typing: Keys to Understanding Productivity in
Tight Gas Sands. Presented at Unconventional Reservoirs Conference, 10-12 February, Keystone, Colorado. SPE-
114164. doi:10.2118/114164-MS.

Sarmiento, M.F.R., Ducros, M., Carpentier, B., Lorant, F., Cacas, M.C., Fiornet, S.P., Wolf, S., Rohais, S. and
Moretti, I. 2013. Quantitative evaluation of TOC, organic porosity and gas retention distribution in a gas shale play
using petroleum system modeling: Application to the Barnett Shale, Elsevier. Marine and Petroleum Geology 45,
315-330 http://dx.doi.org/10.1016/j.marpetgeo.2013.04.003.

Singh P. 2008. Lithofacies and Sequence Stratigraphic Framework of the Barnett Shale, NorthEast Texas. Phd.
Dissertation, Univ. of Oklahoma, Norman, Oklahoma.

Slatt, R., Singh P., Borges, G. et al. 2008. Reservoir Characterization of Unconventional Gas Shales: Example from
the Barnet Shale, Texas, U.S.A. AAPG Annual Convention, San Antonio, Texas, 20 23 April. Search and
Discovery Article #30075.

Sondergeld C.H. and Rai C.S. 1993. A New Concept of Quantitative Core Characterization. The Leading Edge
12(7): 774-779.

Sondhi, N. 2011. Petrophysical characterization of the Eagle Ford shale, Master Thesis, Oklahoma U., Norman,
Oklahoma.

Steinwart, I. and Christmann, A. 2008. Support Vector Machines, Book, Information Science and Statistics, ISBN:
978-0-387-77241-7 (Print) 978-0-387-77242-4 (Online).

Tian, Y. 2014. Occurrence of Multiple Fluid Phases Across a Basin, in the Same Shale Gas Formation Eagle Ford
Shale Example. MS thesis, Texas A&M, College Station, Texas.

Tissot, B., P., and Welte, D., H. 1984. Petroleum Formation and Occurrence, 2 ed.: New York, Springer-Verlag, 699
p.

Thomeer J.H. 1983. Air Permeability as a Function of Three Pore Network Parameters. Journal of Petroleum
Technology, 35(4): 809-814.

Timur, A. 1968. An Investigation of Permeability, Porosity, and Residual Water Saturation Relationship for
Sandstone Reservoirs, The Log Analyst, Vol. 9, No. 4, (July-August 1968), pp. 8.

Tuttle, S. 2010. Oil Resource Plays Examples and Technology. Geological Review, Internal Report, Petrohawk
Energy Corporation, Houston, Texas (September 2010).

Walls, J.D. 1982. Tight Gas Sands Permeability, Pore Structure, and Clay, Journal of Petroleum Technology,
(1982b) 2708-2714.

Wang, F. P., and Gale, J., F., W. 2009. Screening criteria for shale-gas systems: Gulf Coast Association of
Geological Societies Transactions, 59, 779793.

S-ar putea să vă placă și