Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

SPE 124231

Effective Geochemical and Geomechanical Characterization of Shale Gas


Reservoirs from the Wellbore Environment: Caney and the Woodford Shale
D. Jacobi, Baker Hughes; J. Breig, SPE, Newfield Exploration; and B. LeCompte, SPE, M. Kopal, G. Hursan,
SPE, F. Mendez, SPE, S. Bliven, and J. Longo, Baker Hughes

Copyright 2009, Society of Petroleum Engineers

This paper was prepared for presentation at the 2009 SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 4–7 October 2009.

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

Abstract
The successful recovery of hydrocarbons from gas shales requires a fundamental understanding of the reservoir’s rock-matrix
properties. Information about the variable lithologies, mineralogies, and kerogen content is vital to locate favorable intervals
for gas production. Knowledge of the in-situ stresses and porosity of these intervals is essential for developing hydraulic
fracturing strategies to recover the gas in place. Often these properties are established from the analysis of cores extracted
from the wellbore, a time-consuming practice which causes costly delays in well completions and prolonged rig time.
We demonstrate that these reservoir rock properties can be measured and predicted in-situ from the wellbore environment
by a formation evaluation method that employs a combination of measurements made by downhole geochemical, acoustic,
and nuclear magnetic resonance sondes. Using this combination of tool measurements we determine lithology, mineralogy,
and kerogen content. The mineralogy, porosity, acoustic velocities, bulk density, pore pressure, and overburden stress are
then used to compute the unconfined compressive strength, Poisson’s ratio, and horizontal stress for each interval. These
results can then be used to develop hydraulic fracture strategies.
The effectiveness of this approach is shown through characterization of the rock properties of the Caney and the
Woodford Shale from three different wells. The ability to quantify the kerogen content from these formations is emphasized
as there is currently no other direct quantification of carbon from openhole wireline logging available. This approach for
characterization of shale gas reservoirs is also further supported through comparisons of core data that display the
mineralogy, chemistry, kerogen content, and geomechanical properties from the wellbore section.

Introduction
The Woodford and Caney formations comprise a successive series of fissile, carbonaceous, siliceous black shales that are
unconventional, economic gas plays in the Arkoma Basin of eastern Oklahoma (Amsden, 1967; Cardot, 1989, Brinkerhoff,
2007, Schad, 2007). Producing commercial gas from these fine grained lithologies requires the stimulation of a large volume
of rock using hydraulic fracture techniques. The projected azimuth, propagation, and containment of the induced fractures
created using this method are sometimes difficult to predict. Fracture growth is impeded when stimulation stages do not
successfully target siliceous lithofacies with lower fracture gradient. These can often induce extensive intersecting fractures
or contain dormant mineralized fractures that upon reactivation may increase production. Instead, some stages are
inadvertently applied to argillaceous zones that attenuate fracture development due to embedment. Other stages may be
directed toward carbonate facies that have high breakdown pressures. Treatment pressures simply are unable to exceed the
fracture gradient of the rock. Stimulations may also propagate along fault planes intersecting other formations within the
basin leaving much of the reservoir rock unfractured (Vulgamore et al., 2007).
Because of these problems, there can be uncertainty about whether there has been fracture containment within the zone of
interest or whether optimal zones that promote gas recovery have indeed been fully accessed. For example, induced fractures
into the Woodford can pose questions of whether these have been contained within the target area or have grown upward into
the overlying Caney or downward into the underlying Hunton limestone. The differences in geochemical, petrophysical and
geomechanical properties of the lithofacies found in both the Caney and Woodford can be used to improve hydraulic fracture
strategies. Using a combination of logging tool measurements, we determine the kerogen content, porosity, mineralogy, and
the principal stresses of the various lithofacies from the wellbore environment for three wells. Results will show how the
integration of these into a petrophysical model provides reservoir characterization properties comparable to those gained from
core analysis, which has the potential to save money and expedite well completions.
2 SPE 124231

Geology
The Arkoma Basin is a restricted foreland basin that extends from southeastern Oklahoma into northwestern Arkansas (Fig.
1). It formed in succession with other, similar type basins along a fault boundary created by the progressive westward
advancement of the Ouachita thrust front (Forgotson, 2007; Amsden et al., 1967) (Fig. 1). The Woodford Shale is found at
depth in the Arkoma of Oklahoma (Late Devonian – Early Mississipian), separated from the shallower Caney shale by a
thick, silty, clay-rich limestone, called the Mayes (Mississipian) (Boardman and Puckette, 2006) (Fig. 2). Directly below
these stratigraphic contacts, the Woodford overlies a series of heterolithic carbonates called the Hunton whose variable
erosional surface exerts vertical control on the thicknesses of the Woodford section in many areas of the basin (Barrick 2001;
Amsden 1960; Al-Shaieb et al., 2000). Thicker Woodford sections are encountered where channels are present or have
eroded down through the Hunton into the older underlying strata. This often leaves the Woodford unconformably resting
upon the deeper dolomitic-rich Sylvan shale or if incised deeper, the Viola lime (Ordivician) (Amsden 1960; Rottman 2000)
(Fig. 3). Between these channels, preserved plateaus and wedges of the Hunton are found, over which a general thinning of
the Woodford section occurs (Brinkerhoff 2007). Superimposed upon this trend, however is a regional thickening of both the
Woodford and Caney section from northwest to southeast, as the Arkoma’s basin floor deepens as it approaches the Choctaw
fault of the Ouachita thrust system (Amsden 1984) (Figs. 2-3). Along this trend, the thickness of the Woodford transitions
from 50 ft. to 250 ft. on average while the Caney thickens on average from 90 ft. to 220 ft. (Forgotson, 2007).

OKLAHOMA FAYETTEVILLE Shale


4 active (Miss)
NFX rigs
OUACHITA
ARBUCKLE MOUNTAINS
MOUNTAINS

ARKANSAS
TEXAS
BARNETT
Shale
(Miss.)
LOUISIANA

LLANO
UPLIFT SABINE
UPLIFT

Fig. 1: Map showing location of the Arkoma and Fort Worth Basins formed along the Ouachita Frontal Thrust Belt.

The three wells, the Cattle, Cometti, and Newberry, used in this study are good examples of the thickness trends and
variations in the stratigraphy that can occur (Fig. 2). Note the differences in the base of the Woodford for the Cattle compared
to that of the Cometti from the northwest to the southeast located on the basin map (Fig. 2). The well cross sections also (Fig.
3) show both the Caney and the Woodford thickening along that same trend. Also, the Woodford in the Newberry overlies
the Sylvan shale whereas in the other two wells the Woodford rests upon the Hunton (Fig. 3).
Siliceous lithofacies comprising the Woodford are commonly targeted to exploit hydrocarbons from the section due to
their brittleness and tendency to contain mineralized fractures (Cardot 2008). These lithofacies types are found throughout the
section interbedded with organic-rich argillaceous mudstones and dolomitic argillaceous mudstones, all of which contain
phosphatic nodules and pyrite (Buckner et al., 2008; Amsden and Rowland, 1967). Caney lithofacies consists of organic clay
mudstones, calcareous mudstones interbedded with siliceous and silty mudstone intervals that are interrupted periodically by
sections containing calcitic to ferruginous concretions, which in some reports can be up to 12 ft. in diameter (Elias 1956;
Schad 2004; Boardman and Puckette, 2006). The upper siliceous intervals within the Caney similar to that of the Woodford
are also generally targeted for recovery of gas.
SPE 124231 3

Fig. 2: Basin Map of the Arkoma in southeastern Oklahoma showing the variable depth to the base of the Woodford and the location
of the three wells used in this study. The general thickening of the Woodford and Caney from the northwest to the southeast toward
the Ouachita thrust fault system (red) can be seen in the cross sections of the Cattle and Cometti exhibited in Fig. 3.

Caney

Mayes

NFX- CATTLE 5H-13


Woodford

Sylvan
NFX- NEWBERRY 1H-12
NFX- COMETTI 2H-13

Fig. 3: Log cross-sections from three wells used in this study showing the important lithologic units across the Arkoma and their
variable thicknesses. Note the Newberry well where the Hunton is absent and the Woodford is in contact with the older Sylvan
shale.
4 SPE 124231

Reasons for the Integrated Reservoir Characterization Method


In stimulating the Woodford and Caney strata, the goal is to achieve preferential linear extension of hydraulic fractures while
at the same time achieve optimum width and vertical height growth. As the fracture propagates vertically within a section,
changes in minimum horizontal stresses can make it difficult for the fractures to extend laterally to create fairways for
recovering gas. Thus it is important in fracture design to locate potential barriers that control the vertical growth while at the
same time identify those zones that will allow for preferential lateral extension of fractures. Often defining the barriers and
the potential zones of minimum stress requires analysis of core from certain intervals. An analysis of the mineralogy is
essential to this process. It allows for differentiating between lithofacies that would be considered hard and brittle and easily
fractured versus those that are soft and ductile and difficult to fracture. The high quartz content contained in the Woodford
and upper Caney section constitutes a very brittle series of rocks which would in theory contain lower fracture gradient and
natural fractures needed for increasing gas permeability (Economides and Martin, 2007; Cemen et al., 2008). Moreover, the
overlying Mayes based on the clay and calcite composition would be very ductile and thus provide a potential overlying
containment barrier for the Woodford and possibly a bottom barrier for the Caney. Also, the Hunton limestone present in
both the Cattle and the Cometti would serve as an underlying fracture containment barrier for the Woodford, while in the
Newberry well, in the absence of the Hunton, the Sylvan shale would be a potential barrier. Likewise for the Caney in all
three wells, the Mayes would serve as the bottom containment barrier and the section above the “False Caney” which is often
referred to as the Goddard, would serve as the top barrier (Schad 2004).
The proposed barriers and brittle zones however, must be further assessed in relation to the mechanical properties and the
corresponding total organic carbon (TOC) and porosity of the rock matrix. In this case, just because potential brittleness or
ductile behavior has been proposed by the bulk mineralogy or lithology, the rock properties must also be used to establish
degrees of brittleness and ductility within the section. These properties are also typically defined by core analysis. They often
reveal that mineralogy and corresponding acoustic velocities of individual sections within the brittle and ductile bulk zones
vary considerably. Also, the corresponding TOC will vary which also is determined by core analysis. Differences in kerogen
concentrations influence the amount of GIP, increases in TOC often equate to increases in total gas content (Jarvie et al.,
2005) Variations in TOC in both formations from the wells used in this study are observed. Likewise, porosity also varies
considerably. Core analysis is used to determine this property, an important factor for the estimation of gas saturation. These
variations in geomechanical, geochemical, and petrophysical rock properties can be traced to changes in lithofacies within the
bulk sections. Thus, the final goal of any fracture design is to take into account all of these variables to assess the individual
rock integrity of these different lithofacies and establish where they occur within the stratigraphic section.

Formation Evaluation Tools used for the Integrated Reservoir Characterization Method
Complex reservoir properties in both the Woodford and Caney influence recovery and production and must be correlated to
the lithofacies in which they occur to develop effective stimulation strategies. Providing this information requires employing
a number of downhole logging tools. Their measurements allow for an integrated petrophysical characterization of the
reservoir rock. Geochemical sondes that supply formation chemistry (K, Th, U, Si, Al, Mg, C, Ca and Fe; Pemper et al.,
2006) provide the main foundation. This chemistry can be used to first identify the lithofacies that are considered favorable
for gas production, e.g., siliceous mudstones, along with their corresponding mineralogy as well as the total organic carbon
(TOC) which is an approximation of the kerogen in the rock (Jacobi et al., 2007, 2008). The same data can be used to identify
zones that serve as fracture barriers, e.g., dense carbonate sections that possess high fracture closure pressures; or zones that
attenuate fracture energy due to embedment, e.g., organic rich clay mudstones. Porosity is then measured for each, preferably
by nuclear magnetic resonance (NMR), which provides a reasonable estimate of the total porosity to determine gas
saturation. Bulk density is also needed for correlation and computing another organic carbon value using NMR porosity.
Finally, acoustic tools provide compressional velocities and shear velocities needed for deriving geomechanical parameters of
the individual lithofacies. Borehole wall images can also be used to interpret fracture density and azimuth.

Components of the Integrated Method for Reservoir Characterization


Fluctuations in the concentration of U compared to Th/U across the sections in the three wells track the rise and fall in sea
level during the Devonian and Mississipian periods (Ross and Ross, 1987; Johnson et al., 1985) that resulted not only in
changes in basin anoxia, but the deposition of the lithofacies previously discussed; rises in sea level are called “transgressive”
periods, and falls in sea level are called “regressive” periods. The following distinguishes between the two conditions:

(1) Transgressive periods mark very anoxic basin conditions were siliceous, but organic rich mudstones would have
originated from deep marine environments, with basin conditions signified by an abrupt increase in U with Th/U
values below 2.

(2) Regressive periods resulted in shallower, less anoxic water columns where less organic, but clastic, silty mudstones
and carbonate mudstones formed. These conditions reflect transitional marine environments partly influenced by
influx of sediment from continental sources, signified by decreases in U with Th/U values above 2.
SPE 124231 5

This is predicated on the redox sensitivity of U to these changes in sea level versus the Th, which is not redox sensitive
(Adams and Weaver, 1958; Jacobi et al., 2008). Using this relationship along with chemistry from the geochemical tool and
the computed mineralogy, within each section we can differentate between five important lithofacies that are important to
hydraulic fracture strategies in the Caney and the Woodford based on the following criteria:

(1) Siliceous-Organic Mudstones: based on Th/U < 2 vs high U values, high TOC, and high quartz content

(2) Low-Siliceous Organic Mudstones based on Th/U < 2 vs high U values, high TOC, low quartz content

(3) Siliceous Mudstones: based on Th/U > 2 vs. low U values, low TOC, and high quartz content

(4) Carbonate Mudstones: based on the total carbonate mineralogy computed from calcite, dolomite, and siderite.

(5) Low-Organic Mudstones: based on Th/U < 2 versus high U values, low TOC, and low quartz content

Two methods are employed to compute a TOC for each lithofacies. The first is computed using the following equation:

CTOC = Ctotal measured – CCalcite computed - CDolomite computed - CSiderite computed

The second TOC is derived by computing a total grain density using total NMR porosity with the bulk density compared
to an inorganic grain density derived from the computed matrix minerals using the following relationships:

ρ b − ρ fluid φ
ρ gr =
(1 − φ )
ρ m − ρ gr
VTOC =
ρ m − ρ TOC

ρTOC ρ m − ρ gr
mTOC = ⋅
ρ gr ρ m − ρTOC

Where
ρb is the bulk density
ρgr is the total grain density including inorganic and organic matrix constituents
φ is the NMR total porosity
ρfluid is the density of pore filling fluid, determined from NMR fluid typing
VTOC is the volume fraction of organic matrix components
mTOC is the mass fraction of organic matrix components
ρTOC is the density of organic matrix components, determined from core and/or Rockview TOC calibration
ρm is the density of inorganic matrix components, determined using mineralogy from geochemical logs

Compressional and shear velocities are used with the computed mineralogy, TOC, porosity, pore pressure, and
overburden stress in a geomechanical software code called Logging Mechanical Properties (LMP TM). Using these
parameters, the program emulates static rock properties including Poisson’s ratio, Young’s modulus, and compressive
strength at variable confining pressures (Fig. 4). Finally, using the LMP outputs the minimum horizontal stress is computed
according to the following equation:
ν
σ h, min = (σ v − αPp ) + αPp
1− ν
Where ν represents Poisson’s ratio, σv is the overburden, Pp is pore pressure, and α is Biot’s parameter. If the burial
process of sediments is assumed to follow a uniaxial strain compaction process, the equation can be used for minimum
horizontal stress (Economides et al., 1994).
Thus, siliceous lithofacies with low minimum horizontal stress will be designated as favorable for hydraulic fracturing.
Likewise, those facies which exceed about 75% of the minimum horizontal stress value in the fracture barrier layers are
designated as unfavorable. These barriers are typically tight carbonate mudstones or limestones located above or below the
shale sections with very high fracture gradients (Fig. 6). The model developed using this integration of geochemical,
petrophysical, and geomechanical data provides a selective process for planning reservoir stimulations to improve the
6 SPE 124231

recovery of gas reserves. This method was developed to characterize the Barnett Shale in the Fort Worth Basin (Jacobi et al.,
2008). It is has been modified and applied here for characterizing the Caney and Woodford section using three vertical wells
from the Arkoma of Eastern Oklahoma.

Fig. 4: Example workflow for computing static mechanical rock properties in gas shale formations.

Legend for Analyzing Lithology, Mineralogy, and Facies presented on the logs

Fig. 5: Mineralogy Legend for interpreting mineral tracks in Figs. 11-17.


SPE 124231 7

Woodford and Caney Shale Reservoir Characterization Model

Brittle - Frac Target


Siliceous Mudstones
Gas Recovery Zone

Siliceous Organic Mudstones High TOC – Gas Zone

Carbonate Mudstones Fracture Barriers

Lower Organic Mudstones


Fracture Attenuators
Less Siliceous Organic Mudstones

Use geochemical logs to locate silceous lithofacies favorable for hydraulic fracture. Use
lithofacies, mineralogy, TOC, NMR porosity, and acoustic data to compute horizontal stress.

Favorable Frac: Min-Horizontal


Siliceous Mudstones Stress

Must also locate lithofacies that are hydraulic fracture energy barriers. Use mineralogy,
TOC, porosity, and acoustic data to compute horizontal stress

Non-Favorable Frac: Max -


Carbonate Mudstones Horizontal Stress

Fig. 6: Lithofacies and Frac-Zone Legend for the Woodford and Caney integrated reservoir characterization model presented in Figs.
13-18.

Discussion
In the discussion to follow please refer to the legends in Figs. 5-6 when evaluating the data presented. The lithology and
mineralogy cross sections of the three wells exhibited in Figs. 7-8, which were computed from the chemical data from the
geochemical logging sondes, show an impressive correlation between the Lower Pennslyvanian/Mississpian through
Ordivician aged strata of the Arkoma in Southeastern Oklahoma. The Cometti well, the deepest of the three, contains over
600 ft. of combined Caney, Mayes and Woodford section. The mineralogy computed versus the core mineralogy data from
the Cometti (Fig.8) show good correlation. Also the Woodford shale is distinct in all three wells (Fig 7), and the middle and
upper siliceous zones (yellow) are easiliy identifiable across the three as well. The calcareous lime mudstones of the Mayes
and the lower Caney are also easily distinguished in the three. Also, the increase in quartz with subsequent decrease in
carbonate characteristic of the Upper Caney members at the top of the formation is observed in all three sections (Schad
2004). The variable thickness of the Hunton is also correlative in both the Cattle and Cometti, so much so that its
conspicuous disappearance in the Newberry can lead one to to conclude that the Woodford has thickened considerably and
that the limestone located at the bottom of the figure is the Hunton. However, on closer inspection of the dolomitic (blue)
clay rich (gray) mineralogy (which is also present in the Sylvan below the Hunton in both the Cometti and Cattle wells), the
Woodford in the Newberry is resting upon the Sylvan shale. The magnitude of this unconformity is even more apparent with
the realization that the limestone underlying the Sylvan at the bottom of the Newberry log is indeed the Viola limestone that
is often found underlying the Barnett of Mississipian age in some areas of the Fort Worth basin. Keep in mind that the Caney
is interpreted as being age equivalent to the Barnett (Boardman and Puckette, 2006).
8 SPE 124231

Fig. 7: Cross section showing lithologies and mineralogies (Fig. 9) of the various strata including the underlying and overlying
sections bounding the Caney and Woodford section. Both the Mayes and Hunton are considered frac barriers when stimulating the
Woodford Shale. Note the difference in the clay and carbonate content of the overlying Caney compared to the absence of calcite,
lower clay and higher quartz content of the Woodford Shale. Factoring in the topographic elevation of the Cometti and Newberry,
even though the Newberry appears deeper, the Cometti is the deepest of the three wells.
SPE 124231 9

Fig. 8: Results from XRD Mineralogy from the Cometti well for the Upper Woodford and a segment of the overlying Mayes compared
to the mineralogy computed using the geochemical logging tool data.

The data presented in the log tracks in Figs. 7-8 will also be presented in Figs. 9-13, to follow. It is important to note that the
corresponding gas saturations (GIP) presented for both the Cometti and Cattle were calculated and calibrated against core
data. In the Newberry no core data was available, however the GIP relationships established for the other two wells was
applied to calculate gas saturation. The only difference between the Cometti/Cattle versus the Newberry value is the porosity
used to calculate GIP for the former two wells was measured from NMR, while the latter in the absence of NMR data, was
established by density porosity. Adjustments for the Newberry density porosity were made to correct for the effect of TOC by
calculating a bulk grain density from the mineralogy including TOC.
The total porosity measured or calculated was then used with water saturation (Sw) values from core, along with Pickett
plots of deep resistivity, to calibrate the Archie equation in order to calculate Sw for the remaining wellbore section. The
following parameters were utilized for each well:

• Cattle: a = 1, m = 2.17, n=1.17, Rw = 0.025


• Cometti: a=1, m=2.17, n=1.17, Rw = 0.025
• Newberry: a=1, m=2.17, n=1.718, Rw = 0.025
10 SPE 124231

A review of the calculated and core-based Sw data core points exhibited in both the Cometti and Cattle sections show that a
fairly robust correlation was obtained indicating that the porosity measured from NMR was reliable for this purpose (Fig. 9).
However, this calculated Sw procedure is still experimental and needs further validation. The porosity in both the Cattle and
Cometti shows a steady increase from the Woodford to the Caney with the lowest computed Sw found in the Woodford and
the highest in the Caney section.

Fig. 9: Log plots of the Cattle and the Cometti exhibiting the water saturations and measured porosity from NMR versus core data,
the correlation of which indicates a fairly good agreement.

The core porosity values compared to the total porosity measured from NMR are also in fairly good agreement within +/-
2.00 p.u.(porosity units) (Fig.10). However, in the Caney section of the Cattle and Cometti there is disagreement between
NMR log porosity and core porosity. The core bulk densities in this region are systematically lower than the log bulk density.
Since the core total porosity is calculated in part from a bulk volume determined from the bulk density measurement, the core
porosity is also high. The NMR log total porosity is consistent with the log bulk density in these intervals, even though the
measurements are independent. It is possible that the cores were not measured as accurately as those from the Woodford or
that a greater amount of anisotropy exists in the Caney and causes an apparent mismatch. Moreover, the TOC calculated from
the use of NMR porosity compared strongly to TOC core data in both the Cattle and Cometti, further supporting the use of
NMR to determine porosity in these type plays (Figs.11-12). Direct measurement of carbon from the geochemical tool also
allows for a reasonable estimation of TOC compared to core analysis as well (Fig 12).
Gas saturation is presented along with GIP in SCF/ton and BCF/ Sec calculated from gas porosity. This serves as total
GIP in the absence of desorption core analysis which would have allowed for dividing the total GIP into free gas versus
adsorbed gas. A Langmuir isotherm used with the TOC would have allowed for deriving an adsorbed gas component,
however assuming a linear relationship between TOC and the adsorbed gas it is concluded that the adsorbed gas for each well
SPE 124231 11

Fig. 10: Cross plots showing the agreement between NMR measured porosity versus core total porosity for the Caney and
Woodford in the Cattle Well (A) and the Cometti Well (B). The error between the two measurements is approximately +/- 2.00 p.u..
The variability is related in part to the vertical resolution of the NMR measurement correlated to the smaller scale represented by
core data.

Fig. 11: Cross plots showing the agreement between the following: (A) TOC computed from Grain density-NMR-Bulk-Density versus
the TOC from core for the Caney in the Cometti and Cattle; (B) TOC computed from Grain density-NMR-Bulk-Density versus the TOC
from core for the Woodford in the Cometti and Cattle. Computations are based on TOC density of 1.27 g/cc with the understanding
that the TOC density could vary according to maturation. This factor along with the the vertical resolution of the tool compared to
the scale of the core data contributes to the variability in the correlation.
12 SPE 124231

Fig. 12: Log plot from the Cattle showing the components, NMR porosity, inorganic grain density from mineralogy compared to the
core values that were used to compute the TOC (green) in the Woodford-Caney interval using the NMR as well as the geochemical
tool. TOC (RockView TOC) (grey) from the geochemical tool is also included.
SPE 124231 13

would have been similar at a given pressure. However, if such data were provided, then adsorbed gas can be readily
calculated. Finally, the geomechanical moduli derived from traditional acoustic tools which are used to compute the
minimum horizontal stress are shown in their use in computing the confined compressional strength compared to core values
(Fig 13). The derivation of horizontal elastic moduli typically derived from anisotropic formations such as these shales is
problematic. The LMP can be used to address this problem with the ability to emulate these rock properties.

Compressive Strength Simulated and Core Data

50000
45000
40000
Simulated CS (PSI)

35000
30000
25000
20000
15000
10000
5000
0
0 10000 20000 30000 40000 50000
Triaxial Tests (PSI)

Fig. 13: Cross plot showing agreement of compressional strength simulated from LMP using the elastic moduli values versus that
from core analysis from the Cometti well. The same moduli used to compute this property are also used to compute the minimum
horizontal strength for the log well examples.

The previous crossplots and other log plots show how this integrated method can be used for reservoir characterization in
the absence of core data. It also can be beneficial for answering complex questions that arise during stimulation of the Caney
and Woodford Shale. For example, microseismic monitoring suggests that in most cases few lithofacies serve as fracture
barriers within the proximal Woodford (Vulgamore et al., 2007). Thus, the Mayes and Hunton are considered the main
containment barriers. However, observations and studies of microseismic data from the Cometti and the Newberry used in
this study as well as some reported in other studies have suggested that some induced fractures may cross into both the
Mayes/Caney and the Hunton (Vulgamore et al., 2007). Having the ability to predict whether the integrity of these proposed
barriers can contain the induced fractures or could possibly fail during stimulation can help optimize hydraulic fracture plans.
The integrated shale gas reservoir characterization model is designed to provide data to answer these questions. The results
from the model applied to the three wells under study show it can be used for this purpose.
The Woodford section of the Cometti well shows an assortment of lithofacies that are correlative to the other two wells
(Fig. 14). In general, starting at the base of the Woodford, siliceous organic mudstones are prevalent but alternate with more
siliceous mudstone and minor low siliceous mudstones and minor low organic mudstones.
14 SPE 124231

Fig. 14: Log of the Cometti showing model results showing lithofacies of siliceous organic mudstones (black bars), carbonate
mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) and low siliceous mudstones (orange
bars) (see Fig. 6 for legend). Also, see the corresponding favorable fracture zones (green bars) as opposed to potential fracture
barriers (red bars) within both the Woodford, the Hunton and the Mayes formation

Near the top of the section, the Woodford becomes increasingly populated with siliceous mudstones corresponding to a
reduction in clay content with an increase in quartz. Compared to the minimum horizontal stress calculated for each, there are
a number of siliceous mudstones considered as favorable for hydraulic fracture (Fig. 14). Within the section there are no
visible barriers according to the model, however at the top a siliceous interval is found where the model indicates there are
some siliceous mudstones that may have minimum horizontal stresses that are much greater than those below and could be
fracture attenuators. One can follow this trend as the minimum horizontal stress increases up the section. Also located above
this section, the Mayes contains carbonate mudstones that are interbedded with siliceous mudstones. The minimum
horizontal stress indicates these are fracture barriers and may provide the containment for fractures produced in the
Woodford. Likewise, below the Woodford, the Hunton also serves as a barrier. A review of the minimum horizontal stress
curve shows a substantial increase in these intervals. The accumulative thicknesses of favorable siliceous mudstones
compared to the others would suggest that targeting these zones of low minimum horizontal stress would allow for creating
the fracture fairways for recovering gas. However the minimum horizontal strength contrast between the barriers and
favorable zones is on average minimal—about 1200 psi, thus the barriers may be breached during fracturing treatment stages
of the siliceous intervals and access the Caney.
SPE 124231 15

Fig. 15: Log of the Newberry showing model results showing lithofacies of siliceous organic mudstones (black bars), carbonate
mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) and low siliceous mudstones (orange
bars) (see Fig. 6 for legend). Also, see the corresponding favorable fracture zones (green bars) as opposed to potential fracture
barriers (red bars) within both the Woodford, the Hunton and the Mayes formation. Note the Mayes carbonate mudstones serves as
a fracture containment for the Woodford, however interbedded siliceous intervals with minimum horizontal strength may undermine
the rock during the employment of high treatment pressures required to facture the Woodford

The relative minimum horizontal stresses calculated for the favorable siliceous, brittle, mudstones of the Newberry, is
much greater than that in the Cometti (Fig. 15). Treatment pressures for stimulating the rock to produce gas may need to be
increased compared to that of the Cometti. Increasing treatment pressures could challenge the barriers of the Mayes section to
contain the fractures because the interbedded siliceous intervals versus carbonate mudstones comprising the section above
show some of those intervals as brittle, favorable fracture targets. Below the Woodford, the broad steady high minimum
horizontal stress across the Sylvan shale and the Viola should serve as attenuators and barriers. The minimum horizontal
strength contrast between favorable zones versus the barriers is far more significant—on average closer to 2500 psi than that
found in the Cometti well and even in the Cattle (Fig. 16). At the top of the Woodford near the contact of the Mayes the
minimum horizontal strength reaches 8000 psi. Thus, containment of fractures within the Woodford may be more probable
than in the Cometti well.
16 SPE 124231

Fig. 16:. Log of the Cattle displaying model results that show lithofacies of siliceous organic mudstones (black bars), carbonate
mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) and low siliceous mudstones (orange
bars) (see Fig. 6 for legend). Also, see the corresponding favorable fracture zones (green bars) as opposed to potential fracture
barriers (red bars) within both the Woodford, the Hunton and the Mayes formation.

In the Cattle, the Woodford contains similar interbedded lithofacies as the other two with a substantial number of
siliceous mudstone intervals determined by the model to be favorable for fracturing (Fig. 16). The underlying Hunton appears
to serve as a good containment barrier, while the interbedded intervals of the Mayes section just above the Woodford are thin,
but still greater in minimum horizontal stress than the Woodford. This stress however is comparable to that of the Newberry,
indicating possibly greater treatment pressures may need to be employed to stimulate the reservoir. Similar to the Cometti
and Newberry, the siliceous mudstones at the top of the Woodford intervals are not all favorable for fracturing, but present
themselves as fracture attenuators. Major differences between the Cattle, Cometti, and Newbury can also be seen in their
accumulative GIP values. The Newberry has 140 BCF/sec.; the Cometti has 100 BCF/sec.; and the Cattle has 43 BCF/sec. At
the time of this paper, the Woodford and Caney in the Cattle had not been put on production.
SPE 124231 17

Fig. 17: Log of the Cometti displaying model results that show lithofacies of siliceous organic mudstones (black bars), carbonate
mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) (see Fig. 6 for legend). Also, see the
corresponding favorable fracture zones (green bars) as opposed to potential fracture barriers (red bars) within both the Caney and
False Caney. Note the significant thick barriers in the lower Caney that serve as containment for fractures induced into the sparse
favorable zones of the upper Caney section.

The Caney of the Cometti appears to contain minor intervals of siliceous mudstones favorable for fracturing, compared to
the volume of low organic mudstones (Fig. 17). The lower Caney calcareous members appear to possess slightly higher
minimum horizontal stresses that represent major fracture barriers, however the section also contains minor thin carbonate
mudstones in the upper Caney that could also lead to unanticipated stress anisotropy across the section. Sudden drops in the
minmum horizontal stress observed in two areas which correlate with siliceous mudstones could be due to fractures that
could be confirmed by image logs or core analysis. However, these zones may also indicate siliceous facies that have
different rock matrices that possess lower fracture gradients compared to the other siliceous zones. The False Caney located
above the primary Caney section, another zone that can be targeted for hydrocarbons also shows minor siliceous mudstone
intervals that are favorable for fracturing.
In the Newberry, the Caney contains significant siliceous mudstones compared to that of the Cometti (Fig. 18). Again,
these too are designated as favorable fracture targets by the lithofacies model. They also are surrounded by a number of
carbonate mudstones similar to that in the Cometti. This alternating pattern of siliceous mudstone interbedded with carbonate
and organic mudstones are also the type sequences found in the Barnett Shale. In the False Caney a significant number of
siliceous mudstones are designated as favorable, and the lower calcareous members of the Caney form a containment barrier
for the reservoir.
In contrast to the two other reservoirs, the Cattle Caney section contains very few favorable zones to promote gas
recovery (Fig. 19). Instead, carbonate mudstones and lower organic mudstones dominate the section. The clay content
appears to be slightly high compared to the quartz content of the rock, thus such matrices would impede fracture
development. A review of the GIP value also indicates that gas production would not be expected to be high. Moreover, the
rock does not appear, according to the mimimum horizontal stress to be easily accessed to recover hydrocarbons. Similar to
the Cometti and Newberry, the lower calcareous members of the lower Caney serve as a fracture barrier.
18 SPE 124231

Fig. 18: Log of the Newberry showing model results that show location of lithofacies of siliceous organic mudstones (black bars),
carbonate mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) (see Fig. 6 for legend).
Also, see the corresponding favorable fracture zones (green bars) as opposed to potential fracture barriers (red bars) within both
the Caney and False Caney. Note the significant thick barriers in the lower Caney that serve as containment for fractures induced
into the many favorable zones of the upper Caney and False Caney section compared to that was found in the Cometti.

The previous discussion concerning the results from the lithofacies model needs to be tempered with this observation: the
use of the terms “fracture barrier” and “attenuator” do not indicate that these intervals will be unable to be fractured. It means
that based on the minimum horizontal stress it may require greater treatment pressures to initiate the fractures. The model is
intended to be used as guide for developing fracture strategies and in some instances may become the fracture strategy.
Remember, the answers originate from an integration of geochemical, geomechanical, and petrophysical inputs measured
from the wellbore that provide a solution for characterizing the reservoir. In the case of whether fractures will propagate into
the Caney via the stimulation of the Woodford, the data suggest that it could be possible in one well, but not necessarily in
the others. The overall stress contrasts between lithofacies and among lithofacies recorded from well to well will be the key
to answering these questions. The lithofacies model has the potential for finding solutions for completing vertical wells and
providing information on the placement of horizontal wellbores using the integration of the data measured from a wellbore
environment.
SPE 124231 19

Fig. 19: Log of the Cattle showing model results that show lithofacies of siliceous organic mudstones (black bars), carbonate
mudstones (blue bars) siliceous mudstones (yellow bars), low organic mudstones (grey bars) (see Fig. 6 for legend). Also, see the
corresponding favorable fracture zones (green bars) as opposed to potential fracture barriers (red bars) within both the Caney and
False Caney. Note the significant thick barriers in the lower Caney that serve as containment for fractures induced into the very
sparse favorable zones of the upper Caney section and the False Caney where no favorable zones are found. These zones are in
sharp contrast to the more favorable sections found in both the Cometti and the Newberry.

Conclusion
We have demonstrated how the integration of logging measurements produces a model that easily determines the lithofacies,
mineralogy, geomechanical, and petrophysical attributes of gas shales in-situ from the wellbore environment. The model
provides a more selective means to stimulate reservoirs for recovery of gas that has the potential of aiding in the design of
fracture plans. It also has the potential of replacing core data to make those decisions when it is unavailable, saving rig time,
expediting completions and lessening well completion cost. It also provides information on rock properties, such as porosity
and TOC, both of which can be used to determine GIP. The results suggest also that NMR is very useful for measuring total
porosity, providing an additional estimate of organic carbon. Preliminary investigations also suggest that NMR effective
porosity may be able to compute gas saturations based on the examples presented, this will be further validated in future
studies.

Acknowledgements
We are appreciative of Newfield Exploration for allowing us the opportunity to publish the results in this paper. We
acknowledge Baker Hughes for approving this publication as well. We also thank Erika Guerra for the thorough editing of
the manuscript.

References:

Adams, J.S. and Weaver, C.E., 1958, “Thorium-uranium ratios as indicators of sedimentary processes: example of concept of geochemical
facies”, AAPG Bulletin, 42 (2), p. 387-430.
Al-Shaieb, Z., Puckette, J.O., and Blubaugh, P., 2000, The Hunton Group: sequence stratigraphy, facies, dolomitization, and karstification,
Part V: in Rottman, K. et al. (eds.), Hunton Play in Oklahoma (including northeast Texas Panhandle): Oklahoma Geological Survey,
Special Publication 2000-2, p. 39-50.
20 SPE 124231

Amsden T.W., Caplan W.M., Hilpman P.L., McGlasson E.H., Rowland T.L. and Wise O.A., 1967, “Devonian of the Southern
Midcontinent Area, United States” International Symposium on the Devonian System” ed. Oswald D.H. Alta Soc. Petrol Geol., p.
913-932.
Amsden T.W. and Rowland T.L., 1967, “Silurian-Devonian relationship in Oklahoma” International Symposium on the Devonian System,
Calgary 1967 Vol 2 ed. Oswald D.H., Alta Soc Petrol Geol, p. 949-959.
Amsden, T.W., 1960, Hunton stratigraphy: OGS, Bulletin 84, p.1-310.
Amsden, T.W., 1984, “Arkoma Basin model: Middle Ordovician through Early Devonian”,Technical Proceedings of the1981 AAPG Mid-
Continent Regional Meeting, p. 116-118.
Barrick, J.E., 2001, Conodont biofacies and biostratigraphy of Silurian strata of the Hunton Group in Oklahoma, and equivalent units in
west Texas and eastern New Mexico, in Johnson, K.S. (ed.), Silurian, Devonian, and Mississippian Geology and Petroleum in the
Southern Midcontinent, 1999 Symposium: OGS, Circular 105, p. 169.
Boardman, D. and Puckette, J., 2006 “Stratigraphy and Paleontology of the Upper Mississpian Barnett Shale of Texas and the Caney Shale
of Southern Oklahoma”, South Central Section Meeting of the Geological Society of America, OGS Open-File Report 6-2006.
Brinkerhoff, A.E., 2007,” Mapping Middle Paleozoic Erosional and Karstic Patterns with 3-D Seismic Attributes and Well Data in the
Arkoma Basin of Oklahoma”., M.S. Thesis, Brigham Young University.
Buckner , N.,., Slatt, R.M., Coffey B., and Davis R.J.,2008, Stratigraphy of the Woodford Shale from Behind Outcrop Drilling, A.A.P.G.
Search and Discovery Article #50147 Annual Convention San Antonio, Texas
Cardot B.J., 1989, “Thermal Maturation of the Woodford Shale in the Anadarko Basin” , OGS 90 , p.1-15.
Cardot, B., 2008 “Overview of the Woodford Gas Shale Play of Oklahoma, U.S.A.”, A.A.P.G. Search and Discover Article No. 90078,
Annual Convention San Antonio. Texas.
Cemen, I., Ataman, O., Puckette, J. and Boardman, D., 2008, “Natural Fractures in the Woodford Shale, Arbuckle Mountains Oklahoma”
GSA Paper 131-6.
Economides, M. and Martin A.: Modern Fracturing – Enhancing Natural Gas Production, Energy Tribune Publishing Inc., Houston, TX.
(2007).
Economides, M., Hill, D., Ehlig-Economides, C., 1994, Petroleum Production Systems. Prentice Hall, Princeton, NJ.
Elias, M.K., 1956, “Upper Mississipian and Lower Pennsylvanian Formations of South Central Oklahoma,” Hicks, I.C., ed., Petroleum
geology of southern Oklahoma, Vol., 1 A.A.P.G. and Ardmore Geological Society, p. 56-124
Forgotson, J.M., 2007, “Caney Shale, Arkoma Basin Oklahoma”, A.A.P.G. Search and Discovery Article #90071, Rocky Mountain
Meeting, Snowbird Utah
Jacobi, D., Longo, J.M., Sommer, A., and Pemper, R., “A Chemistry-Based Expert System for Mineral Quantification of Sandstones,”
Paper 1.3.45, Trans., PETROTECH 7th International Oil & Gas Conference and Exhibition, New Delhi, India, 2007.
Jacobi, D., Gladkikh, M., LeCompte., Hursan G., Mendez, F., Ong, S., Bratovich, M., Patton, G., and Shoemaker P.,2008, “An Integrated
Petrophysical Evaluation of Shale Gas Reservoirs”, SPE 114925, p. 1-23
Jarvie, D.M., Hill, and R.J., Pollastro, 2005, “Assessment of the gas potential and yields from shales: the Barnett Shale model”, in B.
Cardott, ed., Oklahoma Geological Survey Circular 110: Unconventional Gas of the Southern Mid-Continent Symposium, March 9-
10, Oklahoma City, OK, p. 37-50.
Johnson J.G., Klapper G., Sandberg, C.A. 1985, “Devonian Eustatic Fluctuations in Euramerica”, G.S.A Bull., 96. p. 567-587.,
Pemper, R., Sommer, A., Guo, P., Jacobi, D., Longo, J., Bliven, S., Rodriguez, E., Mendez, F., Han, X., 2006, “A new pulsed neutron
sonde for derivation of formation lithology and mineralogy”, SPE paper 102770.
Ross, C.A., and J.R. P. Ross, 1987, “Late Paleozoic sea levels and depositional sequences” in, C.A. Ross and D. Haman, eds., Timing and
deposition of eustatic sequences: Contrains on seismic stratigraphy: Cushman Foundation for Foraminiferal Research Special
Publication, 24, p. 137-149.
Rottman, K. 2000, “Defining the role of Woodford-Hunton depositional relationships in Hunton stratigraphic traps of Western Oklahoma”
in Johnson K.S. ed., Platform carbonates in the southern Midcontinent 1996 symposium, OGS 101, p. 139-146
Schad S.T., 2004, “Hydrocarbon Potential of the Caney Shale in South Eastern Oklahoma”, M.S. Thesis, University of Tulsa, Tulsa
Oklahoma.
Vulgamore, T., Clawson T.,Pope, C. Wolhart, S. Mayerhofer, M., Machavoe, S. and Waltman C., 2007, “Applying Hydraulic Fracture
Diagnostics to Optimize Stimulations in the Woodford Shale”, SPE 110029, p.1-8

You might also like