You are on page 1of 12

SPE 160862

State-of-the-art Openhole Shale Gas Logging


Javier A. Franquet, SPE; Matt W. Bratovich, SPE; and Richard D. Glass, SPE, Baker Hughes.
Copyright 2012, Society of Petroleum Engineers
This paper was prepared for presentation at the SPE Saudi Arabia Section Technical Symposium and Exhibition held in Al-Khobar, Saudi Arabia, 811 April 2012.
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, as presented, does not necessarily reflect any position of the Society of Petroleum
Engineers, its officers, or members. Papers presented at the SPE meetings are subject to publication review by Editorial Committee of Society of Petroleum Engineers. 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 where and whom the paper was presented. Write Librarian, SPE, P.O. Box
833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.
Abstract
Faced with increasing field maturity and production decline from conventional gas reservoirs, oil companies are shifting their
focus and pursuing new alternatives; one of them being the development of shale and gas plays. To be economically viable,
these low-permeability formations require fracture stimulation. Interval selection within shale reservoirs for hydraulic
fracturing or horizontal laterals are based on several variables: sufficient organic matter or total organic carbon (TOC) and
favorable hydraulic fracturing stimulation. The presence and extent of the natural fracture system can also influence the
performance of a shale reservoir; therefore, natural fractures should be characterized within the shale formation not only from
wireline or LWD borehole images logs but also from cross-dipole deep shear wave imaging which can illuminate fractures up
to 60 ft away that do not intersect the well. To assess these aspects, a mineralogical, structural, and geomechanical
characterization of the shale formation should be conducted. The mineralogical characterization and TOC quantifications
mainly rely on a pulsed neutron spectroscopy and nuclear magnetic resonance (NMR) logs. The processing of capture and
inelastic gamma ray spectra obtained from the pulsed neutron tool quantifies the formations basic elemental composition,
including silicon, calcium, aluminum, iron, sulfur, magnesium, and carbon. Geomechanical characterization is based on
acoustic and density log responses. Variation in the reservoir mineralogy and TOC content affect the rock mechanics
properties. Stress vs. strain curves can be derived from a micro-mechanical model of the rock which enable correlations
between dynamic (obtained from acoustic logs) and static elastic properties (obtained from triaxial compression testing on core
samples). Additionally, the azimuthal and transverse shear wave anisotropies are processed from cross-dipole acoustic logs to
characterize the vertical and horizontal rock stiffness. This anisotropic characterization of the rock enables the evaluation of
the fracture gradient and stress contrast between the target formation and the overlying and underlying formations. The paper
focuses on the interaction between mineralogy, organic content and geomechanical analyses in shale gas reservoir evaluation.
Introduction
Exploitation of unconventional shale gas reservoirs depends on successful horizontal drilling and multi-stage hydraulic
fracturing results. Characterizing rock mineralogy, organic matter content, the natural fracture network characteristics and the
in-situ stress profile play an important role in well stimulation and completion design. Comprehensive formation evaluation
data sets which include acquisition of conventional or rotary core samples are typically acquired in vertical pilot wells. The
integrated shale gas evaluation suite includes mineralogy and lithology characterization logs (natural gamma ray spectral log
in combination with the pulsed neutron gamma ray spectroscopy log), nuclear magnetic resonance (NMR) logging, borehole
imaging logs (wireline acquisition in vertical pilot and LWD in the horizontal lateral boreholes), cross-dipole acoustic logs in
addition to the conventional density-neutron and resistivity logs are also acquired.
Using the variable mineral compositions obtained from the pulsed neutron spectroscopy instrument, various lithofacies are
determined for each shale reservoir. These lithofacies can then be studied in conjunction with the mechanical properties of the
rock, which vary along with variations in the mineralogy. It has been observed that more brittle failure occurs in siliceous
shale intervals than in clay-rich shale intervals when they are compressed in laboratory triaxial loading tests. (Franquet et al.
2010). This observation has been noted in both the Barnett shale (Jarvie et al. 2005) and in the Haynesville shale where
vertical wells that were completed (perforated) and stimulated (hydraulic fractured) in intervals with abundant silica-rich shale
produced better results than wells completed in more clay-rich shale intervals, (LeCompte et al. 2009). The outcome of the
integrated study of lithofacies and mechanical properties is to identify those intervals that are most favorable for fracture
2 SPE SAS 281
stimulation and 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 fracturing results due to proppant embedment, e.g., organic-rich clay
mudstones. These results can be used in the selection of the optimum location for a horizontal lateral.
Comprehensive borehole acoustic logging data are commonly acquired in vertical pilot wells for compressional and shear
wave processing, cross-dipole shear anisotropy, and Stoneley-derived horizontal shear slownesses. These analyses are utilized
to quantify the intrinsic rock anisotropy which can be used to derive anisotropic stress profiles. In-situ horizontal stress
profiling helps to identify potential hydraulic fracture barriers within, above, and below the shale reservoirs. Several shale gas
operators are also characterizing their lateral and horizontal boreholes with LWD images not only for reservoir navigation but
also for better placement of the hydraulic fracturing stages, (Mullen et al. 2010). Recently, deep fracture imaging from cross-
dipole acoustic data enables operators to characterize natural fractures that dont intersect the borehole up to 60 ft away
(Bolshakov et al. 2011) based on the multicomponent dipole SH-wave imaging technique (Tang & Patterson 2009).
Shale lithology and mineralogy logging
The primary technology utilized in the determination of lithology and mineralogy in this integrated approach consists of a
combination of a pulsed neutron and a natural gamma ray spectroscopy tools. These geochemical instruments investigate the
inelastic, capture and natural gamma ray energy spectra to obtain elemental yields and weight fractions of various elements
including Al, C, Ca, Fe, Gd, K, Mg, S, Si, Th, Ti and U. Lithology and mineralogy are then determined by an Expert System
which uses the elemental weight fractions from the geochemical measurements as input. (Pemper et al. 2006 and 2009).
Currently, eighteen minerals are quantified: illite, smectite, kaolinite, chlorite, glauconite, apatite, zeolites, halite, anhydrite,
hematite, pyrite, siderite, dolomite, calcite, K-feldspar, plagioclase, quartz and organic carbon, which is identified either as
kerogen, coal, or oil (Jacobi et al. 2009, Pemper et al. 2009). Fig.1 lists the elements measured in the capture, inelastic and
natural spectrum with the current minerals derived from the elemental weight rock characterization.
Fig. 1: Elements measured by pulsed-neutron geochemical and spectral GR tools along with computed
minerals from elemental weight composition.
The Expert System is a deductive sequential process, which relies solely on geochemistry. It was systematically built on initial
conclusions developed by geochemists, mineralogists and geologists. The system uses a series of over 70 ternary and hybrid
ternary diagrams and applies geologic constraints which assist in obtaining the most probable set of minerals. The approach
begins by identifying the general lithology associated with each tool measurement, based upon the chemical composition of
the formation. Each depth interval is classified into one of the six general categories: Sand, Shale, Coal, Carbonate, Evaporite
or Igneous. These general categories then branch into more specific sub-classifications. A carbonate, for example, is concluded
to be a limestone or dolomite depending upon the amount of Ca or Mg as it is illustrated in Fig. 2. This middle step allows the
analyst to place constraints on the final determination of the subsequent mineral evaluation. The Expert System is constrained
by the principles of mass balance and mineral stoichiometry. Mineral solutions can be determined by optimization or
probabilistic methods which use as input, both conventional log responses and geochemistry logs. These approaches are based
on tool response equations. These methods rely more in fitting (weighted least-squared error) the acquired curves and the
curves derived from the model (theoretical curves) than in application of geological knowledge as opposed to the Expert
System.
SPE SAS 281 3
Fig. 2: Expert System mineralogy determination workflow.
Fig. 3 illustrates the results of the general and specific lithology characterization, mineralogy and element composition of a
typical vertical well drilled in the Barnett shale. The results from the mineralogy logging are compared against the x-ray
diffraction (XRD) and x-ray fluorescence (XRF) core results. The elemental analysis from core is plotted as dots along the
elements tracks while the core mineralogy is presented on the last track. The elements from the pulsed-neutron device are
plotted in pairs:
- K Th (red and dark blue)
- Si Ca (yellow and cyan)
- Mg Fe (dark blue and brown)
- Al S (grey and orange)
- C Ti (green and magenta)
Fi. 3: Elemental weight composition, lithology classification and mineralogy from wireline logging
in the Barnett shale and its comparison against XRD and XRF core analysis.
4 SPE SAS 281
TOC and porosity determination
Quantifying the amount and distribution of the TOC content is a primary objective in shale reservoir evaluation. Several
methods to determine the amount of TOC present have been developed by the industry using empirical relationships to various
conventional log responses such as resistivity, bulk density and total gamma ray. However, the accuracy of these empirical
approaches is often reduced by the variable mineralogy present within a shale reservoir (Jacobi et al. 2008 and Pemper et al.
2009). The integrated petrophysical approach provides two independent measurements of TOC (Jacobi et al. 2008, and 2009).
A direct measurement of TOC is obtained from the elemental carbon weight fraction measured in the inelastic gamma ray
energy spectrum by the pulsed-neutron mineral spectroscopy tool. The amount of the measured carbon which is associated
with the organic material is calculated by subtracting the amount of carbon required as a component of the inorganic minerals
determined by the expert system.
computed Siderite computed Dolomite computed Calcite measured total organic
C C C C C
_ _ _ _
=
The amount of organic carbon is allocated to kerogen, coaly material, bitumen or oil based upon the Th/U ratio, Uranium
content, and other geochemical measurements.
A second TOC calculation is made using NMR porosity, NMR pore fluid density, bulk density, the inorganic matrix density
obtained from the mineralogy and an assumed kerogen density. The calculation is as follows:
( ) |
|

=
1
fluid b
gr
TOC m
gr m
TOC
V

=
TOC m
gr m
gr
TOC
TOC
m

=
Where

b
is the bulk density

gr
is the total grain density including inorganic and organic matrix constituents

TOC
is the kerogen density
is the NMR total porosity

fluid
is the density of pore-filling fluid, determined from NMR fluid typing
V
TOC
is the volume fraction of organic matrix components
m
TOC
is the mass fraction of organic matrix components
Comparison of the TOC weight fractions from the excess carbon and NMR-mineralogy method, described above, can be used
to calculate kerogen density, (Jacobi et al. 2009). Kerogen density has been correlated to thermal maturity in the Marcellus
shale (Ward 2010). The Woodford and Barnett shale formations appear to follow a similar vitrinite reflectance kerogen
density relationship (LeCompte and Hursan, 2010), as it is illustrated in Fig. 4. While this specific relationship should not be
used as a general trend to all type of kerogens from different shale reservoirs, a relationship between kerogen density and
vitrinite reflectance should exist for each shale reservoir but this requires core verification. The thermal maturity of the
kerogen present in the shale is one factor that indicates whether the reservoir will produce dry gas, condensate, or oil.
SPE SAS 281 5
Fig. 4: Kerogen density and vitrinite reflectance relationship obtained from core measurement
in Marcellus, Woodford, and Barnett shale gas wells.
Typical shale gas reservoir porosities are low, often in the range of 3 to 10 %. Porosity calculations using only conventional
log measurements may have significant uncertainties due to the variable mineralogies and the variable amounts of low-density
organic material present in these reservoirs. Uncertainty in the porosity determination can be reduced by utilizing the detailed
mineralogy derived from the Expert System. In the integrated petrophysical approach porosity is preferably derived from the
NMR, which is not affected by lithologic or mineralogic variations. Comparison of NMR total porosities to core porosities in
several shale plays has shown good agreement (Jacobi et al. 2009) as illustrated in Fig. 5.
Examination of SEM images of ion-beam-milled samples reveal a separate nano-porosity system contained within the organic
matter, (Passey et al. 2010 and Sondergeld et al. 2010). Researchers have found that the kerogen contained porosity increases
with increasing kerogen maturity. NMR laboratory measurements of gas shale found that the T2 relaxation time of
hydrocarbons in the organic pores overlap with the T2 response of the clay-bound water contained in the inorganic pores
(Sondergeld et al. 2010). The revised rock model that includes the porosity in the organic material is illustrated in Fig. 4.
Fig. 5: NMR porosity vs. Core porosity comparison and the NMR revised organic shale model.
In the sequential approach, after lithology and mineralogy (including TOC) are obtained, porosity from NMR can be used to
convert mineral weight fractions to volume fractions of the whole rock. These normalized mineral volumes can be associated
to provide matrix and shale inputs to a micromechanical rock model that reproduces the stress-strain behavior such that rock
strength and elastic properties can be obtained.
6 SPE SAS 281
Static mechanical properties
Laboratory testing of core samples is the preferred method for determining quasi-static rock mechanical properties. Triaxial
core experiments, however, could be expensive, time-consuming, and typically only available at few discrete depth levels. A
continuous profile of mechanical properties could be obtained from empirical correlations between core triaxial tests and
logging properties; such as acoustic compression slowness, formation density, or total porosity, (Chang et al. 2006). However,
finding an empirical correlation that always works in all shale gas reservoirs is probably very difficult to find. Analogous rock
types could exhibit different mechanical behavior due to changes in the in-situ conditions, differences in clay content and TOC
(which greatly affect the intrinsic mechanical behavior of shale gas rocks), or changes in diagenetic processes from one basin
to another.
Continuous rock mechanical properties can also be obtained from neural networks (Valbuena et al. 1999), after a
representative set of laboratory results is obtained to effectively train the network. Typically, this type of black-box correlation
would only work within the laboratory sampling range of testing properties, and it would be constrained to the type of
formation and oil field where they came from.
As an alternative to empirical correlations and neural network, an analytical code was developed by Baker Hughes and IKU
(SINTEF) in 1996 to predict static mechanical properties of rocks from logging information, (Raaen et al. 1996). This code
reproduces the micromechanical processes that occur in rocks during triaxial loading. An important part of this model is the
description of the difference between the static and the dynamic moduli of the rock. The code simulates the rock strain
mechanisms that require large deformation to be activated, such as crushing or plastification of grain contacts, pore collapse,
and shear sliding along internal rock cracks. Many of these are not explicitly considered in most empirical correlations.
Recently, shale gas rock deformation had been modeled by this code reproducing the stress-strain behavior under confining
conditions, (Franquet et al. 2010).
Fig. 6 shows the predicted stress-strain curves by the micromechanical code and the laboratory measured curves from
Haynesville shale core samples. The solid curves are the stress-strain curves measured in the lab while the dotted curves are
the predicted stress-strain curves from logging data. The red curves correspond to a core sample taken from a siliceous shale
interval while the green curves correspond to a sample taken from a clay-rich shale interval. QS is the maximum deviatory
stress, CP is the confining pressure, CCS is the confined compressive strength, EMOD is the Youngs modulus, PR is the
Poissons ratio and VSH is the volume of shale.
Fig. 6: Micromechanical stress-strain modeling of triaxial tests from openhole logging data and comparison agaist laboratory
results in two Haynesville shale core samples at 3600 psi of effective confining pressure. Samples with higher silicieous minerals
behave more brittle in comparison to the clay-rich shale sample.
SPE SAS 281 7
Fig. 7: Log-derived static mechanical properties obtained in the Upper Haynesville shale.
Compressive rock strength, confined Youngs modulus (rock stiffness), confined Poissons ratio, and Biots coefficient are
important to characterized shale rocks. An example of the static rock characterization utilizing the micromechanical analytical
code in the Upper Haynesville (also referred as the Boisser) is presented in Fig. 7, (LeCompte et al. 2009). The LMP
simulated results are compared against triaxial compression test performed on core samples (plotted as black dots). The
dynamic mechanical properties are plotted in red next to the log-derived static properties plotted in blue for Youngs modulus
and Poissons ratio. The dynamic Youngs modulus and the dynamic Poissons ratio are higher than the static elastic
properties as obtained from triaxial testing. Shale stiffness determination is critical to accurate estimation of shale brittleness
and hardness from acoustic logging.
The petrophysical volumes were obtained from pulsed-neutron geochemistry logging and the total porosity from NMR
logging. The basic input data for building the micromechanical model are: compressional wave slowness DTC, shear wave
slowness DTS, formation density, volume of shale and total shale porosity. The Lals empirical correlation (Lal 1999)
produced a good prediction of confined compressive strength in the Haynesville shale, as is presented in Fig. 7.
Shale anisotropy and stress profile
Most rocks present some anisotropy in their acoustic properties due to fabric, natural fractures or depositional structure. Shales
have strong transverse isotropy (TI) anisotropy and they could also have azimuthal anisotropy when vertical natural fractures
are present. Shale anisotropy has been studied extensively from seismic (Sayers 2005) and borehole acoustic data (Walsh et al.
2006) to fully characterize its stiffness tensor characteristics. Shale TI anisotropy has been used to improve wellbore
completion in some unconventional reservoirs of North America (Higgins et al. 2008) by characterizing the vertical stress
profile applying the simple ANNIE approximation (Schoenberg & Sayers 1996).
This approximation consists of two assumptions that are combined with three independent acoustic measurements to
characterize the rock stiffness tensor for TI materials. The first assumption defines all non-diagonal terms equal in the stiffness
tensor which only happen in isotropic materials. The second assumption constrains the Delta (o ) Thomsens coefficient
(Thomsen 1986) equal to zero.
8 SPE SAS 281
( ) ( )
( )
44 33 33
2
44 33
2
44 13
2 C C C
C C C C

+
= o
ANNIE approximations:
- C
12
= C
13
- C
13
= C
33
2 C
44
(o = 0)
The major drawback of this ANNIE approximation is the contradictory acoustic measurements obtained in few shale rock
samples where the Delta Thomsens coefficient is not zero and the non-diagonal terms C
12
and C
13
are not equal.
It is possible to use the cross-dipole azimuthal anisotropy in combination with the TI anisotropy to reconstruct the full stiffness
tensor without constraining the model with the ANNIE approximation (Franquet & Rodriguez 2011). This methodology uses
acoustic anisotropies measured in core samples or existing correlations between Epsilon c and Gamma Thomsens
coefficients to build the full (nine terms) orthotropic rock model. Then, the terms of the stiffness tensor are used to estimate in-
situ horizontal stress profiles along the vertical direction, (Franquet et al. 2011).
If the stiffness terms are measured in a vertical borehole drilled toward the z-direction, the effective horizontal stresses acting
in the formation (o
xx
and o
yy
) can be characterized as a function of the stiffness terms C
ij
, pore pressure P
P
and lateral strains
(c
h
and c
H
) from rearranging the first three equations of the above tensor relationship. Three different Biots coefficient could
be needed as this poro-elastic parameter could be also a tensor, particularly when laminations and/or open natural fractures
exist in the shale rock.
( )
H h p x p z v h
C
C C
C
C
C
C P P
C
C
c c o o o o
|
|
.
|

\
|
+
|
|
.
|

\
|
+ + =
33
23 13
12
33
2
13
11
33
13
min ,
( )
h H p y p z v H
C
C C
C
C
C
C P P
C
C
c c o o o o
|
|
.
|

\
|
+
|
|
.
|

\
|
+ + =
33
23 13
12
33
2
23
22
33
23
max ,
An example of the Eagle Ford shale horizontal stress profile is presented in Fig. 8. The TI shear-wave anisotropy was derived
from the Stoneley-wave processing, while the azimuthal rock anisotropy was obtained from the cross-dipole data. The
horizontal shear wave slowness DTSH derived from the Stoneley-wave processing obtained on a vertical well defines the C
66
while the far-field fast and slow shear wave obtained from cross-dipole processing define C
44
and C
55
, respectively. The
vertical compressional wave slowness defines C
33
while the horizontal compressional wave slowness C
11
and C
22
can be
obtained from Epsilon-Gamma correlations or dynamic core measurements. The non-diagonal terms of the stiffness tensor are
the most difficult to obtain in anisotropic rocks since it is required to do compressional or shear wave velocity measurements
on oblique directions respect to the principal anisotropy directions (x, y, and z). For example, C
12
can be obtained from C
66
,
C
22
, C
11
, rock density and the compressional wave velocity V
P-xy45
propagated a 45-degree angle between x and y principal
directions as follow:
( )
66
2
11 22
2
11 22
66
2
45 12
4
1
2 2
2 C C C
C C
C V C
xy P
|
.
|

\
|
=

The stress profile mainly from borehole acoustic logging and the specific lithology characterization from the processing of the
mineralogy logging are the fundamental information for selecting perforating intervals in vertical wells to be stimulated or to
identify siliceous shale intervals to drill horizontal lateral branches and subsequent multi-stage fracturing.
|
|
|
|
|
|
|
|
.
|

\
|

|
|
|
|
|
|
|
|
.
|

\
|
=
|
|
|
|
|
|
|
|
.
|

\
|
xy
zx
yz
zz
yy
xx
xy
zx
yz
zz
yy
xx
C
C
C
C C C
C C C
C C C

c
c
c
t
t
t
o
o
o
66
55
44
33 23 13
23 22 12
13 12 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
SPE SAS 281 9
Fig. 8: Horizontal stress profile from transverse and azimuthal anisotropies acquired from a vertical Eagle Ford Shale well.
LWD horizontal borehole image logging for multi-stage hydraulic fracturing
Initially, multi-stage fracturing in shale gas had been designed by equal spacing and equal fracturing along the lateral
boreholes. Recently, LWD electrical borehole images are used to design the optimum placement of the packers for multi-stage
fracturing in the horizontal braches, (Mullen et al. 2010).
The main objective of the LWD images is to characterize fractures and potential faults that are undetectable from surface
seismic in the horizontal borehole. The goal is to avoid fracturing stages near a fault or pre-existing hydraulic fractures from
previous stimulation jobs performed in nearby horizontal wells. Additionally, the locations of sweet spots containing natural
fractures along the lateral are considered as target zones for hydraulic stimulation. Finally, the borehole-induced fractures
along the well are used to back-calculate the in-situ stress regime and potential implication in hydraulic fracturing stimulation.
Fig. 9: LWD electrical borehole imaging for reservoir navigation and better placement of the hydraulic
fracturing stages along horiozntal wellbore drilled in the Barnett shale.
BarnettShale ForestburgLS
ForestburgLS BarnettShale
Fault Undetectableon3Dseismic
10 SPE SAS 281
Deep shear wave imaging
Deep shear wave imaging consists of using shear body waves generated by a dipole source that radiate away from the
borehole. In the presence of geologic structures with shear impedance, these shear waves can be reflected back to the borehole,
enabling the structure to be imaged. The dipoles azimuthal sensitivity, along with the four-component analysis, enables the
determination of the reflectors azimuth. This principle has been used to trace anhydrite beds into salt dome structures (Tang
2009) and natural fractures in the Haynesville shale.
The shear wave radiation is zero along the dipole source polarization direction because the particle motion is pure
compressional in that direction. However, the radiation is stronger perpendicular to the dipole source polarization. Tang and
Patterson discovered that SH wave radiation pattern (horizontal polarized shear wave) has the best coverage for imaging
structure into the formation (Tang 2009).
Fig. 10 shows clear non-intersecting natural fractures in a vertical well drilled into the Haynesville shale. The identification of
these features is considered very important for wellbore completion, reservoir stimulation operations, and future well
placements.
Fig. 9: Deep shear wave imaging of non-intersecting natural fractures in the Haynesville Shale,
looking out 50 ft from the borehole
Conclusions
Exploitation of unconventional shale gas reservoirs depends on successful hydraulic fracturing and horizontal drilling.
Characterizing rock mineralogy, organic matter content and the in-situ stress profile play an important role for well completion
design in shale gas reservoirs.
Siliceous shale intervals produce more brittle failure than clay-rich layers when they are compressed in triaxial tests; therefore,
the preferred shale gas zones for stimulation are recognized as very silica-rich with high organic content. The NMR porosity
SPE SAS 281 11
gives the best relationship to the core porosity and produces reliable measurements even in complex-mineralogy shale
formations with low porosity.
Borehole cross-dipole and TI anisotropy analysis can be combined with core acoustic anisotropies to fully describe the
orthotropic stiffness tensor without constraining the rock model with the ANNIE approximation. Acoustic core measurements
in some rocks indicate that C
12
is not equal to C
13
and the Thomsen Delta coefficient is not zero.
It is possible to derive horizontal stress profiles using the full characterization of the orthotropic rock stiffness tensor and the
absolute stress magnitude can be back-calculated from borehole breakouts and or induced fractures.
LWD borehole images in horizontal wells not only provide reservoir navigation capability to operators but also enable multi-
stage fracturing placement optimization along the lateral section.
Deep shear wave imaging can image a much larger volume around the borehole and map formation structures that do not
intersect the well and are invisible from surface seismic. Interesting natural fractures have been identified in the Haynesville
shale that might explain high gas productivity zones.
Acknowledgments
The authors would like to thank all operating companies that anonymously released their logging data from Barnett, Eagle
Ford, and Haynesville shales. We would also like to thank Baker Hughes for allowing us to publish this paper.
References
Bolshakov A., D. Patterson, and C. Lan. 201., Deep fracture imaging around the wellbore using dipole acoustic logging, SPE-
146769 presented at the SPE Annual Technical Conference and Exhibition held in Denver, Colorado, 30 October 2
November.
Chang, C., M.D. Zoback, and A. Khaksar. 2006. Empirical relations between rock strength and physical properties in
sedimentary rocks, JPSE, vol. 51, pp 223-237.
Franquet, J.A., C.A. Wolfe, and E.F. Rodriguez. 2010. Micromechanical modeling of rock stress-strain curves from log data,
paper presented at the 1
st
International Conference on Integrated Petroleum Engineering and Geosciences ICIPEG 2010, a
conference of the World Engineering, Science & Technology Congress ESTCON 2010. Kuala Lumpur, Malaysia, 15-17 June.
Franquet, J.A. and E.F. Rodriguez. 2011. Orthotropic horizontal stress characterization from logging and cored-derived
acoustic anisotropies: Presented at the Abu Dhabi Topical Conference - Multidimensional Reservoir Characterization.
SPWLA.
Franquet, J.A., D. Patterson, and D. Moos. 2011. Advanced dipole borehole acoustic processing Rock physics and
geomechanics applications, to be presented at the SEG International Exposition and 81
st
Annual Meeting, San Antonio, Texas,
USA, 19-23 September.
Higgins, S., S. Goodwin, A. Donald, T. Bratton, and G. Tracy. 2008. Anisotropy stress models improve completion design in
the Baxter shale, SPE-115736 presented at the 83
rd
Annual Technical Conference and Exhibition held in Denver, Colorado,
USA, 21-24 September.
Jacobi, D., J.M. Longo, A. Sommer, and R. Pemper. 2007. A chemistry-based expert system for mineral quantification of
sandstones, paper 1.3.45, Trans., PETROTECH 7
th
International Oil & Gas Conference and Exhibition, New Delhi, India.
Jacobi, D., M. Gladkikh, B. LeCompte., G. Hursan, F. Mendez, S. Ong, M. Bratovich, G. Patton, and P. Shoemaker. 2008. An
integrated petrophysical evaluation of shale gas reservoirs, SPE-114925 presented at the CIPC/SPE Gas Technology
Symposium Joint Conference held in Calgary, Canada, 16-19 June.
Jacobi, D., J. Brieg, B. LeCompte, M. Kopal, G. Hursan, F. Mendez, S. Bliven, J. Longo. 2009. Effective Geochemical
Characterization of Shale Gas Reservoirs from the Wellbore Environment: Caney and the Woodford Shale, SPE 12431
presented at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 47 October.
12 SPE SAS 281
Jarvie, D.M., R.J. Hill, and R.M. 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, p 37-50, Oklahoma City, Oklahoma, USA, 9-10 March.
Lal, M. 1999. Shale stability: Drilling fluid interaction and shale strength, SPE-54356 presented at the SPE LACPEC Latin
American and Caribbean Petroleum Engineering Conference, Caracas, Venezuela 21-23 April.
LeCompte, B., F. Mendez, D. Jacobi, and J. Longo. 2008. Defining clay type using NMR and geochemical logging
meassuremnets, paper presented at the SPWLA 49
th
Annual Logging Symposium, Edinburgh, Scotland, 25-28 May.
LeCompte, B., J.A. Franquet, D. Jacobi. 2009. Evaluation of Haynesville Shale vertical well completions with a mineralogy
based approach to reservoir geomechanics, SPE-124227 presented at the SPE Annual Technical Conference and Exhibition
held in New Orleans, Louisiana, USA, 47 October.
LeCompte, B. and G. Hursan. 2010. Quantifying source rock maturity from logs: How to get more than TOC from delta log R,
SPE-133128 presented at the SPE Annual Technical Conference and Exhibition help in Florence, Italy, 19-22 September.
Mullen, M., J. Pitcher, D. Hinz, M. Everts, D. Dunbar, G. Carlstrom, and G. Brenize. 2010. Does the presence of natural
fractures have an impact on production? A case study from the middle Bakken Dolomite, North Dakota, SPE-135319
presented at the SPE Annual Technical Conference and Exhibition held in Florence, Italy, 19-22 September.
Passey, Q., S. Creaney, J. Kulla, F. Moretti, and J. Stroud. 1990. A practical model for organic richness from porosity and
resistivity logs. AAPG Bulletin 74 (12): 1777-1794.
Passey, Q., K. Bohacs, W. Esch, R. Klimentidis, and S. Sinha. 2010. From oil-prone source rock to gas-producing shale
reservoir Geologic and petrophysical characterization of unconventional Shale-Gas reservoir, SPE-131350 presented at the
CPS/SPE International Oil & Gas Conference and Exhibition held in Beijing, China, 8-10 June.
Pemper, R., G. Page, E. Prati. 1996. A new generation natural gamma ray spectroscopy logging system, paper E027, Trans.,
17
th
European Formation Evaluation Symposium, Amsterdam, Netherlands.
Pemper, R., A. Sommer, P. Guo, D. Jacobi, J. Longo, S. Bliven, E. Rodriguez, F. Mendez, X. Han. 2006. A new pulsed
neutron sonde for derivation of formation lithology and mineralogy, SPE-102770 presented at the SPE Annual Technical
Conference and Exhibition held in San Antonio, Texas, USA, 24-27 September.
Pemper, R., Han, X., Mendez, F., Jacobi, D., LeCompte, B., Bratovich, M., Feuerbacher, G., Bruner, M., Bliven, S. 2009. The
direct measurement of carbon in wells containing oil and natural gas using a pulsed neutron mineralogy tool, SPE 124234
presented at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 4-7 October.
Raaen, A.M., K.A. Hovem, H. Joranson, and E. Fjaer. 1996. FORMEL: A step forward in strength logging, SPE-36533
presented at the SPE Annual Technical Conference and Exhibition held in Denver, Colorado, USA, 6-9 October.
Sayers, C.M. 2005. Seismic anisotropy of shales: Geophysical Prospecting, 53, 667-676.
Schoenberg, M., F. Muir, and C.M. Sayers. 1996. Introducing Annie: A simple three parameter anisotropy velocity model for
shales: Journal of Seismic Exploration, 5, 35-49.
Tang, X.M. and D.J. Patterson. 2009. Single-well S-wave imaging using multicomponent dipole acoustic-log data. Geophysics
74: WCA211-WCA223
Thomsen, L. 1986. Weak elastic anisotropy: Geophysics, 51, 1954-1966.
Valbuena, J., J.A. Franquet, and R. Carbonell. 1999. Estimation of rock mechanical properties using a neural network
approach, paper presented at the 2
nd
Symposium on Artificial Intelligence, CIMAF. La Habana, Cuba, 22-26 March.
Walsh, J., B. Sinha, and A. Donald. 2006. Formation anisotropy parameters using borehole sonic data: Presented at the 47
th
Annual Logging Symposium, SPWLA.
Ward, J. 2010. Kerogen density in the Marcellus shale. SPE 131767 presented at the SPE Unconventional Gas Conference,
Pittsburgh, Pennsylvania, USA, 23-25 February.

You might also like