Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Geomechanical Modeling of an Anisotropic Formation-

Bakken Case Study

Mehdi Ostadhassan1, Zhengwen Zeng2, Siavash Zamiran3

1- Dept. of Geology and Geological Engineering, University of North Dakota, Grand Forks, ND, USA,
Email: mehdi.ostadhassan@UND.edu
2- Dept. of Geology and Geological Engineering, University of North Dakota, Grand Forks, ND, USA
3- Department of Civil Engineering, Southern Illinois University Carbondale, IL, USA, Email:
zamiran@siu.edu, zamirans@gmail.com, Website: www.zamiran.net, Phone: +1 (618) 334-4572

Proceeding: 46th US Rock Mechanics/Geomechanics Symposium. Chicago, IL (ARMA 2012), Postprint


version
Publisher: American Rock Mechanics Association
Ostadhassan, M., Zeng, Z., & Zamiran, S. (2012). Geomechanical Modeling of an Anisotropic
Formation- Bakken Case Study. In 46th US Rock Mechanics/Geomechanics Symposium.
Chicago, IL: American Rock Mechanics Association.

ABSTRACT: Production of hydrocarbon causes changes in pore pressure and effective stresses acting on the reservoir
rocks. This will be followed by reservoir compaction, surface subsidence and may lead to fault reactivation, casing or
wellbore failure and closure of micro-cracks. Nonetheless, acquiring a good understanding of rock strength, pore
pressure and in-situ stress will be critical to successful horizontal drilling and hydraulic fracturing. Bakken Formation
of Williston Basin, North Dakota, which is identified by three distinct members, is a huge unconventional, self-
sourced, naturally-fractured reservoir. It is one of the least studied sedimentary rock units in the basin. The over
pressured nature of this formation have made the study of its geomechanical properties even more important. Natural
fractures are also considered as another major source of problems in this reservoir. To investigate these problems,
Mechanical Earth Model (MEM), a numerical representation of the reservoir properties, was built. This enabled to
predict the alterations and changes of the geomechanical properties in the reservoir. The results show that the lower
and upper members are mechanically transverse isotropic whereas the middle member is isotropic. Besides, the
numerical geomechanical modeling demonstrate that the elastic anisotropic characteristics of the upper and lower
members will result in elastic failure of the region around the wellbore following a shear failure phase but the elastic
isotropic middle member will mostly stay in the shear failure state.

Keywords: Williston Basin, Bakken Formation, Unconventional reservoir, Anisotropic shale, FLAC3D

1. INTRODUCTION
Unconventional reservoirs, especially tight shales, are becoming an important target in hydrocarbon
exploration. Oil companies cannot neglect the huge amount of their recoverable reserves. Since the very
low matrix permeability and primary porosity, conventional techniques in drilling, production and
characterization may not be applicable to these reservoirs; new stimulation methods should be applied.
Successful field development and enhanced recovery operation need a comprehensive understanding of
geomechanical properties of the reservoir. Geomechanical study may generate information about rock
strength, pore pressure, in-situ stress and elastic properties. Besides, drilling through the formation will
cause stress alteration around the borehole and even in a radius into the formation, consequently these
changes should be simulated prior to drilling. A good estimation of formation elastic properties will lead to
an accurate stress analysis and can prevent future financial losses.

2. METHODOLOGY
2.1. Mechanical Earth Model (MEM)
To carry out a complete geomechanical study, Mechanical Earth Model (MEM) should be built. MEM is a
numerical representation of reservoir properties in 1D, 2D or even 3D style. MEM contains data related to
the rock failure mechanisms, in-situ stresses, stratigraphy and geologic structure of the reservoir [1-4]. As
aforementioned, MEM should be built any time before the drilling. It will be also upgraded with new
information anytime when drilling is in progress and later during the production. Figure 1 depicts the
flowchart for constructing a proper MEM.

Fig.1. Mechanical Earth Model (MEM) flowchart.

2.2. Mechanical Anisotropy


Anisotropy, the variation of a property with respect to the direction of measurement [5], is an important
concept that has been neglected in constructing the MEM and geomechanical numerical modeling for
decades. In the petroleum industry isotropy assumptions have been frequently applied to geomechanical
modeling, not because they are good approximations, but since anisotropic measurements are not available.
Moreover, isotropy assumptions have led to inaccurate results in geomechanical models [6, 7]. Mechanical
anisotropy means that elastic properties of the rocks such as Young’s modulus, Poisson’s ratio and shear
modulus strongly changes due to the direction of the measurement. Nowadays, recent advancements in
sonic logging technology have made the three dimensional (3D) analysis of mechanical anisotropy possible
around the wellbore with high level of confidence [8, 9].
2.3. Shale Anisotropy
Shales are the dominant elastic component in sedimentary basins [10] and clay minerals as the main
constituent parts of their structure cause them to be highly anisotropic [11] . Anisotropy has made prediction
of shale elastic properties highly challenging. To make shale anisotropy analysis easier, shales could be
described as being transversely isotropic with vertical axis of rotational symmetry [5, 12], thus the study of
their mechanical anisotropy becomes less challenging.
Several years back in the petroleum industry, shales are considered to be only the source or the cap rock of
a petroleum system. Currently the oil from tight shale plays has brought a huge attention to this rock type
where they are playing the role of a reservoir rock, also known as unconventional reservoir. This makes
their mechanical property analysis even more critical. Thus while assessing their mechanical properties
their anisotropic behavior shouldn’t be neglected. The impact of anisotropy on rock mechanical properties
such as Poisson’s ratio, Young’s modulus, rock strength and even the in-situ stresses in tight shale plays is
significant [13]. Accordingly anisotropy shows a great impact on the geomechanical models. In this study
we have included anisotropic characteristics of shales in the geomechanical modeling and elastic properties
determination of the Bakken Formation. Isotropic calculations of such parameters were carried out as well.
The required data was acquired from cross-dipole sonic logging tool in a vertical well located in the
Williston Basin, North Dakota.

3. THEORY
3.1. Stress Determination
3.1.1. Vertical Stress
Total vertical stress is defined as the combination of weight of the rock matrix which is the vertical effective
stress plus the pressure exerted by the fluids in the pore spaces overlying the depth of interest as shown in
Eq. (1):

ℎ = ∫ � � ��� (1)
where ℎ is the total vertical stress at depth ℎ, � � is the density at depth � below the surface and � is
the acceleration due to the gravity. Eq. (1) can also be rearranged in the form of Eq. (2):
� = ℎ −� (2)
Where � is the vertical effective stress and � is the pore pressure which is caused by the fluid in the pore
spaces.
ℎ can be computed by integrating bulk densities from the density logs data. It should be noticed that if
water column exists above the surface, � �� should be added to Eq. (1); � is water density and � is
water depth.
3.1.2. Pore Pressure
A good knowledge of pore pressure is necessary to prevent blow outs and to ensure a safe well design [14].
Pore pressure (� ) can be measured directly from Repeated Formation Test (RFT) and Drill Stem Test
(DST) with high accuracy during drilling. When these data are not available, qualitative pore pressure
estimation from velocity data can be utilized. Several methods for velocity-based pore pressure estimation
have been proposed in the literature [15, 16]. The most widely used one in the petroleum industry is the
Eaton formalism [17]. This approach has been discussed thoroughly in the literature and is based on the
fact that the relationship between the ratio of the observed sonic log value (or slowness) to the normal
velocity value (or slowness) with the pore pressure follows Eq. (3) [18, 19]:

� = − − �ℎ �
(3)

where � is pore pressure (MPa) , �ℎ is hydrostatic pressure (MPa), is total vertical stress (MPa) and �� is
slowness (µs/ft).
3.1.3. Horizontal Stress
Horizontal stress calculations for an isotropic medium have been extensively used in the industry and were
derived from the solution of the linear poro-elastic equation [20]. Horizontal stresses for an isotropic
poroelastic medium under uniform tectonic horizontal strain can be expressed as follows, Eq. (4) and (5),
[6,7]:
� � ��
�ℎ − �� = −�
(� − �� ) + −�
�ℎ + −�
� (4)
� � ��
� − �� = (� − �� ) + � + �ℎ (5)
−� −� −�

where � is Young’s modulus, is Poisson’s ratio, �ℎ is the minimum horizontal stress, � is the maximum
horizontal stress, � is the vertical (overburden) stress, α is Biot’s constant, �ℎ is the minimum horizontal
strain and � is the maximum horizontal strain.
Horizontal stresses for a transversely isotropic medium with vertical axis of symmetry under uniform
tectonic horizontal strain can be developed as [6, 7, 21]:
�ℎ � � � �ℎ � �ℎ � � ℎ �
�ℎ − �� = �� −�ℎ �
[� − � − � ]+ −�ℎ �
�ℎ + −�ℎ �
� (6)
�ℎ � � � �ℎ � �ℎ � � ℎ �
� − �� = �� −�ℎ �
[� − � − � ]+ −�ℎ �
� + −�ℎ �
�ℎ (7)

where �ℎ is Young’s modulus in the plane of isotropy, � is Young’s modulus along the axis of
symmetry which is the direction of anisotropy, ℎ is Poisson’s ratio in the plane of isotropy, is
Poisson’s ratio along the axis of symmetry. α is Biot’s constant and is the poroelastic constant. It isclear
that the variants of Eqs. (4) and (5) are very well improved from isotropic � and to anisotropic ones in
Eqs. (6) and (7).

3.2. Elastic Parameters


3.2.1. Isotropic Elastic Parameters
Elastic parameters are the main input parameters which should be calculated for evaluating the horizontal
stress. Conventional sonic logging enables measurement of the dynamic isotropic Young’s modulus and
Poisson’s ratio along the wellbore by Eqs. (8) and (9) [22, 23]:

= −
(8)

� −
�= −
(9)

where � is the formation density, � is compressional wave velocity and � is shear wave velocity.
3.2.2. Anisotropic Elastic Parameters
For an anisotropic elastic medium assuming vertical transverse isotropy (VTI), by applying linear Hook’s
law expressed in Eq. (10) and taking � as the axis of rotational symmetry the stiffness matrix becomes Eq.
(11) in the conventional two-index Nye notation [25]:
� =� � (10)
where � is the stress tensor, � is the forth rank stiffness tensor and � is strain tensor.
 C 11 
 
C 12 C 13 0 0 0
C 21 C 22 C 23 0 0 0 
 C 31 
C ij   
C 32 C 33 0 0 0 (11)
 0 0 
 0 0 
0 0 C 44 0

 
C 66 
0 0 0 C 55
 0 0 0 0 0

From Eq. (11) it’s found that five independent nonvanishing elastic stiffness coefficients plus
� governing on the elastic transverse isotropic medium are: � = � , � , � = � , � =
� −�
� =� =� ,� =� and � = [24], in the conventional two-index Nye notation [25].

The inverse version of Eq. (10) known as compliance tensor in which the strain is expressed as the linear
function of stress can be written as � = � where = � − . Thus the nonvanishing elastic compliance
coefficients for a linear elastic transverse isotropic medium become: = , , =
, = = = , = and = − in the two index notation.
The compliance tensor of a vertical transverse isotropic medium can be expressed in terms of Young’s
modulus and Poisson’s ratio as follows [26, 27]:
−� −�
� � �
−� −�
� � �
−� −�
� � �
= (12)
µ

[ µ ]
If � is taken as the axis of vertical rotational symmetry and � � the plane of symmetry then � = �ℎ
is horizontal Young’s modulus in the plane of symmetry, � = � is the vertical Young’s modulus
along the axis of symmetry, = = ℎ is the horizontal Poisson’s ratio in the plane of symmetry
(isotropy) and = = is the vertical Poisson’s ratio along the vertical axis of symmetry. The
anisotropic Young’s modulus and Poisson’s ratio for a vertically transverse isotropic medium in terms of
elastic stiffness coefficients � , Eqs. (13)- (16) are obtained by mathematical calculations after re-

inversing the compliance tensor Eq. (12), � = . Thus the horizontal and vertical � and will be as
follows [6, 13, 23, 28]:
� −� � � − � +� �
�ℎ = (13)
� � −�


� =� −� +�
(14)

� � −�
ℎ =� � −�
(15)

=� +�
(16)

To evaluate the anisotropic elastic parameters from Eqs. (13-16), elastic stiffness coefficients (� ) are
needed. Cross-dipole sonic log is capable to capture those required elastic stiffness coefficients (� ) of a
transverse isotropic medium azimuthally and radially deep into the formation [24, 29, 30, 31].

4. DISCUSSION AND RESULTS


Sonic Scanner log (Mark of Schlumberger) was acquired through the Bakken Formation in a vertical well
in the Williston Basin, North Dakota. The Bakken Formation is primarily a shaly unit which is aged late
Devonian to early Mississippian. The Bakken Formation consists of three members: upper shale member,
middle clastic and carbonate member and lower shale member [32]. Due to the very low porosity and
permeability of this formation, along with the presence of fractures, it is categorized as unconventional,
naturally fractured reservoir. As a result in order to enhance the production from this tight formation,
hydraulic fracturing should be performed. Therefore a comprehensive mechanical earth modeling and
geomechanical numerical modeling is necessary for success in such operations. Due to the presence of
possible vertical and sub-vertical fractures and high clay volume content of this formation, isotropic
assumptions for geomechanical study will be misleading; hence anisotropic characteristics of the Bakken
Formation [24] should be included in the mechanical earth modeling and geomechanical study. Figure 2
represents the corresponding Neutron-Density log (first track), gamma ray log (second track) and formation
lithology (third track) through the Bakken Formation. This formation can be easily identified from the
overlying Lodgepole Formation and underlying Three Forks Formation from its very distinct high gamma
ray signature. Likewise the upper and lower shale members can be distinguished from the middle member
with their much higher gamma ray response.
4.1. Stress Profile Determination
From Figure 2 the depth intervals for Upper Bakken (UB), Middle Bakken (MB) and Lower Bakken (LB)
are 9712ft to 9728ft, 9728ft to 9776ft and 9776ft to 9814ft, respectively. The chosen well is highly
productive. This well is located far from any significant geologic structures in the study area and the layers
are found to be horizontal. Figure 3 depicts the stress profile through the Bakken interval. The total vertical
stress is obtained from Eq. (1). Due to the availability of complete density data in the well, density trend
analysis was unnecessary [15, 33].
To calculate the pore pressure Eq. (3) was applied [17, 19]. For such purpose, compressional wave velocity
(slowness) trend throughout the well was obtained and is shown in Figure 4. A close look into the slowness
variation trend (black line), a deflection from the normal velocity-compaction trend is observed in the
Bakken interval (red oval). This represents the overpressure nature of this layer. Eaton relation, Eq. (3),
was used to fit the slowness-depth data to delineate the normal velocity (slowness) trend that would be
created only due to the compaction [15].
Fig.2. Conventional logs through the Bakken Formation. Neutron-Density (first track), Gamma Ray (second track)
and formation lithology and clay volume (third track).

Fig.3. Stress profile through the Bakken Formation.


Fig. 4. Compressional wave slowness trend (black curve) and deflection from the normal velocity (slowness) is
obvious in the Bakken (red oval) due to the overpressure nature of this formation.
This deviation from the normal velocity (slowness) trend can be explained by unloading effects which were
originated from the decrease in the effective stress acting on the rock frame. The decrease in the effective
stress normally takes place because of increase in the pore pressure in a specific interval [15, 34].
Considering the petroleum system governing on the Bakken, this overpressure behavior can be explained
by the conversion of kerogen to hydrocarbon in the upper and lower shale members. In addition the high
clay volume content of the middle member (Figure 2) in the vicinity of clastic and carbonate facies is caused
a high degree of heterogeneity. Consequently, migration of the generated and expelled hydrocarbon from
the upper and lower members into the middle section through vertical fractures can illustrate the abnormal
overpressure characteristics of the Middle Bakken. Pore pressure profile is represented in Figure 3.
Ultimately we can declare that the highly compacted shales of UB and LB along with high clay volume
content of MB are the primary reasons for the high overpressure nature of these intervals.
Minimum horizontal stress profile under isotropic assumptions in the Bakken Formation is shown in Figure
3. This stress profile (green curve in Figure 3) was calculated using Eq. (4). Minimum horizontal stress for
transverse isotropic medium was estimated from Eq. (6) and plotted in Figure 3(purple line). Comparing
these two minimum horizontal stress profiles we can notify a better improvement when anisotropy
assumptions are involved in the calculations. It is concluded that anisotropy will change the magnitude of
minimum horizontal stress towards more accurate approximations.
In order to evaluate the horizontal stress magnitude from Eqs. (4-7), lateral tectonic strains, �ℎ , � should
be gleaned. It’s often assumed that after fluid withdrawn from a reservoir where the vertical stress is the
major acting principal stress, lateral strain is inhibited by the rock adjacent to the reservoir, thus uniaxial
strain assumptions maybe acceptable [27] so uniaxial strain assumption was taken into account. This means
that the only strain occurred was in the vertical direction and the lateral tectonic strains, �ℎ , � in Eqs. (4-
7) were assumed zero [7]. However adding the tectonic strain values from the core calibration will result in
more accurate outcomes [6, 7, 35]. Since the Williston is considered as a cratonic basin and is currently
under tectonic equilibrium, setting the lateral strains, �ℎ , � equal to zero was reasonable.
To calculate horizontal stress from Eq. (4) and Eq. (6) Biot’s constant� = and poro-elastic constant =
. The measurement of Biot’s constant for the Bakken samples is still in progress at the North Dakota
University Petroleum Research Lab (PRL). If true values of Biot’s constant are applied, it is expected to
improve the predicted minimum horizontal stress.
4.2. Derivation of Elastic Parameters
Isotropic dynamic Poisson’s ratio and Young’s modulus were derived from Eqs. (8) and (9) with
compressional and shear wave velocities. The calculated Poisson’s ratio and Young’s modulus are
considered as the dynamic elastic parameters; and these values should be transformed to the static ones to
better represent the geomechanical characteristics of the formation. The dynamic Poisson’s ratio was not
converted to the static value since the laboratory measurements denote that the static and dynamic Poisson’s
ratios won’t differ dramatically. Dynamic isotropic Young’s modulus was transformed to the static one by
Wang correlation [36] for the soft rocks [1]:
� � = . � � − . (17)
where �is Young’s modulus in GPa. Isotropic static and dynamic Poisson’s ratio and Young’s modulus
logs for the whole Bakken interval are presented in Figures 5 and6. Anisotropic static Young’s modulus in
the direction of the vertical axis of rotational symmetry which was set to be the borehole axis is plotted in
Figure 7. The anisotropic Young’s modulus in the horizontal direction (the plane of symmetry) can be found
in Figure 7. The dynamic Young’s modulus was converted to the static value using Eq. (17). Dynamic
anisotropic Poisson’s ratio log was developed through Eqs. (15) and (16) in the horizontal and vertical
directions respectively; and the corresponding log is shown in Figure 8. The dynamic anisotropic Poisson’s
ratio was not transformed to the static one because the differences are not significant.

Fig. 5. Isotropic dynamic Poisson's ratio of the Bakken Formation.


Fig. 6. Isotropic static and dynamic Young's modulus of the Bakken Formation.

Fig. 7. Anisotropic horizontal and vertical static Young's modulus of the Bakken Formation.
Fig. 8. Anisotropic Poisson's ratio of the Bakken Formation.
It’s evident that the Young’s modulus values measured in the horizontal direction (in the plane of symmetry)
are much greater than that in the vertical direction (along the axis of symmetry) for both upper and lower
shales (Figure 7). This phenomenon reflects their VTI behavior. Accordingly the dynamic Poisson’s ratio
measured in the vertical direction was slightly larger than that in the horizontal direction for the upper and
lower shale members. These two evidences strongly demonstrate the anisotropic behavior of shales and the
presence of horizontal planes of weakness. This signifies that a VTI material is typically stiffer in the plane
of isotropy than in the direction of anisotropy [27]. In the Middle Bakken overlapping the horizontal and
vertical static Young’s modulus curves for most part of this interval stands for its isotropic behavior. This
represents that Young’s modulus and Poisson’s ratio are not dependent to the direction of measurement
(Figure 7). However a closer look to the lower section of the Middle Bakken (9760ft to 9780ft), the cross-
over of the vertical and horizontal static Young’s modulus curves (green and blue curves in Figure 7)
indicates the occurrence of anisotropy. This type of anisotropy which is known as Horizontal Transverse
Isotropy (HTI) makes the vertical Young’s modulus greater than the horizontal Young’s modulus. It was
interpreted as the existence of the vertical fractures in this specific zone. In order to confirm this
interpretation, the available drilling cores from this zone were analyzed; obviously the existence of vertical
fractures was confirmed (red circle in Figure 9). The presence of vertical fractures in this specific zone
makes it a perfect candidate for hydraulic fracturing operations. To evaluate the mechanical properties of
this specific section, horizontal transverse isotropy (HTI) assumptions should be applied. HTI is a type of
anisotropy which exists in a medium with vertical parallel fractures [5].Therefore the plane of symmetry
will be vertical and the axis of rotational symmetry is set to be horizontal.
Isotropic static shear modulus (Figure 10) and the anisotropic static shear modulus in the horizontal plane
(Figure 11) for the lower and upper members were derived from Eq. (18) [27]. This equation defines the
relationship between static Young’s modulus, Poisson’s ratio and the static shear modulus:

�= +�
(18)

where � is static Young’s modulus, � is static shear modulus in and is Poisson’s ratio.

Fig. 9. Core view from the well with the existence of a vertical fracture in the Middle Bakken.

Fig. 10. Isotropic static shear modulus of the Bakken Formation.


Anisotropic static shear modulus (Figure 11) of lower and upper VTI members along the axis of symmetry
was calculated using Eq. (19) [37]:

�� �ℎ �
� =� (19)
ℎ � + �� +��

where �ℎ ,� are static Young’s moduli in the plane of symmetry and in the direction of anisotropy
(vertical), respectively and is Poisson’s ratio in the direction of anisotropy.

Fig. 11. Anisotropic static shear modulus of the Bakken Formation in horizontal and vertical directions.
Tables 1 and 2 are the summary of the average values of the calculated static elastic parameters in the three
members of the Bakken Formation under anisotropic and isotropic assumptions respectively.

Table 1. Anisotropic elastic properties for the Bakken Formation.


Evert (GPa) Gvert (GPa) νvert
UB 5.01 2.53 0.18
MB 19.93 7.94 0.24
LB 5.83 2.77 0.16
EHorz(GPa) GHorz(GPa) νHorz
UB 8.44 3.62 0.16
MB 19.92 7.97 0.24
LB 10.09 4.08 0.15
Table2. Isotropic elastic parameters of the Bakken Formation.
G (GPa) E (GPa) ν

UB 1.84 4.67 0.26

MB 8.02 19.90 0.24

LB 2.07 5.46 0.24

It should be emphasized that due to the equal values of the average elastic parameters in horizontal and
vertical directions of MB (Table 1), it is considered isotropic in the numerical modeling (section 5).

5.GEOMECHANICAL MODELING (CODING)


The calculated in-situ stress, pore pressure and the calculated mechanical elastic parameters enabled the
geomechanical modeling (coding) of the Bakken Formation under isotropic and anisotropic approaches.
This task was performed using FLAC3D software a product of Itasca Consulting Group. Three main steps
were taken to accomplish the anisotropic and isotropic geomechanical modeling. The steps are presented
in Figure 12.

Fig 12. Flowchart of the geomechanical modeling.


The first step was to make the model and characterize this medium with assigning elastic parameters to an
undrilled Bakken Formation. The average elastic properties that were assigned to the model are from in
Tables 1 and 2 for anisotropic and isotropic cases respectively. In this beginning step the calculated isotropic
and anisotropic horizontal stress, vertical stress and the pore pressure were assigned to the isotropic and
anisotropic models correspondingly. The program was executed to bring the model to its mechanical
equilibrium state conditions. The model is set to behave elasto-plastically. The second step was to drill the
formation. In this step a cylindrical hole was made through the model to configure the drilled Bakken
Formation. In order to evaluate the results precisely around the borehole, it was decided to increase the
number of grids in this specific zone. The program was then executed for the second round; and new
mechanical equilibrium was achieved. In this step, the deformations and displacements in the formation
take place. Displacements in meter around the borehole in elastic isotropic and anisotropic input parameters
for UB, MB and UB-MB interface are shown in Figures 13 and 14.

Fig 13. Displacements in meters in UB, MB and UB-MB interface under isotropic assumptions.

Fig 14. Displacements in meters in UB, MB and UB-MB interface under anisotropic assumptions.
The third step was to run the model under the elasto-plastic Mohr-Coulomb failure criterion. The common
Mohr-Coulomb formulation is expressed as follows [27, 38]:
= + � tan � (20)

� = � + � tan ˚+ (21)

In Eq. (20), � is the normal stress and is the shear stress, is cohesion and � is the angle of internal
friction. In Eq. (21), � is the maximum principal stress and � is the minimum principal stress. � the
unconfined compressive strength is given by [1,27]:

� = tan ˚+ (22)

In order to model the medium under elasto-plastic Mohr-Coulomb failure the unconfined compressive
strength (UCS,� ) is required to be established. To achieve this goal, two separate methodologies were
tested. The first method is to acquire the compressional sonic velocity or slowness [1, 39], while the second
method uses Young’s modulus to predict the UCS [38].
In the first method, for the Lower and Upper Bakken Shale intervals Eq. (23) was applied to generate the
UCS profile [39]:
.
� = . � (23)
Eq. (23) is defined between compressional sonic velocity (km/s) and UCS (MPa) for shales [39]. For the
Middle Bakken, core lithology investigation like acid treatment was carried out and indicated that the
dominant lithology in MB is limestone; so a relevant correlation, Eq. (24) between compressional sonic
slowness (μsec/ft) and UCS for limestone was applied to calculate the UCS [40]:
⁄� .8
� = (24)

The second method uses the static Young’s modulus (GPa) to predict the UCS (MPa). Separate empirical
correlations were acquired: one for upper-lower shales, Eq. (25) [38] and the other one for middle limestone,
Eq. (26) [38]. Figure 15 is the UCS profile of the Bakken Formation generated from each of those methods.
.
� = . � (25)
.
� = . � (26)
Based on the fact that the generated UCS log from the Young’s modulus is more reliable than that from the
velocity data [41] and also to keep consistency of using static elastic data until the end of this study it was
decided to use Young’s modulus to create the unconfined compressive strength profile. The average UCS
values (Table 3) from Eqs. (25) and (26) were set as the input parameters in Mohr-Coulomb failure analysis
in the geomechanical modeling.
Table 3. Elastic parameters used for elasto-plastic Mohr-Coulomb failure criterion in the Bakken Formation.
Cohesion UCS from
�˚
(MPa) E (MPa)

UB 6.4 22.63 31

MB 15.63 63.11 37.3

LB 6.72 23.91 31.3


Fig. 15. Unconfined compressive strength from velocity data and Young's modulus for the Bakken Formation.

The average friction angle in Table 3 was calculated from two separate correlations: Eq. (27) for the lower-
upper members [42] and Eq. (28) for the middle member [38].
.
�= . � (27)
�− � ⁄µ ℎ + �ℎ − � ⁄µ
� = tan− � ℎ − �
(28)

In Eq. (27), � is the internal angle of friction in degree and � is the compressional wave velocity in km/s.
In Eq. (28), � is the gamma ray readings of the formation. � � and � ℎ� are the gamma ray
reference values for sand and shale layers, respectively. µ ℎ� and µ � are the internal friction
coefficients defined as µ = tan � for pure sand and pure shale. Eq. (28) principally uses gamma ray log for
the estimation of the internal friction angle. As gamma ray measures the amount of clay volume content of
a formation, Eq. (28) implies that a shalier rock possess a lower value of internal friction angle. It’s also
identified that internal friction angle decreases with increase of porosity, clay volume or the sum of both
properties within a rock [41]. It should be mentioned that relative low values of friction angle was
interpreted due to the presence of hydrocarbon in the middle member and high Total Organic Content
(TOC) of the upper and lower shale members.
The following figures display the final results of the anisotropic and isotropic geomechanical modeling.
These figures are the outcomes of the third step of the modeling process. Figures 16 and 17 depict the
variation of principal stresses in Pascal around the well in UB, MB and UB-MB interface under anisotropic
medium stress assumptions. Figures 18 and 19 show the variations of principal stresses (Pa) around the
well under the isotropic medium stress calculations. It should be noticed that due to the similar elastic
behavior of UB and LB, only the results from UB were displayed.

Fig. 16. Stresses (Pa) in UB, MB and UB-MB interface Fig.17. Stresses (Pa) in UB, MB and UB-MB interface
under anisotropic stress assumptions. under anisotropic stress assumptions.

Fig.18. Stresses (pa) in UB, MB and UB-MB interface Fig.19. Stresses (pa) in UB, MB and UB-MB interface
under isotropic stress assumptions under isotropic stress assumptions.

Plastic region around the borehole in the formation under isotropic UB, anisotropic UB and isotropic MB
are shown in Figures 20-22:

.
Fig. 20. Plastic region around the borehole in isotropic MB.

Fig. 21. Plastic region around the borehole in UB under anisotropic medium assumptions.

Fig.22. Plastic region around the borehole in UB under isotropic medium assumptions.
Figure 21 indicates that using anisotropic parameters such as elastic mechanical properties and the stresses
for the UB will cause the formation behave elastically after undergoing a shear failure in the vicinity of the
borehole (brown region). On the opposite, Figure 22 points out that under isotropic assumptions for the UB,
blocks will stay in the shear failure state (red region).
6. SUMMARY
Mechanical Earth Model (MEM) of the Bakken Formation was built. This task was performed by
calculating the elastic parameters and the effective stresses in the three different members of the Bakken
Formation. These parameters were generated by running cross-dipole sonic log in a vertical well in North
Dakota. The dynamic elastic moduli—Young’s modulus and Poisson’s ratio—were converted to static
ones through empirical correlations. These values were calculated under isotropic and anisotropic formulae.
It’s found that UB and LB are highly VTI with �ℎ >� and > ℎ . Although MB is assumed
to follow mechanical isotropy in most portions of the interval, in some deeper sections it shows � >
�ℎ and ℎ > . This warrants the HTI behavior of this portion which is related to the presence
of vertical fractures. The existence of the vertical fractures was also proved by core analysis.
In-situ stressws and pore pressure profile were generated under isotropic and anisotropic conditions. It was
well understood that UB-LB is highly overpressured due to the kerogen to hydrocarbon transformation in
low porosity-permeability shales. Middle Bakken was found to be less overpressured due to the migration
of hydrocarbon from UB and LB in to this member. In addition, anisotropic stress model was found to
better describe the stress conditions in these two layers.
Geomechanical modeling (coding) was performed under isotropic and anisotropic medium conditions in
three main steps to evaluate the deformations and horizontal stresses around the borehole. As a result of the
final step -- the Mohr-Coulomb failure analysis-- it was found that in anisotropic models for UB, the
formation will undergo a shear failure followed by an elastic behavior compared to the isotropic models
that show typically shear failure. Finally we found that anisotropic approximations will highly improve the
geomechanical modeling results and better represent the true behavior of the Bakken Formation.
It’s noteworthy that performing laboratory experiments on the core plugs is highly recommended to confirm
the calibration of dynamic to static transforms and to better represent the elastic parameters.

5. ACKNOWLEGEMENT
The authors would like to thank the following sponsors for financial support: US DOE through
contract of DE-FC26-08NT0005643 (Bakken Geomechanics), North Dakota Industry Commision
together with five industrial sponsors (Denbury Resources Inc., Hess Corporation, Marathon Oil
Company, St. Mary Land & Exploration Company, and Whiting Petroleum Corporation) under
contract NDIC-G015-031, and North Dakota Department of Commerce through UND’s Petroleum
Research, Education and Entrepreneurship Center of Excellence (PREEC). The authors would like
to express special appreciation to Jack Breig from Whiting Petroleum Corporation for providing
the data and permission to publish.

REFERENCE
1. Sayers, C., C. Russel and M. Pelorosso, J. Adachi, J. Pastor, V. Singh, K. Tagbor and P. Hooyman, 2009.
Determination of rock strength using advanced sonic log interpretation techniques. In Proceedings of the
SPE ATCE, New Orleans, 4 – 7 October, SPE 124161.
2. Sayers, C., S. Kisra, K. Tagbor, A. D. Taleghani and J. Adachi, 2007. Calibrating the mechanical properties
and in-situ stresses using acoustic radial profiles. In proceedings of the SPE ATCE, Anaheim, 11-14
November, SPE 110089.
3. Plumb, R. A., S. Edwards, G. Pidcock, D. Lee and B. Stace, 2000. The mechanical earth model and its
application to high risk well construction projects. In Proceedings of IADC/SPE Drilling Conference, New
Orleans, 23-25 February, IADC/SPE 59128.
4. Plumb, R.A., P. Hooyman, D. Vineengen, N. Dutta, G. Ritchie and K. Bennaceur, 2004. A new geomechanics
process reduces operational risks from exploration to production. In proceedings of the NARMS, Houston, 5-
9 June, ARMA/NARMS 04-616.
5. Tsvankin, I. 2005. Seismic signatures and analysis of reflection data in anisotropic media. 2nd ed. Elsevier
Science.
6. Higgings, S., S. Goodwin, Q. Donald, A. Donald, T. Bratton, and G. Tracy, 2008, Anisotropic stress models
improve completion design in the Baxter shale, In Proceedings of SPE ATCE, Denver, 21-24 September,
SPE 115736.
7. Thiercelin, M.J. and R.A. Plumb, 1994. Core based predictions of lithologic stress contrasts in east Texas
formations. J. SPE Formation Evaluation. 9: 14, 251-258.
8. Pistre, V., T. Kinoshita, T. Endo, K. Schilling, J.Pabon, B.Sinha, B.,T.Plona, T. Ikegamiand D. Johnson,
2005, A New modular wireline logging sonic tool for measurement of 3D (Azimuthal, radial and Axial)
formation acoustic properties, In Proceedings of SPWLA 46th annual logging symposium, New Orleans, June
26-29.
9. Walsh, J., B. Sinha, and A. Donald, 2006, Formation anisotropy parameters using borehole sonic data, In
Proceedings of SPWLA 74th annual logging symposium, Colorado Springs, 4-7June.
10. Sayers, C., 1994. The elastic anisotropy of shales, J. Geophys. Res.Solid Earth: 99, 767-774.
11. Hornby, B.E., 1994.The elastic properties of shales: PhD thesis, University of Cambridge.
12. Sondrgeld, C. and C. Rai, 2011. Elastic anisotropy of shales. TLE March 2011, 324- 331.
13. Lo, T., K. Coyner and N. Toksoz, 1985. Experimental determination of Berea Sandstone, Chicopee Shale
and Chelmsford Granite. J. Geophys. 51:1, 164-171.
14. Sayers, C., Z. Nagy, J. Adachi, V. Singh, K. Tagbor and P. Hooyman, 2009. Determination of in-situ stress
and rock strength using borehole acoustic data. In Proceedings of SEG 2009 International Exposition and
Annual Meeting, Houston, 23-30 October.
15. Sayers. M., 2006. An introduction to velocity-based pore pressure estimation. TLE December, 1496-1500.
16. Gutierz, M., N. Braunsdor and B. Couzens, 2006. Calibration and ranking of pore pressure prediction models.
TLE December, 1516-1523.
17. Eaton. B., 1972. Graphical method predicts geopressures worldwide. World Oil, 182, 51-56.
18. Mouchet, J. and A. Mitchell, 1989. Abnormal pressure while drilling. Elf Acquitane Manuels Techniques 2,
Boussens, France.
19. Ruth, P. and R. Hillis, 2000. Estimating pore pressure in the Cooper Basin, South Australia: Sonic log method
in an uplifted basin. J. Expl. Geophys. 31, 441-447.
20. Thomson, M and J.R., Willis, 1991. A reformulation of anisotropic poro-elasticity, J. Geophys. Res.:58, 612
21. Waters, G, R. Lewis and D. Bently, 2011. The effect of mechanical properties anisotropy in the generation
of hydraulic fractures in organic shales. In Proceedings of SPE ATCE, Denver, 30 Oct- 2 Nov, SPE 146776.
22. Mavko, G., T. Mukerji and J. Dvorkin, 1998. The rock physics handbook, 1st ed. Cambridge University Press,
Cambridge.
23. Wendt, A., M. Kongslien, B. Sinha, B. Vissapragada, A. Newton, E. Skomedal, L. Renlie and E. Pedersen,
2007. Enhanced mechanical earth modeling and wellbore stability calculations using advanced sonic
measurements—A case study of HPHT Kvitebjon Field in the Norwegian North Sea. In Proceedings of SPE
ATCE, Anaheim, 11-14 November, SPE 109662.
24. Ostadhassan, M., Z. Zeng, H. Jabbari, Anisotropy analysis in shales by acquiring advanced sonic data, 2012.
In Proceedings of AAPG ACE, Long Beach, CA, 22-25 April.
25. Nye, J. 1985. Physical properties of crystals. Oxford University Press.
26. Amadei, B. W. Savage and H. Swolfs, 1987. Gravitational stresses in anisotropic rock masses. Int. J Rock
Mech. and Mining sc. and Geomech. Abstracts, 24:1, 5-14.
27. Jeager, J., N. Cook, and Zimmerman, R., 2007. Fundamentals of rock mechanics. 4th ed. Blackwell
Publishing.
28. King, M. 1969. Static and dynamic moduli of rocks under pressure, in Somerton, W., and H. Ed. Rock
mechanics-theory and practices. In Proceedings of 11th symposium of rock mechanics. University of Calif.,
Berkeley, 329-351.
29. Sinha, B., B. Vissapragada, L. Renlie and S. Tysse, 2006. Radial profiling of the three formation shear moduli
and its application to well completions. J Geophys.,71:6, E65-E77.
30. Sinha, B., B. Vissapragada, L. Renlie and E. Skomedal, 2006. Horizontal stress magnitude estimation using
the three shear moduli—A Norwegian Sea case study. In Proceedings of SPE ATCE, San Antonio, 24-27
September, SPE 103079.
31. Sinha, B., Vissapragada, B., Wendt, A., Kongslien, M., Eser, H., Skomedal, E., Renile, L. and Pedersen, E.,
2007. Estimation of formation stresses using radial variation of three shear moduli- a case study from a high-
pressure, high-temperature reservoir in Norwegian continental shelf, In Proceedings of SPE ATCE, Anaheim,
11-14 November, SPE 109842.
32. Price, L. and LeFever, J., 1994, Dysfunctionalism in Williston Basin: The Mid-Madison/Bakken petroleum
system, Bulletin of Canadian Petroleum Geology, 42, 187-218.
33. Traugott, B., 1997. Pore/Fracture determination in deep water. World Oil, 1997.
34. Bowers, G., 1995. Pore Pressure Estimation from Velocity Data: Accounting for Overpressure Mechanisms
besides Under-compaction. J SPE Drilling and Compl.,10:2, 89-95.
35. Fjar, A., R. Holt, A. Raaen, R. Risnes and P. Horsud, 1992, Petroleum related rock mechanics, Elsevier
Publishing.
36. Wang, Z., 2000. Dynamic vs. static properties of reservoir rocks. In Wang, Z., A. Nur and A. Eds, Seismic
and acoustic velocities in reservoir rocks. Volume 3: Recent developments, Published by SEG.
37. Lekhnitskii, S., 1981. Theory of elasticity of an anisotropic body. Moscow: Mir Publishers.
38. Chang, C., M. Zoback and A. Khaksar, 2006. Empirical relations between rock strength and physical
properties in sedimentary rocks, J. Pet. Sc. and Eng., 51, 223-237.
39. Horsud, P., 2001. Estimating mechanical properties of shales from empirical correlations. J. SPE Drilling
and Compl.,68-73, SPE 56017.
40. Millitzer, H. and R. Stoll, 1973. Einige Beitrageder geophysics zur primadatenerfassung im Bergbau, Neue
Bergbautecknik, Lipzig 3, 21-25.
41. Plumb, R., 1994. Influence of composition and texture on the failure properties of clastic rocks. In Proceeding
Eurock SPE/ISRM Rock Mechanics in Petroleum Engineering Conference, Delft, 29-31 August,SPE 28022.
42. Odunlami, T., H. Soroush, P. Kalathingal and J. Somerville, 2011. Log based rock property evaluation—A
new capability in a specialized log data management platform. In Proceedings SPE/DGS Saudi Arabia
Technical symposium and Exhibition, Al khobar, 15-18 May, SPE 149050.

You might also like