You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/324538159

Geomechanical wellbore stability analysis for the reservoir section in J-NC186


oil field

Conference Paper · March 2018

CITATIONS READS

0 3,835

4 authors, including:

Tareq Albukhari
University of Tripoli
1 PUBLICATION 0 CITATIONS

SEE PROFILE

All content following this page was uploaded by Tareq Albukhari on 16 April 2018.

The user has requested enhancement of the downloaded file.


Geomechanical wellbore stability analysis for the reservoir section in J-
NC186 oil field

TAREQ. M. ALBUKHARI1, GHAITH K. BESHISH2, MOFTAH M. ABOUZBEDA3, ABDASALAM


MADI4

1
M. Sc. student of Mining Geomechanics, University of Tripoli, Tripoli, Libya
2
Prof. Rock Fracturing, University of Tripoli, Tripoli, Libya
3
Prof. Geotechnical Engineering, University of Tripoli, Tripoli, Libya
4
Lecturer. Rock Engineering, University of Tripoli, Tripoli, Libya

Corresponding Author Email: t.albukhari@uot.edu.ly

ABSTRACT: Wellbore instability has been the primary cause of financial losses in boreholes in the drilling
industry. Wellbore instabilities are responsible for 10-20% of the total drilling cost. It has been estimated that
wellbore instability causes an annual economic loss of $ 1 - 6 billion in the oil industry across the world. One
type of these wellbore instabilities was a whole enlargement problem encountered in the J-NC186 Oil field
southern of Libya. This paper examines one side to solve this problem. so a geomechanical wellbore stability
analysis was implemented using Schlumberger Techlog software to effectively plan the future drilling opera-
tions in the field to maximize the drilling margin for the future wells drilled, and to optimize the future field
development. 1-D geomechanical models show that the horizontal wells are more stable than the vertical
ones. The models also show that drilling to the orientation of east-west can minimize the wellbore instabili-
ties. The models suggested the optimum mud weight window to drilling the reservoir section safely, and rep-
resented the results using stereographic projection. The outcomes of this study can be utilized for further field
developments for enhancement of the hydrocarbon production (e.g. hydraulic fracturing, open-hole comple-
tions, and enhanced hydrocarbon recovery).

1 INTRODUCTION rounding the hole must take the load that was previ-
ously taken by the removed rock. As a result, the in
Wellbore stability is a term used in the Oil and gas situ stresses are significantly modified near the
industry to describe usable condition of the borehole borehole wall. This is presented by a production of
during drilling operations. A usable hole must suc- an increase in stress around the wall of the hole, that
cessfully accommodate logging or any open-hole is, a stress concentration. The stress concentration
evaluation, casing run and any other drilling activi- can lead to rock failure of the borehole wall, depend-
ties. ing up on the existing rock strength. The basic prob-
The Oil and gas industry continues to fight bore- lem is to know, and to be able to predict, the reaction
hole problems. The problems include: hole collapse, of the rock to the altered mechanical loading. This is
tight holes, stuck pipes, poor hole cleaning, hole en- a classical, though not very easy rock mechanics
largement, plastic flow, fracture, and lost circulation. problem (Al-Ajmi, 2006).
Most of these borehole problems that drive up drill- In order to avoid borehole failure, drilling engi-
ing costs are related to wellbore stability. These neers should adjust the stress concentration properly
problems are mainly caused by the imbalance creat- through altering the applied internal wellbore pres-
ed between the rock strengths and induced stresses sure (i.e., mud pressure) and the orientation of the
after wellbore drilling. borehole with respect to the in situ stresses. In gen-
Wellbore stability is dominated by the in situ eral, the possible alteration of the borehole orienta-
stress system. When a well is drilled, the rock sur- tion is limited. It is therefore obvious that wellbore
instability could be prevented by mainly adjusting log, and porosity log), drilling reports and formation
the mud pressure. Traditionally, the mud pressure is tests (Jimenez et.al.,2007).
designed to inhibit flow of the pore fluid into the
well, regardless of the rock strength and the field
stresses. In practice, the minimum safe overbalance
pressure (well pressure − pore pressure) of typically
100-200 psi, or a mud density of 0.3 to 0.5 lb/gal
over the formation pore pressure, is maintained (e.g.,
French and McLean, 1992; Awal et al., 2001). This
may represent no problem in competent rocks, but
could result in mechanical instability in weak rocks.
In general, the mud pressure required to support the
borehole wall is greater than that required to balance
and contain fluids, due to the in situ stresses which
are greater than the formation pressure (Al-Ajmi,
2006). Figure 2. Wellbore stability model building up (Jimenz
Hole enlargement or collapse due to brittle rock et.al.2007)
failure of the wall. Symptoms of this condition are
poor cementing, difficulties with logging response
and log interpretation, and poor directional control. Primary assumptions during this stage are the validi-
Poor cementing of the casing could lead to problems ty of linear elastic theory for porous media, the uni-
for perforating, sand control, production and stimu- formity of rock formations, the representativity of
lation. Furthermore, when the hole starts to collapse, formation test and well logging. Based on these, the
small pieces of the formation may settle around the geomechanical model building up consists of the
drill string and pack off the annulus (i.e., hole pack- next items:
off), while medium to large pieces fall into the bore- 1- To choose the correlation wells.
hole and might jam the drill string (i.e., hole bridg- 2- To determine the lithologies at wells.
ing). These may prevent pulling the string out of the 3- To assess shale content at different lithologies
hole (i.e., stuck pipe), and so the planned operations (Vshale).
are suspended. Stuck pipe problems due to borehole 4- To calculate shear and compressive acoustic
collapse is illustrated in Figure 1 (Al-Ajmi, 2006). velocity on wells.
5- To compute elastic moduli, rock strength and
Mohr-Coulomb failure envelope.
6- To get well in situ stresses (overburden stress,
minimum horizontal stress, maximum hori-
zontal stress, orientation of horizontal stresses
and pore pressure).
7- To evaluate the mudweight window, which
has three parameters:
I) minimum mudweight,
II) maximum mudweight and,
III) optimum mudweight while drilling
(Jimenez et.al.2007).

3 CASE STUDY
Figure 1. Stuck pipe problem due to borehole collapse. (a) Hole
pack-off. (b) Hole bridge.
In this paper we’ve analyzed wellbore stability for
wells that encountered a hole-enlargement within J-
NC186 Oil field in southern of Libya which be-
2 GEOMECHANICAL MODEL BUILDING UP
longed to Repsol Exploration Murzuq S.A (REM-
SA).
The methodology to build up the model based on the
correlation wells available in (Figure 2) which are
sited next to the studied well and have similar be-3.1 Field Overview
havior respect to it. Further- more, correlation wells
J-Field is located within the 3D seismic cube in
must have as much information as possible to obtain
block NC186 which is located on the western north
the geomechanical model. This means well logging
margin of Murzuq Basin (Figure 3). The Murzuq
(density log, sonic log, gamma ray log, resistivity

2
Basin is one of the several intracratonic basins of * Initializations of well logs are defined as follow:
North Africa. It covers an area of over 350000 km2, CAL: Caliper, DT: Sonoc ∆T, GR: Gamma ray, SP:
and its sedimentary fill reaches a maximum thick- Spontaneous Potential, RES: Resistiivity, Neutron,
ness in the basin centre of about 4000m., comprising RHOB: Density, CKH: Core Permeability, CPOR:
a predominantly marine Palaeozoic section overlaid Core Porosity, TEMP: Temperature, FMI: For-
by a continental Mesozoic section. The present bor- mation microimage, RFT: Repeated Formation Test.
ders of the Murzuq basin are delimited by tectonic
uplifts. The present day structural framework is a
3.3 In-situ stress determination
consequence of the cumulative tectonic movements
ranging from Palaeozoic through to Tertiary times. In-situ stress magnitudes can have a large effect on
The main tectonic events are related to the Caledo- wellbore stability for all wellbore orientations. For
nian (late Silurian-early Devonian), Hercynian (end this reason, it is important to model the stress condi-
Carboniferous-Permian) and Alpine (early Tertiary) tions over the entire length of the interested section
orogenies. Although the present-day structuring is in order to effectively predict when wellbore failure
related mostly to Hercynian compressional move- will occur.
ments and the later tectonic activity, the regional lin-
eaments (NW-SE) are probably related to late Pre- 3.3.1 Vertical Stress (Overburden Pressure), Sv
cambrian Pan-African fault systems, which largely The magnitude of Sv is equivalent to integration of
controlled the early Palaeozoic structural and depo- rock densities from the surface to the depth of inter-
sitional evolution of this basin. est, z. In other words,
J-NC186
𝑧
𝑆𝑣 = ∫0 𝜌(𝑧)𝑔𝑑𝑧 ≈ 𝜌̅ 𝑔𝑧 (1)

Where ρ(z) is the density as a function of depth, g is


gravitational acceleration and ρ is the mean overbur-
den density.
Figure 4.3 shows the log of bulk density from one
of the exploratory wells. Here, the depth indicated
by the dashed lines represents the area between the
surface and the beginning of the bulk density logged
Figure 3. J-Field location Map

3.2 Geomechanical Data from Correlation wells


The geomechanical model developed for this inves-
tigation includes data collected from four reference
wells located in this field. Table 1 shows an over-
view of the data from these wells.

Well
Well logs and tests*
Name
HCAL - DT4P - DT4S - GR - SP -
DEEPRES - MICRORES - SHALRES -
J02
TNPH - RHOB - CKH - CPOR - TEMP -
FMI - RFT
CALI - DTCO - DT4S - SP - GR - Shall -
J04 RES - DEEP - RES - MICRO - RHOZ -
TNPH - CKH - CPOR - TEMP - FMI - RFT
HCAL - GR - SP - DT4S - DTCO - MI-
J21i CRORES - SHALLRES - DEEPRES -
TNPH - RHOZ - TEMP - FMI - RFT
CAL - DT - SP - GR - DEEPRES - MI-
J24i CRORES - SALLOWRES - TNPH - RHOZ
- TEMP - RFT
Table 1. Overview of data collected from exploratory wells in
the study field interval.
Figure 4. Vertical Stress log and assumed stress for shallow
formations lacking log data.
Bulk density logs from the exploratory wells 3.4 Rock Deformation Properties
these showed densities above 2.35 g/cc as shallow as Acquisition of formation mechanical properties,
1430 ft. Therefore, for this analysis, the average such as Poisson’s ratio and Young’s modulus is of
density for shallow zones, where log data was not
great importance in conducting wellbore-stability
available, was assumed to be the same as the aver- analysis. Typically, extensive laboratory experi-
age of the first 300 ft where results from bulk densi- ments need to be conducted to acquire these for-
ty logs where available. As such, the bulk density mation properties. However, since no core was re-
was assumed to be 2.3 g/cc for the zone from the leased for this study at the time of the completion of
surface to the beginning of the logged interval. From this paper, all the parameters were obtained utilizing
this assumption of shallow formation density, the well-log data and appropriate empirical approaches.
overburden/vertical stress gradient was estimated to
be ≈ 1 psi/ft. It is possible that this method may 3.4.1 Young’s Modulus and Poisson’s Ratio
overestimate the bulk density, in turn overestimating Using the concept of elastic moduli equations de-
the overburden stress as well. scribed by Clark (1966), the dynamic compressive
modulus (M), shear modulus (Gs), bulk modulus (K)
3.3.2 Formation Pore Pressure, PP along with the dynamic Young modulus ( Ed) and
Several well testing methods exist for determining Poisson’s ratio (ϑ) have been calculated. The results
the pore pressure within a formation. The investiga- of the calculations are shown in Figure 6. The elastic
tion of these techniques is beyond the scope of this modulus equations used are shown in Equations Eq.
paper. The pore pressure measurements for this 2 Eq. 3
study were taken from operator’s pore pressure tests
for each well named Repeated Formation Test (RFT) 𝑉𝑝 2
(Figure 5), which is done by a tool that isolates small (( ) −2)
𝑉𝑠
sections of the reservoir and removes fluid samples 𝑣= 𝑉𝑝 2
(2)
2∗(( ) −1)
and takes pressures. It is important to note that the 𝑉𝑠
pore pressure gradient is an extremely important in-
put in the wellbore stability calculations, and if the 4
pore pressure prognosis is changed, the stability es- 𝐾 = 𝜌 ∗ (𝑉𝑝2 − 3 ∗ 𝑉𝑠2 ) (3)
timates will be affected.
𝐸 = 3 ∗ 𝐾 ∗ (1 − 2 ∗ 𝑣) (4)

Static elastic moduli are preferred over moduli ob-


tained using dynamic approaches (Eissa and Kazi,
1988). This preference is based on the theory of the
pseudo-static behavior of rock. Therefore, many
studies were conducted to correlate these two modu-
li. Analysis of the literature of a dynamic to static
modulus correlation was briefly discussed in the fol-
lowing paragraph.
 The dynamic Young’s modulus is typically
greater than the static Young’s modulus.
 The dynamic Poisson’s ratio is generally
lower than the static one, and finding a
correlation between these two is a chal-
lenging task due to the typically lower re-
solution of lateral deformation mea-
surements in calculating the radial strains.
 The ratio between the dynamic and static
moduli approaches typically to unity as
the confining pressure increases.
The correlation used to determine the static
Young’s modulus in this study was Morales’s corre-
lation (Morales and Marcinew 1993). The correla-
tion is retrieved based on several high-permeability
Figure 5. Pore Pressure measurements by RFT sandstones ranging from very fine-grained to coarse-
grained from several fields - Kuparuk, Alaska;
Elmworth and Mitsue, Canada; Punta Benitez and
Boqueron, Venzuela. This is applicable to all three

4
porosity groups: consolidated (10%-15%), moder- The Coates-Denoo algorithm was introduced in
ately consolidated (15%-25%), and weakly consoli- the late 1960’s and is based on the Deere and Mil-
dated (>25%).For porosity 0%-10%, the equation is ler’s sandstone and shale data (1963). It suggests
not defined (Schlumerger Techlog manual, 2015). that you can predict the strength of shaly sands by
combining the sandstone and shale correlations of
Deere and Miller (Schlumerger Techlog manual,
2015).

3.5.2 Tensile Formation Strength


Tensile strength of rock is an important parameter in
calculating and constraining the minimum and max-
imum horizontal stresses. Typically, for unconsoli-
dated formations tensile strength is assumed to be
zero. That assumption is based on the disturbance of
an intact condition of rock by bit penetration
(kadyrov, 2012). However, in highly compacted and
strong formations under high in-situ stresses, as in
our study field, the tensile strength might be non-
zero. To obtain a reliable tensile strength value, Bra-
zilian laboratory measurements should be conduct-
ed. In the absence of the core measurement data, the
tensile strength is usually estimated at 10-12% of the
UCS for all facies. This approach might be some-
what misleading considering that tensile strength is
typically impacted by the lithology type, compaction
Figure 6. Formation Deformation properties with Gamma level, lamination orientation, and presence of micro-
ray as reference cracks (Hobbs, 1964). The relationship between
UCS and tensile strength for different rock types
was reported by Hobbs (1964).
3.5 Rock Strength Properties Most of massive rocks such as limestone and
Preliminary rock strength properties were deter- sandstone have higher ratio of the tensile to UCS
mined through the use of well log analysis. than in the laminated rocks. Therefore, it is reasona-
ble to use 0.15 ratio for massive and strong rocks
(dolomite, limestone, and sandstone) and ratio of
3.5.1 Uniaxial Compressive Strength 0.05 for shale formations in our study. A lower
Uniaxial compressive strength (UCS) is a critical pa- boundary of the SHmax and Shmin was estimated when
rameter in both constraining the maximum horizon- the tensile strength at the wellbore was taken as zero
tal stress magnitude and obtaining an appropriate (kadyrov, 2012).
rock-failure envelope for the formation of interest.
Typically, this parameter is obtained from laboratory 3.5.3 Angle of Internal Friction
core measurements under uniaxial loading stress Relating the angle of internal friction to petrophysi-
conditions. Core samples should represent various cal parameters is quite limited. Because even weak
facies. Then, the obtained UCS is usually correlated formations can have a high friction angle (Zoback,
to the different petrophysical and geomechanical pa- 2010), and the most appropriate approach is to ob-
rameters such as compressional velocity, porosity, tain friction angle is when a uniaxial compressive
clay volume, Young’s modulus. This type of correla- strength test is conducted. Again, in the absence of
tion may enhance a prediction of UCS from the core measurements in our study field, internal fric-
well-log derived properties without expensive and tion angle was obtained using empirical correlation
time-consuming laboratory experiments (kadyrov, published by Schlumberger Techlog software de-
2012). signers (kadyrov, 2012).
There is a high possibility that the empirically ob- This mothod maps Gamma Ray to Friction Angle
tained UCS values would contain some errors com- with a linear correlation. A cutoff is applied to Fric-
pared to the actual one. Definitely, further validation tion Angle (Figure 7). With default parameters, GR
and calibration using laboratory measurement data 120 gAPI is mapped to FANG 20 dega and GR 40
should be applied. In this study, since core samples gAPI is mapped to FANG 35 dega. If the calculated
are not available yet, the empirical correlations of FANG is less than 15 dega, it is forced to 15 dega. If
Coates-Denoo were used to constrain UCS in J-186 it is greater than 40 dega, it is forced to 40 dega. One
Field. also can change the default parameters (Schlumerger
Techlog manual, 2015).
well in order to effectively predict when wellbore
failure will occur.

3.6.1 Orientations of the Principle Horizontal


Stresses
One of the important factors affecting wellbore fail-
ure is an orientation of the principle horizontal
stresses (Barton et al. 1997). According to Barton et
al. (1997), breakouts will be observed if the hoop
stress is most compressive at the direction of the
minimum horizontal stress and when the stress con-
centration overwhelms the rock strength. By con-
trast, the circumferential stress has the least com-
pression at the orientation of the maximum principle
Figure 7. Determining Friction Angle from Gamma ray data
horizontal stress, causing the drilling-induced frac-
tures. Therefore, the orientation of the wellbore
breakouts and tensile fractures is the clear indication
3.5.4 Biot’s Coefficient
of the horizontal stress azimuths, assuming that the
Biot’s coefficient, which is also a stress dependent
well is vertical. This approach was utilized using the
parameter, was used in a numerical wellbore-
available FMI log data for the Well J2-NC186 in the
stability model. Since no laboratory measurement
8 ½ inch section. The interpreted interval is 5102 –
data for Biot’s coefficient was available for this
4254 ft. The FMI log was scanned for visible drill-
study, a possible range of Biot’s coefficient for this
ing induced tensile fractures and breakouts. Figure 7
study was taken as 0.6-1 with the mean value of 0.8.
shows a sample of these breakouts appeared between
Analysis of the planned XLOT analysis in the study
(4620 - 4630).
field might be helpful in further constraining this pa-
rameter (kadyrov, 2012).

Figure 9. FMI log with breakouts and calliper fluctuation.

Figure 8. Formation Deformation properties with Gamma Also, the reports obtained from the service company
ray as reference interpreted the orientation as following:
“Borehole enlargement occurred in the direction N-S
over the interval between 4650 – 4450 ft. Drilling
induced fractures are developed at places. There-
3.6 Horizontal Stresses fore, the maximum horizontal stress direction can be
In-situ stress magnitudes and orientation can have a estimated as perpendicular to the direction of
large effect on wellbore stability for all wellbore ori- breakout i.e. E-W.”
entations. For this reason, it is important to model
the stress conditions over the desired length of the

6
3.6.2 Magnitudes of the Minimum and Maximum
Horizontal Stresses (𝑈𝐶𝑆+2𝑃𝑝 +∆𝑃+𝑇° )−𝑆ℎ𝑚𝑖𝑛 (1+2 𝑐𝑜𝑠 2𝜃𝑏 )
𝑆𝐻𝑚𝑎𝑥 = (7)
Magnitudes of the minimum horizontal stress are es- 1−2 𝑐𝑜𝑠 2𝜃𝑏
sential parameters in the determination of a stress
regime. According to Oort et al. (2001), the most ac- Where 𝟐𝜽𝒃 equals π-wb . wb is the width of a
curate value of the minimum horizontal stress corre- breakout in degrees. The UCS is also in psi.
sponds to the fracture closure pressure during the
Extended Leak-off Tests (XLOT). Since there are no Moos and Zoback (1990) described in detail the ap-
available XLOT data in J-NC186 Field (except one proach for constraining the uncertainty of the maxi-
hydrofrac test (Figure 8.)), the Eaton method (Equa- mum horizontal stress magnitude obtained using
tion 5) was used to calculate the extended fracture Equations 6. and 7. The approach is based on the de-
propagation pressure (Mitchell, 1995), that was as- termination of allowable stress conditions (stress
sumed to be equal to the magnitude of the minimum polygon) in which tensile failures and/or breakouts
horizontal stress. can occur. The parameters required to build the
stress polygon are as follows:
𝜈  Coefficient of internal friction
𝑆ℎ𝑚𝑖𝑛 = (1−𝜈) (𝑆𝑣 − 𝑃𝑝 ) + 𝑃𝑝 (5)
 Pore pressure
 Mud weight
 Magnitudes of overburden stress
 Magnitude of minimum horizontal stress
In this study most of the calculated values from
Equations 6. and 7. are in the range of constrained
using the stress polygon approach. The uncertainty
of the magnitude would be less in intervals where
both breakouts and tensile fractures occurred.
As Equation 7. is sensitive to a breakout width val-
ue, it is critical to have borehole image data with a
good quality to successfully implement this ap-
proach. The misinterpretation of the breakout width
by +/- 20° might result in the fluctuation of the mag-
nitudes by 1 – 1.5 g/cc.

Figure 10. hydrofrac job data at depth between 4919-4975 ft.

Unless FMI log and multi-caliper log data are avail-


able, the determination of the maximum horizontal
stress magnitude is often a highly challenging task
(Kadyrov, 2012). Peska and Zoback (1995) de-
scribed two methods of utilizing the borehole image
and caliper data to obtain magnitudes of the SHmax.
The methods are based on the identification and ap-
plication of wellbore tensile cracks and breakouts
using the following equations (Zoback, 2010). Equa-
tion 6 was utilized to constrain the magnitude of the
using wellbore tensile fractures:
Figure 11. Stress polygon at the depth 4919 ft with 3 differ-
ent stress regime areas (NF=Normal Faulting, SS=Strike-Slip
𝑆𝐻𝑚𝑎𝑥 = 3𝑆ℎ𝑚𝑖𝑛 − 2𝑃𝑃 − ∆𝑃 − 𝑇0 (6)
and RF=Reverse Faulting).

Where ∆P is the difference between the wellbore


pressure (swab pressure) and pore pressure (PP), and Wellbore breakout and tensile failure events were
T0 = Tensile Strength (in psi). A width of breakouts plotted as illustrated in Figure 12. It is evident that
were used in Equation 6. to constrain the magnitude the stress regime in the entire interval is in the Nor-
of the SHmax. mal/strike-slip regime (Sv=SHmax>Shmin).
Tensile wellbore-failure events are additional val- gime, where we can get both strike-slip faults and
idations of this stress regime interpretation. It is in- normal faults active at the same exactly at the same
teresting to note that It has been suggested, on the time (Zoback 2014).
basis of Landsat data, that basement in the Murzuq
Basin shows evidence of strike-slip faulting formed
as a result of intraplate stresses, which exerted a
controlling influence on later tectonic trends (Hallet
2002).

Figure 12. Stress polygon at the depth 4919 ft: Constraining


the magnitude of SHmax.
Figure 13. Magnitude of pore pressure and principle stress-
es as a function of true vertical depth. It’s clear that the stresses
It’s clear that the magnitude of SHmax can’t be lower at basement of the reservoir tend to be more Strike-Slip fault-
than 4485 psi because, the maximum horizontal ing
stress can’t be lower the minimum horizontal stress,
which we get its value from hydrofrac if that provid-
er or by using Eaton equation (Equation 5). Also we 3.7 Geomechanical Model Results and Discussion
can put another pounds on magnitude of SHmax from The Mohr-Coulomb failure criteria have been uti-
the observations of breakout and DITF from FMI lized in the geomechanical model to evaluate the
logs with their equations (Equation 6 and 7). We critical mud weights to prevent breakouts and tensile
recognize that at 4919 ft we have no breakouts, so fractures in vertical and arbitrarily oriented well-
the magnitude of SHmax can’t exceed the value of bores.
9197 psi, we can get this value either by using Eq. 6. The geomechanical model was applied to the res-
or from the graph from the intersection of Shmin line ervoir interval (Hawaz Formation), from its top 4236
(the green line) with the DITF line (the orange line) . ft to depth 5100 ft, in a hole with diameter of 8 1/2”,
Also as we have no breakouts at that depth, the at the vertical Well J2-NC186, which was drilled us-
magnitude of SHmax can’t reach the value of 6077 psi ing a mud weight of 9.2 ppg.
(see the intersection of Shmin line (the green line) Different depths (cases) were selected within the
with the breakout (the blue line). So we can say the interval of interest (see appendix A&B). Cases 1
magnitude of SHmax is: represent TVD at which wellbore breakouts oc-
curred. Case 2 corresponds to the depth where no
4485 < 𝑆𝐻𝑚𝑎𝑥 < 6077 or 5281 ± 795 breakouts were observed.
Also different drilling scenarios was proposed to
It’s interesting to know that the vertical stress at this drill the reservoir, based on drilling deviated well in
depth is 5260 psi, so its magnitude is comparable the different azimuth and different orientations either
magnitude of maximum horizontal stress which is (see appendix A&B). We used the stereographic
about 5281 psi. This state of stress is extremely projection to represent the proposed optimum mud
common in the earth and it was seen in many reser-
window related to those scenarios to predict the op-
voirs. They call it Normal/Strike-Slip faulting re- timum mud weight window (see appendix A). At the

8
end of this chapter we represented the suggested the might reach formation breakdown pressure at 14
safe mud window, optimum well deviation and azi- ppg.
muth to drill the entire reservoir section safely, and
preventing any wellbore instability problems (see
appendix B).

3.7.1 Breakout interval


FMI log of the well J2-NC186 showed that this sec-
tion of the well represented a severe of breakouts
from 4450 ft to 4710 ft, due to high stress concentra-
tion around the well, where the difference between
SHmax and Shmin reached about 5950 psi. Fundamen-
tally, the variation of stress around the wellbore wall
amplifies the far-field stress concentration by a fac-
tor of 4 (Zoback, 2010) as follow:
At the point of minimum compression around the
wellbore (i.e. parallel to Shmin) at θ = 0◦, 180◦,
Kirsch equations reduce to:
Figure 15. Optimum mud window, with insitu stresses and rock
𝑚𝑖𝑛
𝜎𝜃𝜃 = 3𝑆ℎ𝑚𝑖𝑛 − 𝑆𝐻𝑚𝑎𝑥 − 𝑃° (8) properties data, and Gamma ray as shale flag.

Whereas at the point of maximum stress concentra-


tion around the wellbore (i.e. parallel to SHmax) at θ = We also noted that high SHmax is combined with
90◦, 270◦, shale streaks, like that at 4335 ft to 4380 ft, which
also has a relative high compression strength.
𝑚𝑎𝑥
𝜎𝜃𝜃 = 3𝑆𝐻𝑚𝑎𝑥 − 𝑆ℎ𝑚𝑖𝑛 − 𝑃° (9)
3.7.2 No breakout observed intervals
The other section of the well seemed to be stable if
Such that the difference between the two is: the mud weight doesn’t go under the lower bound, 8
ppg for breakouts (or shear failure in yellow) and 7.8
𝑚𝑎𝑥 𝑚𝑖𝑛
𝜎𝜃𝜃 − 𝜎𝜃𝜃 = 4(𝑆𝐻𝑚𝑎𝑥 − 𝑆ℎ𝑚𝑖𝑛 ) (10) ppg for kick (which in fact it increases gradually as
the pore pressure increase with depth) (which is rep-
Which corresponds to the amplitude of the sinusoi- resented in grey color).
dal variation of hoop stress around the wellbore For the upper bound of the mud weight, which
shown in Figure 14 and helps explain why observa-
represents the beginning of mud losses, or in other
tions of wellbore failures so effectively indicate far-
field stress directions. words when we begin hydrofrac the formation,
which indicates that we reach the magnitude of least
principal stress, which is generally Shmin (in some
depths Sv) and represented in light purple, especially
in the section where we have normal/strike-slip
faulting regime (SHmax=Sv) at the section from 4720
ft to 4920 ft, where we might get a mud losses if the
weight reach 15 ppg, in this section the breakdown
pressure less severe compared to the breakout sec-
tion, where we start to have problematic mud weight
if the mud weight reached 14 ppg.
In the upper and lower sections of the well we
have no drilling problems problem, so that we have
a wide mud window from 8 ppg to 17 ppg, especial-
ly in the lower part of the well when we have a rela-
Figure 14. Variation of effective principal stresses, σθθ , σrr tive high compressive strength sandstone.
and σzz around a vertical wellbore as a function of azimuth.

We can see that theses stress concentrations makes 3.7.3 The effect of deviation and azimuth in well-
the safe mud weight to prevent breakouts increased bore stability
to about 11 ppg compared to the other sections. Also Stereographic projection can considered to be the
we can see that if the mud weight increased, we best tool to represent the effect of deviation and az-
imuth in wellbore stability. Thus we took selected
depths where we had some abnormal conditions es- The conclusions of this study are as follows:
pecially in areas where we have stress perturbations 1- The stress regime in the reservoir section of J-
and high compression/extension environments. NC186 Field was found to be mainly strike-slip,
which is in agreement with the tectonic history of
3.7.4 Stress perturbations the region.
When we took a quick review to the stress environ- 2- The geomechanical models shows that the hori-
ment in the reservoir section, we recognize that we zontal wells in the interested section are more
stable than the vertical ones, that’s because of the
have some stress perturbation and transition zones,
strike-slip faulting stress regime.
this may be due to old earth crust movements and 3- The models shows that the drilling to the orienta-
activities, so we can divide the stresses in this sec- tion of east-west in tge interested section (i.e. pa-
tion like in the table 2. rallel to the maximum horizontal stress) can mi-
nimize the wellbore instabilities.
Depth Stress regime 4- The wellbore stability model developed in this
study for the J-NC186 can be potentially applied
Normal/Strike Slip to other fields in Murzug Basin using a similar
4255 - 4317
Faulting approach which might be adjusted to the particu-
4317 - 4428 Strike-Slip Faulting lar field specifications and requirements.
Strike-Slip / Reverse
4428 - 4450 Even though we are confident with the obtained in-
Faulting
4450 - 4460 Strike-Slip Faulting put data for this wellbore stability analysis, laborato-
Normal / Strike Slip ry core measurements to obtain geomechanical for-
4460 - 4480 mation properties for different facies in the J-NC186
Faulting
Field would help increase the confidence level of the
4480 - 4523 Strike Slip Faulting obtained study results.
4523 - 4532 Reverse Faulting The recommendations for future work from this
Normal/Strike-Slip study are as follows:
4532 - 4601
Faulting 1- The geomechanical formation properties should be
4601 - 4650 Strike-Slip Faulting obtained under the true-triaxial core measurements
Normal/Strike-Slip for various facies of the J-NC186 Field.
4650 - 4750 2- The obtained laboratory geomechanical parameters
Faulting
should be correlated to the petrophysical parameters
4750 - 4832 Normal Faulting to derive these geomechanical parameters from well
Normal/Strike-Slip logs and to reduce costly geomechanical laboratory
4832 - 4906
Faulting measurements in the life cycle of the field.
4906 - 5105 Strike-Slip Faulting 3- Extended leak-off tests should be conducted at va-
Table 2. Stress regimes in the reservoir section rious intervals to calibrate the calculated magnitude
of the minimum horizontal stress.
We took the middle of each interval as repre- 4- Probabilistic analysis techniques such as Monte-
Carlo Quantitative Risk Assessment (QRA) should
sentative depth to make sensitive analyze of well-
be done to reduce degree of uncertainties in rock
bore stability. properties and in-situ stresses obtained from geophy-
Sensitivity analysis results using wellbore infor- sical logs.
mation (azimuth and deviation), Wellbore parame-
ters (mud weight, mudcake coefficient), and geome-
chanical data (elastic properties, rock strength, pore
pressure, stress) as input and perform wellbore sta-
bility analysis with respect to a given depth and sen-
sitivity analysis with respect to a specific input pa-
rameters is detailed in the Appendixes A and B.

4 CONCLUSION AND FUTURE WORK

A geomechanical wellbore stability analysis study


was implemented to effectively plan the future drill-
ing operations in the J-NC186 Field, to maximize
the drilling margin for the future wells drilled, and to
optimize the future field development.

10
5 REFERENCES Manshad, A. K., Jalifar, H., and Aslannejad, M., 2014, Analy-
sis of vertical, horizontal and deviated wellbores stability
by analytical and numerical methods, J Petrol Explor Prod
Aadnoy, B.S., 2011. Petroleum Rock Mechanics: Drilling Op- Technol (2014) 4:359–369.
erations and Well Design. Elsevier Reference.
Pašić, B., Gaurina-Međimurec, N. and Matanović, D., 2007,
Aadnoy, B.S., 2010, Modern Well Design, 2nd edn, Taylor & Wellbore Instability: Causes And Consequences. Rudar-
Francis Group, London, UK. sko-geološko-naftni zbornik, Vol. 19, str. 87 - 98 Zagreb.

Al-Ajmi, A. M. and Zimmerman, R. W., 2006 Wellbore Stabil- Rabia, H., 2002, Well Engineering and Constructions, Entrac
ity Analysis Based on a New True-Triaxial failure criteri- Consulting.
on. Dissertation, University of KTH, Oklahoma, pp.1-5.
Reynolds, S. E., 2001, Characterization and Modelling of the
Elyasi, A., Goshtasbi, K., Saeidi, O. and Torabi, S. R., 2013, Regional In-situ Stress Field of Continental of Australia, A
Stress determination and geomechanical stability analysis thesis submitted to Adelaide University, Australia.
of an oil well of Iran. Sadhana, Vol. 39, Part 1, February
2014, pp. 207–220. Shakib, J. T., Jalalifar, H., and Akhgarian, E., 2013, Wellbore
Stability in Shale Formation Using Analytical and Numer-
Fjaer, E., Holt, R. M., Horsrud, P., Raaen, A. M., and Risnes, ical Simulation. Journal of Chemical and Petroleum Engi-
R., 1992. Petroleum Related Rock Mechanics, Elsevier, neering, University of Tehran, Vol. 47, No.1, Jun.2013,
Amsterdam. PP. 51-60.

Gholami, R., Moradzadeh, A., Rasouli, V., Hanachi, J., (2014), Zoback, M., 2014, Reservoir Geomechanics, Free Open Online
Practical application of failure criteria in determining safe Course, Stanford University.
mud weight windows in drilling operations. Journal of
Rock Mechanics and Geotechnical Engineering 6 (2014) Zoback, M., 2010, Reservoir Geomechanics: New York, Cam-
13–25. bridge University Press.

Goodman, R.E.. 1989. Introduction to Rock Mechanics. 2nd Zoback, M. D., Moos, D., Larry Mastin, L. and Anderson, R.
ed. New York City: Wiley. N., 1985, Well Bore Breakouts and in Situ Stress. U.S.
Geological Survey, Menlo Park, California. journal of ge-
Hallet, D., 2002, Petroleum Geology of Libya, 2 nd ed. Elsevier ophysical research, vol. 90, No. b7, pages 5523-5530, June
Reference. 10, 1985.

Hilgedick, S. A., 2012, Investigation of wellbore stability in a


North Sea field development, A thesis submitted to Mis-
souri University of Science and Technology, Missouri.

Himmelberg, N. C. 2014, Numerical simulations for wellbore


stability and integrity for drilling and completions, A the-
sis submitted to Missouri University of Science and Tech-
nology, Missouri.

Jimenez, J. M. C., Lara, L. C. V., Rueda, A. and Trujillo, N. F.


S., 2007, Geomechanical Wellbore Stability Modelling of
Exploratory Wells-Study Case at Middle Magdalena Ba-
sin. CT&F Ciencia, Technologia y Futuro.

Kadyrov, T., 2012, Integrated Wellbore Stability Analysis For


Well Trajectory Optimization And Field Development:
The West Kazakhstan Field, A thesis submitted to Colora-
do School of Mines, Colorado.

Kalu, I. E., 2013, Numerical Modelling Of Wellbore Instability


(Shear Failure) Using Mechanical Approach, A thesis
submitted to the African university of science and technol-
ogy, ABUJA, NIGERIA, pp.1.

Khaksar, A. and Fang, Z., 2009, Rock Strength from Core and
Logs: Where We Stand and Ways to Go. SPE EU-
ROPEC/EAGE, Amsterdam, The Netherlands, 8–11 June.
SPE 121972.
Appendix A
Geomechanical Models for the selected depths

Every pair of stereographic projection display:

1- The Shear Failure Minimum Mud Weight


(Breakout) Vs. Borehole Orientation plot is a
polar plot where the azimuth indicates the bo-
rehole azimuth and distance from the origin
indicates the borehole deviation magnitude.
This plot shows wellbore breakout as a func-
tion of borehole orientation (azimuth and de-
viation), and the color shading indicates the
wellbore damage mud weight.

2- The Breakdown Mud Weight Vs. Orientation


plot is a polar plot where the azimuth indi-
cates the borehole azimuth and distance from
the origin indicates the borehole deviation
magnitude. The color shading on the plot in-
dicates the breakdown mud weight. This plot
shows breakdown mud weight as a function of
borehole orientation (azimuth and deviation).
Figure A.2. Stereographic Projection represent the effect of
well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (4371
ft)

Figure A.3. Stereographic Projection represent the effect of


well deviation and azimuth in Breakouts (upper) and DITFs
Figure A.1. Stereographic Projection represent the effect of (lower) as a function of mud weight window at depth of (4439
well deviation and azimuth in Breakouts (upper) and DITFs ft)
(lower) as a function of mud weight window at depth of (4286
ft)

12
Figure A.4. Stereographic Projection represent the effect of Figure A.6. Stereographic Projection represent the effect of
well deviation and azimuth in Breakouts (upper) and DITFs well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (4455 (lower) as a function of mud weight window at depth of (4501
ft) ft)

Figure A.5. Stereographic Projection represent the effect of


Figure A.7. Stereographic Projection represent the effect of
well deviation and azimuth in Breakouts (upper) and DITFs
well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (4470
(lower) as a function of mud weight window at depth of (4528
ft)
ft)
Figure A.8. Stereographic Projection represent the effect of Figure A.10. Stereographic Projection represent the effect of
well deviation and azimuth in Breakouts (upper) and DITFs well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (4567 (lower) as a function of mud weight window at depth of (4700
ft) ft)

Figure A.11. Stereographic Projection represent the effect


Figure A.9. Stereographic Projection represent the effect of of well deviation and azimuth in Breakouts (upper) and DITFs
well deviation and azimuth in Breakouts (upper) and DITFs (lower) as a function of mud weight window at depth of (4791
(lower) as a function of mud weight window at depth of (4626 ft)
ft)

14
Figure A.12. Stereographic Projection represent the effect of
well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (4869
ft)

Figure A.13. Stereographic Projection represent the effect of


well deviation and azimuth in Breakouts (upper) and DITFs
(lower) as a function of mud weight window at depth of (5005
ft)
Appendix B
Mud weight window as a function of well devia-
tion and azimuth for the selected depths

Four situations has been chosen for well devia-


tion: 0◦,30◦,60◦ and 90◦.

1- The Mud Weight Window vs. Deviation plot


shows a safe mud weight window as a func-
tion of deviation at a given azimuth.
2- The Mud Weight Window vs. Azimuth plot
shows a safe mud weight window as a func-
tion of azimuth at a given deviation.

Figure B.1. The safe mud weight window as a function of well Figure B.2. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views) inclination (the upper view) and azimuth (the last three views)
at depth of (4286 ft) at depth of (4371 ft)

16
Figur B.3. The safe mud weight window as a function of well Figure B.4. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views) inclination (the upper view) and azimuth (the last three views)
at depth of (4439 ft) at depth of (4455 ft)
Figur B.5. The safe mud weight window as a function of well Figure B.6. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views) inclination (the upper view) and azimuth (the last three views)
at depth of (4470 ft) at depth of (4501 ft)

18
Figur B.7. The safe mud weight window as a function of well Figure B.8. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views) inclination (the upper view) and azimuth (the last three views)
at depth of (4528 ft) at depth of (4567 ft)
Figure B.9. The safe mud weight window as a function of well Figur B.10. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views) inclination (the upper view) and azimuth (the last three views)
at depth of (4626 ft) at depth of (4700 ft)

20
Figure B.11. The safe mud weight window as a function of Figur B.12. The safe mud weight window as a function of well
well inclination (the upper view) and azimuth (the last three inclination (the upper view) and azimuth (the last three views)
views) at depth of (4791 ft) at depth of (4869 ft)
Figur B.13. The safe mud weight window as a function of well
inclination (the upper view) and azimuth (the last three views)
at depth of (5005 ft)

22
View publication stats

You might also like