Anda di halaman 1dari 9

Tectonophysics 492 (2010) 192200

Contents lists available at ScienceDirect

Tectonophysics
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / t e c t o

Subsurface fracture analysis and determination of in-situ stress direction using FMI
logs: An example from the Santonian carbonates (Ilam Formation) in the Abadan
Plain, Iran
Mojtaba Rajabi a,, Shahram Sherkati b, Bahman Bohloli a, Mark Tingay c
a
b
c

School of Geology, University College of Science, University of Tehran, Tehran, Iran


National Iranian Oil Company (NIOC), Exploration Directorate, Tehran, Iran
Tectonics, Resources and Exploration (TRaX), Australian School of Petroleum, University of Adelaide, Adelaide, Australia

a r t i c l e

i n f o

Article history:
Received 28 March 2010
Received in revised form 7 June 2010
Accepted 17 June 2010
Available online 25 June 2010
Keywords:
Borehole breakouts
FMI logs
Fractures
Ilam Formation
In-situ stress
SW Iran

a b s t r a c t
The relationship between the present-day stress eld and natural fractures can have signicant implications
for subsurface uid ow. In particular, fractures that are aligned in orientations favourable for reactivation by
either shear or tensile failure in the in-situ stress eld often exhibit higher hydraulic conductivities. The Ilam
Formation of southwestern Iran is an important hydrocarbon reservoir containing numerous natural
fractures. However, little is known about the state of stress in this region, or any of Iran's petroleum
provinces. We conducted analysis of the present-day maximum horizontal stress orientation and the density,
orientation and hydraulic conductivity of natural fractures in the Ilam carbonates using high resolution
Formation Micro Imager resistivity logs in two wells. A total of 51 breakouts with an overall length of 215 m
were observed in the two wells, indicating a maximum horizontal stress orientation of 68N (7.6) in well
A and 58N (6.3) in well B. Furthermore, the wellbore-derived stress orientations determined herein are
consistent with those inferred from nearby earthquake focal mechanism solutions, indicating that stresses in
the sedimentary cover are linked to the resistance forces generated by ArabiaEurasia collision. Furthermore,
the correlation between stress orientations estimated from earthquake focal mechanism solutions and
breakouts indicates that focal mechanism solution data, which is often considered to be unreliable for stress
eld analysis near transform margins, may provide reliable information on the stress orientation near
continental collision zones. The image log data also reveals three sets of open, and presumably hydraulically
conductive, fractures with strikes of (i) 160170N, (ii) 110140N and (iii) 070080N. Fracture set (iii) is
consistent with being formed and open in the present-day stress eld. However, fracture sets (i) and
(ii) strike at a high angle to the present-day maximum horizontal stress, and are interpreted herein to be the
result of either pre- or syn-folding related forces. The observation that different sets of open fractures in the
eld can be either sensitive or insensitive to the present-day stress is critical for improving hydrocarbon
recovery.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Knowledge of the present-day tectonic stress is an essential issue
in petroleum exploration and production, and, in particular, is a key
parameter in:

borehole stability;
reservoir drainage and ooding patterns;
uid ow in naturally-fractured reservoirs;
hydraulic fracture stimulation, and;
seal breach by fault reactivation (Tingay et al., 2009).
Corresponding author. Tel./fax: +98 2166491623.
E-mail addresses: m.rajabi@hotmail.com, rajabigeo@yahoo.com (M. Rajabi).

0040-1951/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2010.06.014

The present-day state of stress is described by determination of the


stress tensor. It is commonly assumed that one principal stress acts
vertically in sedimentary basins and thus the stress tensor can be
simplied to consist of four components, the magnitudes of the vertical,
maximum horizontal and minimum horizontal stresses in addition to
the orientation of the maximum horizontal stress (Bell, 1996; Tingay
et al., 2009). Of these four components, determination of the maximum
horizontal stress (SHmax) orientation has received extensive attention in
recent 20 years, particularly with regards to the control of in-situ
stresses on subsurface uid ow and fault reactivation (Barton et al.,
1995; Sibson, 1996; Jones and Hillis, 2003; Tingay et al., 2010a).
Fractures that are most susceptible to tensile or shear failure in the
present-day stress, typically those striking approximately parallel or
within 30 of the maximum principal stress orientation, are often

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

observed to transmit the largest volumes of uids (Barton et al., 1995;


Sibson, 1996). Furthermore, extensive analysis of ooding operations
has observed that uid ow is enhanced and pumping rates more
strongly correlated between well pairs that are located parallel to the
present-day SHmax orientation (Heffer et al., 1997).
The scientic importance of understanding the present-day
maximum horizontal stress orientation is further highlighted by the
ndings of the World Stress Map (WSM) Project, which has spent over
20 years building an extensive freely-available repository of presentday stress information. The 2008 release of the World Stress Map
Project contains 21,750 present-day stress indicators from all over the
world and reveals the complexity of the global stress pattern
(Heidbach et al., 2010). Early studies of the present-day stress eld
revealed that the primary, plate-scale stress eld is controlled by plate
boundary forces such as ridge push, slab pull and resistance at
continental collision zones coupled with large intra-plate forces such
as gravitational body forces near mountain ranges (Zoback, 1992;
Reynolds and Hillis, 2000). However, more recent studies have
highlighted the signicance of smaller-scale perturbations in the
stress eld, superimposed upon the plate-scale stress pattern, that are
often observed at the basin to eld scale (Heidbach et al., 2007; Tingay
et al., 2009, 2010b). Furthermore, it is knowledge of stress orientations
at smaller basin and eld scales that has critical importance for
petroleum applications such as wellbore stability, and hydraulic
fracture stimulation (Bell, 1996; Tingay et al., 2005).
Knowledge of the present-day stress orientation is particularly
important in Iran, which has an extensive and mature petroleum
exploration and production industry, and is also prone to stressrelated geohazards such as earthquakes. Yet, the 2008 World Stress
Map database contains very little present-day stress information for
Iran and no stress data from petroleum wells (Heidbach et al., 2009).
Indeed, all of the stress data currently available for Iran is derived from
earthquake focal mechanism solutions from events that are typically
at depths of ten kilometres or more, and which might not be relevant
for petroleum applications, particularly in areas possibly detached by
salt or low-angle faults. Furthermore, the majority of these earthquake focal mechanism solutions are located along the boundary
between the Arabian and Eurasian plates, and there are concerns
surrounding the reliability of stress information derived from earthquakes near plate boundaries (Heidbach et al., 2010). For example,
stress orientations derived from earthquake focal mechanism solutions along the San Andreas Fault Zone and Great Sumatran Fault are
often highly inconsistent with those obtained from more reliable
petroleum industry data (Zoback et al., 1987; Mount and Suppe, 1992;
Heidbach et al., 2010).
In this study we examine resistivity image logs to determine the
present-day stress orientation in the Santonian carbonates (Ilam
Formation) of the Abadan Plain in SW Iran. This area of Iran is located
adjacent to the boundary of ArabianCentral Iran plates, thus allowing
comparison between exiting stress data from earthquake focal
mechanism solutions within a plate boundary region. Finally, we
compare the stress orientation with hydraulically conductive fractures
in the Ilam Formation to examine the implications of the present-day
stress orientation on enhanced petroleum production in the region.
2. Geological setting
The study area is located in the Abadan Plain in the western Dezful
Embayment (SW Iran), which is part of the Zagros fold-and-thrust belt
(Fig. 1; McQuarrie, 2004; Alavi, 1994, 2004). The Zagros mountain range
is a collisional belt between the Iranian block (belonging to Eurasia) and
the Arabian plate, whose convergence started at the beginning of the
Late Cretaceous (Ricou, 1974; Berberian and King, 1981) and accelerated during the Late Miocene and Pliocene (Stocklin, 1968). The
convergence is still active at the present-day, in a NNESSW direction
(DeMets et al., 1990, 1994, 2010; McClusky et al., 2000, 2003; Sella et al.,

193

2002; Nilforoushan et al., 2003; Vernant et al., 2004). The Zagros orogen
generally consists of three parallel belts (Fig. 1a). 1) To the northeast, the
UrumiehDokhtar magmatic assemblage (UDMA), 2) southwest of
UDMA, the Zagros imbricate zone (ZIZ) includes both the Sanandaj
Sirjan zone and the Zagros thrust zone of Stocklin (1968) and 3) The
Zagros foldthrust belt (ZFTB), (the Zagros simple folded zone of
Falcon, 1974), extends parallel and to the southwest of the ZIZ (Fig. 1;
Alavi, 2004, 2007). The Abadan Plain is located in the western end of the
Dezful Embayment and is positioned between the Precambrian Arabian
Shield to the southwest and the ZFTB to the northeast (Fig. 1; Abdollahie
Fard et al., 2006).
The Upper Cretaceous Ilam Formation of the Bangestan Group is
composed of light gray (locally buff to white) shallow-marine
Hippurites-bearing limestones (grainstone, pelletal packstone, dark
bioclastic wackestone) with intercalations of black ssile shale and
broken by several intraformational disconformities (Alavi, 2004). The
Ilam Formation is a proven Upper Cretaceous reservoir and is
stratigraphically positioned between the CampanianMaastrichtian
Gurpi Formation (marls to marly limestones) and Lafan Formation
(Coniacian shales; Fig. 2). The two wells examined in this study are
located in anticlines that have fold axes trending NWSE (well A;
Zagros mountain trend) and NS (well B; Arabian plate trend).
Differences and characteristics of these major trends in the Zagros
Mountains are described in detail by Abdollahie Fard et al. (2006).
3. Methodology: Determination of present-day maximum
horizontal stress orientation
The present-day SHmax orientation was determined herein from
borehole breakouts interpreted from Formation Micro Imager (FMI)
logs. When a borehole is drilled, the material removed from the
subsurface is no longer supporting the surrounding rock (the wellbore
wall). As a result, the stresses become concentrated in the wellbore wall
(Kirsch, 1898). Borehole breakouts are stress-induced elongations of the
wellbore and occur when the wellbore stress concentration (circumferential stress) exceeds that required to cause compressive failure of
intact rock (Bell and Gough, 1979). The elongation of the cross-sectional
shape of the wellbore is the result of compressive shear failure on
intersecting conjugate planes, which causes pieces of the borehole wall
to spall off (Bell and Gough, 1979). The maximum circumferential stress
around a vertical borehole occurs perpendicular to SHmax (Kirsch, 1898).
Hence, borehole breakouts are elongated perpendicular to the presentday SHmax direction (Bell and Gough, 1979).
The FMI is a resistivity imaging tool consisting of two perpendicular
pairs of caliper arms, with the end of each arm hosting a pad and
attached ap. The pads and aps contain a number of resistivity sensors
(typically 24 on each pad or ap), the data from which can be processed
to build up a picture of the wellbore wall based on resistivity contrasts
(Ekstrom et al., 1987). The FMI tool also collects information on the
shape of the wellbore cross-section and the hole geometry, in addition
to the resistivity image data. Breakouts appear on resistivity image logs
as broad, parallel, often poorly resolved conductive zones separated by
180 and exhibiting caliper enlargement in the direction of the
conductive zones (Fig. 3; Bell, 1996; Tingay et al., 2008). FMI images
can also be used to interpret drilling-induced fractures (DIFs) which are
oriented parallel to the in-situ SHmax orientation (Bell, 1996). However,
no DIFs were observed in this study and thus are not discussed further
herein.
The breakouts observed in each well are used to determine an
average SHmax orientation using standard circular statistical methods
(Mardia, 1972). The average SHmax direction is then given a qualityranking according to the World Stress Map criteria that takes into
account the standard deviation of breakout orientations in addition the
total length and number of breakouts observed (Heidbach et al., 2010).
The World Stress Map quality-ranking scheme assigns stress indicators
a ranking from A-quality (highest reliability, SHmax accurate to 12 and

194

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

195

Fig. 2. Cretaceous and Tertiary stratigraphy of the study area (modied from Motiei 1993, Alavi 2004, and Abdollahie Fard et al. 2006). The FMI image logs examined herein are
predominately collected within the Upper Cretaceous formations.

observed in a signicant volume of rock) through to D-quality (SHmax


accurate to 40) and E-quality (no breakouts observed or breakout
orientations too statistically scattered).
In addition to breakouts, the FMI logs herein were also used to
interpret the geometry of all natural fractures. Fractures appear on
resistivity logs as sinusoidal zones that may be either electrically
resistive or conductive (Fig. 4). Herein, we have been particularly
focussed on electrically-conductive fractures that, due to the invasion
of conductive drilling mud, are conventionally assumed to be open
and hydraulically conductive (Barton et al., 1995).
4. Results and discussion
4.1. Horizontal in-situ stress direction
A total of 22 borehole breakouts (BO) with an overall length of
65 m were interpreted in well A (Table 1). The mean BO azimuth in

this well is 158N and therefore, the average SHmax orientation


interpreted for well A is 068N. The standard deviation of breakouts in
well A is 7.6 and thus the SHmax orientation interpreted for well A is
ranked as B-quality according to the WSM criteria.
A total of 29 BO with an overall length of 150 m were interpreted
in well B (Table 2). The mean breakout azimuth is 148N and thus
SHmax is interpreted to be oriented towards 058N in well B. Breakouts
in well B have a standard deviation of 6.3 and thus the SHmax
orientation derived for well B is ranked as A-quality.
The overall low standard deviations of the breakout orientations
highlights that breakout orientations are consistent with depth
throughout each well. Furthermore, the largely ENEWSW SHmax
orientation observed in wells A and B is broadly consistent with both
SHmax orientations derived from nearby earthquake focal mechanism
solutions and with the absolute plate motion direction of the Arabian
plate (Fig. 5). Hence, the largely ENEWSW SHmax orientation observed
herein is likely to be associated with forces driving the Arabian plate

Fig. 1. (a) Subdivisions of the Zagros Belt (modied from Alavi, 2007). Abbreviations: AD Arak depression, DR Dezful recess, EAF East Anatolian Fault, FS Fars salient, GKD
Gav Khooni depression, KR Karkuk recess, LS Lorestan salient, MAC Makran accretionary complex, MFF Mountain front exure, MZT Main Zagros Thrust, OL Oman Line,
PTCCCS Paleo-Tethyan continentcontinent collisional suture, SD Sirjan depression, SRRB SavehRafsanjan retroforeland basin, SSZ SanandajSirjan zone, ZTZ Zagros
thrust zone, UDMA UrumiehDokhtar magmatic assemblage, ZDF Zagros deformational front, ZFTB Zagros foldthrust belt, ZIZ Zagros imbricate zone, ZS Zagros suture.
(b) Main structural subdivisions of the Simply Folded Zagros Belt (after Falcon 1961). The study area and wells (A and B) are located within the Abadan Plain region (inclined stripes).

196

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

Fig. 3. A typical breakout observed on an FMI log in well B. The breakout is identied as a pair of poorly resolved conductive zones observed on opposite sides of the borehole
(outlined in bold) and showing caliper enlargement in the same direction (note caliper 2 greater than caliper 1). The breakout pictured herein is oriented approximately NNWSSE
and thus indicates a present-day SHmax orientation of approximately ENEWSW.

motion (e.g. ridge push in the Red Sea and Gulf of Aden) in addition to
resistance forces generated at the ArabiaIran continental collision
zone.
The correlation between SHmax orientations derived from breakouts
herein and earthquake focal mechanism solutions in the WSM project is
also scientically signicant. Earthquake focal mechanism solutions
make up 72% of the 2008 WSM database (Heidbach et al., 2010).
However, the reliability of using focal mechanism solutions near plate
boundaries as present-day stress indicators has recently been brought
into question (Heidbach et al., 2010). Stress orientations inferred from
focal mechanism solutions assume that the earthquake motion is along
faults that are sub-optimally oriented with the present-day stress
orientation. Yet, analysis of stress orientations from different methods
near plate boundaries has revealed that some plate boundaries, most
notably the San Andreas Fault Zone and Great Sumatran Fault, are
mechanically weak (low coefcient of friction) and may thus be
reactivated by non-optimal and highly-oblique stress elds (Zoback
et al., 1987; Mount and Suppe, 1992). Hence, there exists a higher
likelihood for errors in stress orientations derived from earthquake focal
mechanism solutions near plate boundaries and these data must be

considered as potentially unreliable (Heidbach et al., 2010). The


potential for error in present-day SHmax orientations derived for focal
mechanism solutions is, theoretically, different in normal, strikeslip
and thrust faulting stress regimes and Heidbach et al. (2010) argue that
the potential for errors in SHmax orientations derived from earthquake
focal mechanism solutions is minimal in continental collision zones,
such as in Southwestern Iran. The consistency between SHmax orientations derived from more reliable breakouts herein and earthquake focal
mechanism solutions in the WSM database provides some validation to
this hypothesis.

4.2. Fracture analysis


The overwhelming majority of Ilam Formation fractures observed in
wells A and B were electrically (and presumably hydraulically)
conductive. A total of 72 fractures were observed in well A (70 conductive
and 2 non-conductive) and 33 fractures in well B (29 conductive and 4
non-conductive). The open fractures observed in both wells can be
classied into three major sets (i): 160170N (in both well A and well

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

197

Fig. 4. An example of an electrically resistive natural fracture (left; assumed to be hydraulically non-conductive or closed) and an electrically-conductive fracture (right; assumed
to be hydraulically conductive or open) in FMI logs from well A.

B), set (ii): 110140N (in well A) and set (iii) 070080N (in well A;
Fig. 6).
Several studies have been published on fractures in Iranian
carbonate reservoirs (e.g., McQuillan, 1973, 1974, Gholipour 1998,
Rezaie and Nogole-Sadat 2004, Ahmadhadi et al. 2007, 2008, Khoshbakht et al., 2009) but most of these studies are focused on Asmari
Formation; the most important oil reservoir in Iran. Based on
Ahmadhadi et al. (2007, 2008), regional fractures in the Asmari
Formation are mostly sub-vertical and interpreted to be initiated before
the main Mio-Pliocene folding phase of the sedimentary cover. Hence,
the fracturing observed in the Asmari Formation is interpreted to be a
consequence of reactivation of deep basement faults associated with the
continental collision of the Arabian plate with Central Iran plate.
The origin of the open fractures observed in the Ilam Formation
herein is uncertain. Only set (i) is observed in any statistically signicant

Table 1
Results of BOs analysis in well A. Most BOs examined herein are predominately
collected within the Upper Cretaceous Ilam, Lafan and Sarvak Formations.
Formation

Top
(m)

Bottom
(m)

Caliper1
(mm)

Caliper2
(mm)

Orientation
of BO

Ilam

2902.700
2907.000
2911.008
2914.726
2923.041
2929.001
2946.965
2964.576
2966.301
2968.304
2971.262
2974.020
3023.179
3046.345
3047.771
3053.360
3057.610
3064.983
3114.732
3145.358
3156.986
3186.945

2905.980
2907.500
2913.622
2918.631
2928.106
2934.490
2951.521
2965.443
2967.132
2968.743
2973.200
2977.100
3030.999
3047.248
3049.949
3054.510
3057.900
3078.730
3115.187
3145.982
3161.264
3188.244

303
307
308
304
307
303
306
311
308
308
307
307
308
305
305
305
305
309
306
304
306
308

313
319
336
310
330
359
331
329
329
319
332
338
340
340
329
327
353
358
338
332
351
368

168.0
160.0
168.4
153.0
165.4
165.4
157.4
164.9
159.0
165.4
163.9
135.1
150.9
157.0
154.0
140.4
157.0
158.0
154.0
162.0
157.0
155.0

Lafan
Sarvak

way in well B. The 160170N strike of fractures observed in well B is


inconsistent with the fractures forming in the present-day ENEWSW
SHmax orientation and is also inconsistent with the forces that formed
the Zagros mountain ranges. However, well B is located within a NS
trending anticline (Arabian trend) and the fracturing is broadly
consistent with axial-planar parallel fracturing or with fracturing that
might have occurred before the main MiocenePliocene folding event.
Three different fractures sets are observed in well A that range
across 90 from ENEWSW to SSENNW. Fracture set (i) is also

Table 2
Results of BOs analysis in well B. Most BOs examined herein are predominately
collected within the Upper Cretaceous formations.
Formation

Top
(m)

Bottom
(m)

Caliper1
(mm)

Caliper2
(mm)

Orientation
of BO

Ilam

2445.053
2550.921
2568.418
2661.633
3065.266
3400.292
3405.869
3415.336
3439.572
3443.887
3454.211
3460.213
3469.398
3625.082
3634.810
3641.237
3654.30
3669.362
3685.191
3694.621
3712.368
3723.028
3734.857
3737.593
3749.684
3768.246
3787.776
3798.519
3802.451

2447.220
2559.814
2597.577
2663.812
3072.343
3402.514
3409.269
3417.192
3441.540
3448.499
3456.481
3462.434
3471.443
3627.390
3638.409
3645.838
3658.410
3684.015
3691.248
3696.934
3715.262
3726.191
3736.411
3739.408
3752.174
3779.768
3791.706
3801.120
3814.820

228
231
260
215
247
220
219
242
268
265
222
224
258
260
218
218
258
255
250
238
220
245
249
242
213
216
245
245
276

238
211
310
260
270
232
250
221
216
217
239
250
220
219
265
272
217
216
215
215
266
215
220
220
248
230
215
220
216

151.3
149.7
145.1
164.8
151.2
142.2
144.1
148.3
157.2
145.0
149.7
140.2
153.2
152.8
124.2
159.7
149.0
155.7
148.7
140.0
155.0
154.2
145.2
143.1
144.2
143.5
143.2
154.2
148.2

Sarvak
Kazhdumi

Darian
Gadvan

198

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

Fig. 5. Maximum horizontal stress orientations in Iran from the World Stress Map database and from the wells analysed herein. Symbols and different colours indicate the method of
measurement (circles are focal mechanism solutions, inward-facing arrows are breakouts) and the stress regime (NF = normal faulting stress regime; SS = strikeslip faulting stress
regime; TF = thrust faulting stress regime; black = undened stress regime). Length of the lines indicates quality of data. Only earthquake focal mechanism solution data, typically
from 10 km or more depth was previously available for Iran (Heidbach et al., 2008). The NEENE SHmax orientations observed from borehole breakouts herein are consistent with the
focal mechanism solutions observed in the Zagros Mountain belt. A and B shows location of studied wells and heavy arrow shows motion of the Arabian plate relative to the Eurasia.

dominant in well A and is again interpreted herein to represent a prefolding (Arabian) fracture set. Fracture set (ii) is broadly parallel to
the axis of the NWSE (Zagros trend) anticline in which well A is
located, and thus this fracture set may represent axial-planar
fracturing associated with this folding. Fracture set (iii) is oriented
approximately sub-parallel or at a low-angle (b22) to the presentday SHmax orientation and is thus interpreted herein as being the
result of recent post-folding fracturing. We also suggest that the
increased open fracture density in well A is related to curvature of this
anticline during the main Zagros folding phase.

4.3. Relation between open fractures and maximum horizontal in-situ


stress
There are two main hypotheses about relation of in-situ stress and
open fractures.
1) Stress-sensitive fractures, in which fractures are hydraulically
conductive due to being suitably oriented for failure (shear or
tensile) in the present-day stress tensor (e.g., Barton et al. 1995,
Sibson, 1996; Finkbeiner et al., 1997; Major and Holtz 1997).

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

199

Fig. 6. (Upper) Orientation of open fractures in well A (Right) and B (Left). Three trends of natural fractures are observed in these wells striking in (i) 160170N, (ii) 110140N and
(iii) 070080N. (Lower) Comparison of present-day SHmax orientation (lower) with the strike of open fractures (upper) in the Ilam Formation. Only fracture set iii is oriented subparallel to the present-day stress orientation and is interpreted herein to genetically related to the recent stress state and to be hydraulically conductive in response to the presentday stress regime. Fracture sets i and ii are interpreted herein to have been formed by previous tectonic events and are likely to be stress-insensitive and be hydraulically
conductive due to other processes, such as partial mineralization.

2) Stress-insensitive fractures, in which fractures are propped


open due to partial mineralization or from fracture wall asperities
and thus may be oriented in any direction and have no relation to
the present-day stress tensor (Laubach et al., 2004).
The observation herein that three open fracture sets exist in the
Ilam Formation, of which only one (set iii) may be genetically linked
to the present-day stress orientation while two sets (set i and ii) are
oriented at a high angle to the SHmax orientation, suggest that both
stress-sensitive and stress-insensitive fracture sets occur and have
signicant inuence on hydraulic conductivity (Fig. 6).
No detailed production data was available for this study, and thus
we are unable to assess the productivity of each fracture set. However,
optimal production from naturally-fractured reservoirs is, in essence, a
function of drilling wells to intersect as many hydraulically conductive
fractures as possible. Hence, a horizontal well drilled approximately
towards the NESW or NNESSW would be expected to encounter the
greatest number of open fractures in both wells within the study
region (Fig. 6). Furthermore, horizontal wells deviated towards the
present-day SHmax orientation are typically the most stable in thrust
faulting stress regimes (SHmax N Shmin N Sv), as is expected to exist
within the study region. Deviation of wells towards the present-day
SHmax orientation in thrust faulting stress regimes reduces both the
absolute stress and differential stresses acting on the borehole,
resulting in the least possible circumferential stress concentration
around the wellbore and thus reducing the likelihood of breakout and

associated drilling problems such as struck pipe and wellbore collapse


(Peska and Zoback, 1995). Hence, we suggest that drilling of
production wells that are highly-deviated or horizontal and oriented
approximately towards the NE or SW is likely to both intersect more
hydraulically conductive fractures and reduce wellbore instability
problems.
5. Conclusions
The present study focused on subsurface fracture analysis and
direction of in-situ stress in the Santonian carbonate (Ilam Formation),
SW of Iran. Analysis of 51 breakouts with a total length of 215 m in two
wells, indicate a SHmax of 68N (7.6) in well A and 58N (6.3) in
well B. The correlation between SHmax orientations derived from
breakouts and focal mechanism solution in this region, suggesting that
focal mechanism solution data near continental collision zones may
provide reliable information of the stress orientation.
Results of fracture analysis shows three major sets of open fractures
in Ilam Formation with strikes of (i) 160170N, (ii) 110140N and
(iii) 070080N. Fracture set (i) or pre-folding fracture set present in
both wells, but fractures sets (ii) and (iii) observed only in well A.
These results show that increasing fracture density in well A is possibly
the result of major Zagros folding phases.
Relation between open fractures and horizontal in-situ stress show
that only fracture set (iii) is consistent with the present-day stress
orientation while two sets (set i and ii) are oriented at a high angle to

200

M. Rajabi et al. / Tectonophysics 492 (2010) 192200

SHmax orientation, conrms that both stress-sensitive and stressinsensitive fracture sets can occur and have signicant inuence on
hydraulic conductivity. We suggest that production wells in the
region should be drilled highly deviated towards the NE or SW to both
maximize the number of hydraulically conductive fractures intersected by the borehole and to reduce the likelihood of wellbore
instability problems.
Acknowledgment
The authors wish to thank Petroleum Engineering and Development
Company (PEDEC) for sponsorship, data preparation and permission to
publish the data. We are grateful to M. Mohammadinia for his
cooperation. The authors also wish to thank Chris Morley and Oliver
Heidbach for their insightful and constructive reviews. Mark Tingay's
contribution forms TRaX record #108.
References
Abdollahie Fard, I., Braathen, A., Mokhtari, M., Alavi, S.A., 2006. Interaction of the Zagros
FoldThrust Belt and the Arabian-type, deep-seated folds in the Abadan Plain and
the Dezful Embayment, SW Iran. Petrol. Geosci. 12, 347362.
Ahmadhadi, F., Lacombe, O., Daniel, J.M., 2007. Early reactivation of basement faults in
Central Zagros (SW Iran): evidence from pre-folding fracture populations in the
Asmari Formation and Lower Tertiary paleogeography. In: Lacombe, O., Lave, J.,
Verge's, J., Roure, F. (Eds.), Thrust Belts and Foreland Basins; From Fold Kinematics
to Hydrocarbon Systems, Frontiers in Earth Sciences. Springer Verlag, pp. 205208.
Chapter 11.
Ahmadhadi, F., Daniel, J.M., Azzizadeh, M., Lacombe, O., 2008. Evidence for pre-folding
vein development in the Oligo-Miocene Asmari Formation in the Central Zagros
Fold Belt, Iran. Tectonics 27 (TC1016), 122.
Alavi, M., 1994. Tectonics of the Zagros orogenic belt of Iran: new data and
interpretations. Tectonophysics 229, 211238.
Alavi, M., 2004. Regional stratigraphy of the Zagros fold-thrust belt of Iran and its
proforeland evolution. Am. J. Sci. 304, 120.
Alavi, M., 2007. Structures of the Zagros fold thrust belt in Iran. Am. J. Sci. 307,
10641095.
Barton, C.A., Zoback, M.D., Moos, D., 1995. Fluid ow along potentially active faults in
crystalline rock. Geology 23 (8), 683686.
Bell, J.S., 1996. Petro Geoscience In situ stresses in sedimentary rocks (part 1):
measurement techniques. Geosci. Can. 23, 85100.
Bell, J.S., Gough, D.I., 1979. Northeastsouthwest compressive stress in Alberta:
evidence from oil wells. Earth Planet. Sci. Lett. 45, 475482.
Berberian, M., King, G.C.P., 1981. Towards a paleogeography and tectonic evolution of
Iran. Can. J. Earth Sci. 18, 210265.
DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophys. J.
Int. 101, 425478.
DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1994. Effect of recent revisions to the
geomagnetic reversal time scale on estimates of current plate motions. Geophys.
Res. Lett. 21, 21912194.
DeMets, C., Gordon, R.G., Argus, D.F., 2010. Geologically current plate motions. Geophys.
J. Int. 181, 180.
Ekstrom, M.P., Dahan, C.A., Chen, M.Y., Lloyd, P.M., Rossi, D.J., 1987. Formation imaging
with microelectrical scanning arrays. Log Anal. 28, 294306.
Falcon, N.L., 1961. Major earth-exuring in the Zagros Mountain of southwest Iran. J.
Geol. Soc. Lond. 117, 367376.
Finkbeiner, T., Barton, C.A., Zoback, M.D., 1997. Relationships among in-situ stress,
fractures and faults, and uid ow: Monterey Formation, Santa Maria Basin,
California. AAPG Bull. 81 (12), 19751999.
Gholipour, A.M., 1998. Patterns and structural positions of productive fractures in the
Asmari Reservoirs, Southwest Iran. J. Can. Pet. Technol. 37 (1), 4450.
Heffer, K.J., Fox, R.J., McGill, C.A., Koutsabeloulis, N.C., 1997. Novel techniques show
links between reservoir ow directionality, earth stress, fault structure and
geomechanical changes in mature wateroods. Soc. Petrol. Eng. J. 2, 9198 Paper
Number 30711.
Heidbach, O., Reinecker, J., Tingay, M.R.P., Mller, B., Sperner, B., Fuchs, K., Wenzel, F., 2007.
Plate boundary forces are not enough: second- and third-order stress patterns
highlighted in the World Stress Map database. Tectonics 26 (TC6014), 19. doi:10.1029/
2007TC002133.
Heidbach, O., Tingay, M., Barth, A., Reinecker, J., Kurfe, D., Mller, B., 2008. The World
Stress Map database release 2008. doi:10.1594/GFZ.WSM.Rel2008.
Heidbach, O., Tingay, M.R.P., Barth, A., Reinecker, J., Kurfe, D., Mller, B., 2009. The
World Stress Map based on the database release 2008, 1. Commission of the
Geological Map of the World, Paris, p. 46. doi:10.1594/GFZ.WSM.Map2009. M.

Heidbach, O., Tingay, M.R.P., Barth, A., Reinecker, J., Kurfe, D., Mller, B., 2010. Global
crustal stress pattern based on the World Stress Map database release 2008.
Tectonophysics 482, 315.
Jones, R.M., Hillis, R.R., 2003. An integrated, quantitative approach to assessing fault-seal
risk. AAPG Bull. 347, 189215.
Khoshbakht, F., Memarian, H., Mohammadnia, M., 2009. Comparison of Asmari, Pabdeh
and Gurpi formation's fractures, derived from image log. J. Pet. Sci. Eng. 67, 6574.
Kirsch, V., 1898. Die Theorie der Elastizitt und die Bedrfnisse der Festigkeitslehre.
Zeitschrift des Vereines Deutscher Ingenieure 29, 797807.
Laubach, S.E., Olson, J.E., Gale, J.F.W., 2004. Are open fractures necessarily aligned with
maximum horizontal stress? Earth Planet. Sci. Lett. 222, 191195.
Major, R.P., Holtz, M.K., 1997. Identifying fracture orientation in Mature Carbonate
Platform reservoir. AAPG Bull. 81 (7), 10631069.
Mardia, K.V., 1972. Statistics of Directional Data: Probability and Mathematical
Statistics. Academic Press, London. 357 pp.
McClusky, S., Balassanian, S., Barka, A., Demir, C., Ergintav, S., Georgiev, I., Gurkan, O.,
Hamburger, M., Hurst, K., Kahle, H., Kastens, K., Kekelidze, G., King, R., Kotzev, V.,
Lenk, O., Mahmoud, S., Mishin, A., Nadariya, M., Ouzounis, A., Paradissis, D., Peter, Y.,
Prilepin, M., Reilinger, R., Sanli, I., Seeger, H., Tealeb, A., Toksoz, M., Veis, G., 2000.
Global positioning system constraints on plate kinematics and dynamics in the
eastern Mediterranean and Caucasus. J. Geophys. Res. 105, 56955719.
McClusky, S., Reilinger, R., Mahmoud, S., Ben Sari, D., Tealeb, A., 2003. GPS constraints
on Africa (Nubia) and Arabia plate motions. Geophys. J. Int. 155, 126138.
McQuarrie, N., 2004. Crustal scale geometry of the Zagros fold-thrust belt. Iran. J. Struct.
Geol. 26, 519535.
McQuillan, H., 1973. Small-scale fracture density in Asmari Formation of Southwest Iran
and its relation to bed thickness and structural setting. AAPG Bull. 47 (12),
23672385.
McQuillan, H., 1974. Fracture patterns on Kuh-e Asmari Anticline, Southwest Iran. AAPG
Bull. 58 (2), 236246.
Motiei, H., 1993. Stratigraphy of Zagros. Treatise on the Geology of Iran. Geological
Survey of Iran, Tehran. in Persian.
Mount, V.S., Suppe, J., 1992. Present-day stress orientations adjacent to active strikeslip
faults: California and Sumatra. J. Geophys. Res. 97, 1199512013.
Nilforoushan, F., Masson, F., Vernant, P., Vigny, C., Martinod, J., Abbassi, M., Nankali, H.,
Hatzfeld, D., Bayer, R., Tavakoli, F., Ashtiani, A., Doeringer, E., Daignires, M.,
Collard, P., Chry, J., 2003. GPS network monitors the ArabiaEurasia collision
deformation in Iran. J. Geod. 77, 411422.
Peska, P., Zoback, M.D., 1995. Compressive and tensile failure of inclined well bores and
determination of in situ and rock strength. J. Geophys. Res. 100, 1279112811.
Reynolds, S.D., Hillis, R.R., 2000. The in situ stress eld of the Perth Basin, Australia.
Geophys. Res. Lett. 27 (20), 34213424.
Rezaie, A.H., Nogole-Sadat, M.A., 2004. Fracture modeling in Asmari Reservoir of Rag-e
Sed Oil-Field by using Multiwell Image Log (FMS/FMI). Iranian Int. J. Sci. 5 (1),
107121.
Ricou, L.E., 1974. L'volution gologique de la rgion de Neyriz (Zagros iranien) et
l'volution structurale des zagrides, Thse. Universit d'Orsay, France.
Sella, G.F., Dixon, T.H., Mao, A., 2002. REVEL: a model for recent plate velocities from
space geodesy. J. Geophys. Res. 107 (B4), 11-111-32 ETG.
Sibson, R.H., 1996. Structural permeability of uid-driven fault-fracture meshes. J.
Struct. Geol. 18, 10311042.
Stocklin, J., 1968. Structural history and tectonics of Iran: a review. AAPG Bull. 52,
12291258.
Tingay, M., Muller, B., Reinecker, J., Heidbach, O., Wenzel, F., Flecknstein, P., 2005.
Understanding tectonic stress in the oil patch: the World Stress Map Project. Lead.
Edge 24 (12), 12761282.
Tingay, M., Reinecker, J., Mller, B., 2008. Borehole breakout and drilling-induced
fracture analysis from image logs. World Stress Map Project Stress Analysis
Guidelines. available online at: www.world-stress-map.org.
Tingay, M.R.P., Hillis, R.R., Morley, C.K., King, R.C., Swarbrick, E., Damit, A.R., 2009.
Present-day stress and neotectonics of Brunei: implications for petroleum
exploration and production. AAPG Bull. 93 (1), 75100.
Tingay, M.R.P., Morley, C.K., Hillis, R.R., Meyer, J., 2010a. Present-day stress orientation
in Thailand's basins. J. Struct. Geol. 32, 235248.
Tingay, M.R.P., Morley, C.K., King, R.E., Hillis, R.R., Hall, R., Coblentz, D., 2010b. The
Southeast Asian Stress Map. Tectonophysics 482, 92104.
Vernant, P., Nilforoushan, F., Hatzfeld, D., Abbassi, M.R., Vigny, C., Masson, F., Nankali, H.,
Martinod, J., Ashtiani, A., Bayer, R., Tavakoli, F., Chry, J., 2004. Present-day crustal
deformation and plate kinematics in the Middle East constrained by GPS
measurement in Iran and northern Oman. Geophys. J. Int. 157, 381398.
Zoback, M.L., 1992. First- and second-order patterns of stress in the lithosphere: the
world stress map project. J. Geophys. Res. 97, 1170311728.
Zoback, M.D., Zoback, M.L., Moun, V.S., Suppe, J., Eaton, J.P., Healy, J.H., Oppenheimer, D.,
Reasenberg, P., Jones, L., Raleigh, C.B., Wong, I.G., Scotti, O., Wentworth, C., 1987.
New evidence on the state of stress of the San Andreas Fault system. Science 238,
11051111.

Anda mungkin juga menyukai