You are on page 1of 18

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

net/publication/330074094

Estimation of pore pressure, tectonic strain and stress magnitudes in the


Upper Assam basin: A tectonically active part of India

Article  in  Geophysical Journal International · January 2019


DOI: 10.1093/gji/ggy440

CITATIONS READS

17 1,710

3 authors:

Jenifer Alam Rima Chatterjee


Indian Institute of Technology (ISM) Dhanbad Indian Institute of Technology (ISM) Dhanbad
7 PUBLICATIONS   27 CITATIONS    110 PUBLICATIONS   1,580 CITATIONS   

SEE PROFILE SEE PROFILE

Sumangal Dasgupta
PETRONAS
9 PUBLICATIONS   106 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Master Thesis View project

Preparation of Indian Stress Map for Analysis of Crustal Motion and Seismotectonics View project

All content following this page was uploaded by Jenifer Alam on 09 January 2019.

The user has requested enhancement of the downloaded file.


Geophys. J. Int. (2019) 216, 659–675 doi: 10.1093/gji/ggy440
Advance Access publication 2018 October 25
GJI Geodynamics and tectonics

Estimation of pore pressure, tectonic strain and stress magnitudes in


the Upper Assam basin: a tectonically active part of India

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Jenifer Alam, Rima Chatterjee and Sumangal Dasgupta
Department of Applied Geophysics, IIT(ISM), Dhanbad 826004, India. E-mail: rima c 99@yahoo.com

Accepted 2018 October 24. Received 2018 September 11; in original form 2017 October 16

ABSTRACT
Conventional well logs serve as effective tools for estimation of pore pressure (PP), tectonic
strain and stress magnitudes in normal and thrust faulted parts of the Upper Assam basin.
Several models have been developed for prediction of PP and stress magnitudes over the years.
Acquiring tectonic strain in active area enhances the accuracy of in situ stress estimate in
the Upper Assam. This paper discusses Eaton’s sonic model for PP estimation and assesses
stress magnitudes incorporating tectonic strain in poroelastic models. Eaton’s exponent is
varied from 0 to 3 to get best-fit results for PP prediction validated with available modular
dynamic tester and mud weight data. PP is estimated for geological formations: Dhekiajuli,
Girujan, Tipam, Barail, Kopili, Sylhet, Langpur to the granitic basement in normal and thrust
faulted areas of the Upper Assam basin using Eaton’s exponent of 1.0. The predicted tectonic
strain using uniaxial model varies from −0.0044 to 0.0061. The computed tectonic strain using
biaxial poroelastic model ranges from 0.000099 to 0.00077 along NE–SW and from −0.001 to
−0.0019 along NW–SE directions for normal faulting regime, respectively. In thrust faulting
regime, the tectonic strains are obtained as 0.00092 and 0.00053 along NE–SW and NW–
SE directions, respectively. The vertical (SV ), maximum (SH ) and minimum (Sh ) horizontal
stress gradients range 21.7–23.8, 14.9–23.1 and 12.0–21.4 MPa km−1 , respectively. Moreover,
we have proposed a multiple linear regression model for evaluation of Sh from independent
parameters, namely, PP, Poisson’s ratio and SV from three wells and dependent parameter as
leak-off test (LOT) data. This is further tested with other three wells in the study area. The
predicted Sh is found to have satisfactory correlation with observed LOT for these wells.
Key words: Geomechanics; Downhole methods; Numerical Modelling.

The Upper Assam basin occurring in the Paleocene–Eocene con-


1 I N T RO D U C T I O N
tinental shelf of the Indian plate is overthrust by the Himalayas on
The northeastern part of India developed complicated tectonics due the north–northwest and by the Naga Thrust (NT) on the southeast
to northward convergence followed by rotation against Eurasia and (Molnar & Qidong 1984; Banerjee et al. 2008; Vernant et al. 2014;
oblique convergence against the southeast Asia since the Cretaceous Fig. 1). The northeastern portion of the Indian plate is deformed
(Fitch 1970; Seeber et al. 1981; Maung 1987; Khan 2005; Khan and is seismically active, with the remarkable exception of the As-
et al. 2011). This part is one of the most tectonically active regions sam Gap (Khattri et al. 1983). There are only four earthquakes with
in the world since Mesozoic time (Gowd et al. 1998). Knowledge of magnitudes between 4.5 and 5.9 and one with 8.6 that has occurred
contemporary stresses can provide important information regarding in the Upper Assam basin in the past 200 years (Gowd et al. 1998;
intraplate deformation, geodynamic and neotectonic processes, as Rajendran & Rajendran 2011). Previous researchers have already
well as localized processes, with implications for mining, petroleum discussed about the seismotectonic activity in the northeastern India
and geothermal industries (Bell 1996; Hillis & Reynolds 2003; through earthquake focal mechanism and global positioning system
Rajabi et al. 2016; Bailey et al. 2017). Patterns of stress orientations (GPS) survey (e.g. Datta Gupta et al. 2015; Gupta & Singh 1986;
in the Indian subcontinent have been analysed previously (Gowd Gupta et al. 1986). This paper aims at reliable stress magnitude
et al. 1992; Srivastava & Bharali 2000; Chatterjee & Mukhopadhay determination from well logs, measured pore pressure (PP) and
2001, 2002, 2003; Singha & Chatterjee 2014, 2015). However, stress leak-off test (LOT) data for six wells distributed under the study
magnitude data for some locations in India including the Upper area. Being the first commericially producing basin of India, esti-
Assam basin have not yet been analysed and/or incorporated so far. mation of tectonic strain and stress magnitudes of the Upper Assam


C The Author(s) 2018. Published by Oxford University Press on behalf of The Royal Astronomical Society. 659
660 J. Alam, R. Chatterjee and S. Dasgupta

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 1. (a) Study area with respect to India, (b) brief tectonic setting of northeastern India (after Nandy 2001; Raoof et al. 2017) and (c) location of six wells
under the study area. NT, Naga Thrust; DT, Disang Thrust.

basin has vast application in geodynamics as well as conventional Three wells, namely, M1, M2 and M3 are located in the Assam
and unconventional hydrocarbon exploration and exploitation. Main Gap below a tributary of the Bramhaputra river while the rest three
objectives of our study is to estimate PP, tectonic strain and in situ N1, N2 and J1 are located in the Belt of Schuppen below the NT
stress magnitudes using poroelasic models. An alternative method is (Fig. 1). Major geological horizons through M1–M3 and N1–N2
proposed to estimate minimum horizontal compressive stress using are displayed through the seismic sections for normal faulted and
multiple linear regression model. thrust faulted region (Figs 2a and b). The wells, M1, M2 and M3
penetrate geological formations from top Alluvial through Dhekia-
juli, Girujan, Tipam, Barail, Kopili, Sylhet, Langpur to the granitic
2 B R I E F T E C T O N I C S O F S T U DY A R E A basement (Fig. 2a), while the wells, N1 and J1 in thrusted zone
cross through Girujan Suprathrust, Tipam Suprathrust, Argillaceous
The study area (between 27◦ N–28◦ N and 95◦ E–96◦ E) falls in Barail Suprathrust, Barail Subthrust and then reach the NT. The well
the Eastern Himalayan zone that comprises of Sikkim Himalaya, N2 reaches Tipam top near Margherita Thrust (Fig. 2b). A series of
Bhutan Himalaya, Arunachal Himalaya and the Eastern Himalayan N–S striking normal faults has been found to exist in the seismic gap
Syntaxis (EHS) (e.g. Molnar 1984; Molnar & England 1990; Yin & between 92◦ E–94◦ E and 26◦ N–28◦ N (Armijo et al. 1986; Bendick
Harrison 2000; Yin 2006; Raoof et al. 2017). EHS is composed of & Bilham 2001; Mullick et al. 2009).
thrusted older crystalline complex along the main boundary thrust
and main central thrust (MCT; GSI 2000). EHS is the meeting place
of the great Himalayan arc and the Indo–Burmese Arc. The main
3 E S T I M AT I O N O F S T R E S S
features of this syntaxis zone are Mishmi Thrust, Tidding Suture,
M A G N I T U D E S A N D O R I E N TAT I O N
Lohit Thrust and a part of Disang Thrust (DT) and NT. The NE–SW
trending ‘Belt of Schuppen’ is bounded by the NT in the northwest Lithospheric stresses have been defined in terms of the orientation
and the DT in the southeast, extending from Lower Assam in the and magnitude of three orthogonal principal stresses, namely: max-
south to the Mishmi Hills in the north, running parallel to MCT imum, minimum and intermediate. The vertical stress (SV ) caused
belt (Angelier & Baruah 2009). The Upper Assam basin forms due to the mass of overburden is assumed to be one of the three
a southeast-dipping shelf in the foreland part of Assam–Arakan principal stresses. The two remaining principal stresses are referred
Basin. This is bordered by the shield of Mikir hills towards its west as the maximum (SH ) and minimum (Sh ) horizontal stresses acting
and Mishmi hills along its northeastern boundary. About 4–6 km in the horizontal plane. Tectonic regimes such as: extensile, com-
thick sediments are deposited belonging to Tertiary and Quaternary pressive and strike-slip are determined by the relative magnitudes
age overlying basement (Sawai & Sar 2006; Saha 2011). Major of these three principal stresses (Zoback et al. 1989). The aim of
amount of conventional oil reserves are found in this foreland basin this study is to characterize the local stress regime under the study
producing for more than a century. area using poroelastic model.
Tectonic Strain and Stress Magnitude in Assam 661

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 2. Seismic section (a) passing through wells M1 and M3 in normal faulted sediments and (b) traversing through wells N1 and N2 in thrust faulted
sediments. Major geological horizons have been delineated in these sections along with thrusts that are acting in their vicinity. Normal faulted and thrusted
sediments are shown in panels (a) and (b). Naga Thrust and Margherita Thrust have been annotated in panel (b) (after Dakua et al. 2008).

3.1 Estimation of vertical stress proposed the following empirical relation:


Stress magnitudes are constrained using wellbore geophysical logs    U1  B
and well test data. For estimation of stress magnitudes and PP study, σ
V = V0 + A σmax , (2b)
well logs such as: gamma ray (GR), density (ρ), compressional σmax
sonic traveltime (t), ratio of compressional and shear sonic veloc-
ity (Vp /Vs ), formation microimager (FMI) logs and well test data: where,
modular dynamic tester (MDT), mud weight (MW), LOT have been   B1
υmax − 1500
used for six wells under the study area. Vertical stress magnitudes σmax = , (2c)
from six wells: M1, M2, M3, N1, N2 and J1 are derived from density A
logs using the following eq. (1) (Plumb et al. 1991):
where, U is the unloading parameter. σ max and υ max represent the
 z values of maximum effective stress and maximum velocity at the
SV = ρ (z) gdz, (1) onset of unloading, respectively (Bowers 1995).
0
Eaton’s sonic equation (Eaton 1972, 1975) to compute PP is
where z = depth at point of measurement, given below:
ρ (z) = bulk density of the rock, which is function of the depth
and PP = SV − (SV − Ph ) ∗ (tn /t)n . (3)
g = acceleration due to gravity
where, Ph = hydrostatic pressure gradient is assumed as 10 MPa
km−1 ,
t = compressional sonic traveltime,
tn = traveltime computed from normal compaction trend (NCT)
3.2 Estimation of pore pressure and
The fluid pressure within pore space of a rock is referred to as PP. It is n = Eaton’s exponent,
necessary to determine the PP correctly as PP and horizontal stress Both methods depend on pressure mechanism and require knowl-
are interconnected. The pressure that is exerted by a static column edge of depositional history of the basin (Azadpour & Manaman
of fluid is called the hydrostatic pressure and it varies with the 2015). However, Eaton’s equation is directly calibrated in depth–
fluid density (Swarbrick & Osborne 1998). There are two methods pressure space where the data are collected whereas, Bower’s equa-
commonly used to determine PP from well logs: Eaton method and tion requires transforming the data to VES–velocity space. Gutier-
Bowers method (Eaton 1972, 1975; Bowers 1995). The Eaton’s rez et al. (2006) show that the selection of one PP prediction model
sonic method is quick and directly applicable to estimate PP. The over another primarily depends on convenience and performance
computed PP based on the compaction trend can easily be calibrated in a particular geological setting. Both models are expected to per-
with the measured pressure from MDT and MW. Another popular form well in this setting and Eaton’s sonic equation (3) is chosen to
method, the Bowers method employs vertical effective stress (VES). compute PP for six wells under this area for simplicity.
In compaction disequilibrium conditions, Bowers (1995) pro- The first step in determining the PP from sonic log is to esti-
posed an empirically determined method to calculate the effective mate the NCT. Consolidation and compaction of clastic sediments
stress as follows: due to its burial causes loss of porosity. Compressional sonic trav-
eltime (t) in formations depends on rock lithology, mineralogy,
V = V0 + Aσ B (2a) porosity, clay content, fluid saturation, stresses and temperature (Lei
et al. 2012). The compaction property observed in sonic traveltime
where, V is the velocity at a given depth and V0 stands for the surface against low-permeable shale follows a linear/exponential trend with
velocity (normally 1500 m s−1 ), σ represents the VES, and A and B depth and is termed as NCT. The behaviour of NCT is observed with
are the parameters obtained from calibrating regional offset velocity loss of porosity as sediments are loaded with hydrostatic pressure
versus effective stress data. In unloading conditions, Bowers (1995) and only the trend from such sediments is used to establish the
662 J. Alam, R. Chatterjee and S. Dasgupta

direct pressure measurement data is not available (Law & Spencer


1998; van Ruth et al. 2002). MDT data are available to calibrate the
estimated PP from t log. A chi-square statistic has been obtained
to bring out the relation between the measured pressure from MDT
and the computed PP using Eaton’s sonic equation. A chi-square
statistic is a single number that provides information regarding how
much difference exists between measured data and computed data.
The formula used is (Koch & Link 1970)

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
n (PPMi − PPEi )2
χ2 = , (4)
i=1 PPEi
where χ 2 is the chi-square statistic,
PMi = the measured PP from the MDT and
PPEi = PP estimated from Eaton’s equation.
Figure 3. The normal compaction trend (NCT) for combining three wells: Table 1 lists the chi-square statistic (χ 2 ) for fitting between PPMi
M1, M2 and M3 in the normal faulted regime. Single exponential trend and PPi with varying Eaton’s exponent from 0 to 3 for six wells. A
describes the compaction trend. Shale points are used to draw NCT for the low value of chi-square is inferred as a higher correlation between
depth interval 500–1600 m, which is under hydrostatic condition. two data sets as compared to a high value of chi-square statistic. It is
observed that in wells M1, M2, M3 and N2, χ 2 is lowest for PP es-
timation with Eaton’s exponent 1. In wells N1 and J1, it is observed
baseline behaviour for the Eaton method. Selection of depth in- that χ 2 is minimum for PP estimation with Eaton’s exponent 0.5
terval against shale with similar lithological properties is the most following nearly hydrostatic condition at shallow depth. As we lack
critical criteria to draw NCT from t log of individual wells under MDT values at deeper depth for wells N1 and J1, we use Eaton’s
the study area. GR logs are used for delimiting shale points from exponent 1 for further PP determination under the study area.
well logs. Setting GR log values from 70 to 120 API, we delimit For the sake of simplicity, PP predicted with Eaton’s exponent 0,
shale points from all wells. NCT for PP modelling using Eaton’s 1, 2 and 3 are shown for four wells: M1, M3, N1 and J1, respectively
method considers the depth ranges that are hydrostatic. The NCT (Figs 5a–d). The Eaton’s exponent is essentially an ‘amplification
in t log has been found to follow an exponential relationship for factor’ that links the magnitude of the sonic velocity anomaly ob-
selected depth intervals against shale points for six wells. served to the magnitude of the overpressure. Large value of ex-
It is noted that a single NCT can be fitted for the shale points ponent ‘n’ magnifies the effect of changes in compressional sonic
combining three wells, M1, M2 and M3 that are in close proximity traveltime implying that large ‘n’ corresponds to a small or weak
located in normal faulting regime. The normally compacted shale sonic response to changes in PP. An exponent of 3.0 or higher is
points (depth interval: 500–1600 m), from these three wells have the normal relationship between overpressures generated by dise-
been used to draw the combined NCT as shown in Fig. 3. However, quilibrium compaction and their response on the sonic log (Tingay
separate NCTs are drawn to fit shale points of wells N1, N2 and et al. 2003). Places such as tertiary deltas show very large sonic
J1, respectively, in thrust faulting regime (Figs 4a–c). These wells velocity anomalies in association with overpressure, and they need
are quite distant from each other. It is observed from Fig. 4 that to be amplified by an exponent of 3 to match the predicted PP (e.g.
individual exponential equations fit the compaction trends in well Swarbrick & Osborne 1998; Tingay et al. 2003, 2008; Singha &
N2 and J1, while two exponential relations are required to fit the Chatterjee 2014; Dasgupta et al. 2016). We require to use an expo-
traveltime against shale in well N1. A single trend is fitted through nent of 1.0 because the sonic log anomalies used to predict PP herein
the shale points upto the depth of 1538 m (thrusted sediments) are much lesser than what is normally observed with disequilibrium
and another trend is drawn to explain the compaction from 1538 compaction. Small value of ‘n’ such as 1 used in this case study di-
to 3775 m in well N1. Therefore, two separate NCTs are used to minishes the amplification effect and suggests the sonic response is
illustrate the shale compaction for the deep well, N1 penetrating larger than typical for a comparable change in PP. A very low sonic
sediment upto 3775 m (Fig. 4a). The seismic section through well Eaton’s exponent would suggest that PP follows the hydrostatic and
N1 shows thrusted sediments in the upper part of the well till depth is dependent on the sonic velocity in these formations.
of 1538 m, which may be the reason for dual NCT in this well The estimation of horizontal stress magnitude requires the infor-
(Fig. 4a). Figs 4(b) and (c) show the NCTs for the wells N2 and J1, mation on PP, vertical stress and tectonic strain under the study area.
respectively in the thrusted sediments (Figs 4b and c). It is very com- Conventional techniques for estimating the maximum and minimum
mon to use a single exponential trend, normally applied for shales horizontal stresses are based on analysing borehole breakouts (BOs)
deposited of similar age, environment and provenance throughout and leak-off pressure necessary to fracture the surrounding forma-
the basin (Tingay et al. 2003). However, it is best to use separate tion, respectively (Gough & Bell 1981; Bell & Gough 1982; Zoback
NCT for varying depositional environments, unconformities and et al. 1985; Moos & Zoback 1990; Raaen et al. 2006). We shall es-
uplifted sediment section (Bjørlykke 2014). There is a deviation of timate horizontal stress magnitudes using frictional limiting theory
t from NCT amounting 15–18 μs ft−1 at about 4000 m in well and poroelastic models in the next section.
N1, which may be attributed to lithology change or disequilibrium
compaction. Other wells show shale points more or less coinciding
with the NCT with slight deviation from NCT.
3.3 Estimation of horizontal stress magnitudes
PP has been calculated varying Eaton’s exponent (n) as 0.5, 1,
1.5, 2 and 3 using t log. The most reliable and direct pressure mea- The Earth’s crust contains different types of structural variation like,
surement can be obtained mainly from the drill stem test, MDT and distributed fault systems, fractures and discontinuities of varying
repeat formation test. MW can be used as a proxy for the PP where scales and orientations (Chester & Chester 2000; Zoback 2007).
Tectonic Strain and Stress Magnitude in Assam 663

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 4. Presents the NCT for three wells (a)N1, (b)N2 and (c)J1 which belong to the thrust faulting regime. NCT for well N1 is fitted with two exponential
trends while for well N2 and J1, it is individually fitted with a single exponential trend.

Table 1. Chi-square statistic (χ 2 ) for varying pore pressure with varying Eaton’s exponent.
χ 2 for Pore pressure χ 2 for Pore pressure χ 2 for Pore pressure χ 2 for Pore pressure
χ 2 for Pore pressure calculated with calculated with calculated with calculated with χ 2 for Pore pressure
calculated with Eaton’s exponent Eaton’s exponent Eaton’s exponent Eaton’s exponent calculated with
Well Eaton’s exponent 0 0.5 1.0 1.5 2.0 Eaton’s exponent 3.0
M1 1.43 0.69 0.39 0.49 1.13 3.11
M2 0.66 0.73 0.05 0.20 0.55 0.72
M3 0.60 0.31 0.10 0.13 0.17 0.56
N1 2.04 0.19 0.52 1.19 2.16 6.27
N2 1.76 0.60 0.41 1.72 1.53 4.37
J1 0.34 0.28 0.47 0.74 1.16 2.10

Frictional strength of faults, fractures and discontinuity limit the where, μ is the coefficient of internal friction, assumed to be
stress magnitudes at depth. Several observations around the world 0.6 (Byerlee 1978; Zoback & Healy 1984) for the Upper Assam
show that stress magnitudes in the crust are in equilibrium with basin.
its frictional strength (e.g. Zoback & Healy 1984; Townend & Eqs (5) and (7) have been used for computation of lower and
Zoback 2000). Frictional limiting theory suggests the relationship upper bounds of Sh and SH , respectively. The lower bound of Sh in
between the effective stress ratio and frictional strength of rock normal faulting regime varies from 14.1 to 15.0 MPa km−1 while in
under following fault regimes (Anderson 1951; Jaeger & Cook reverse faulting regime the upper bound of SH ranges from 47.8 to
1979): 51.3 MPa km−1 . Frictional limiting theory (Jaeger & Cook 1979)
considers that the maximum horizontal stress is limited by the fric-
 2
SV − PP  2 1/ tional properties of rock. This theory overestimates the horizontal
Normal faulting ≤ μ +1 2 +μ , (5)
Sh − PP stress magnitudes (Bailey et al. 2017).
Rocks may behave inelastically for significant periods in their
history with material properties that vary with time depending
 2 on consolidation and diagenesis (Blanton & Olson 1999).The
SH − PP  1 Blanton–Olson approach determines tectonic strain assuming
Strike − slip faulting ≤ μ2 + 1 /2 + μ and (6)
Sh − PP porous medium and uses static Young’s Modulus, Poisson’s Ratio,
SV , PP and LOT data. They assume that the strain in the direction
orthogonal to the applied tectonic stress is zero. Blanton & Olson
 2 (1999) have developed new constants (say, C1 and C2 ) involving
SH − PP  1
Reverse faulting ≤ μ2 + 1 /2 + μ , (7) static Young’s modulus (Ys ) and Poisson’s ratio (ν) characterizing
SV − PP
664 J. Alam, R. Chatterjee and S. Dasgupta

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019

Figure 5. The hydrostatic pressure (Ph ), predicted pore pressure (PP) from Eaton’s sonic equation with varying exponent of 1 (PP1), 2 (PP2) and 3 (PP3)
along with vertical stress (SV ), mud weight (MW) and modular dynamic tester (MDT) data for wells (a) M1, (b) M3, (c) N1 and (d) J1, respectively. Here, PP
using Eaton’s exponent 0 presents hydrostatic pressure.
Tectonic Strain and Stress Magnitude in Assam 665

the rock properties. using eq. (14):


Ys Sh = ν C1 εtect + C2 . (14)
C1 = (8)
1 − ν2
Only a few LOT data point (one from each well, total six) are used
to calculate tectonic strain, the rest LOT points (eight from all six
νSV + (1 − 2ν) α.PP + αT .Ys .T wells) are used for validation of estimated Sh . A good match is found
C2 = , (9)
1−ν between the LOT and Blanton–Olson model predicted Sh (Singha
where C1 , C2 = Blanton–Olson constants, & Chatterjee 2014). It is observed from Table 2 that tectonic strain

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
α = Biot’s constant, (εtect ) varies from −0.0032 to −0.0044 in normal faulted region
αT = thermal coefficient of expansion and while it varies from 0.0024 to 0.0061 in thrust faulted region. The
T = geothermal gradient considered as 68◦ F km−1 (Barman thrust faulted regime is characterized by a higher value of tectonic
et al. 2015). strain as compared to the tectonic strain of normal faulted regime
Biot’s constant has been assumed as 1.0 (Bachman et al. 2011). (Table 2).
In the absence of measured value for the thermal coefficient of Linear poroelastic model considers horizontal strain in two Carte-
expansion α T , 5.56E-6/F can be used for sandstones, 5.00E-6/F for sian directions. The maximum (εx ) and minimum (ε y ) tectonic strain
shale (Song 2012). The effect of the thermal term is reflected in the is obtained by solving the linear poroelastic equations for the Upper
computation of C2 in the 5th decimal place. Therefore due to its Assam basin (Al-Qahtani & Rahim 2001):
negligible effect, thermal term is discarded. Eq. (9) is a simple form
1 − νs2 νs
of generalized poroelastic equation when one lateral strain is set to νs εx + ε y = Sh − (SV − PP) − PP , (15)
Ys 1 − νs
zero.
Static values of elastic constants are based on the measurement
1 − νs2 νs
of deformation induced in a material by a known force unlike the νs ε y + εx = SH − (SV − PP) − PP , (16)
dynamic constants that involve measure of ultrasonic body wave Ys 1 − νs
velocities (Zamora et al. 1994; Fjaer 2009). There are several empir- where, εx and εy are the horizontal strain along NE–SW and NW–
ical relations to convert dynamic moduli to static moduli. Amongst SE directions, respectively. The tectonic strains are necessary in
these, we consider the relation proposed by Wang (2000). Static poroelastic model to compute horizontal stress magnitudes using
Young’s modulus (Ys ) is calculated from dynamic Young’s modulus eqs (15) and (16). First, we use the measured Sh from LOT and SH
(Yd ) using the following relation (10) (Wang 2000): from BO data for computation of tectonic strains at specific depths.
The Sh magnitude is constrained using data from LOT (Zhang
Ys = 0.4142 ∗ Yd − 1.0593, (10a)
& Roegiers 2010). LOT measurements are not available at the BO
depths of wells M1, M3 and J1. Hence, available LOT values for
well M1, M3 and J1 are plotted against depth showing loci of LOT
νs = νd . (10b) for these wells. This LOT trend is used to estimate the value of
The static (ν s ) and dynamic (ν d ) values of Poisson’s ratio are Sh for three wells, at depths where actual measurement of LOT is
same (Wang 2000). unavailable (Figs 6a–c).
Yd and ν d have been calculated from Vp /Vs ratio and density logs Image log evaluation is preferred to constrain maximum horizon-
for six wells under the study area using the following relations (11) tal stress (SH ) at the BO interval (Chhajlani et al. 2002). BOs are
and (12) (e.g. Das & Chatterjee 2017): wellbore enlargements caused by stress-induced failure of a well

2 occurring 180◦ apart. Stress-induced BOs observed on FMI logs
Vp
1 Vs 2 − 2 are shear failures of the borehole wall in response to horizontal
νd =
2 , (11) compression. This failure occurs at the azimuth of Sh , which is
2 Vp − 1
Vs 2 the azimuth at which the circumferential compressive stress (Hoop
Stress) is highest (Zoback et al. 1985; Barton et al. 1988; Bell 1996;
 Tingay et al. 2008). FMI log data are available for two wells M1
ρVs 2 3Vp 2 − 4Vs 2
Yd = (12) and M3 only. SH magnitude can be obtained from BO interval using
Vp 2 − Vs 2
FMI log with known information of Sh and unconfined compressive
The Ys and ν s are used for computation of tectonic strain for all strength (UCS) of rocks (Plumb et al. 2000) for these wells. The
six wells under the study area. C1 and C2 are calculated for the BO intervals such as: 3990–4032 m from well M1 and 2658–2800
entire length of all wells. m from well M3 are chosen from FMI logs with interpolated LOT
The tectonic strain and minimum horizontal stress is given by values from Fig. 6. Figs 7(a) and (b) show the typical examples of
(Blanton & Olson 1999; Song 2012) BO for well M1 and M3, respectively (Fig. 7). The FMI log data for
well J1 are not available to us but Srivastava & Bharali (2000) have
Sh  − C2 
εtect = , (13) reported BO interval of 2180–2830 m for well J1 in this thrusted
ν C1  area.
where, εtect = tectonic strain. The magnitude of SH at the BO intervals has been computed
The primes in eq. (13) indicate that Sh , C1  and C2  are the values using interpolated LOT (as Sh ), PP and UCS data using eq. (17),
of Sh , C1 and C2 , respectively, associated with a specific depth where maximum circumferential compressive stress (Hoop Stress)
at which LOT is available. LOT at a particular depth is generally is equal to UCS as (Barton et al. 1988)
assumed to be equivalent to the Sh value (Bailey et al. 2017). The
3SH −Sh −2PP−P = UCS, (17)
values of Sh , C1  , C2  and εtect have been listed in Table 2 along with
the depth points at which they are calculated. This tectonic strain where, P is the difference between wellbore pressure and forma-
is further considered to compute Sh for the total length of the well tion fluid pressure.
666 J. Alam, R. Chatterjee and S. Dasgupta

Table 2. Lists the leak off test (LOT) values at specific depth points with Poisson’s ratio (ν), C1 , C2 and uniaxial tectonic strain (ε tec ) for six wells, Upper
Assam basin.
Blanton & Olson
(1999) method N1 N2 J1 M1 M2 M3
Depth (m) 1827 950 1312 2711 2207 2791
LOT (MPa) 41.36 21.25 28.50 35.03 33.02 33.00
ν 0.35 0.35 0.28 0.32 0.37 0.33

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
C1 5502.74 4968.61 12238.45 6200.19 3498.26 7139.43
C2 29.66 17.04 18.74 42.22 37.10 43.26
ε tect 0.0061 0.0024 0.0027 −0.0036 −0.0032 −0.0044

Figure 6. The leak-off test (LOT) trend considered for the computation of tectonic strain in poroelastic model for three wells: (a) M1 (b) M3 and (c) J1,
respectively.

Traditionally, rock strengths are obtained directly from labora- Similarly, the strain εy ranges from −0.00081 to −0.0011 in well
tory tests on core samples. But in absence of core samples, rock M1, −0.0010 to −0.0019 in well M3 and from 0.00051 to 0.00053
strength parameter and elastic moduli can be obtained from empir- in well J1. We have used the smallest as well as largest values of
ical relationship based on well log data. For determination of SH , εx and εy in each well to calculate respective values of horizontal
the value of UCS is obtained from Vp for these three wells using stress magnitudes. The stress magnitudes in each case (smallest and
eq. (18) in BO regions using FMI log (Horsrud 2001): largest values of εx and εy ) are found to differ by about 2 MPa, so
we have assumed the largest value of tectonic strains (Table 3b) for
UCS = 0.77∗Vp 2.93 . (18) estimation of Sh and SH for entire length of individual wells M1,
The formation where BO is encountered in Barail formation and M3 and J1. Table 3b lists interpolated LOT as Sh  and BO derived
is mostly composed of Argillaceous shale with minor surges of silt, SH  and largest value of tectonic strains at specific depths for three
intercalation of sand and shale layers belongs to Oligocene age. We wells. The largest value of strains ε x and εy vary from 0.000099 to
have core derived UCS values for this formation adjacent to wells 0.00077 and from −0.0011 to −0.0019 along NE–SW and NW–
M1 and M3 only for validation of UCS value obtained from log SE directions, respectively, in normal faulted regime whereas ε x is
data (Oil India Ltd, unpublished report). Therefore, magnitude of 0.00092 along NE–SW and εy is 0.00053 along NW–SE direction
SH has been computed at the BO intervals for three wells shown for well J1 in thrust faulted regime. Higher lateral strain is observed
in Table 3a. Table 3a shows the information on lithology with their in thrust faulting regime as compared to the lateral strain of normal
geological age, ranges of UCS and computed SH for the BO interval faulted regime (Table 3b). The computed tectonic strain in thrust
for these wells. faulted regime is positive NE–SW as well as NW–SE direction
Tectonic strains are computed at the BO intervals as mentioned in whereas tectonic strain in normal faulting regime is negative along
Table 3a. The strain εx ranges from 0.00040 to 0.00077 in well M1, NW–SE direction. The well M2 is in the vicinity of M1 and M3
0.00005 to 0.000099 in well M3 and 0.00069 to 0.00092 in well J1. lying in the normal faulted region while N1 and N2 share the same
Tectonic Strain and Stress Magnitude in Assam 667

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 7. The FMI log for wells (a) M1 and (b) M3 depicting breakout (BO) intervals for derivation of SH.

Table 3a. Breakout interval for three wells with lithology, Unconfined compressive strength (UCS) and maximum horizontal stress (SH ).
Well M1 M3 J1
Breakout Interval (m) 3990–4032 2658–2800 2370–2473
Lithology Agillaceous Barail dominated unit Barail shale with minor surges of silt Barail shale–shaly sand
shale sand intercalation
Age Oligocene Oligocene Oligocene
UCS (MPa) 43.73–49.34 20.09–29.28 35.17–49.79
SH (MPa) 63.42–63.95 38.29–40.57 45.37–48.99

Table 3b. Results of poroelastic modelling for obtaining tectonic strain along x- and y-axes for three wells in the Upper
Assam basin.
Poroelastic model M1 M3 J1
Sh (MPa) 50.48 32.54 44.99
SH (MPa) 63.42 39.68 48.45
v 0.25 0.34 0.28
Ys (GPa) 13.38 5.20 11.75
ε x (maximum tectonic strain) 0.00077 0.000099 0.00092
ε y (minimum tectonic strain) −0.001 −0.0019 0.00053
Sh = minimum horizontal stress, SH = maximum horizontal stress, v, Poisson’s ratio, Ys = static Young’s modulus, ε x
and ε y = tectonic strain in NE–SW and NW–SE direction.

thrusted regime as in well J1. These tectonic strains are further used with MDT, MW, LOT and BO derived SH for all wells (Fig. 8). The
along with other parameters in eqs (19) and (20) to derive minimum magnitude of Sh is validated with LOT data for all wells while that
and maximum horizontal stresses for six wells, M1, M2, M3, N1, of SH is validated with BO derived SH for three wells. Poroelastic
N2 and J1 in the Upper Assam basin: model driven Sh has a good match with LOT. This predicted SH is in
good agreement with the BO derived SH . Owing to non-availability
νs νs Ys Ys
Sh = (SV − PP) + PP + εx + ε y and (19) of image or dipmeter logs for wells, M2, N1 and N2 falling under
1 − νs 1 − νs 2 1 − νs 2
the study area validation of SH is not possible but LOT values serve
as a check on Sh values for these three wells. The poroelastic model
νs νs Ys Ys derived Sh magnitude for well N1 underestimate the LOT values
SH = (SV − PP) + PP + εy + εx . (20)
1 − νs 1 − νs 2 1 − νs 2 throughout the wells. A mismatch is observed between the LOT
Fig. 8 displays the hydrostatic pressure, PP, SV , SH and Sh along and model predicted Sh values at 2071 m of well M1. However, it
668 J. Alam, R. Chatterjee and S. Dasgupta

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 8. The comparative pressure profiles of pore pressure (PP), hydrostatic modular dynamic tester (MDT), mud weight (MW), leak-off test(LOT),
maximum horizontal stress (SH ) derived from breakout interval, minimum horizontal stress (Sh ) derived from Blanton & Olson (1999) model, Sh and SH from
biaxial poroelastic model and vertical stress (SV ) for wells (a) M1, (b) M2, (c) M3, (d) N1, (e) N2 and (f) J1. For poroelastic modelling of horizontal stress,
well M2 uses strain values calculated from well M1 while N1 and N2 use strain values calculated for well J1 as they share similar stress regimes.

shows quite encouraging match at other calibration points where with gradient varying from 21.7 to 23.8 MPa km−1 . The horizontal
LOT is available. Blanton–Olson derived Sh shows a slightly closer stresses show higher value than the SV (SH > Sh > SV ) for wells N1,
match with LOT. Both the models are estimating Sh satisfactorily N2 and J1 for upper part of the well due to lateral compression in
for all wells except well N1 (Fig. 8). Presence of hydrofrac or BO the thrusted zone. The rate of increase of horizontal stress is less
data would have assisted to explain the lowering of Sh magnitude as compared to the rate of increase of SV with depth. It is observed
for well N1. from the poroelastic model that Sh gradient varies from 12.0 to
Table 4 shows the PP, stress gradients, stress ratio and inferred 15.2 MPa km−1 in normal faulted region while it reaches maximum
faulting regimes for these six wells. PP is observed to be hydro- of 21.4 MPa km−1 in thrust faulted region. Similarly, SH gradient
static for geological formations: Dhekiajuli, Girujan, Tipam, Barail varies from 14.9 to 16.4 MPa km−1 in normal faulted region while
and Kopili formations. PP gradient of wells, M1, M2 and N2 be- in thrust faulted region it varies from 16.7 to 23.1 MPa km−1 . Using
comes slightly high ranging from 10.7 to 11.9 MPa km−1 in Sylhet the biaxial poroelastic model, it is observed that the average Sh /SV
and Langpur formations. The SV magnitude increases with depth varies from 0.86 to 1.02 and from 0.68 to 0.72 in thrust and normal
Tectonic Strain and Stress Magnitude in Assam 669

Table 4. Lists pore pressure and other stress gradients from different models under different stress regimes.
Sh (Blanton &
Stress gradient from Olson, 1999)
Well and depth PPGradient SV Gradient Frictional theory gradient Poroelastic stress
interval (MPa/km) (MPa/km) (MPa/km) (MPa/km) gradient (MPa/km) Sh /SV SH /SV Faulting Regime

Sh SH Sh SH
N1 (690–4309 m) 10.10 23.30 51.30 17.30 16.10 16.70 0.86 0.92 Thrust to Normal

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
faulting
N2 (236–1510 m) 11.60 23.80 49.90 20.80 21.40 23.10 1.01 1.09 Thrust faulting
J1 (600–1300 m) 10.30 22.40 47.80 17.00 17.80 19.70 1.02 1.18 Thrust to normal
faulting
M1 (408–4189 m) 10.70 22.20 14.40 11.70 12.50 16.40 0.72 0.82 Normal faulting
M2 11.9 21.70 15.00 14.80 15.20 16.10 0.72 0.83 Normal faulting
(1109–2835 m)
M3 (435–2799 m) 10.50 21.80 14.10 12.80 12.00 14.90 0.68 0.76 Normal faulting
PP=pore pressure, SV =Vertical stress, Sh = minimum horizontal stress, SH =maximum horizontal stress.

faulted regions, respectively. An average SH /SV varies from 0.92 to groups in this analysis and the alternative hypothesis simply states
1.18 and from 0.76 to 0.83 in thrust and normal faulted regions, that at least one of the parameters mj = 0, j = 1, 2,3.
respectively. It is noted that the lower bound of Sh calculated from The t-test is implemented to compare individual means. The
frictional limiting theory does not match with the measured Sh from two variables are linked using an unstandardized coefficient us-
LOT in the normal faulting regime. The frictional limiting theory ing linear regression. This can be generalized to multiple regres-
predicted Sh and SH gradients overestimate the value of Sh and SH sions taking into consideration several covariates and simultane-
computed from poroelastic model. ously trying to explain their collective relationship to the outcome.
Fig. 9 displays the Blanton–Olson predicted Sh versus LOT and The model summary Table 5b shows the F-statistic is equal to
poroelastic model predicted Sh versus LOT plots showing variation 957.227/95.907 = 9.981. The distribution is F (3, 6), and the prob-
of goodness of fit (R2 ) from 0.98 to 0.99, respectively. Standard ability of observing a value greater than or equal to 9.981 and
deviation (sd) of residual error is minimum (±1.39) for prediction significance (sig.) of this test has value of 0.010. There is strong ev-
of poroelastic model derived Sh . For Sh , predicted using Blanton– idence that m1 is not equal to zero. The ratio between sum of squares
Olson model the sd of residual error is ±1.43. This implies that for regression model (SSM) and total sum of squares (SST), that
Sh estimated using poroelastic model fits better with the measured is, SSM/SST = R2 is known as the squared multiple correlation co-
LOT data. efficient (Table 5c). Therefore, R2 = 2871.680/3447.124 = 0.833.
Adjusted R2 in Table 5c is computed using the formula 1 − ((1 –
R2 )∗(N − 1))/(N − 3 − 1)) where N is 10. Standard (Std.) error
of the estimate is referred to as the root mean squared error. It is
4 M U LT I P L E R E G R E S S I O N A N A LY S I S the square root of the mean square for the residuals in Table 5c.
FOR PREDICTION OF Sh In this linear regression model the estimated raw or unstandardized
regression coefficient for a predictor variable (referred to as B in
Another estimate of Sh is demonstrated to constrain stress magni-
Table 5d) is interpreted as the change in the predicted value of the
tude and tectonic strain of any faulted regime. It is observed that the
dependent variable for one unit increase in the predictor variables:
Sh varies with PP, vertical stress and the ratio of seismic velocities
SV , PP and (Vp /Vs )2 . If B = 1, for every unit increase in predictor,
of the medium (Eberhart-Phillips et al. 1989; Hillis 2000; Vernik &
the dependent variable also increases by 1. Measure of dependent
Kachanov 2010). Amongst six wells, the wells N1, N2 and M1 have
variable (Sh ) when all predictors are set to zero is got by the constant
maximum numbers of LOT data as compared to the rest of the wells.
term (Table 5d). The intercept or constant term (Table 5d) provides
The data from these three wells are used for multiple regression anal-
the predicted value of the dependent variable when all predictors are
ysis for computation of Sh . A single regression model is proposed
set to zero. The constant coefficient (20.025) is this simple mean of
for a dependent variable (Sh ) using three independent variables: SV ,
group means. Impact of each predictor is reflected by knowing the
PP and (Vp /Vs )2 for the above-mentioned three wells distributed in
values of standardized coefficient, beta as well as the significance
the Upper Assam basin. Table 5a lists the well names, LOT with
(sig.) for each predictor (Table 5d).
respective depths, SV , PP and (Vp /Vs )2 for the statistical analysis
The beta value measures the effect of each predictor variable on
and is showing the values for the dependent and independent vari-
Sh . For example, a beta value of 0.459 of SV indicates that a change
ables. The technique, analysis of variance (ANOVA), evaluates the
of one sd in the predictor variable will result in a change of 0.459 sds
relationship between a dependent variable and three independent
in the dependent variable. So, a higher beta value indicates greater
variables. This is carried out using IBM SPSS Statistical Software,
effect on dependent variable.
version 21.0.
Multiple linear regression attempts to model the relationship be-
The ANOVA calculations for multiple regressions as shown in
tween three independent variables and a dependent variable (Sh ) by
Table 5b, the model degrees of freedom (df) are equal to 3, the error
fitting a linear equation to observed data:
(residual) df, that is, the df for residual are equal to 6 and the total df
are 9. The null hypothesis for ANOVA (Koch & Link 1970) states
that means of group such as m1 , m2 and m3 for SV , PP and (Vp /Vs )2 ,
 2
respectively are equal to zero, that is, m1 = m2 = m3 = 0. The in- Vp
Sh = 20.025 + 0.317SV + 0.533PP − 2.117 . (21)
dependent variables namely; SV , PP, and (Vp /Vs )2 are three different Vs
670 J. Alam, R. Chatterjee and S. Dasgupta

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 9. The regression plot between (a) Blanton–Olson and (b) poroelastic model predicted minimum horizontal stress (Sh ) and leak-off test (LOT),
combining data from wells M1, M3 and J1 with best-fit regression line and standard deviation (sd) of residual error.

Table 5a. Independent and dependent variables utilized for generation of multilinear regression model.
Well name Depth (m) LOT (MPa) SV (MPa) PP (MPa) (Vp /Vs )2
N1 691 9.30 12.94 7.22 3.99
N1 1827 41.36 39.45 18.55 4.27
N1 3140 62.36 70.06 30.56 3.47
N2 239 3.15 5.37 2.97 7.64
N2 950 21.01 22.06 10.94 4.41
N2 1471 31.11 34.47 19.42 4.24
M1 686 9.10 13.94 7.05 9.74
M1 2711 35.03 58.10 28.50 3.85
M1 3897.79 54.05 85.64 44.55 4.12
M1 2071 29.02 43.83 22.93 5.29

Table 5b. Results from the Multiple Regression Model.


ANOVAa
Sum of
Model squares df Mean square F Sig.
1 Regression 2871.680 3 957.227 9.981 .010b
Residual 575.443 6 95.907
Total 3447.124 9
a Dependent variable: LOT.
b Predictors: (constant), (Vp /Vs )2 , PP, SV .

Table 5c. Model Summary.


Adjusted Std. Error of
Model R R2 R2 the Estimate
1 .913a .833 .750 9.79322
a Predictors: (constant), (Vp /Vs )2 , PP, SV .

Table 5d. Coefficients from Multiple Regression Analysis.


Coefficientsa
Standardized
Model Unstandardized Coefficients Coefficients t Sig.

B Std. Error Beta


1 (Constant) 20.025 13.974 1.433 .202
SV .317 .225 .459 1.409 .209
PP .533 .487 .345 1.095 .316
(Vp /Vs )2 −2.117 1.960 −.217 -1.080 .322
a Dependent Variable: LOT.
Tectonic Strain and Stress Magnitude in Assam 671

4.1 Result and validation magnitudes SV > SH > Sh , suggests association of normal fault in
deeper sections of wells N1 and J1. The well N2 reaches 1521 m in
Eq. (21) has been used to estimate Sh for another three wells: M2,
purely reverse faulting regime (Fig. 2b) showing the order of stress
M3 and J1 for testing of models under study. Fig. 10 displays the
magnitudes SH > Sh ≈ SV (Fig. 8e). The wells, N1 and J1, being
regression results between the predicted Sh with LOT, for wells M2,
deeper wells cross-cuts the NT on its way and characterize two dif-
M3 and J1 with R2 = 0.83. The sd of residual error in this regression
ferent faulting regimes (e.g. Fig. 2b for well N1). However, delving
model is found to be ± 4.44. The predicted Sh matches in a very
deeper into the wells we observe SV > SH > Sh suggesting a change
good manner with four LOT values whereas there is a mismatch
in regime to normal faulting system. This may be attributed to the

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
of about 5 MPa with one LOT value (Fig. 10). This new Sh model
fact based on previous observations that SV is not exactly the mini-
is now applied for these wells for computation of Sh with depth.
mum stress because the principal stresses in thrust systems can be
Fig. 11 illustrates the predicted Sh from regression model along
rotated near active thrust faults or detachments (Hafner 1951; Last
with LOT values and computed Sh using poroelasic model. As the
& McLean 1996). On the rotation of stress axes, the minimum stress
sd of residual error is minimum for Sh predicted using poroelastic
will be less than the vertical stress/overburden pressure (Couzens-
model, we have compared the regression model predicted Sh with
Schultz & Chan 2010). Though rotation of principal stress axes is
that predicted from poroelastic model. Eq. (19) (with strain set to 0)
likely to be present but it is unlikely to be the major driver from
is a specific instance of a function of SV , PP and (Vp /Vs )2 . Eq. (21)
a shallow thrust faulting regime to a deep normal faulting regime.
is expressed as a function of SV , PP and (Vp /Vs )2 . It is observed
More detailed subsurface information is required to substantiate the
that predicted Sh using regression model mostly agrees well with
cause of change in stress regime.
the Sh, trend computed from poroelastic model with a maximum
Spatial distribution of tectonic strain has widespread application
mismatch of 10 MPa in deeper sections of well M3. The model
for studying the movement of the Earth’s crust, wellbore failures
predicted Sh equals satisfactorily with the poroelastic model driven
in hydrocarbon exploration as well as pillar designing in coal min-
Sh for wells M2 and J1. Maximum mismatch of 2.75 and 2.5 MPa
ing area in active region (Gray & See 2007; Gupta & Gahalaut
is observed in wells M3 and J1, respectively. Regression model
2014; Hayavi & Abdideh 2016; Molaghab et al. 2017). Gray &
predicted Sh diverges from given LOT values at deeper sections of
See (2007) had measured maximum and minimum tectonic strains
wells M3 and J1. It is noted that the model predicted Sh shows an
along maximum and minimum principal stress direction using over-
elevated value of 10 MPa at 2791 m and 2.5 MPa at 2430 m for
coring technique in an Australian coalfield. The tectonic strain rate
wells M3 and J1, respectively. Without prior information on LOT
is generally measured using GPS observation for continental dy-
values, conventional log data may be utilized to estimate Sh using
namic research. The occurrence of an earthquake locally depends
eq. (21) in this area satisfactorily.
on the absolute level of strain, but a complete knowledge of past
earthquakes history for a sufficiently long time interval is, in most
The parameters used in this regression model incorporate con-
cases, inaccessible. A minimum estimate of the local accumulated
ventional log data from normal and thrust faulted regions. The
strain can, however, be obtained by the point wise integration of
mismatch upto 10 MPa observed from Fig. 11 may cause signifi-
seismic moment accumulation and release, for the time interval
cant error for detailed well design but may be adequate for regional
covered by sufficiently complete information (D’Agostino 2014).
work such as crustal stress anisotropy and overall stress estimation.
The compression rate ranging on an average between ∼50 and 100
Earlier model discussed in previous sections (eqs 14 and 19) require
nanostrain yr−1 is the major component of deformation in the EHS
information on tectonic strain to compute Sh . There is no real differ-
area. Also transverse extension of an average of ∼50 nanostrain
ence between the requirements for the Blanton–Olson, poroelastic
yr−1 is found to exist around 28◦ N, between 92◦ E and 94◦ E, ex-
and regression models described in this paper. The first two models
plained by existence of a series of N–S striking normal faults (Datta
do use tectonic strains but mathematically they are simply fitting
Gupta et al. 2015).
parameters that are never measured. The regression approach does
The tectonic strain computed from log data may certainly add to
not explicitly reference strain but the same data are being fit and the
the scientific value of understanding geodynamics of this region.
linear fitting parameters serve the same mathematical purpose as
The calculated strain thus indicates the sediment deformation at the
the strain parameters in the prior methods. Though the approaches
well site in thrusted and normal faulted region of the Upper Assam
have different mathematical forms but they fit the same dependent
basin. Blanton–Olson model predicted tectonic strain ranges from
variables while using the same independent variables. The only
−4.4 × 10−3 to 6.1 × 10−3 . The maximum and minimum horizon-
advantages of a particular approach must lie in the accuracy of pre-
tal strains along NE–SW and NW–SE from poroelastic model vary
diction and likely validity of extrapolation to conditions outside of
from 9.9 × 10−5 to 9.2 × 10−4 and −1.9 × 10−3 to 5.3 × 10−4 , re-
this particular data set. Comparing the sd of residual errors for Sh
spectively. The comparison between regression model predicted Sh
estimated using all three models it is observed that sd is minimum
and poroelastic model predicted Sh show good corroboration. The
for Sh estimation using poroelastic model. However, eq. (21) pro-
advantage of regression model is that minimum horizontal stress
vides a gross estimate of Sh in this area when data for more detailed
magnitude can be estimated in normal as well as thrust faulted
estimates are lacking.
regimes even without prior knowledge of tectonic strain under the
study area. The discrepancy of 10 MPa between the regression
model predicted Sh and LOT values in deeper section of wells M3
5 DISCUSSION
and J1 will cause inaccuracies in detailed studies of well design, cas-
In a compressive system as observed in the NT area for wells N1, N2 ing design, wellbore stability analysis and production scheduling.
and J1, it is inferred that SV is minimum in the upthrusted sediments However, this model may be used for regional stress studies where
but in deeper parts the SV is found to be more than the magnitudes of LOT, minifrac or hydrofrac data are not available. The Sh gradient
SH and Sh especially in wells N1 and J1. Initially in the upper parts derived from Blanton–Olson and poroelastic models substantiates
of the wells (upto 1635 m in N1 and 1119 m in J1) thrust faulting each other for all six wells. In this compressive stress fields it is
mechanisms are followed with SH > Sh > SV . The order of the stress expected that all LOT values will be between 90 and 100 per cent
672 J. Alam, R. Chatterjee and S. Dasgupta

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Figure 10 The regression plot between predicted minimum horizontal stress (Sh ) and leak off test (LOT), combining data from wells M2, M3 and J1 is shown
with the best fit regression line and standard deviation (sd) of residual error.

Figure 11 Shows the regression model (ANOVA) predicted minimum horizontal stress (Sh ), and poroelastic model predicted Sh with depth for wells (a)M2,
(b)M3 and (c)J1 respectively. Hydrostatic pressure (Ph), pore pressure (PP), mud weight (MW), leak off test (LOT) and vertical stress (SV ) are also plotted.

of SV . However in deeper section such as in well N1, LOT falls to M1, M2 and N2. This paper demonstrates different methodologies
as low as 70 per cent of SV . This may be interpreted as existence of to estimate stress magnitudes in tectonically active part of this basin.
normal faulting stress regime in these thrust systems (Dakua et al. The frictional limiting model overestimates the magnitude of SH and
2008). Geologic evidence for this active compressional thrust sys- Sh . Blanton–Olson model is useful for computation of Sh magnitude
tem (Belt of Schuppen) under the study area is supported by GPS for six wells under the study area. Using poroelastic model, the
motions and focal mechanism studies (Couzens-Schultz & Chan maximum and minimum tectonic strains are observed in normal and
2010; Kayal et al. 2010; Baruah et al. 2013; Gupta & Gahalaut thrust faulting regimes. Employing these strain parameters, Sh and
2014). SH magnitudes are supported with LOT and BO derived SH values,
respectively. Average stress ratio, Sh /SV varies from 0.70 in normal
faulting regime to 0.96 in thrust faulting regime whereas SH /SV
6 C O N C LU S I O N S varies from 0.80 in normal faulting regime to 1.06 in thrust faulting
regime, respectively. The order of stress magnitudes describe the
PP has been estimated using Eaton’s sonic model in the Upper
normal faulting and thrust faulting regime that corroborate with
Assam basin for six wells with an exponent of 1.0 calibrated with
previous studies based on earthquake focal mechanism and GPS
MDT data. PP gradients mostly follow hydrostatic trend with a
survey. A new regression model has been attempted to predict Sh in
slight high value ranging from 10.7 to 11.9 MPa km−1 for wells
Tectonic Strain and Stress Magnitude in Assam 673

this area with very good R2 = 0.83 using well log data. This paper Bjørlykke, K., 2014. Relationships between depositional environments,
discusses the application of three different models: two physical burial history and rock properties. Some principal aspects of diagenetic
models with different levels of complexity and one ad hoc linearized process in sedimentary basins, Sedimentary Geol., 301, 1–14.
Blanton, T. & Olson, J., 1999. Stress magnitudes from logs: effects of tectonic
form. From Figs 9 and 10 we can infer that amongst the three models,
strains and temperature, SPE Reservoir Eval. Eng., 5, 62–68.
the computed Sh using poroelastic biaxial strain model follows the
Bowers, G.L., 1995. Pore pressure estimation from velocity data: account-
measured LOT most faithfully. However, complex geological terrain ing for pore pressure mechanisms besides undercompaction, SPE Drill.
and inaccessible areas pose as constraints for estimating Sh due to Completion, 10, 89–95.

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
non-availability or lacking of LOT, minifrac and hydrofrac data. Byerlee, J.D., 1978. Friction of rock, Pure appl. Geophys., 116, 615–626.
This paper has major highlights of finding of tectonic strain, stress Chatterjee, R. & Mukhopadhyay, M., 2001. Stress-induced borehole elon-
magnitudes and delineation of stress regimes using well log data, gation up to 3.6 km depth in the Cauvery basin, India, SPWLA 42nd
which can further be modified with larger data set covering other Annual Logging Symposium, in SPWLA 42nd Annual Logging Sympo-
tectonic parts of northeastern India. sium, Houston, TX.
Chatterjee, R. & Mukhopadhyay, M., 2003. Stress modeling for the oil and
gas fields of Krishna–Godavari and Cauvery basins, India, using finite
element technique, Petrophysics, 44(5), 244–252.
Chatterjee, R. & Mukhopdhyay, M., 2002. In-situ stress determination using
AC K N OW L E D G E M E N T S well log data for the oil fields of the Krishna–Godavari basin, Petrophysics,
Authors are thankful to Oil India Limited, Duliajan for provid- 43, 26–27.
ing well log, seismic data and geological information. Ministry Chester, F.M. & Chester, J.S., 2000. Stress and deformation along wavy
frictional faults, J. geophys. Res., 105, 23 421–23 430.
of Earth Science, India is acknowledged for funding the project
Chhajlani, R., Zheng, Z., Mayfield, D. & MacArthur, B., 2002. Utilization of
(MoES/P.O. (Seismo)/1(273)/2015). We thank Department of Ap-
geomechanics for Medusa field development, deep-water Gulf of Mexico,
plied Geophysics, Indian Institute of Technology (Indian School of in SPE Annual Technical Conference and Exhibition, San Antonio, TX.
Mines), Dhanbad in helping us carry out the research work. Couzens-Schultz, B.A. & Chan, A.W., 2010. Stress determination in active
thrust belts: an alternative leak-off pressure interpretation, J. Struct. Geol.,
32, 1061–1069.
D’Agostino, N., 2014. Complete seismic release of tectonic strain and earth-
REFERENCES quake recurrence in the Apennines, Geophys. Res. Lett., 41, 1155–1162.
Al-Qahtani, M. & Rahim, Z., 2001. A mathematical algorithm for modeling Dakua, K.K., Gogoi, N.B. & Bharali, B.R., 2008. Structural features and its
geomechanical rock properties of the Khu and Pre-Khu reservoirs in brief study in and around Margherita-Ledo Area, Upper Assam Basin, in
Ghawar field, in Proceedings of the SPE Annual Technical Conference 7th International Conference and Exposition on Petroleum Geophysics,
and Exhibition, pp. 123–129, Bahrain. Hyderabad, India.
Anderson, E.M., 2nd, 1951. The Dynamics of Faulting and Dyke Formation Das, B. & Chatterjee, R., 2017. Wellbore stability analysis and prediction
with Applications to Britain, Oliver and Boyd, Edinburgh, 206. of minimum mud weight for few wells in Krishna–Godavari Basin, India,
Angelier, J. & Baruah, S., 2009. Seismotectonics in northeast India: a stress Int. J. Rock Mech. Min. Sci., 93, 30–37.
analysis of focal mechanism solutions of earthquakes and its kinematic Dasgupta, S., Chatterjee, R. & Mohanty, S., 2016. Magnitude, mechanisms
implications, Geophys. J. Int., 178(1), 303–326. and prediction of abnormal pore pressure using well data in the Krishna–
Armijo, R., Tapponnier, P., Mercier, J.L. & Tangling, H., 1986. Quaternary Godavari Basin, east coast of India, AAPG Bull., 100(12), 1833–1855.
extension in southern Tibet: field observations and tectonic implications, Datta Gupta, T., Riguzzi, F., Dasgupta, S., Mukhopadhyay, B., Roy, S. &
J. geophys. Res., 91, 13 803–13 872. Sharma, S., 2015. Kinematics and strain rates of the Eastern Himalayan
Azadpour, M. & Manaman, N.S., 2015. Determination of pore pressure from Syntaxis from new GPS campaigns in Northeast India, Tectonophysics,
sonic log: a case study on one of Iran carbonate reservoir rocks, Iran. J. 655, 15–26
Oil Gas Sci. Technol., 4(3), 37–50. Eaton, B.A., 1972. Graphical method predicts geopressures worldwide,
Bachman, R.C., Sen, V., Khalmanova, D., Okouma Mangha, V. & Settari, A., World Oil, 182, 51–56.
2011. Examining the effects of stress dependent reservoir permeability Eaton, B.A., 1975. The equation for geopressure prediction from well logs,
on stimulated horizontal montney gas wells, in Canadian Unconventional in Fall Meeting of the Society of Petroleum Engineers of AIME, Dallas,
Resources Conference, Calgary, Canada. TX.
Bailey, A., Tenthorey, E. & Ayling, B., 2017. Characterising the present-day Eberhart-Phillips, D., Han, D.H. & Zoback, M.D., 1989. Empirical rela-
stress regime of the Georgina Basin, Aust. J. Earth Sci., 64(1), 121–136. tionships among seismic velocity, effective pressure, porosity, and clay
Banerjee, P., Burgmann, R., Nagarajan, B. & Apel, E., 2008. Intraplate content in sandstone, Geophysics, 54, 82–89.
deformation of the Indian subcontinent, Geophys. Res. Lett., 35. Fitch, T.J., 1970. Earthquake mechanisms in the Himalayan, Burmese, and
Barman, B.J., Kashyap, B., Das, R. & Mitra, D.S., 2015. On the geothermal Andaman Regions and continental tectonics in central Asia, J. geophys.
gradient anomalies of hydrocarbon entrapment in parts of Rudrasagar and Res., 75(14), 2699–2710.
Charali fields in Upper Assam, in 11th Biennial International Conference Fjær, E., 2009. Static and dynamic moduli of a weak sandstone, Geophysics,
& Exposition, Jaipur, India. 74, WA103–WA112.
Barton, C.A., Zoback, M.D. & Burns, K.L., 1988. In-situ stress orientation Gough, D.I. & Bell, J.S., 1981. Stress orientations from oil well fractures in
and magnitude at the Fenton Hill geothermal site, New Mexico, deter- Alberta and Texas, Can. J. Earth Sci., 18, 638–645.
mined from wellbore breakouts, Geophys. Res. Lett., 15, 467–470. Gowd, T.N., Srirama Rao, S.V. & Chary, K.B., 1998. Seismotectonics of
Baruah, S., Baruah, S. & Kayal, J.R., 2013. State of tectonic stress in north- northeastern India, Curr. Sci., 74(1), 75–80.
east India and adjoining south Asia region: an appraisal, Bull. seism. Soc. Gowd, T.N., Srirama Rao, S.V. & Gaur, V.K., 1992. Tectonic stress field in
Am., 103, 894–910. the Indian subcontinent, J. geophys. Res., 98, 11 879–11 888.
Bell, J.S., 1996. Petro geoscience. 1. In situ stresses in sedimentary rocks Gray, I. & See, L., 2007. The measurement and interpretation of in situ
(part 1): measurement techniques, Geosci. Can., 23(2), 85–100. stress using an overcoring technique from surface, in 1st Canada-US
Bell, J.S. & Gough, D.I., 1982. The use of borehole breakouts in the study Rock Mechanics Symposium, pp. 721–727, eds Eberhardt, E., Stead, D. &
of crustal stress, 82–1075, 539–557, Open-File Rep, U.S. Geol. Surv. Morrison, T., Taylor & Francis.
Bendick, R. & Bilham, R., 2001. How perfect is the Himalayan arc? Geology, GSI, 2000. Seismo-Tectonic Atlas of India and its Environs, Geological
29, 791–794. Survey of India.
674 J. Alam, R. Chatterjee and S. Dasgupta

Gupta, H.K. & Gahalaut, V.K., 2014. Seismotectonics and large earthquake Plumb, R., Edwards, S., Pidcock, G., Lee, D. & Stacey., B., 2000. The
generation in the Himalayan region, Gondwana Res., 25(1), 204–213. mechanical Earth model concept and its application to high-risk well
Gupta, H.K., Rajendran, K. & Singh, H.N., 1986. Seismicity of the construction projects, in Proceedings of IADC/SPE Drilling Conference,
north-east India region. Part I: the data base, J. Geol. Soc. India, 28, New Orleans, LA.
345–365. Plumb, R.A., Evans, K.F. & Engelder, T., 1991. Geophysical log responses
Gupta, H.K. & Singh, H.N., 1986. Seismicity of the north-east India region. and their corelation with bed-to-bed stress contrasts in Paleozoicrocks,
Part II: earthquake swarms precursory to moderate magnitude to great Appalachian plateau, New York, J. geophys. Res., 96, 14 509–14 528.
earthquakes, J. Geol. Soc. India, 28, 367–406. Raaen, A.M., Horsrud, P., Kjørholt, H. & Økland, D., 2006. Improved routine

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
Gutierrez, M.A., Braunsdor, N.R. & Couzens, B.A., 2006. Calibration and estimation of the minimum horizontal stress component from extended
ranking of pore-pressure prediction models, Leading Edge, 25(12), 1516– leak-off tests, Int. J. Rock Mech. Min. Sci., 43, 37–48.
1523. Rajabi, M., Tingay, M. & Heidbach, O., 2016. The present-day stress field
Hafner, W., 1951. Stress distributions and faulting, Bull. geol. Soc. Am., 62, of New South Wales, Australia, Aust. J. Earth Sci., 63, 1–21.
373–398. Rajendran, K. & Rajendran, C.P., 2011. Revisiting some significant
Hayavi, M.T. & Abdideh, M., 2016. Estimation of in-situ horizontal stresses earthquake sources in the Himalaya: perspectives on past seismicity,
using the linear poroelastic model and minifrac test results in tectonically Tectonophysics, 504, 75–88.
active area, Russ. J. Earth Sci., 16. Raoof, J., Mukhopadhyay, S., Koulakov, I. & Kayal, J.R., 2017. 3-D seismic
Hillis, R., 2000. Pore pressure/stress coupling and its implications for seis- tomography of the lithosphere and its geodynamic implications beneath
micity, Explor. Geophys., 31, 448–454. the northeast India region, Tectonics, 36, 962–980.
Hillis, R.R. & Reynolds, S.D., 2003. In situ stress field of Australia, Hillis, Saha, D., 2011. Integrated analysis of Gravity and Magnetic data in the Upper
R.R. & Muller, R.D.in Evolution and Dynamics of the Australian Plate, Assam shelf and adjoining Schupen belt area: a critical review, in The
372, pp. 49–58 , Geological Society of America. 2nd South Asian Geoscience Conference and Exhibition, GEOIndia2011,
Horsrud, P., 2001. Estimating mechanical properties of shale from empirical Greater Noida, India.
correlations, SPE Drill. Completion, 16(2), 68–73. Sawai, S. & Sar, D., 2006. Analysis of gravity data for estimating thickness
Jaeger, J.C. & Cook, N.G.W., 1979. Fundamentals of Rock Mechanics, 3rd of Gondwana sediments in Dhansiri valley of Upper Assam of Assam–
edn, Chapman & Hall. Arakan Basin, in 6th International Conference & Exposition on Petroleum
Kayal, J.R. et al., 2010. The 2009 Bhutan and Assam felt earthquakes (Mw Geophysics, Kolkata, India.
6.3 and 5.1) at the Kopili fault in the northeast Himalaya region, Geomatics Seeber, L., Armbruster, J. & Quittmeyer, R., 1981. Seismicity and con-
Nat. Hazards Risk, 1(3), 273–281. tinental collision in Himalayan arc, in Zagros, Hindu-Kush, Himalaya:
Khan, P.K., 2005. Mapping of b-value beneath the Shillong Plateau, Geodynamic Evolution, Vol. 3, pp. 215–242, eds Gupta, H.K. & Delany,
Gondwana Res., 8, 271–276. F.M., American Geophysical Union.
Khan, P.K., Ghosh, M., Chakraborty, P.P. & Mukherjee, D., 2011. Seismic Singha, D.K. & Chatterjee, R., 2014. Detection of overpressure zones
b-value and the assessment of ambient stress in northeast India, Pure and a statistical model for pore pressure estimation from well logs in
appl. Geophys., 168(10), 1693–1706. the Krishna–Godavari Basin, India, Geochem. Geophys. Geosyst., 15(4),
Khattri, K., Wyss, M., Gaur, V.K., Saha, S.N. & Bansal, V.K., 1983. Local 1009–1020.
seismic activity in the region of the Assam Gap, northeast India, Bull. Singha, D.K. & Chatterjee, R., 2015. Geomechanical modeling using finite
seism. Soc. Am., 73, 459–470. element method for prediction of in-situ stress in Krishna–Godavari basin,
Koch, G.S. & Link, R.F., 1970. Statistical Analysis of Geological Data, Vol. India, Int. J. Rock Mech. Min. Sci., 73, 15–27.
1, John Wiley & Sons, Inc. Song, L., 2012. Measurement of minimum horizontal stress from logging
Last, N.C. & McLean, M.R., 1996. Assessing the impact of trajectory on and drilling data in unconventional oil and gas, MSc thesis, University of
weIls drilled in an overthrust region, J. Pet. Technol., 48, 620–626. Calgary, Calgary.
Law, B.E & Spencer, C.W, 1998. Abnormal pressure in hydrocarbon envi- Srivastava, S.K. & Bharali, B., 2000. Stress regimes in parts of Upper Assam
ronments, AAPG Memoir, 70, 1–11. basin: borehole breakout data analysis and its comparison with earthquake
Lei, T., Sinha, B.K. & Sanders, M., 2012. Estimation of horizontal stress focal mechanism, J. Ind. Geophys. Union, 4(1), 59–68.
magnitudes and stress coefficients of velocities using borehole sonic data, Swarbrick, R.E. & Osborne, M.J., 1998. Mechanisms that generate abnormal
Geophysics, 77(3), 181–196. pressures: an overview, in Abnormal Pressures in Hydrocarbon Environ-
Maung, H., 1987. Transcurrent movements in the Burma–Andaman Sea ments: AAPG Memoir, Vol. 70, pp. 13–34, eds Law, B.E., Ulmishek, G.F.
region, Geology, 15(10), 911–912. & Slavin, V.I., American Association of Petroleum Geologists
Molaghab, A., Taherynia, M.H., Aghda, S.M.F. & Fahimifar, A., 2017. Tingay, M., Hillis, R., Swarbrick, R., Morley, C. & Damit, A., 2008. Origin
Determination of minimum and maximum stress profiles using wellbore of overpressure and pore-pressure predictionin the Baram Delta province,
failure evidences: a case study—a deep oil well in the southwest of Iran, Brunei, AAPG Bull., 93, 51–74.
J. Pet. Explor. Prod. Technol., 7, 707–715. Tingay, M.R.P., Hillis, R.R., C.K., M., Swarbrick, R.E. & Okpere, E.C.,
Molnar, P., 1984. Structure and tectonics of the Himalaya: constraints and 2003. Variation in vertical stress in the Baram Basin, Brunei: tectonic and
implications of geophysical data, Annu. Rev. Earth Planet. Sci., 12, 489– geomechanical implications, Mar. Pet. Geol., 20, 1201–1212.
518 Townend, J. & Zoback, M.D., 2000. How faulting keeps the crust strong,
Molnar, P. & England, P., 1990. Temperatures, heat flux, and frictional stress Geology, 28(5), 399–402.
near major thrust faults, J. geophys. Res., 95, 4833–4856. van Ruth , P.J, Hillis, R.R. & Swarbrick, R.E, 2002. Detecting overpres-
Molnar, P. & Qidong, D., 1984. Faulting associated with large earthquakes sure using porosity-based techniques in the Carnarvon basin, Australia,
and the average rate of deformation in central and eastern Asia, J. geophys. APPEA, 42:559–569.
Res., 89, 6203–6227. Vernant, P.R. et al., 2014. Clockwise rotation of the Brahmaputra Valley
Moos, D. & Zoback, M.D., 1990. Utilization of observations of well bore relative to India: tectonic convergence in the eastern Himalaya, Naga
failure to constrain the orientation and magnitude of crustal stresses’ Hills, and Shillong Plateau, J. geophys. Res., 119(8), 6558–6571.
application to continental, Deep Sea Drilling Project, and ocean Drilling Vernik, L. & Kachanov, M., 2010. Modeling elastic properties of siliciclastic
program boreholes, J. geophys. Res., 95(B6), 9305–9325 rocks, Geophysics, 75, 171–182.
Mullick, M., Riguzzi, F. & Mukhopadhyay, D., 2009. Estimates of motion Wang, H.F., 2000. Theory of Linear Poroelasticity, Princeton University
and strain rates across active faults in the frontal part of eastern Himalayas Press.
in North Bengal from GPS measurements, Terra Nova, 21, 410–415. Yin, A., 2006. Cenozoic tectonic evolution of the Himalayan orogen as
Nandy, D.R., 2001. Geodynamics of Northeastern India and Adjoining Re- constrained by along-strike variation of structural geometry, exhumation
gion, ACB Publications. history, and foreland sedimentation, Earth-Sci. Rev., 76, 1–131.
Tectonic Strain and Stress Magnitude in Assam 675

Yin, A. & Harrison, M., 2000. Geologic evolution of the Himalayan–Tibetan Zoback, M.D., 2007. Reservoir Geomechanics, Cambridge Univ. Press.
orogen, Annu. Rev. Earth Planet. Sci., 28, 211–280. Zoback, M.D. & Healy, J.H., 1984. Friction, faulting, and ‘in situ’ stresses,
Zamora, M., Sartoris, G. & Chelini, W., 1994. Laboratory measurements Ann. Geophys., 2, 689–698.
of ultrasonic wave velocities in rocks from the Campi Flegrei volcanic Zoback, M.D., Moos, D., Mastin, L. & Anderson, R.N., 1985.
system and their relation to other field data, J. geophys. Res., 99(B7), 13 Well borehole breakouts and in situ stress, J. geophys. Res., 90,
553–13 561. 5523–5530.
Zhang, J. & Roegiers, J., 2010. Discussion on Integrating borehole-breakout Zoback, M.L., Zoback, M.D., Adams, J., Assumpcao, M., Bell, S., Bergman,
dimensions, strength criteria, and leakoff test results,to constrain the state E.A. & Fuchs, K., 1989. Global patterns of tectonic stress, Nature,

Downloaded from https://academic.oup.com/gji/article-abstract/216/1/659/5144768 by National Geophysical Research Institute user on 09 January 2019
of stress across the Chelungpu Fault, Taiwan, Tectonophysics, 492, 295– 341(6240), 291–298.
298.

View publication stats

You might also like