You are on page 1of 6

Stress characterization in deep boreholes using acoustoelasticity

J. A. Donald
Schlumberger, Kuala Lumpur, Malaysia

R. Prioul, T. Lei, B. Sinha


Schlumberger-Doll Research Center, Cambridge, United States

ABSTRACT: Horizontal stress magnitudes are determined from dynamic elastic moduli, effective vertical stress and a
wellbore stress model using nonlinear elasticity in rock formations with medium to high porosity (15 to 40%). Fullwaveform borehole acoustic waves (flexural and Stoneley) and the near-wellbore stress distribution are used to estimate
the third-order elastic constants required to relate changes in the far-field shear moduli to changes in stress magnitudes.
A case study from offshore Malaysia has been completed in which the results indicated a normal stress regime and horizontal stress anisotropy of 8 to 12%. Estimates of the h are consistent with the LOT data and the formations did not
show any borehole breakout, yet H was found to be greater thanhmin. The third-order elastic constants c144 and c155
were found to be consistent with laboratory values in the published literature. Stress-velocity relationships using these
results are shown to be suitable for analysis of stress path effects for time-lapse seismic surveys.

1 INTRODUCTION
It has become a standard practice within the petroleum industry to construct wellbore geomechanical
models for applications such as designing safe mud
windows for drilling, predicting sand production and
designing stimulation treatments in the form of hydraulic fracturing. A particular need is better quantifying of geomechanical properties, i.e. the in-situ
stress field, pore pressure, material properties (elastic, yield or quasi-brittle failure, hardness, rock-fluid
sensitivity), their anisotropic nature and their spatial
heterogeneities, as well as the presence of discontinuities (such as natural fractures or geological layering). With the evolution of the unconventional resource market, much advancement has been made to
quantify the impact of layering anisotropy within
shale rocks for stimulation design (Higgins et al.
2008, Prioul et al. 2011).
It is known that in many offshore environments
where deepwater exploration is increasing, linear
elasticity is not always sufficient as a constitutive
law (e.g. stress-sensitive high porosity rocks), yet
simple empirical stress models (Eaton 1969, Matthews & Kelley 1967) are still being used today.
Even though these models are robust and require only limited input data, they do not account for unbalanced tectonic stress or formation anisotropy. Case
studies have shown that these two factors should be
considered; otherwise, well integrity issues may
arise (Kozlowski et al. 2011). The minimum horizontal stress can be determined through extended
leak-off tests or mini-fracs. In deep wellbores the

maximum horizontal stress is difficult to measure directly, and therefore borehole failure models have
been traditionally employed to determine its magnitude. However, within most offshore wells, synthetic oil-based mud systems are commonly used, resulting in fewer observations of wellbore failure.
Furthermore, in formations that are considered nonelastic, these failure models are not applicable (Zoback et al. 1985).
Linear-elastic stress models derived from static
core measurements and integrated with borehole
sonic dynamic moduli have been employed to generate a continuous stress profile along the wellbore
(Plumb et al. 2000). However, uncertainty with these models is accepted when using both dynamic
elastic properties measured from borehole sonic data
and those calibrated with static properties measured
in the laboratory. Rocks are generally not considered purely linear elastic, and considerations for laboratory and logging conditions must be well understood (Fjr & Holt 1999, Fjr et al. 2005).
Stress characterization around boreholes has been
studied using acoustic methods with the theory of
acoustoelasticity (Norris et al. 1994) in the laboratory (Winkler et al. 1998) and using field data (Sinha
et al. 2000, Plona et al. 2000). The variation of the
elastic wave velocities in a propagating medium subject to externally applied stresses is defined as
acoustoelasticity (Thurston & Brugger 1964, Sinha
1982, Norris et al. 1994). In addition to the bulk and
shear moduli used in linear elastic theory (also referred to as the second-order elastic constants),
higher order terms called the third-order elastic constants are used to describe the change of effective
elastic constants or velocity with the change in stress

(Sinha et al. 2006). All rocks have some degree of


nonlinear elasticity (Winkler & Liu 1996). However,
this phenomenon is more likely to be observed this
phenomenon within medium- to high-porosity rocks
given the current accuracy of borehole acoustic logging technology to resolve changes in slowness with
stress.
A technique has been developed that uses fullwaveform borehole acoustic waves (monopole, dipole and Stoneley) and the near-wellbore stress distribution in combination with an acoustoelastic
model and knowledge of the vertical effective stress
to estimate the two horizontal stress magnitudes and
third-order elastic constants (see Lei et al. 2012 for a
comprehensive review). An example from offshore
Malaysia is presented illustrating this technique. We
also show how in-situ calibrated velocity-stress
transforms can be used for understanding stress effects on time-lapse seismic data.
2 BACKGROUND

2.2 Borehole sonic measurements


Borehole sonic measurements have evolved over the
past two decades with the understanding of how anisotropy affects both near-field and far-field regions
around wellbores (Franco et al. 2006). With the advent of flexural dipole logging to estimate shear in
slow formations, much effort has been focused on
the acoustics processing techniques to accurately estimate the shear slowness. Through these efforts, it
was observed by Plona et al. (2000) that the prediction of the flexural wave behavior in the laboratory
was indeed verified in the field for identifying
stress-induced anisotropy (Sinha et al. 2000). Fullwaveform sonic logs are acquired as described by
Pistre et al. (2005) to obtain compressional, crosseddipole and Stoneley waveforms. The dipole sources
are orientated orthogonally to each other and processed to determine the fast and slow shear wave
slownesses (slowness = 1/velocity) and the polarization azimuth of the far-field fast shear wave (Plona
et al. 2000, Alford 1986, Bose et al. 2007). An example of the crossed-dipole anisotropy processing
results is shown in Figure 1.

2.1 Laboratory experiments


Laboratory experiments were conducted to understand the effect on elastic wave velocities surrounding the wellbore and the impact on flexural wave
(dipole) measurements (Winkler et al. 1998, Sinha &
Winkler 1999).
Experiments were designed to illustrate the impact of the applied far-field stress on the nearwellbore region. Winkler et al. (1998) built an apparatus in which a small-diameter hole was drilled
through a block of Berea sandstone and the block
was exposed to uniaxial external load. P- and Swave transducers were placed in the regions that
correspond to the areas of maximum and minimum
compression in the near wellbore and in the far field.
Shear and compressional velocities in the nearwellbore region with highest compression were
greater than those in the region with least compression. Shear wave velocities with polarization parallel
to the stress direction are greater than the velocities
with polarization perpendicular to the stress direction. With the increase in stress, the differences between shear wave polarizations in the far field became larger.
A second experiment was then conducted for a
borehole drilled in a larger Berea sandstone block.
The borehole was water sealed, and a set of transducers and receivers were placed and aligned either
parallel or perpendicular to the uniaxial stress direction. Results of this experiment verified theoretical
prediction of the crossing dipole dispersion as an indicator of stress-induced anisotropy.

At discrete depths, the fast and slow flexural and


Stoneley wave train data are transformed to the frequency-slowness domain, where the slowness is
plotted against the frequency; this is termed slowness-dispersion analysis as shown in Figure 2.
The crossover of the dipole dispersions from the
fast (red) and slow (blue) flexural waves measured
in the field verifies the same signature as was shown
in the laboratory experiments. The third mode is the
Stoneley wave (cyan) which is a borehole-guided
mode which is sensitive to formation mobility, mud
velocity, hole size and the shear modulus in the
cross-sectional plane intersecting the borehole. The
solid lines represent the theoretical homogeneous
isotropic dispersion for each wave, taking into account the borehole fluid bulk modulus and far-field
formation moduli (shear and bulk), borehole diameter and the presence of the sonic tool in the wellbore.
The difference between the theoretical model dispersion and the measured dispersions as a function of
frequency yields a dynamic shear modulus as a function of wavelength. The shear modulus can be extracted from the sand face into the far field up to
seven borehole radii away, yielding a shear radial
profile (Sinha et al. 2006). An example of the shear
radial profiles is shown in Stress-dependent acoustics
Assuming that one principal stress is vertical, V, we
can define a coordinate system with X3 pointing to
the vertical axis, X1 pointing to the azimuth of maximum horizontal stress H, and X2 pointing to the azimuth of minimum horizontal stress h. When the
rock is stress-sensitive, the sonic velocities change

as a function of incremental changes in effective


stress above and beyond a reference state. In nearvertical wellbores where there are indications of
stress-induced anisotropy from the dipole dispersions, the slow, fast and Stoneley shear provide estimates of the three shear moduli c44, c55 and c66
where cij = (1/DTshear)2bulk. The vertically propagating compressional wave velocity yields the compressional modulus c33.

Figure 1: Crossed-dipole anisotropy processing of flexural


wave data to determine fast and slow shear slownesses, along
with the polarization direction of the fast shear wave.

Then the three compressional (P-) wave moduli and


three shear (S-) wave moduli can be expressed in
terms of the diagonal elements of elastic stiffness
tensor as follows:
2
2
2
c11 bV11 , c22 bV22 , c33 bV33 , and,
c44 bV32 , c55 bV31 , c66 bV12
2

(1)

where b denotes the formation bulk density and Vij


(i, j = 1; 2; 3) denotes the velocity of a wave traveling along axis Xi and having polarization along Xj.
Following acoustoelasticity theory (Thurston &
Brugger 1964, Sinha 1982) applied to borehole conditions (Norris et al. 1994), the equations of motion
for elastic waves in prestressed formations referred
to the statically deformed state lead to relationships
between elastic stiffnesses and the effective principal stresses in the formation (Lei et al. 2012, equations 20 to 25):
Stress sensitivity coefficients for the compressional moduli and rely on Mref, ref, c111 and c112,
whereas the stress sensitivity coefficients for the
shear moduli and rely on Mref, ref, c144 and c155. Mref
and ref are the two independent second-order elastic
constants in a hydrostatically loaded reference stress
(the rock is assumed isotropic in the unstressed or
hydrostatically loaded state). There are three independent third-order elastic constants, c111, c112 and
c123, with c144 = (c112 c123)/2 and c155 = (c111
c112)/4. Equations referred to the statically undeformed state suitable for laboratory conditions can
be found in Prioul et al., 2004.
As shown by Pistre et al. (2009) and Sun & Prioul
(2010), the stress regime, or Q factor can be related
to the relative ranking of the shear moduli as shown
in Table 1.
Table 1: Shear moduli, stress regimes and Q factor.
Shear moduli
Stress
Q factor
ranking
regime

Figure 2: Slowness-dispersion analysis indicating stressinduced anisotropy with classical crossover behavior.

c55 c44
1
c55 c66

c55 > c44 > c66

Normal
faults

c55 > c66 > c44

Strikeslip
faults

c55 c66 2c 44
2
c55 c 44

c66 > c55 > c44

Thrust
faults

3c66 2c 44 c55
3
c66 c44

2.3 Application to field data


For a given zone that shows stress-induced anisotropy for a normal faulting regime, the ratio of shear
moduli to the corresponding formation stresses
yields:
c c
c c
D 55 66 55 44
(2)
V h H h
Figure 3: Shear radial profiles at a single depth.

where the acoustoelastic parameter D = 3/2+(c155


c144)/2 With measurements of the three shear
moduli, the overburden stress and the minimum horizontal stress directly measured from XLOTs or
minifracs, Equation 2 can be rearranged to solve for
the maximum horizontal stress directly as

c55 c 44



c c
H

55

(3)

shear radial profiles are shown in Figures 1 through


3. The results of the stress magnitude inversion are
presented in Figure 4.
The choice of each zone requires that there is a positive indication for dispersion crossover, intrinsic
formation isotropy (no layering) and consistent formation properties over a minimum length of 3 m and
a minimum of 2% shear slowness anisotropy between fast and slow dipole sources.

66

0.5

If both the minimum and maximum horizontal


stresses are unknown, then the D parameter must be
solved independently. Subsequent work by Lei et al.
(2012) shows a method of obtaining D independently from the dipole radial profiles combined with a
borehole stress model.
Inversion of the measured dipole dispersions
yields a radial profile of the shear modulus from the
sand face into the far field. In conjunction with the
measurements, an equivalent isotropic model of the
simulated dipole dispersions for each direction
(maximum and minimum horizontal stress directions) can be generated in an anisotropic stress environment. The dipole measurements are affected by a
combination of the near-wellbore stresses (axial, radial and tangential) as well as the far-field stresses
(vertical, maximum and minimum horizontal). By
combining the elastic solution from Kirsch (1898)
with the effect of the dipole measurements in the
near and far field by Sinha & Kostek (1996), we obtain the following relationships (Lei et al. 2012):
c55 (r , ) 0 m1

a2 3
a4
(c 55 c 44 ) 4 c 55
2
r
2
r

c55 (r , ) / 2 m2

a2 3
a4
(c 55 c 44 ) 4 c 44
2
r
2
r

(4)
(5)

where a is the distance from the wellbore wall, r is


the radius of the wellbore and m1 and m2 are functions of c144, c155 and the reference moduli. The full
derivation is shown in Lei et al. (2012) for the m1
and m2 terms. The model radial profiles from Equations 4 and 5 are compared with the measured radial
profiles from the fast (related to c55) and slow (related to c44) dipoles, respectively. Once the nonlinear
elastic constants are known, then the minimum and
maximum horizontal stress magnitudes can be obtained independently.

0.5

Pore Pressure Gradient


g/cm3
Overburden Stress Gradient
G/C3
Maximum Horizontal Stress Gradient

0.5
1
6
GR/Baseline

MD
(m)
1:2000

Gamma Ray
0

gAPI 150

Bulk Density
g/cm3
3
Bit Size
in
16
Caliper 2
in
16
Caliper 1
in
16

2.5
2.5
2.5

Minimum Horizontal Stress Gradient


0.5
0
0
0

C44
GPA
C55
GPA
C66
GPA

0.5

0.5

0.5

2.5
Mud Weight
g/cm3
Pore Pressure Points
g/cm3
LOT
g/cm3

2.5
C155
2.5

-10000 GPa 10000

2.5

-10000 GPa 10000

C144

2350

2400

2450

2500

2550

2600

2650

Figure 4: Stress magnitude analysis from offshore Malaysia.

The results of the three shear moduli indicate that


this section is a normal stress regime, where c55 > c44
> c66 and the stress Q factor is computed to be 0.66.
A result of an LOT is plotted at the upper part of the
interval along with the equivalent mud weight used
to drill the well. Both values are consistent with the
results of the minimum horizontal stress values. The
difference between minimum and maximum horizontal stresses ranges from 2.5 to 3.5 MPa, or 8 to
12%. The sediments are relatively high in porosity,
and the change in the three shear moduli with depth
is very evident. As an example, we report the complete stress determination in Table 2 for a depth of
2500 m. It should be noted that the dual axis caliper
measurements show no ovality over the logged section. Difference in the shear moduli is evidence that
the principal formation stresses are different.
Table 2. In situ stress magnitudes for well from offshore Malaysia at depth 2500 m.
v
H
h
Pp

MPa
MPa
MPa
MPa

35.3
34.0
31.3
26.5
0.9

3 CASE HISTORY
The case history is from offshore Malaysia where
sonic data is gathered for seismic ties, wellbore stability, pore pressure analysis and completion design.
The dipole crossover from the dispersion analysis
indicates that the dominant mechanism of anisotropy
is differential horizontal stress. Examples of the
shear anisotropy, slowness dispersion analysis and

4 ROCK PHYSICS MODEL FOR


GEOMECHANICS
Once all three stress magnitudes and the stresssensitivity coefficients of the zone of interest have
been determined, we have in-situ calibrated velocityto-stress transforms that can be readily used for

time- lapse seismic reservoir geomechanics. The coefficients for the case study are reported in Table 3.
Table 3. Reference moduli and stress-sensitive constants used
for stress determination at 2500 m
Mref
c144
c155
ref
ref
GPa
GPa
MPa
GPa
GPa
12.77
3.69
9.7
3.26.103
5.47.103

During primary depletion, vertical effective stress


and horizontal effective stress increase within the
reservoir because of pore pressure decrease, whereas
in the caprock, vertical effective stress decreases and
horizontal stresses may increase (Herwanger &
Horne 2009). The stress path K is a convenient way
to characterize tensor stress changes from an initial
stress state by a single parameter. It is defined as the
ratio between the change in minimum horizontal effective stress and the change in vertical effective
stress:
K

h
V

(6)

We assume here that the maximum and minimum


horizontal stresses are changing in the same way
(H =h ). We consider several modes of deformation, i.e., different values of K, to illustrate how
the seismic velocities will change as a function of
the different stress paths. For example, under hydrostatic stress changes such as pore pressure changes,
the horizontal and vertical stresses are increased
simultaneously by equal amounts, i.e., K=1. If the
deformation of the reservoir is constrained by a nolateral-deformation boundary condition (such as in
uniaxial strain experiments), elasticity theory tells us
that vertical stress changes are associated to horizontal stress changes as 0 < K = (/1-) < 1. For laterally unconstrained compression using only vertical
force (i.e. horizontal stress changes as H
=h=0)), the stress path will be K = 0. Negative
stress paths are predicted for overburden stretching
(K < 0). Figure 5 shows the variation of the vertical
fast ( V31Shmax SV ) and slow ( V32Shmin SV ) shear velocities as a
function of the effective vertical stress for K=-0.5, 0,
0.5 and 1. Since the non-linear model was calibrated
near a reference stress, we analyze only perturbations within 10 MPa of the vertical stress of the considered depth. For comparison on those plots, we reported the classical empirical Eberhardt-Phillips et
al. (1989) that depend on porosity (26%) and Vclay
(5%) as well as an effective stress (velocities in km/s
and stress in kilobars):
V
3.7 4.94 1.57 V 0.361( P e
)
(7)
For a visual comparison of stress-sensitivity effects, we arbitrarily shifted VS EP 89 at the reference
vertical stress. We make several observations: (i) the
two vertical shear velocities vary very differently
16.7 Pe

EP 89

with vertical stress depending on the stress path because velocities depend on the three principal stress
magnitudes (and shear velocities are sensitive to
stress both in the propagation and polarization directions); (ii) the empirical model cannot capture differences due to stress path because it relies on only
one stress; (iii) the stress sensitivities are significantly stronger than the empirical VS EP 89 for all considered
stress paths (K = 0.5, 0, 0.5 and 1); (iv) our model
is calibrated for in-situ conditions whereas the empirical model had to be artificially calibrated to the
in-situ conditions.
When compressional stress-sensitivities are
known, the full elastic stiffness-to-stress transforms
would be known, and fluid substitution on the anisotropic orthorhombic medium could be easily pursued
for advanced time-lapse seismic scenario analysis.

clay

Shmax SV

Shmin SV

Figure 5: Vertical fast ( V31


) and slow ( V32
) shear velocities as a function of the effective vertical stress for K=-0.5,
0, 0.5 and 1 for Malaysia.

4 CONCLUSIONS
Horizontal stress magnitudes and third-order elastic
constants were determined using full-waveform
borehole acoustic waves along with the effective
vertical stress and an acoustoelastic model based on
nonlinear elasticity. A review of the theory, laboratory work, and field methods illustrated the application of using an acoustoelastic model for stress characterization.
An example from Malaysia was
presented where rock formations exhibited measurable stress-sensitivity to acoustic waves, and this
technique provided estimates of stress magnitudes
consistent with the field observations. The stress
characterization has direct applications for confirming the present-day geological setting, providing input to wellbore stability models and completion designs for a safe pore pressure drawdown without
producing sand and hydraulic fracturing operations.
When all three stress magnitudes and the stresssensitivity coefficients of the zone of interest are
known, the in-situ calibrated velocity-to-stress trans-

forms can be used to understand the stress path effects on velocities and could be used for time-lapse
seismic reservoir geomechanics.
5 REFERENCES
Alford, R. M. 1986. Shear data in the presence of azimuthal anisotropy: 56th Annual International Meeting, SEG, Expanded Abstracts, 476479.
Bose, S., Sinha, B.K., Sunaga, S., Endo, T., & Valero, H.P.
2007. Anisotropy processing without matched crossdipole
transmitters: SEG Technical Program Expanded Abstracts
2007: pp. 114-118.
Eaton, B. A. 1969. Fracture gradient prediction and its application in oilfield operations. Journ. Pet. Tech., October
1969 : 1353-1360.
Eberhardt-Phillips, D., Han, D.-H. & Zoback, M. 1989. Empirical relations among seismic velocity, effective pressure,
porosity, and clay content in sandstone. Geophysics 54(1):
82- 89.
Franco, J. L. A., De, G. S., Renlie, L. & Williams, S. 2006.
Sonic Investigations In and Around the Borehole. Oilfield
Review, Spring 2006.
Fjr, E. & Holt, R.M. 1999. Stress and stress effects on acoustic velocities from cores, logs and seismics; Proc. SPLWA
40th Annual Logging Symposium Proceedings, May 30June 3, 1999.
Fjr, E., Larsen, I. & Scheldt, T. 2005. What is the Poissons
ratio of soft rock? ARMA 40th U.S. Symposium on Rock
Mechanics, June 25-29, 2005.
Herwanger, J. & Horne, S. 2009. Linking reservoir geomechanics and time-lapse seismics: Predicting anisotropic velocity changes and seismic attributes, Geophysics, 74,
W13-W33.
Kozlowski, K., Fidan, M., Donald, A., Shotton, P., Nielsen, H,
Anderson, N. & Jocker, J. 2011. Overburden characterization for geomechanics and geophysical applications in the
Eldfisk field: A North Sea case study; Proc. SPWLA 52nd
Annual Logging Symposium, May 14-18, 2011.
Higgins, S., Goodwin, S., Donald, A., Bratton, T. & Tracy, G.
2008. Anisotropic stress models improve completion design in the Baxter Shale; Proc. SPE Annual Technical
Meeting, 21-24 October, 2008.
Kirsch, G. 1898. Die Theorie der Elastizitt und die Bedrfnisse der Festigkeitslehre. Zeit Verein Deutsch Ing;42:797
807.
Lei, T., Sinha, B.K. & Sanders, M. 2012. Estimation of horizontal stress magnitudes and stress coefficients of velocities using borehole sonic data: Geophysics 77(3): WA181WA196.
Matthews, W. R. & Kelly, J. 1967., How to predict formation
pressure and fracture gradient: Oil and Gas Journal, 92
106.
Norris, A. N., Sinha, B. K. & Kostek, S. 1994, Acoustoelasticity of solid/fluid composite systems: Geophys. J. Int.,118,
439-446.
Pistre, V., Kinoshita, T., Endo, T., Schilling, K. & Pabon, J.
2005. A modular wireline sonic tool for measurements of

3D (azimuthal, radial, and axial) formation acoustic properties: Presented at the 46th Annual Logging Symposium,
Society of Petrophysicists & Well Log Analysts, SPWLA,
113.
Pistre, V., Yan, G.R., Sinha, B., Prioul, R. & Vidal-Gilbert, S.
2009. Determining stress regime and Q factor from sonic
data. Proc. SPWLA 50th Annual Logging Symposium, June
2124, 2009.
Prioul, R., A. Bakulin, and V. Bakulin, 2004, Non-linear rock
physics model for estimation of 3D subsurface stress in anisotropic formations: Theory and laboratory verification:
Geophysics, 69(2): 415425
Prioul, R., Karpfinger, F., Deenadayalu, C. & Suarez-Rivera,
R. 2011. Improving Fracture Initiation Predictions on Arbitrarily Oriented Wells in Anisotropic Shales: Proc.
CSUG/SPE SPE-147462, Canadian Unconventional Resources Conference held in Calgary, Alberta, Canada, 1517 November, 2011
Plona, T. J., Kane, M. R., Sinha, B. K. & Walsh, J. 2000. Using
acoustic anisotropy; Proc. SPWLA 41st Annual Logging
Symp., June 47, 2000.
Plumb, R., Edwards, S., Pidcock, G., Lee, D. & Stacey, B..
2000. The Mechanical Earth Model Concept and Its Application to High-Risk Well Construction Projects; Proc.
IADC/SPE Drilling Conference, 23-25 February, 2000
Sinha, B. K.. 1982. Elastic waves in crystals under a bias; Ferroelectrics, 41(1): 6173.
Sinha, B.K. & Kostek, S. 1996. Stress-induced azimuthal anisotropy in borehole flexural waves; Geophysics 61(6):
18991907.
Sinha, B. K., Kane, M. R. & Frignet, B. 2000. Dipole dispersion crossover and sonic logs in a limestone reservoir; Geophysics 65(2): 390407.
Sinha, B. K., Vissapragada, B., Renlie, L. & Tysse, S. 2006.
Radial profiling of the three formation shear moduli and its
applications to well completions. Geophysics 71(6); E65E77.
Sinha, B.K. & Winkler, K.W. 1999. Formation nonlinear constants from sonic measurements at two borehole pressures;
Geophysics, 64 (6): 1890-1900.
Sun, H. & Prioul, R. 2010. Relating shear sonic anisotropy directions to stress in deviated wells; Geophysics 75(5):
D57-D67.
Thurston, R.N. & Brugger, K., 1964, Third-order elastic constants and the velocity of small amplitudes elastic waves in
homogeneously stressed media: Physical Review, A33,
1604-1610.
Winkler, K.W., Sinha, B.K. & Plona, T.J. 1998. Effects of
Borehole Stress Concentrations on Dipole Measurements;
Geophysics 63 (1): 11-17.
Winkler, K.W. & Liu, X. 1996. Measurements of third-order
elastic constants in rocks: J. Acoust. Soc. Am., 100( 3):
1392-1398.
Zoback, M. D., Moos, D., Mastin, L. & Anderson, R.N. 1985.
Well bore breakouts and in situ stress. J. Geophys. Res.
90(B7): 55235530.

You might also like