You are on page 1of 173

Vladimir Grechka obtained his

Diploma in Geophysical Pros-


Applications of Seismic
pecting from Novosibirsk State
University in 1984. He worked at
the Tomsk branch of the Institute
Anisotropy in the Oil and
of Geology and Geophysics (which
was reorganized into the Institute
of Geophysics) from 1984 to 1994, where he received his
Gas Industry
PhD degree in 1990 with a thesis on ray tracing in aniso-
tropic media.
E D U C AT I O N T O U R S E R I E S CIS (O T EAGE)
Vladimir spent one year at the University of Texas at Dallas
before joining the Center for Wave Phenomena (CWP) at the
Colorado School of Mines as a post-doc (1995). He later
became a Research Associate Professor and a co-leader
of the Center for Wave Phenomena. His research at the
CWP was devoted to traveltime inversion and anisotropic
velocity-model building.

In 2001, Grechka joined Shell in Houston, Texas. While


there, he has been involved in a number of research and
technology development projects, including those on multi-
3
component and wide-azimuth seismic, reservoir simulation
and monitoring, vertical seismic profiling, fracture charac-
terization and microseismics.

To date, Vladimir has published more than 300 research


papers and one book. He is a member of SEG and EAGE. He
received the Clarence Karcher Award from the SEG (1997)
and two Honorable Mentions for papers published in Geo-
physics in 2000 and 2003. His paper on the NMO ellipses
(1998) is the most cited paper published in Geophysics in
1998. He was recently nominated to serve as the Editor-in-
Chief of Geophysics for 20092011. V. Grechka

12815-OTEIII Grechka Boek cover.indd 1 28-07-2009 10:10:52


Applications of
Seismic Anisotropy in the
Oil and Gas Industry

Vladimir Grechka

12815-OTE III Gretchka.indd 1 28-07-2009 10:13:41


2009 EAGE Publications bv

All rights reserved. This publication or part hereof may not


be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopy, recording, or
any information storage and retrieval system, without the prior
written permission of the publisher.

ISBN 978-90-73781-68-9

EAGE Publications bv
PO Box 59
3990 DB HOUTEN
The Netherlands

12815-OTE III Gretchka.indd 2 28-07-2009 10:13:41


CONTENTS
Contents

List of Figures 5

List of Tables 13

Introduction 15

Acknowledgements 17

1 Definitions and examples 19


1.1 Simple examples 19
1.2 Definition of seismic anisotropy 21
1.3 Seismic examples 22
1.4 Physical causes of seismic anisotropy 30

2 Plane waves 37
2.1 Strain, stress and the equation of motion 37
2.2 Hookes law 40
2.3 Symmetry classes 43
2.4 Christoffel equation 50

3 Rays and traveltimes 61


3.1 High-frequency asymptotics of seismic wavefields 61
3.2 Ray-tracing equations and techniques 63
3.3 Examples 69

4 Thomsen parameters 73
4.1 Phase and group velocities in VTI media 73
4.2 Thomsen notation 76
4.3 Extensions of Thomsen parameterization 83

5 Normal moveout 87
5.1 NMO velocity in a single VTI layer 87
5.2 Anisotropic Dix equation 91
5.3 NMO ellipse 97
5.4 3D Dix equation 103

6 Nonhyperbolic moveout 109


6.1 NH moveout of P-waves in a single VTI layer 109
6.2 NH moveout in stratified VTI media 115
6.3 Azimuthal variation of quartic moveout 117
6.4 Layered orthorhombic media 120

12815-OTE III Gretchka.indd 3 28-07-2009 10:13:41


Contents

7 Anisotropy estimation from VSP data 123


7.1 Laterally homogeneous overburden 123
7.2 Laterally heterogeneous overburden 126

8 Fracture characterization 135


8.1 Effective media theories 136
8.2 Comparison of theoretical predictions 143
8.3 Numerical modeling of effective elasticity 146
8.4 Governing parameters for vertical cracks 152
8.5 Seismic characterization of multiple fracture sets 153

9 Bibliography 159

12815-OTE III Gretchka.indd 4 28-07-2009 10:13:41


LIST
List OF FIGURES
of Figures

1.1 Block made of layers of steel (white) and rubber (black). The apparent stiffness of the
block under a compressive load depends on the direction of the load. 19
1.2 Fractures in sandstones of Checkerboard Mesa (Zion National Park, Utah, USA). 20
1.3 Anisotropic and isotropic depth-migrated sections obtained in (a) exact anisotropic
model, (b) anisotropic model estimated from surface reflection data, (c) isotropic
model constructed via conventional processing, and (d) isotropic model with the cor-
rect vertical velocity (after Han et al., 2001). 23
1.4 Common-image gathers after prestack depth migration with (a) isotropic and (b) ani-
sotropic velocity model (after Sarkar and Tsvankin, 2004). 25
1.5 P- (a) and shear-wave (b) gathers after the azimuthally-invariant (isotropic) NMO
correction. The traces are sorted by the source-to-receiver azimuth. The apparent
cosine-type dependence of the residual moveout is indicative of azimuthal anisotropy.
The jitter obscuring this dependence is caused by slightly unequal offsets at different
azimuths. The events at approximately 1.27 s (a) and 2.35 s (b) are the reflections from
the bottom of the Rulison reservoir (after Vasconcelos and Grechka, 2007). 26
1.6 Isotropic (a) and anisotropic (b) PS-wave common-conversion-point stacks (after
Grechka et al., 2002c). 27
1.7 Errors in (a) the lateral position of the PS-wave conversion point and (b) the incidence
angle of the P-leg for PP- and PS-waves at the Siri reservoir caused by neglecting ani-
sotropy (after Grechka et al., 2002c). 27
1.8 Three-component record (a, b, c) of a shear microseismic event and the rotated 3C
trace (d, e, f) that reveals shear-wave splitting. 28
1.9 Fast (a) and slow (b) shear reflections from the Austin Chalk horizon (after Mueller,
1992). 29
1.10 Shale under a microscope (after Hornby et al., 1994). 31
1.11 Sets of well-developed joints in Caithness, Scotland, UK (source: Wikipedia, http://en.
wikipedia.org/wiki/File:Joints Caithness. JPG). 32
1.12 Equivalence of finely layered isotropic and homogeneous anisotropic solids for propa-
gation of long seismic waves. 33
1.13 Sonic (black), shear (blue), and density (red) logs acquired over a 500 m interval in a
deepwater oil field in the Gulf of Mexico (after Bakulin and Grechka, 2003). 34
2.1 Small tetrahedron whose three faces lie in the coordinate planes. 39
2.2 Normal (red) and shear (black) components of the stress tensor. 42
2.3 Deformation of (a) isotropic and (b) anisotropic solids under a compressive load. 43
2.4 Two systems of vertical cracks embedded in azimuthally isotropic host rock at oblique
directions with respect to each other generally result in monoclinic symmetry with a
horizontal symmetry plane (after Bakulin et al., 2000). 46
2.5 Fractured sandstones near Paria Canyon in Arizona, USA. 46
2.6 Orthorhombic media have three mutually orthogonal planes of mirror symmetry. One
of the orthotropic models contains vertically fractured fine horizontal layers (after
Rger, 1998). 47
2.7 Orthogonal fracture sets in sandstones of Cedar Mesa, Utah, USA (source: Jon Olsons
home page, http://www.pe.utexas.edu/~jolson/Welcome.html). 47

12815-OTE III Gretchka.indd 5 28-07-2009 10:13:41


List of Figures

2.8 Sketch of an HTI model. Shear waves polarized parallel and normal to the
isotropy plane (cracks) propagate with different vertical velocities. The velocity of the
-wave is greater than that of (after Rger, 1997). 49
2.9 Typical geometry of phase-velocity surfaces in orthorhombic media. Red arrow points
to the shear-wave singularity (after Grechka et al., 1999a). 52
2.10 Slowness (dashed blue) and phase velocity (solid red) of the SV wave in a VTI
medium with the density-normalized stiffness coefficients ,
, and . The vertical and horizontal lines
indicate the symmetry axis and the isotropy plane, respectively. 53
2.11 Slowness (dashed blue) and phase velocity (solid red) shown in Figure 2.10, and
group velocity (solid black). The red, green and black arrows indicate the vectors ,
and , respectively. The gray arrows point to the phase-velocity extrema. 56
2.12 The same slowness (dashed blue), phase-velocity (solid red), and group-velocity (solid
black) surfaces as in Figure 2.11. The dark and light blue arrows are normals to the
slowness surface. Their lengths are equal to the group velocities . The corresponding
wavefront normals are shown with red and brown arrows. 57
2.13 Shear-wave slowness (a, b) and group-velocity (c, d) surfaces in an orthorhombic solid
with density-normalized stiffness coefficients (in Voigt notation)
and
. The units of axes are s/km in (a, b) and km/s in (c, d). 58
3.1 Seismic wavefront at two consecutive time moments t and . The dots indicate
elementary sources, the red arcs are the elementary wavefronts. 64
3.2 Rays emerging at observation surface typically do not arrive at predetermined receiver
locations (squares). 65
3.3 Ray trajectory through a sequence of homogeneous anisotropic layers. 66
3.4 Kinematics of reflection-transmission problem in the local -plane. If the inci-
dent wave propagates in the upper half-space, the group-velocity vectors of reflected
waves point up, , while the group-velocity vectors of transmitted waves point
down, . 67
3.5 Reflection traveltimes (top row) and ray trajectories (bottom row) of SV-waves in
two-layer VTI models. The relevant density-normalized stiffness coefficients in
Voigt notation are: in the top layer and
in the bottom layer. Two models differ in
the values of stiffness coefficient in the top layer. It is equal to =4.50 in (a) and
=2.36 in (b). The units of all are km2/s2. Red horizontal lines in the bottom
panels are the model interfaces. 70
3.6 Detail of the multivalued traveltime in Figureg 3.5b. 70
3.7 Same as Figure 3.5 but for 71
4.1 P-wave phase velocity function (solid) for positive and negative . The dashed line is
a circular arc whose radius is equal to the vertical velocity . 79
4.2 SV-wave phase velocity function (solid) for a positive . The dashed circular arc has a
radius equal to . 79
4.3 Dependence of the exact P-wave phase velocity (calculated with equation 4.17) on
in VTI media with . The legend indicates the ratios. The
corresponding vertical shear-wave velocities are , 1.6 km/s, and 2.4 km/s. 82
5.1 Ray trajectories in the reflector dip plane at zero and nonzero offsets. The vertical
-plane is assumed to be a symmetry plane so that the strike components of vec-
tors and are equal to zero (modified from Tsvankin, 1995). 88

12815-OTE III Gretchka.indd 6 28-07-2009 10:13:41


List of Figures

5.2 Time-migrated seismic line. The gray bar indicates the locations of CMP gathers
examined in Figures 5.3 and 5.4 (after Alkhalifah and Tsvankin, 1995). 90
5.3 Constant-velocity stacks after the conventional isotropic NMO-DMO sequence. The
velocity values correspond to the NMO velocities of the subhorizontal reflectors (after
Alkhalifah and Tsvankin, 1995). 91
5.4 Anisotropic constant-velocity stacks. The velocities correspond to the stacking veloci-
ties of the subhorizontal reflectors (after Alkhalifah and Tsvankin, 1995). 91
5.5 Stack of horizontal layers. 92
5.6 (a) Average P-wave vertical velocity from a check shot (gray), the effective NMO veloc-
ity (black), and (b) the interval Thomsen coefficient (after Alkhalifah et al., 1996). 93
5.7 Zero-offset ray (blue) reflected from a dipping interface beneath a layered anisotropic
medium. The dashed lines indicate nonexisting interfaces that would generate zero-
offset reflected rays with exactly the same trajectories as the one shown with the solid
blue line (modified from Alkhalifah and Tsvankin, 1995). 94
5.8 Influence of transverse isotropy on the P-wave NMO velocity. Calculations are done
with equation 5.6 normalized by (after Alkhalifah and Tsvankin, 1995).
The horizontal axes approximately correspond to sin ; the dip angles range from 0 to
70. (a) Models with the same but different and , (solid
black); (gray); and (dashed). The three curves in (a)
nearly overlap. (b) Models with different values: (solid black); (gray);
and (dashed). 96
5.9 Parameters of a 2D VTI model: (a) the P-wave vertical velocity , the Thomsen
coefficients (b) and (c) and (d) the anellipticity coefficient (after Han et al.,
2001). 97
5.10 Interval (a) P-wave NMO velocity and (b) anellipticity coefficient (solid lines) as func-
tions of the two-way vertical time estimated from subhorizontal and dipping reflections
at midpoints between 4.9 km and 6.7 km in Figure 5.9. The dashed line in (b) indicates
the true values of (after Han et al., 2001). 97
5.11 Same as Figure 5.10 but for midpoints in the range 11.0 km 16.8 km (after Han et
al., 2001). 98
5.12 Ray trajectories and pure-mode reflection traveltime in wide-azimuth CMP geometry
(after Grechka et al., 1999b). 99
5.13 NMO ellipse that has the azimuth of the major axis and the semi-axes =
and = . 99
5.14 NMO ellipse in a dipping VTI layer. 102
5.15 The strike-line P-wave NMO velocity (normalized by the zero-dip NMO
velocity) as a function of the ray parameter for dips ranging from 0 to 90 (after
Grechka and Tsvankin, 1998b). (a) Different models with the same but differ-
ent and (solid); (dashed); ,
(dotted). (b) Models with different values: (solid); (dashed);
(dotted). 102
5.16 Dipping reflector beneath a horizontally layered overburden. The effective NMO
ellipse in this model can be obtained from the generalized Dix equation 5.40 (after
Grechka et al., 1999b). 102
5.17 Plan view of the source and receiver positions (small circles) in a single superbin.
The superbin contains approximately 400 source-receiver pairs with a common-
midpoint scatter of about 80 m or 2% of the maximum offset (after Grechka and
Tsvankin, 1999b). 105

12815-OTE III Gretchka.indd 7 28-07-2009 10:13:42


List of Figures

5.18 Semblance curves obtained by the conventional velocity analysis, which ignores the
azimuthal dependence of moveout velocity (dashed) and by the azimuthal veloc-
ity analysis (solid). Arrows indicate the major reflection evens (after Grechka and
Tsvankin, 1999b). 105
5.19 Effective NMO eccentricities at (a) s and (b) s (after Grechka and
Tsvankin, 1999b). 106
5.20 Interval NMO eccentricities for the horizon between 2.14 s and 2.57 s (after Grechka
and Tsvankin, 1999b). 106
6.1 Hyperbolic (solid) and nonhyperbolic (dashed) moveout terms normalized by their
values at the offset-to-depth ratio equal to two. Equation 6.7 with the parameters
, was used for this computation. 112
6.2 (a) Synthetic seismogram of the P-wave reflected from the bottom of a VTI layer
described by parameters , and . The spread
length is equal to two reflector depths; the source pulse is a Ricker wavelet with central
frequency of 40 Hz. (b) Semblance contours in the coordinates calculated
with equation 6.11 for . (c) Same as (b) but plotted in the coordinates
(after Grechka and Tsvankin, 1998a). 113
6.3 Semblance contours for the model from Figure 6.2 after addition of linear traveltime
error that changes from +4 ms at zero offset to at the offset . The sem-
blance maximum corresponds to (after Grechka and Tsvankin, 1998a). 114
6.4 Traveltimes for different source positions (dots) recorded by a single downhole receiver
and the best-fit nonhyperbolic moveout curve (solid) computed using equation 6.11.
The offsets and traveltimes are doubled to simulate a reflection experiment (after
Grechka and Tsvankin, 1998a). 114
6.5 RMS time residuals (in ms) calculated with equation 6.11 for different pairs .
The residual for the best-fit model at the center of the contours is 6.7 ms (after Grechka
and Tsvankin, 1998a). 114
6.6 Time-migrated seismic line (after Alkhalifah, 1997). 116
6.7 Semblance analysis for CMP 300 in Figure 6.6 at different zero-offset times . Gray
lines correspond to contours of the effective . The maximum offset in the data is
4.3 km (after Alkhalifah, 1997). 116
6.8 Interval and (black) as functions of at CMP 300 in Figure 6.6. The gray curves
show the error bars caused by the picking uncertainties (after Alkhalifah, 1997). 117
6.9 Magnitude of the azimuthally-varying quartic coefficient computed using
equation 6.13 for a VTI layer above a planar reflector with the dip
, and . The arrows mark the azimuth of the reflector dip plane
(after Pech et al., 2003). 118
6.10 Source-receiver geometry for a superbin used for azimuthal NH move- out analysis.
Each blue dot marks a source-receiver pair; the polar angle corresponds to the source-
receiver azimuth from the north (see the numbers on the perimeter), whereas the
radius is the offset. The total fold for this superbin is 2,491 (after Vasconcelos and
Tsvankin, 2006). 120
6.11 Seismic data at superbin in Figure 6.10 after a conventional hyperbolic NMO correc-
tion. The gather is sorted by increasing offset and contains all azimuths, which are
irregularly sampled (Figure 6.10). Non- hyperbolic moveout manifests itself through
the curvature of the NMO-corrected events at long offsets (so-called hockey sticks).
The jittery character of the reflections suggests the presence of traveltime variations
with azimuth (after Vasconcelos and Tsvankin, 2006). 121

12815-OTE III Gretchka.indd 8 28-07-2009 10:13:42


List of Figures

6.12 Results of nonhyperbolic moveout inversion for the reflections from the Viking horizon
(the maximum offset-to-depth ratio is 2.5), the Blairmore (ODR = 2.0), the Lower
Vanguard (ODR = 1.9) and the Mississippian Unconformity (ODR = 1.8). The arrows
mark the estimated direction of the semi-major axis of the NMO ellipse. The number
by each arrow is the azimuth of the axis with respect to the north. All parameters are
the effective values for a given reflection event (after Vasconcelos and Tsvankin, 2006). 121
7.1 First-break P-wave times (in ms) recorded by a geophone placed at a depth of 4,509 ft
(1,374 m) in a vertical well at Rulison Field, Colorado, USA. The data were acquired
by the Reservoir Characterization Project, Colorado School of Mines (after Grechka et
al., 2007). 123
7.2 Measurements carried out for estimating anisotropy in a typical VSP geometry. The
traveltime difference, , between geophones (dots) located at a distance along a
wellbore defines the apparent slowness, . Three-component traces recorded
by each downhole geophone yield the direction of particle motion, , defined by the
polar polarization angle, , and azimuth, . The azimuths of the horizontal compo-
nents and of the geophones usually vary along the tool string and should be
treated as unknown unless independently measured (after Grechka et al., 2007). 124
7.3 P-wave slowness surface (in s/km) constructed from 3D VSP data (after Dewangan and
Grechka, 2003). 125
7.4 Horizontal projections of the P-wave polarization vectors (tick lengths are proportional
to the magnitudes of the projections) plotted at the source locations with respect to the
well, which is placed at the coordinate origin (after Dewangan and Grechka, 2003). 125
7.5 Exact functions for various ratios but fixed parameters
and (symbols). The dashed line shows the
WAA of given by equation 7.2 (after Grechka and Mateeva, 2007). 128
7.6 Geometry of walkaway VSP (after Grechka and Mateeva, 2007). Surface shot locations
are shown in red. Positions of 3C receivers in the borehole are marked in cyan. Depth-
migrated surface seismic data are displayed on the background of the isotropic P-wave
depth-velocity model (white color corresponds to the P-wave velocity in water, magenta
to the velocity in salt). 129
7.7 Particle-motion hodogram for a typical source-receiver pair used in the inversion. The
geophone axis is vertical; the axes and are horizontal but their azimuths
are unknown. Open circles indicate the picked first-break time, dots mark the particle
motions at 2 ms time increments. The hodogram corresponds to approximately one
quarter of the dominant period of P-waves (after Grechka and Mateeva, 2007). 130
7.8 P-wave traveltimes (a) and polarization angles (b). The dots indicate the quantities picked
from several traces in a common-shot gather. The solid lines are the linear traveltime fit (a)
and the mean polarization angle (b). They give the slowness-polarization pair, , which
is one data point for the inversion. The dashed lines correspond to one standard devia-
tion from the best-fit lines. These standard deviations are 0.25 ms for the times and 1 for
the polarization angles (after Grechka and Mateeva, 2007). 130
7.9 Slowness-polarization data and estimated parameters (a) in the salt at a depth of 18,500
ft (5,639 m) and (b) beneath the salt at a depth of 21,750 ft (6,629 m). Crosses associated
with each data point (black dot) indicate the standard deviations in the picked and
values. The solid circles correspond to the best-fit VTI functions, open circles to
those functions one standard deviation. The obtained P-wave vertical velocities in SI
units are (a) and (b) The solid red lines show
the isotropic dependencies implied by those (after Grechka and Mateeva, 2007). 131

12815-OTE III Gretchka.indd 9 28-07-2009 10:13:43


List of Figures

7.10 Vertical velocities (a) and anisotropy coefficients (b). The thin solid lines in (b) are the
standard deviations of and estimated from the uncertainties in the and picks
(after Grechka and Mateeva, 2007). 132
7.11 Anisotropy coefficient (a) and gamma-ray log for subsalt sediments (b). Thin lines
in (a) show the standard deviation of (after Grechka and Mateeva, 2007). 132
7.12 Azimuthal variation of the fitted vertical slowness (in s/kft) corrected for isotropy as a func-
tion of the P-wave horizontal polarization components for the best-fit orthorhombic model
in the depth range 4,510 ft4,910 ft (1,375 m1,497 m). The white circles indicate the slow-
ness variations expected in the absence of azimuthal anisotropy (after Grechka et al., 2007). 134
8.1 Effective stiffness coefficients and for a single set of dry cracks. The back-
ground velocities are , and density is ;
they yield the Lam coefficients . Symbols indicate dif-
ferent theoretical predictions: the first-order Hudsons (equations 8.20 and 8.21),
the second-order Hudsons (equations 8.21 and 8.26 8.28), Schoenbergs
(equations 8.4, 8.12, 8.13, and 8.15), and the NIA (equations 8.4 and 8.19), which
takes into account nonzero crack aspect ratios ( for all fractures). The bars cor-
respond to the 95% confidence intervals (the mean values two standard deviations)
of the numerically computed stiffness coefficients obtained for 100 random realiza-
tions of the fracture locations (after Grechka and Kachanov, 2006c). 144
8.2 Anisotropy coefficients of effective HTI media. The symbols
are the same as those in Figure 8.1 (after Grechka and Kachanov, 2006c). 144
8.3 Horizontal cross-section of the stress component through a model containing dry frac-
tures (wire spheroids). The crack density is . The arrow indicates the direction of
applied remote load, whose magnitude is 1 MPa (after Grechka and Kachanov, 2006c). 146
8.4 Horizontal cross-sections of the stress component for arrays of (a) non-intersecting
and (b, c, d) intersecting fractures. The aspect ratios of the fractures lie in the range
. The arrows indicate the directions of applied uniaxial remote load
whose magnitude is 1 MPa (after Grechka and Kachanov, 2006d). 148
8.5 Effective anisotropy coefficients of fractured media. The bars correspond to the 95%
confidence intervals (the mean two standard deviations) of the numerically comput-
ed coefficients. The triangles indicate their values for models with intersecting cracks.
The predictions of the linear-slip theory, which ignores the nonzero crack aspect ratios
(equations 8.4, 8.12, 8.13, and 8.15) and the non-interaction approximation, which
accounts for them (equations 8.4 and 8.19), are shown with , respectively. All
numerical effective models are triclinic but only Tsvankins orthorhombic coefficients
are displayed (after Grechka and Kachanov, 2006d). 149
8.6 Fracture geometries created to study the influence of crack shape on the effective properties.
All fractures are vertical and planar; their normals are parallel to the -axis. Geometries
4, 5, and 6 contain rock islands inside the cracks and model partially closed fractures
(after Grechka et al., 2006). 150
8.7 Misfits (equation 8.38) for the six fracture shapes in Figure 8.6 (after Grechka et
al., 2006). 151
8.8 Models containing three sets of vertical rectangular cracks. Fractures that intersect
their neighbors are shaded, isolated cracks are transparent. The background Poissons
ratio is (after Grechka et al., 2006). 151
8.9 Relative deviations (equation 8.39) from orthotropy of the numerically computed
effective stiffness tensors for intersecting crack arrays such as those shown in Figure 8.8
(after Grechka et al., 2006). 152

10

12815-OTE III Gretchka.indd 10 28-07-2009 10:13:43


List of Figures

8.10 P-wave seismic section at Rulison Field. The arrows mark the reflection events used for
azimuthal velocity analysis (after Vasconcelos and Grechka, 2007). 154
8.11 Output of the fracture characterization: the background velocities and of P- (a)
and S-waves (b), and the principal crack densities . The directions of the
principal fracture sets are shown with ticks; their lengths are proportional to the eccen-
tricities of the interval P-wave NMO ellipses (Figure 8.14b). The star indicates the well
location from where the FMI log shown in Figure 8.12 was acquired (after Vasconcelos
and Grechka, 2007). 154
8.12 Fracture count (blue) in well shown with star in Figure 93 and the 90% confidence
interval (dashed red) corresponding to the azimuth of the fracture set with the density
estimated from seismic data (after Vasconcelos and Grechka, 2007). 155
8.13 Vertical velocities and anisotropy coefficients ,
at Rulison reservoir (after Vasconcelos and Grechka, 2007). 156
8.14 The shear-wave splitting coefficient and the eccentricity of the P-wave NMO
ellipses at Rulison reservoir (after Vasconcelos and Grechka, 2007). 156

11

12815-OTE III Gretchka.indd 11 28-07-2009 10:13:44


12815-OTE III Gretchka.indd 12 28-07-2009 10:13:44
LIST
List OF TABLES
of Tables

3.1 Summary of kinematic ray-tracing methods. 69


4.1 Phase and group velocities of P-, SV- and SH-waves in the vertical symmetry-axis direc-
tion, , and in the horizontal isotropy plane, . 76
4.2 Average values of anisotropic coefficients measured on core samples by Wang (2002). 80

13

12815-OTE III Gretchka.indd 13 28-07-2009 10:13:44


12815-OTE III Gretchka.indd 14 28-07-2009 10:13:44
INTRODUCTION
Introduction

Numerous natural materials exhibit elastic anisotropy or the dependence of velocities of seismic
waves on the direction of wave propagation. The presence of elastic anisotropy on all spatial scales
ranging from the rock-forming minerals to the upper mantle and the inner Earths core is docu-
mented in thousands of publications. For us who are involved in exploration and development of
hydrocarbon resources, the relevant seismic anisotropy spans the range of scales from a microimage
of a core sample (~ 10-5 m) to a land or marine survey (~ 10+4 m) and primarily pertains to sedi-
mentary formations. In the mid-1980s, the mounting evidence of the ubiquity of seismic anisotropy
precipitated into the energy industry and made us both aware and concerned.
The concerns of practicing geophysicists are well founded. They are based on a clear understand-
ing that about anything we do to improve our data (for instance, employing longer arrays, acquiring
wide-azimuth seismic, utilizing multicomponent geophones) almost invariably makes the data more
susceptible to anisotropy. Consequently, ignoring anisotropy becomes less forgiving and more det-
rimental to our data-acquisition efforts. The same pertains to seismic data processing. With more
powerful computers at our disposal, we might, for example, take advantage of the vector nature of
seismic wavefields for imaging purposes. This, however, requires including anisotropic phenomena
because particle motions of body waves recorded by our geophones usually do not comply with the
isotropy-prescribed behavior of being either normal or orthogonal to the propagating wavefronts.
There are two well-recognized objectives in dealing with anisotropy in the oil and gas industry
and they are different at the exploration and field-development stages. In exploration, we typi-
cally would like to estimate anisotropy from the available seismic data, incorporate the obtained
estimates into the velocity model and migrate the data using this model in a hope to improve the
image of the target horizons compared to that in the best isotropic velocity field. At this stage, we
usually do not ask what causes seismic anisotropy and how it relates to the rock properties; as long
as our anisotropic image is superior to its isotropic counterpart, our job is done.
The anisotropy-related objectives at the development stage are quite different. There, we do
want to find out the physical reasons for the measured anisotropy, especially those pertaining to the
reservoir fluids and in-situ rock conditions. For instance, distinguishing shale and sand packages,
fracture characterization, or interpretation of time-lapse anisotropy would be typical goals pursued
at the field development stage.

Accomplishing those goals requires:


a sound understanding of basic principles of seismic wave propagation in anisotropic media;
the ability to properly model at least the main characteristics of propagating waves;
knowledge of plausible physical causes of anisotropy and their relations to geology, sedimen-
tology, and tectonics;
good seismic-processing skills; and
expertise in the inverse theory because it is extensively used to solve anisotropic parameter-
estimation problems.

This book exposes readers to these issues and explains how geophysicists in the energy industry
measure, interpret and use seismic anisotropy.

15

12815-OTE III Gretchka.indd 15 28-07-2009 10:13:44


Introduction

We begin with general definitions and intuitive explanations of the subsurface features that
could manifest themselves as seismic anisotropy. We then proceed with a discussion of plane waves
and rays in the presence of anisotropy and relate them to similar objects in more familiar isotropic
media. Next, we talk about Thomsen anisotropy parameters and emphasize their significance
for seismic data processing. After that, our discussion turns to the reflection traveltimes that are
routinely used to build seismic velocity models. We explain the influence of anisotropy on the con-
ventionally measured normal-moveout velocity and on the less important for velocity-model build-
ing but commonly observed nonhyperbolic moveout. Estimation of the relevant parameters from
traveltime data will be the focus of our discussion. The book concludes with two chapters describing
more specialized applications of seismic anisotropy to vertical seismic profiling (VSP) and fracture
characterization.
This book is written in conjunction with the OTE III course that the European Association of
Geoscientists & Engineers (EAGE) offers in 20092010. As I also teach the Society of Exploration
Geophysics (SEG) short course Seismic Anisotropy: Basic Theory and Applications in Exploration
and Reservoir Characterization with Ilya Tsvankin and an internal Shell course Introduction to
Seismic Anisotropy, the material presented in the book partially overlaps that covered in my other
courses.
The one-day format of the OTE III course necessitated leaving out several anisotropy topics that
are nevertheless important for certain applications. They include parameter estimation and imag-
ing for tilted transverse isotropy, anisotropic amplitude-versus-offset (AVO) signatures, converted
waves, sonic and laboratory anisotropy measurements and perhaps some others. To fill in those
holes the reader is referred to the bibliography, which lists many original contributions.
I would particularly recommend the well-written textbook by Tsvankin (2001), the course notes
by Thomsen (2002), as well as the more mathematically and rock-physics oriented books by Helbig
(1994) and Carcione (2001), which in some sense complement the classical crystal-acoustics texts
by Fedorov (1968), Musgrave (1970) and Auld (1973). The topics of anisotropic AVO and VSP are
thoroughly discussed in Rger (2002) and MacBeth (2002), respectively.
The list of books available in Russian is more limited. To the best of my knowledge, the only
existing textbook on seismic anisotropy was recently written by Goldin (2008). In addition to the
original Fedorovs (1965) book, there is another crystal-physics text by Sirotin and Shaskolskaya
(1978), two books on observations and applications of seismic anisotropy in exploration (Nevsky,
1974) and global seismology (Chesnokov, 1977) and a comprehensive anisotropic ray-tracing book
by Petrashen (1980).

16

12815-OTE III Gretchka.indd 16 28-07-2009 10:13:44


ACKNOWLEDGEMENTS
Acknowledgements

I feel indebted to numerous people. They include many geophysicists with whom, over the years,
I discussed various anisotropy-related issues. Those discussions shaped my current understand-
ing of seismic anisotropy. I am grateful to the colleagues who co-authored papers with me and
thus directly influenced my research path. My work as a Geophysics Associate Editor on Seismic
Anisotropy exposed me to some unconventional views and broadened the area of my interest. I am
especially thankful to Irina Obolentseva who introduced me to the subject of seismic anisotropy in
the early 1980s and to Ilya Tsvankin who was my senior colleague at Colorado School of Mines in
the late 1990s and with whom we obtained a number of the results described in this book.
Finally, while I believe that all relevant publications have been properly referenced and that the
equations in the book are correct, I sincerely apologize for any possible omissions and errors.

17

12815-OTE III Gretchka.indd 17 28-07-2009 10:13:44


12815-OTE III Gretchka.indd 18 28-07-2009 10:13:44
DEFINITIONS
1 Definitions AND EXAMPLES
and examples

1.1 Simple examples


Before defining seismic anisotropy formally, it makes sense to understand its meaning on an intui-
tive level. We will do it with examples.

1.1.1 Isotropy
Suppose we throw a rock in a lake. Naturally, we expect to see circles originating at the point where
our rock hits water. The outer circle is the front of the wave propagating along the water surface.
We ask why this wavefront has a circular shape? One possible answer is that the wavefront is circular
because the waves have the same speed in all directions. The property of a wave to propagate with a
velocity that does not depend on direction is called isotropy. Consequently, we say that the surface of
water is isotropic with respect to wave propagation.

1.1.2 Anisotropy
Let us consider a different experiment. Suppose we are given a block made of steel and soft rubber
layers which are glued together (Figure 1.1). We do not know the internal structure of the block but
we can squeeze it and observe its behavior. If we apply a load along the layers, that is, in the direc-
tion shown with white arrows in Figure 1.1, we discover the block to be stiff just a little bit softer

Figure 1.1: Block made of lay-


ers of steel (white) and rubber
(black). The apparent stiffness
of the block under a compres-
sive load depends on the direc-
tion of the load.

19

12815-OTE III Gretchka.indd 19 28-07-2009 10:13:44


Chapter 1

Figure 1.2: Fractures in sandstones of Checkerboard Mesa (Zion National Park, Utah, USA).

than steel itself.(1) If we squeeze it across the layers in the direction of black arrows in Figure 1.1, we
find the block to be much softer almost as soft as rubber. Thus, we will conclude that the stiffness
of our block depends on the direction of the applied load. In other words, the elastic behavior of the
block exhibits anisotropy.
Anisotropy can be also understood geometrically. For instance, let us count the number of the
steel-rubber pairs that compose a yardstick of a fixed length (red in Figure 1.1). This number is
going to depend on the direction of the yardstick, whereas the number of water molecules in the
previous example is approximately the same in any direction.
The experiment with the yardstick gives a qualitative insight into the directional dependence
of wave-propagation velocities in layered structures. As the wave velocity in rubber is smaller than
that in steel, it is plausible that each layer of rubber would slow down a wave propagating through
it. Taking our yardstick to be equal to the wavelength leads to the obvious conclusion that the wave-
propagation velocity is smaller across the layers than along them. This purely qualitative result is
known to be rigorously correct when the wavelength is at least three times greater than the thickness
of the combined steel-rubber pair; such a wavelength is needed to cancel out multiple reflections
occurring in the layered sequence (Rytov, 1956; Helbig, 1984).
There exists a practical question of whether the model of stiff and soft layers is realistic. Sometimes
it is. For example, a photo of sub-parallel cracks in the sandstones of Checkerboard Mesa (Figure 1.2)
looks very similar to our model. By analogy, we would expect fractures to slow down waves propagat-
ing across them (that is, horizontally in Figure 1.2) compared to waves propagating along the cracks
(vertically in Figure 1.2). This qualitative statement coincides with the conclusion of the popular
linear-slip theory (Schoenberg, 1980) developed to describe wave propagation in fractured media.

1 The discussion is intentionally kept qualitative here. In fact, the problem of obtaining the overall static
properties of layered materials has classic solutions (Riznichenko, 1949; Postma, 1955; Backus,1962). A
more modern treatment that accounts for a finite size of the layers is given by Grechka and Rojas (2007).

20

12815-OTE III Gretchka.indd 20 28-07-2009 10:13:44


Definitions and examples

1.2 Definition of seismic anisotropy


The discussed examples imply how seismic anisotropy could be defined. While several slightly dif-
ferent definitions exist (e.g., Helbig, 1994; Winterstein, 1990), velocities seem to be of the outmost
importance for anything done with seismic data. Therefore, our definition is:

Elastic media, where seismic velocities depend on the direction of wave propagation at some physical
points, are called anisotropic.

This definition brings together the notions of direction and physical point. They refer to dis-
tinctly different characteristics of elastic media but play equally important roles here. A dependence
on the direction specified, for instance, by the unit vector means anisotropy. A medium is said to be
anisotropic when seismic velocity depends on , that is, . In contrast, a velocity dependence
on physical point or on the spatial location , is called heterogeneity. For instance, an elastic medium
is heterogeneous when . In practice, we typically encounter media that are simultaneously
anisotropic and heterogeneous. As a result, seismic velocities there are functions of both arguments,
. Thus, at least conceptually, we can distinguish four different types of media:
homogeneous isotropic = const,
homogeneous anisotropic = ,
heterogeneous isotropic = and
heterogeneous anisotropic = .

Consequently, two physical experiments can be proposed to discriminate isotropy from anisotropy
and homogeneity from heterogeneity. Isotropy implies that a medium remains the same for any rotation.
This is not so for anisotropic media. A test for homogeneity is translation rather than rotation.
A block of elastic medium is homogeneous if for any translation within this block
and heterogeneous otherwise. Because rotation and translation are not derivable from each other, ani-
sotropy and heterogeneity characterize fundamentally different properties of elastic media.

1.2.1 Anisotropy versus heterogeneity


Yet, it is not difficult to see that anisotropy and heterogeneity are related. Let us return to our steel-
rubber model depicted in Figure 1.1. Now suppose that we are allowed to examine the internal
structure of the composite. It would reveal heterogeneity as the cause of anisotropy that we experience
squeezing the block in different directions. Thus, heterogeneity on the scale of individual layers appears
as anisotropy on the scale of the entire block. Generalizing our observation, we might say that

ordered heterogeneity on microscale results in anisotropy on macroscale.

This statement turns out to be very general. Indeed, any solid is heterogeneous on the scale
comparable with the distances between its molecules or atoms. Heterogeneity and ordering on a
molecular scale, which are responsible for the anisotropy of minerals, comprise the subject of crys-
tal acoustics (e.g., Musgrave, 1970). This type of heterogeneity however, does not necessarily entail
anisotropy observed on a seismic scale because the rock-forming minerals might be oriented ran-
domly in the Earth, leading to seismic isotropy. Examples of this sort are well known. For instance,
although most of the minerals composing sandstones are anisotropic, the sandstones themselves
are nearly isotropic on scales of a core sample (Wang, 2002) and larger. Likewise, while salt (NaCl)
crystals are distinctly anisotropic (Sun, 1994), VSP data acquired in salt bodies in the Gulf of Mexico
unambiguously point to the salt isotropy on seismic scale (e.g., Grechka and Mateeva, 2007).

21

12815-OTE III Gretchka.indd 21 28-07-2009 10:13:45


Chapter 1

An important conclusion that emerges from our discussion is that the property of materials to
be isotropic or anisotropic depends on the scale of investigation or on the applied yardstick. For
wave propagation, this yardstick is the wavelength. Therefore, when we talk about anisotropy, our
statements implicitly refer to a certain wavelength and thus imply a certain size of the physical
point. In this book the physical point usually corresponds to the seismic wavelength, whose linear
size is about 10+2 m. To investigate anisotropy and separate it from heterogeneity on a smaller
scale, we would normally use shorter wavelengths. For instance, the sequence

seismic sonic core sample microtomographic image

represents a conventional set of decreasing scales from some 10+2 10+4 m to 10-5 m.(2)

1.3 Seismic examples


While formal definition of seismic anisotropy given in the previous section was needed to specify
our subject area, it tells us neither how anisotropy manifests itself in seismic data nor how we might
know whether the subsurface is anisotropic. We will discuss those issues next.

1.3.1 Depth misties of seismic data


The fundamental implication of anisotropy for wave propagation arises from the very fact that
seismic wavefronts excited by point sources in the Earth are usually non-spherical (for comparison,
refer to the example described in section 1.1.1). Their deviation from spheres has the following
consequences for seismic imaging:

1. The migration velocity, which is usually derived from stacking-velocity analysis, is different from
the vertical or check-shot velocity. As a result of this discrepancy, the imaged horizons are mis-
positioned in depth; they are typically too deep because the stacking velocities are often greater
than the check-shot velocities.
2. If one uses the true vertical velocity to position reflectors correctly, the image is poorly focused
for the same reason: the vertical velocity is different from the stacking velocity and thus results
in a suboptimal image.
3. Wavefronts of conventional P-waves usually deviate not only from a sphere but also from an ellip-
soid. This means that isotropic migration is no longer capable of imaging different dips on the
same section because focusing of those dips requires different isotropic velocities. Also, if conflict-
ing dips (e.g., intersections of faults and horizons) are present in the section, their images become
shifted with respect to each other, which might result in geologically implausible structures.

We illustrate these features with a series of depth migrations of a 2D synthetic data set generated by
J. Leveille and F. Qin (pers. comm.) with an anisotropic finite-difference code. The model is shown
in Figure 5.9; the details of the processing sequence and imaging can be found in section 5.2 and
in Han et al. (2001). Figure 1.3 compares four prestack depth images of a portion of the model
containing a dipping fault.
The images in Figures 1.3a and 1.3b have similar overall quality except for a slight deteriora-
tion in the focusing of the fault plane in Figure 1.3b, which can be explained by our inability to

2 Let us also mention that modern techniques make it possible to obtain some information about hetero-
geneity even from static (infinite wavelength) measurements (Huet, 1990; Ostoja-Starzewski, 2008). This
information, however, is necessarily incomplete.

22

12815-OTE III Gretchka.indd 22 28-07-2009 10:13:45


Definitions and examples

reconstruct the fine details of spatial variations of velocity and anisotropy from reflection seismic
data. Image 1.3a also contains a multiple arrival (marked with the black arrow) that was only par-
tially attenuated. The main difference between Figures 1.3a and 1.3b are the horizon depths (white
arrows). As reflection data do not constrain the true vertical velocity, the stacking velocity was used
in its place for the anisotropic migration in Figure 1.3b. Evidently, this has led to the distorted
depth scale of the entire image all horizons are too deep.
Figure 1.3c can be considered as an ideal output of the conventional isotropic processing
sequence. The isotropic velocity model used to generate Figure 1.3c is based on the correct stack-
ing velocity that may be obtained, for instance, from noise-free semblance analysis. Comparing the
subhorizontal reflectors in Figure 1.3c (white arrows) with those in Figure 1.3a, we notice that they
are mispositioned, which was also the case in Figure 1.3b. The overall quality of the image however,
is comparable to those in Figures 1.3a and 1.3b. For example, there are no conflicting dips in the
vicinity of the fault, which could be indicative of the presence of anisotropy. The continuity and
crispness of the fault-plane reflection is somewhat inferior to those in Figure 1.3a but the difference

Figure 1.3: Anisotropic and isotropic depth-migrated sections obtained in (a) exact anisotropic model,
(b) anisotropic model estimated from surface reflection data, (c) isotropic model constructed via con-
ventional processing, and (d) isotropic model with the correct vertical velocity (after Han et al., 2001).

23

12815-OTE III Gretchka.indd 23 28-07-2009 10:13:45


Chapter 1

is minor. A good quality of this isotropic image is explained by the small values of the wavefront
anellipticities in the original model that control the dip dependence of the stacking velocity in ani-
sotropic media (section 5.2). A greater anellipticity would degrade the image quality.
Another option for choosing the isotropic migration velocity is illustrated in Figure 1.3d. This
time the data were migrated with the correct vertical velocity, which may be obtained from check
shots or well logs but generally not from the surface reflection data. As expected, all subhorizontal
reflectors in Figure 1.3d are correctly positioned in depth (white arrows). However, the quality of
the image is considerably lower than in Figures 1.3a 1.3c. The difference between the vertical and
the correct stacking velocity causes misstacking of the subhorizontal events throughout the section.
As this difference also distorts the dip dependence of the stacking velocity, it mispositions the fault
causing the intersections of reflectors with different dips (white circles in Figure 1.3d). Clearly, just
about any seismic interpreter would conclude that the data in Figure 1.3d were migrated with an
inappropriate velocity.
The presented analysis was possible because all pertinent model parameters were known. While
this is not the case in reality, the misties, that is, discrepancies between the depths of seismic hori-
zons and well markers are routinely observed. In many cases, their origin is the difference between
the stacking and vertical velocities, that is, seismic anisotropy.

1.3.2 Nonhyperbolic moveout


One of the best known manifestations of seismic anisotropy can be seen in reflection data recorded at
large offsets that exceed the reflector depth by a factor of two or more. At such long spreads, the P-wave
moveout is often nonhyperbolic. Although several reasons can be put forward to explain the moveout
nonhyperbolicity (e.g., Fomel and Grechka, 2001), the most common is the subsurface anisotropy that
makes the shapes of seismic wavefronts anelliptic (Alkhalifah and Tsvankin, 1995; Alkhalifah, 1997).
If long-spread data are migrated isotropically, the so-called hockey sticks are observed (Figure
1.4a). The same hockey sticks are typically present on common-midpoint gathers corrected using
the conventional hyperbolic moveout equation. To remove them, we need to either mute out the
large-offset data (and lose the stacking power and the ability to do a subsequent AVO analysis at
long offsets) or apply a nonhyperbolic moveout correction (e.g., Alkhalifah and Tsvankin, 1995; Siliqi
and Bousqu, 2000). Figure 1.4b shows that such a correction, which is a part of anisotropic migra-
tion, successfully removes the hockey sticks and flattens the data for the entire offset range.

1.3.3 NMO ellipse


The popularity of multi-azimuth and wide-azimuth seismic data acquisition has recently surged due
to the ability of such data to illuminate and image targets beneath a complex overburden (subsalt
imaging in the Gulf of Mexico would be a typical application area). Acknowledging this growing
trend, both The Leading Edge (2007) and First Break (2008) devoted their pages to special sections
on land and marine wide-azimuth seismic. Once we open up the registration azimuth, however, we
make our data sensitive to azimuthal anisotropy. In its presence, which could be due to preferen-
tially-oriented cracks or dipping shale-sand sequences, we expect the conventional stacking or the
normal moveout (NMO) velocity to vary azimuthally. This variation is known to be elliptical in the
horizontal plane (Grechka and Tsvankin, 1998b) and termed the NMO ellipse.
Figure 1.5 shows the P- and shear-wave NMO ellipses observed in the tight-gas Rulison Field in
Colorado, USA. Clearly, we need to apply azimuthally-varying NMO velocity to flatten such moveouts.
Failing to do so results in suboptimal stacking and blurry images. In addition, the orientations
and eccentricities of the NMO ellipses can be related to the fracture parameters. This important
information for tight-gas development would be lost if the data are processed isotropically and the
azimuthal anisotropy is ignored.

24

12815-OTE III Gretchka.indd 24 28-07-2009 10:13:46


Definitions and examples

Figure 1.4: Common-image gathers


after prestack depth migration with
(a) isotropic and (b) anisotropic
velocity model (after Sarkar and
Tsvankin, 2004).

1.3.4 Converted waves


Multicomponent data often allow us to establish the presence of anisotropy in the subsurface with
relatively little processing effort. In the marine environment, the word multicomponent invariably
implies converted (PS) waves because reliable and efficient shear-wave marine sources are unavail-
able. Thus, waves excited with air guns propagate down to a reflector as P-modes, whereby they
generate two reflected waves: P and S. Both PP and PS reflections are recorded at the sea bottom
with three-component (3C) geophones.
Because the PP- and PS-waves are influenced by the rock properties and fluids differently, they
usually provide complementary information about the subsurface. Having this information is thus
helpful for both exploration and reservoir characterization. To make use of these potential benefits,
the PP and PS images should tie in depth. This obvious requirement, however, becomes problematic
if we migrate the PP and converted-wave data isotropically.
To understand the nature of the problem we need to realize that co-depthing of the PP and
PS images imposes a certain ratio on the vertical times and the vertical velocities of P- and shear-
waves. Fixing this ratio in isotropic media means that only one stacking velocity (say, the P-wave
velocity) needs to be estimated from the data, while the other (the S-wave velocity) is implied
by the velocity ratio. Therefore, once a conventional isotropic P-wave velocity model is built, the
model for converted-wave migration can be immediately constructed based on the co-depthing
requirement.
Unfortunately, PS-wave images obtained under the assumption of isotropy are usually noisy
and poorly focused. Figure 1.6a displays such a common-conversion-point stack obtained over the
Siri reservoir in the North Sea. The reason for suboptimal image is the effective anisotropy of the

25

12815-OTE III Gretchka.indd 25 28-07-2009 10:13:46


Chapter 1

Figure 1.5: P- (a) and shear-wave (b) gathers after the azimuthally-invariant (isotropic) NMO correc-
tion. The traces are sorted by the source-receiver azimuth. The apparent cosine-type dependence of the
residual moveout is indicative of azimuthal anisotropy. The jitter obscuring this dependence is caused
by slightly unequal offsets at different azimuths. The events at approximately 1.27 s (a) and 2.35 s (b)
are reflections from the bottom of the Rulison reservoir (after Vasconcelos and Grechka, 2007).

subsurface. Essentially, it makes the ratio of P-to-S vertical velocities, which honor the image co-
depthing, different from the ratio of the P- and S-wave NMO velocities needed for optimal stacking
of the converted waves. Consequently, isotropy becomes too restrictive for imaging of both PP- and
PS-wave reflection data.
Alternatively, when we deem the subsurface anisotropic, estimate its anisotropy and incorpo-
rate it into the velocity model, the image quality increases dramatically. Figure 1.6b attests to that.
Comparing the isotropic (Figure 1.6a) and anisotropic (Figure 1.6b) stacked sections reveals a sig-
nificant improvement achieved by accounting for anisotropy. First, application of accurate NMO
velocities in the anisotropic model boosts higher frequencies in the stacked reflections and, there-
fore, increases the temporal resolution. Second, anisotropic processing provides a crisp picture of
faulting (outlined with a dashed box) in the shallow part of the section and significantly improves
the image of the top of the reservoir (white arrow). Finally, a poor PS-wave moveout approximation
under the assumption of isotropy creates noise that does not stack out (Figure 1.6a). The anisotropic
section is much cleaner (Figure 1.6b) because the converted-wave moveout is better reproduced in
the anisotropic model.

1.3.5 PP- and PS-wave AVO


Figure 1.7 sheds additional light on the problems of isotropic converted-wave imaging. To accu-
rately image the faults in the shallow part of the section in Figure 1.6, it is necessary to account
for anisotropy in computing the common-conversion-point (CCP) trajectories. Poor focusing and
positioning of the fault-plane reflections on the isotropic section (Figure 1.6a) can be explained
by the conversion-point smearing due to vertical heterogeneity and unaccounted anisotropy. This

26

12815-OTE III Gretchka.indd 26 28-07-2009 10:13:46


Definitions and examples

Figure 1.6: Isotropic (a) and anisotropic (b) PS-wave common-conversion-point stacks (after Grechka
et al., 2002c).

Figure 1.7: Errors in (a) the lateral position of the PS-wave conversion point and (b) the incidence
angle of the P-leg for PP- and PS-waves at the Siri reservoir caused by neglecting anisotropy (after
Grechka et al., 2002c).

smearing at the target level exceeds 500 m for the largest offset in the data (Figure 1.7a) and reaches
about 340 m for the maximum offset (2,600 m) used to produce the stacks in Figure 1.6.
Such shifts of the conversion points not only reduce the lateral resolution but also bias the iso-
tropic AVO responses for PS-waves because each CCP gather includes reflections from a wide range
of the subsurface locations. In addition, neglecting anisotropy in AVO analysis introduces errors in
the offset-to-angle transformation that reach 8 for converted waves at an offset of 3 km but remain
relatively small for the PP reflections (Figure 1.7b).

27

12815-OTE III Gretchka.indd 27 28-07-2009 10:13:46


Chapter 1

Figure 1.8: Three-


component record (a, b,
c) of a shear microseismic
event and the rotated 3C
trace (d, e, f) that reveals
shear-wave splitting.

1.3.6 Shear-wave splitting


Perhaps the most direct evidence for seismic anisotropy is shear-wave splitting. As only one body
S-wave can propagate in isotropic media, no splitting occurs there. The presence of anisotropy gives
rise to two S-waves that generally have different polarizations and velocities. Hence, one shear-wave
arrives to the geophones faster than another. This phenomenon is termed splitting (or birefrin-
gence) because the shear waveforms recorded by three-component (3C) geophones appear to split
into the fast and slow shear modes as their travel distances increase.
Numerous records of shear waves exhibit splitting. It has been extensively observed in the earth-
quake seismology (e.g., Crampin et al., 1984), multicomponent surface reflection data (Alford, 1986;
Lynn and Thomsen, 1986; Mueller, 1992) and VSP (Winterstein and Meadows, 1991; Winterstein
et al., 2001). Nowadays, when producing fields are often instrumented with passive 3C listening
devices, split shear-waves are frequently identified in the recorded microseismicity.
Figure 1.8 presents an example of such a microseismic event. The left column shows 200 ms of
raw 3C data. The zero time corresponds to the estimated occurrence time of the event. The geo-
phone axes and are directed to the north and east, respectively, while the -axis is vertical. The
right column in Figure 1.8 exhibits the same data, which were rotated within the shear plane to
separate the expected shear arrivals into the - and -components. By definition, the shear plane
is normal to the -component (Figure 1.8f) that points in the direction of the minimum energy
of the particle motion in the selected time window (200 ms in this example). We observe that the
- and -components in Figures 1.8c and 1.8f look very similar. This happens because the exam-
ined shear-waves propagate approximately vertically and their particle-motion directions are close
to the horizontal plane. Therefore, the shear plane is nearly horizontal.
To separate the recorded shear-waves into the fast S1 and slow S2 modes, the original data were
rotated to the new directions and (Figures 1.8d and 1.8e), which maximize the normalized
cross-correlation, ( t), of the rotated traces. Its maximum, ( t) = 0.92, is located at the time shift
t = 26 ms. This t is the estimated shear-wave splitting time in ms. The blue dots in Figures
1.8d and 1.8e show the time-shifted traces (t+ t) and (t t), respectively, and make it pos-
sible to visually assess the similarity of the split shear waveforms.

28

12815-OTE III Gretchka.indd 28 28-07-2009 10:13:47


Definitions and examples

The time delay t allows us to calculate the S-wave splitting coefficient, , which is approxi-
mately equal to the ratio of t and the total shear-wave propagation time. This yields 1.9%,
which is a typically observed magnitude of the shear-wave splitting coefficient (e.g., Winterstein et
al., 2001).

1.3.7 Shear-wave amplitudes


Two shear-waves propagating in anisotropic media differ not only in their velocities. Having
approximately orthogonal particle motions or polarizations, they also interact differently with the
subsurface heterogeneities. For instance, in naturally fractured formations containing a dominant
set of oriented cracks, the fast shear-wave S1 is polarized along the fracture planes. As a result, it is
relatively insensitive to the fractures and its reflectivity is primarily controlled by the elastic prop-
erties of the unfractured formation. In contrast, the reflectivity of the slow shear-wave S2, whose
polarization generally points at an oblique direction to the fractures, is expected to have certain
sensitivity to the fractures themselves. This allows us to envision that relative amplitudes of two split
shear-waves might help in diagnosing the presence of oriented cracks in the subsurface.
Mueller (1992) demonstrated the feasibility of such an approach. He applied Alford (1986)
rotation to shear-wave data acquired over the naturally fractured Austin Chalk trend in Texas,
USA and showed that, indeed, the reflectivity of S2-waves is influenced by the cracks. Figure 1.9a
displays the fast shear-wave section at the Austin Chalk level, which is dipping from 4.85 s on
the left to 4.95 s on the right. The S1 section exhibits a consistent reflection amplitude in marked
contrast to the slow shear-wave section (Figure 1.9b), which shows a considerable amplitude vari-
ation. The area in the middle of the S2-section, where the slow shear-wave amplitude is weak,
is approximately 250 m wide and corresponds to oil-bearing, near-vertical fractures (Mueller,
1992). Clearly, ignoring the shear-wave splitting and processing the data as if the subsurface were
isotropic would make fracture characterization difficult, if not impossible, and could result in a
missed field-development opportunity.

Figure 1.9: Fast (a) and slow (b) shear reflections from the Austin Chalk horizon (after Mueller, 1992).

29

12815-OTE III Gretchka.indd 29 28-07-2009 10:13:47


Chapter 1

To conclude this section, it suffices to say that the subsurface anisotropy manifests itself in many
types of seismic data and in a variety of forms. As it influences both the traveltimes and amplitudes
of seismic waves, ideally, it should be incorporated in both imaging and AVO workflows. Perhaps the
only option to get away with isotropic processing is to limit data registration to P-waves and acquire
them in conventional-offset, narrow-azimuth geometries. The seismic industry, however, is heading
in the opposite direction.

1.4 Physical causes of seismic anisotropy


While the anisotropic phenomena discussed in the previous section might seem diverse and dif-
ficult to classify, they are caused by essentially the same mechanism. It has already been identified
in section 1.2.1:

ordered heterogeneity on microscale appears as anisotropy on macroscale.

Thus, to understand the physical reasons for anisotropy observed in seismic data, we need to
find some widespread, ordered, small-scale heterogeneities in the subsurface. If those heterogenei-
ties happen to be intrinsically anisotropic, their alignment would tend to preserve the intrinsic
anisotropy on a larger scale.

1.4.1 Shales
Such small, anisotropic, aligned particles in sedimentary rocks were recognized a long time ago.
They are the clay minerals (e.g., illite, biotite, kaolinite and muscovite) that comprise a significant
portion of shales. All clay minerals have platelet-type shapes and are known to be extremely aniso-
tropic. According to the existing laboratory measurements (Alexandrov and Ryzhova, 1961; Belikov
et al., 1970; Katahara, 1996), the ratios of the P-wave velocities along and across the platelets often
exceed 1.5, while the same ratios for the shear-waves polarized in the platelet planes are usually
greater than 2 and can reach 3.5.
The alignment of clay platelets in shales is thought to be related to the sedimentation process. As
sedimentation takes place in the gravity field, the platelets are deposited approximately horizontally.
Deviations from their horizontal orientation can be attributed to the presence of round particles.
Figure 1.10 shows both those particles and the clay platelets in a typical shale. The platelet misalign-
ment reduces the magnitude of shale anisotropy compared to that of clays. Still, as long as the orien-
tations of the short platelet axes are not completely random, some portion of the intrinsic anisotropy
of clays survives in their mixture and exhibits itself as the effective anisotropy of shales.
The final step in our upscaling sequence

mineral rock stratum

rests on geologic evidence that shales are probably the most common rock type in the first few
kilometers below the Earths surface. In fact, by some estimates, shales compose up to 75% of oil-
and gas-producing sedimentary basins worldwide. This makes shale anisotropy a major contributor
to the observed seismic anisotropy. As a confirmation of this statement, we quote the study of Banik
(1984) who examined several data sets from the North Sea and found a clear correlation between
the seismic depth misties, which quantify the differences between the vertical and stacking velocities
and therefore provide a measure of anisotropy, and the fraction of shales in the subsurface.

30

12815-OTE III Gretchka.indd 30 28-07-2009 10:13:47


Definitions and examples

1.4.2 Fractures
Another ubiquitous feature of rocks is the presence of fractures. For our purposes, a fracture can be
defined as a compliant planar inhomogeneity. That is,
(i) one fracture dimension the width is much smaller than the other two and
(ii) the fracture infill is much more compliant than the host rock.

For example, the rubber layers in the steel-rubber model in Figure 1.1 can be treated as fractures
provided that their linear dimensions are much greater than their thicknesses. Because their infill
is elastic, geologists would call these fractures veins. In seismic practice, however, it is customary
to distinguish between the liquid-filled and dry (or gas-filled) fractures.
It is widely accepted that the fracture lengths cover the entire range of scales of exploration interest
(e.g., Narr et al., 2006). To emphasize this length variety, a special terminology has been introduced
in geologic literature. It includes such terms as microcracks, fractures, joints and faults. Although
undoubtedly helpful for geologic description of outcrops, this terminology has relatively little rel-
evance for us and we will use the terms fractures and cracks interchangeably. The insignificance
of geologic terminology stems from the fact that the influence of fractures on seismic waves depends
not on the absolute fracture sizes but rather on the ratios, r, of those sizes to the wavelength.
In general, three physically different regimes can be recognized. First, when r << 1, contributions of
multiple fractures are averaged by a propagating wave and the entire rock behaves as if it were homo-
geneous. Second, when r~1, the dispersion and scattering phenomena prevail, giving rise to extremely
complex wavefields and thus presenting undeniable evidence for the medium-scale heterogeneity. Third,
when r >> 1, individual fractures generate distinct reflections and appear as faults on seismic sections.
Here we are concerned only with the first regime, r << 1, when a rock mass is effectively homo-
geneous and might be anisotropic. Its anisotropy is primarily controlled by the fracture orientations,
or more specifically, by the angular distribution of the directions of normals to the fracture faces.
The more pronounced peak this distribution has around a certain spatial direction, the stronger the
crack-induced anisotropy. One end member of this distribution a single set of cracks (Figure 1.2)
was qualitatively discussed in section 1.1. The opposite end member comprises randomly oriented
fractures that result in the overall isotropy.

Figure 1.10: Shale under


a microscope (after
Hornby et al., 1994).

31

12815-OTE III Gretchka.indd 31 28-07-2009 10:13:48


Chapter 1

In reality, however, these end members are observed infrequently. The reason is that gravity
usually creates nonhydrostatic stresses in rocks. As those stresses are compressive, they tend to close
any preexisting cracks. Therefore, open fractures are most likely to be oriented normally to the
minimum stress direction if it remains unchanged for a sufficiently long time. When the subsurface
structure is not very complicated, this stress direction is close to the horizontal plane and conse-
quently, nearly vertical fractures are often encountered.
Predicting the azimuths of those fractures from any available geologic or stress data proves to be
much more difficult than predicting the polar angles of normals to their faces. The difficulty arises
from the variability of the minimum stress direction in the course of geologic history of a given
formation. This history usually includes at least several changes in the tectonic regime which are
arguably accompanied by the acts of opening and closure of differently oriented fracture sets. As a
result, the contemporary minimum stress direction correlates poorly with the modern-day fractures
(Laubach et al., 2004; Narr et al., 2006). The absence of this simple correlation shows up explicitly
in the existence of multiple fracture sets (such as those displayed in Figure 1.11), contrary to the
notion that only a single set of cracks oriented normally to the minimum stress direction is permit-
ted. In fact, numerous well-image log data indicate that multiple fracture sets are the rule rather
than an exception. Still, the approximately vertical orientations of those sets make the subsurface
effectively anisotropic. We will discuss crack-induced anisotropy in more detail in Chapter 8.

1.4.3 Fine layering


The isotropic fine (compared to the wavelength) layering is probably the least important cause
for anisotropy observed in seismic data. This might seem odd because finely layered structures
were examined historically first and shown to give rise to effective anisotropy (Riznichenko, 1949).
The problem of describing layer-induced anisotropy is usually considered to be solved by Backus
(1962), who proved that any isotropic, finely layered composite is equivalent to a homogeneous

Figure 1.11: Sets of well-developed joints in Caithness, Scotland, UK (source: Wikipedia, http://en.
wikipedia.org/wiki/File:Joints_Caithness.JPG).

32

12815-OTE III Gretchka.indd 32 28-07-2009 10:13:48


Definitions and examples

Figure 1.12: Equivalence


of finely layered isotropic
and homogeneous ani-
sotropic solids for the
propagation of long seis-
mic waves.

anisotropic medium with a symmetry axis oriented normally to the layers. This is schematically
shown in Figure 1.12. Thus, the seismic exploration community had a good knowledge of anisot-
ropy induced by fine layering by the early 1960s, when the first paper devoted to crack-induced
anisotropy was published by Bristow (1960).
It is, therefore, instructive to understand why anisotropy due to fine layers happens to be weaker
and, hence, less important than that due to cracks in concentrations significant for the oil industry.
Figure 1.1 is helpful in this respect. We already mentioned that thin rubber layers can be treated as
cracks filled with a compliant material. So what is the difference between the cracks and the layers?
This difference has a purely quantitative nature. The fractures of interest for exploration and field-
development are filled with fluids whose shear moduli are equal to zero (or nearly zero if a non-
vanishing fluid viscosity and finite frequencies of propagating waves are taken into account). Those
low shear moduli make fluid-filled cracks extremely high-contrast inclusions in the surrounding
rock. Clearly, the contrast in the shear moduli of the rock and the fluid infill of a fracture is almost
equal to the rock shear modulus itself. On the other hand, when we talk about fine layers we usually
refer to variations in the elastic properties encountered in well logs (Figure 1.13). Those variations
are obviously much smaller.
To demonstrate why contrasts in the elasticities of different constituents of a composite matter,
we refer to the paper published by Bakulin and Grechka (2003). Let us describe the relevant local
elastic parameters by the vector , , whose components generally
depend on the spatial coordinates , and the corresponding effective parameters by the vector
. For upscaling problems, where isotropic heterogeneity causes the over-
all anisotropy, all local components quantifying anisotropy are equal to zero, =0; they
are just the placeholders in vector .
Bakulin and Grechka (2003) showed that any effective parameter has the form

(1.1)

where is the mean value or the volume average of are the spatial fluctuations,
, of this (when ) and all other (when ) local components of ,
and the symbol denotes a quantity that has the order of . If the magnitudes of all fluctua-
tions are small,

33

12815-OTE III Gretchka.indd 33 28-07-2009 10:13:48


Chapter 1

(1.2)

equation 1.1 implies that approximation

(1.3)

is sufficiently accurate because of the quadratic dependence of on . In other words, under


conditions 1.2, local fluctuations make only second-order contributions to the effective properties.
The result expressed by equations 1.1 and 1.3 and inequality 1.2 has a direct bearing on the
magnitude of anisotropy induced by isotropic fine layering and by cracks located in an otherwise
isotropic host rock. If and are the parameters quantifying the local and effective anisot-
ropy, = 0 and, in accordance with equation 1.1,

(1.4)

Therefore, the magnitude of effective anisotropy is governed solely by the strength of the fluctua-
tions in the local bulk and shear moduli.
We note that inequality 1.2 is clearly violated for liquid- and gas-filled fractures. In fact, ~
max for the shear modulus , and the crack-induced anisotropy described by equation
1.4, is not necessarily weak. In contrast, for typical layered sequences, inequality 1.2 is often valid,
implying a weak layer-induced anisotropy.
The logs displayed in Figure 1.13 illustrate this point. While the vertical velocity heterogeneity is
sizeable, Backus (1962) averaging applied to the entire interval yields fractional differences between
the vertical and horizontal velocities that do not exceed 1% for both P- and shear-waves. Although
those differences can be made greater if Backus averaging is restricted to the most heterogeneous
narrow depth intervals (this has been shown, for example, by Liner and Fei, 2006), the overall con-
clusion drawn from the logs in Figure 1.13 remains valid:

Figure 1.13: Sonic (black),


shear (blue) and density (red)
logs acquired over a 500 m
interval in a deepwater oil field
in the Gulf of Mexico (after
Bakulin and Grechka, 2003).

34

12815-OTE III Gretchka.indd 34 28-07-2009 10:13:49


Definitions and examples

the layer-induced anisotropy is usually weak much weaker than anisotropy routinely observed in
seismic data.

In summary, we saw that seismic anisotropy is mainly caused by the presence of shales and
oriented fractures in the subsurface. Anisotropy due to layering usually plays a subsidiary role.
Sometimes, people point to nonhydrostatic compressive stresses as to another, independent of the
already mentioned physical cause of anisotropy. Such stresses are known to result in stress-induced
anisotropy, whose analysis in crystals and metals comprises a branch of materials science (e.g.,
Murnaghan, 1951; Thurston, 1965; 1974 and references therein). Changes of stresses in rocks,
however, are invariably accompanied by the opening and closure of preexisting fissures and small
fractures. This makes it possible to satisfactorily describe stress-induced anisotropy in terms of
crack-induced anisotropy (Shapiro and Kaselow, 2005).

35

12815-OTE III Gretchka.indd 35 28-07-2009 10:13:49


12815-OTE III Gretchka.indd 36 28-07-2009 10:13:49
PLANE
2 Plane WAVES
waves

After discussing seismic anisotropy qualitatively in Chapter 1, we turn our attention to its quantita-
tive treatment. Our goal in this chapter is to describe the simplest wave object a plane wave and
analyze its properties in anisotropic media. We will mostly be concerned with the velocities and
slownesses of plane waves because they are necessary for describing the kinematics of seismic rays
that play a critically important role in anisotropic velocity-model building and imaging.

2.1 Strain, stress and the equation of motion


There are many excellent books on linear elasticity. They make it possible to keep our discus-
sion brief and refer for detail to classic texts by Love (1944), Landau and Lifshitz (1959), Aki and
Richards (1980) and Ben-Menahem and Singh (1981). Nevertheless, this section contains perhaps
the most complex math in the book. Therefore, a mathematically challenged reader might skip to
the next section and go back only if needed.
The basic quantities used to describe a state of deformation (or distortion) of a solid are strain
and stress. To introduce the strain tensor, we consider two closely spaced points that have the
Cartesian coordinates and + d before deformation. Once the solid is deformed, the point that
initially was at moves to the new location + ( ). The vector is called the displacement. The
displacement of a neighboring point, + d , is given by ( ) + d or, equivalently, by ( + d ).
Assuming that |d | is vanishingly small, we can expand the displacement ( + d ) in a Taylor
series and write

(2.1)

where dots denote negligible terms that have order |d |2 and higher. The quantity is known as
the displacement gradient. It follows from equation 2.1 that d is equal to the dot product . d ,

(2.2)

or in components,

(2.3)

Hereafter, summation with respect to repeated low-case Roman indexes from 1 to 3 is assumed.
Equation 2.3 has an equivalent form (Aki and Richards, 1980; Ben-Menahem and Singh, 1981),

(2.4)

37

12815-OTE III Gretchka.indd 37 28-07-2009 10:13:49


Chapter 2

where the first term,

(2.5)

is the strain tensor. By definition, it is the second-order symmetric tensor. The word symmetric
means that

(2.6)

When the components of the displacement gradient are infinitesimal, that is,

(2.7)

the second term in equation 2.4 describes a rigid rotation of a small volume around point .
Inequality 2.7 is an important assumption of the linear theory of elasticity. It allows us to fully rep-
resent elastic deformation of any small element of a solid by the strain tensor.
The strains introduced by equations 2.5 are caused by forces applied to a solid. Those forces
are conventionally subdivided into the body forces, , which are distributed over a volume, and the
surface (or contact) forces, , that act along external and internal surfaces in the medium. Let us
examine the force, , applied to a small surface element, , whose orientation is character-
ized by the unit normal . This force gives rise to the traction vector, , defined as

(2.8)

We can now apply Newtons second law to an arbitrary volume in our solid:

(2.9)

where is the mass density, t is the time and the surface integral is taken over the external surface
of V. Equation 2.9 is used to define the stress tensor. To do this, we describe our volume V by its
characteristic linear size L so that and . Therefore, as L and V approach zero,
equation 2.9 reduces to

(2.10)

because the first two volume integrals approach zero faster than the surface integral when L 0.
If V is a small thin disc with a negligible edge area, equation 2.10 implies that

38

12815-OTE III Gretchka.indd 38 28-07-2009 10:13:49


Plane waves

(2.11)

or

(2.12)

which is simply the third Newtons law.

Next, suppose that V is a small tetrahedron


shown in Figure 2.1. Then, equation (2.10) yields

(2.13)

Figure 2.1: Small tetrahedron whose three


Using equation 2.12 and noting that the trian- faces lie in the coordinate planes.
gles OBC, AOC and AOB in Figure 2.1 are the
projections of ABC on the coordinate planes
and that ratios of the areas of triangles OBC, AOC, and AOB to the area of ABC are equal to the
components of vector , we obtain

(2.14)

Equation 2.14 allows us to introduce the second-order stress tensor . By definition,

(2.15)

Similarly to the strain tensor, the stress tensor is symmetric,

(2.16)

Its symmetry can be proven by equating the rate of change of angular momentum of volume V to
the moment of forces acting upon V (see Aki and Richards, 1980, for details). It then follows that
the components of traction applied to a surface element with normal are given by

(2.17)

While equations 2.12, 2.15, and 2.17 might seem trivial in statics, the presented derivation shows
that they are also valid during accelerations. This allows us to substitute equation 2.17 into Newtons
second law (equation 2.9), apply the Gauss divergence theorem to the surface integral there, and
arrive at the equation of motion:

(2.18)

39

12815-OTE III Gretchka.indd 39 28-07-2009 10:13:50


Chapter 2

2.2 Hookes law


To solve the equation of motion 2.18 in a unique fashion, we need to supplement it with a relation-
ship between two so-far independent quantities: the displacement vector and the stress tensor .
Their well-known relation for linear elasticity is given by Hookes law. In modern literature, it is
written as

(2.19)

where is the fourth-rank stiffness tensor. With Hookes law and definition 2.5 of the strain tensor,
the equation of motion 2.18 can be solved for the remaining unknown the displacement vector .
However, we postpone our discussion of the equation of motion to later sections in this chapter and
instead concentrate on the stiffness tensor.

2.2.1 Stiffness tensor


Equality 2.19 is one of the possible definitions of the stiffness tensor. Among other things, it allows
us to establish important symmetry properties of . Since the strain and stress tensors are symmetric
(equations 2.6 and 2.16), the stiffness tensor has to possess the symmetry with respect to interchang-
ing of its first two and last two indexes,

(2.20)

It also has another type of symmetry,

(2.21)

which follows from the existence of the scalar strain-energy function, , defined for linearly elas-
tic media as (e.g., Landau and Lifshitz, 1959; Aki and Richards, 1980; Ben-Menahem and Singh,
1981)

(2.22)

The requirement for energy associated with any nonzero deformation to be positive, >0, implies
the positive definiteness of . This is the only constraint imposed by physics on the components of the
stiffness tensor for generally anisotropic solids.
The symmetry relations 2.20 and 2.21 reduce the number of independent elements of the stiff-
ness tensor from 3333=81 to 21 and make it possible to replace a pair of indexes ij with a
single index I according to the so-called Voigt recipe

(2.23)

where ij is the Kronecker delta; ij = 1 when i = j and ij = 0 otherwise. The explicit form of rela-
tions 2.23 is

40

12815-OTE III Gretchka.indd 40 28-07-2009 10:13:50


Plane waves

(2.24)

In Voigt notation, the stress and strain tensors become six-dimensional vectors with the components

(2.25)

where factors 2 in front of the off-diagonal elements of the strain tensor appear because those ele-
ments are doubled in the tensorial formulation of Hookes law given by equation 2.19 (see Helbig,
1994, for a discussion of other options for taking this doubling into account). Voigt notation proves
especially useful for the stiffness tensor . This fourth-rank tensor is now represented by a 66
symmetric matrix

(2.26)

We will use this representation throughout the book.

Combining equations 2.25 and 2.26, we can write Hookes law in a matrix (as opposed to tenso-
rial) form

(2.27)

Summation from 1 to 6 with respect to repeated upper-case Roman indexes is assumed in this equa-
tion and henceforth.

41

12815-OTE III Gretchka.indd 41 28-07-2009 10:13:50


Chapter 2

Figure 2.2: Normal (red) and shear (black)


components of the stress tensor.

Expanding equations 2.27 and presenting them as

(2.28)

reveals the block structure of the stiffness matrix and makes clear which stiffness components enter
the relations between various components of the strain and stress tensors. For instance, the normal
stress components , , and (or , no summation with respect to i) shown in Figure 2.2 relate
to the normal strain components , , and via the upper-left block of the stiffness matrix, ,
. Those stiffnesses are responsible for the changes in shape and volume of a solid
under a normal load. They generally do not vanish in either isotropic or anisotropic media(3) (Figure
2.3). Likewise, the diagonal stiffnesses , , and in the lower-right block, which relate shear
strains and stresses, can be equal to zero only in fluids and not in solids.
The meaning of s occupying the upper-right and lower-left blocks of the stiffness matrix
in equation 2.28 is different. They relate the normal stresses to shear strains and vice versa.
Their nonzero values imply, for example, that a normal stress causes a shear strain, as is sche-
matically shown in Figure 2.3b. As we know from our everyday experience, this does not hap-
pen in isotropic media, therefore, those stiffness coefficients have to be equal to zero for isot-
ropy. Finally, the off-diagonal elements , , and in the lower-right block relate the shear
stress components to different shear strain components .
A nonzero , for instance, implies that a horizontal shear stress (Figure 2.2)
results in a shear deformation in the vertical -plane. Clearly, such shear deformations can
occur solely due to anisotropy.

3 Equalities c12 = c13 = c23 = 0 are satisfied for isotropy only when the Poissons ratio, , is zero. Although
artificial materials with have been manufactured, the author is unaware of the existence of natural
solids that have nonpositive Poisson's ratios.

42

12815-OTE III Gretchka.indd 42 28-07-2009 10:13:51


Plane waves

Figure 2.3: Deformation of (a) isotropic and (b) anisotropic solids under a compressive load.

2.3 Symmetry classes


The stiffness tensor, , is introduced in its full generality by matrix 2.26. This matrix does not
explicitly indicate how many stiffness elements are nonzero and how many of them are independent
of each other. This information, however, is of obvious relevance in many applications because fewer
nonzero entries of the stiffness tensor tend to simplify mathematical manipulations with it. Yet more
importantly, if the stiffness coefficients are unknown and have to be estimated from some meas-
urements, which is commonly the case in geophysics, their larger number usually entails a higher
uncertainty in some of the obtained stiffness components. Therefore, establishing the structure of
the stiffness tensor, if such a structure exists, and reducing the number of nonzero s is beneficial
for solving many practical problems.
The structure of the stiffness tensor is derived from the notion of symmetry. To understand what
this means, we need to realize that numeric values of the stiffness components, , as well as those
of any other tensor, depend on a coordinate frame chosen to describe the tensor. Thus, it makes
sense to introduce two Cartesian coordinate frames that have
the same origin and relate to each other via the rotation matrix

(2.29)

Let us suppose that the stiffness components are given in the coordinate frame . The coordi-
nate rotation specified by matrix changes them to (e.g., Auld, 1973; Helbig, 1994)

(2.30)

An analogous transform for the stiffness elements in Voigt notation is called the Bond trans-
formation. The explicit form of the Bond matrix can be found in Auld (second edition, 1990, p. 74)
and Carcione (2001, p. 8).
It is apparent from equation 2.30 that numerical values of components of tensor generally
differ from those of and depend on the directions of coordinate axes with respect to . For
example, a component = 0 in the coordinate frame might become nonzero in a different
frame . If possesses certain types of structures, it might happen that some rotation matrices
do not alter its elements. Then, tensor is said to be symmetric with respect to those transforma-
tions or to belong to a certain symmetry class.

43

12815-OTE III Gretchka.indd 43 28-07-2009 10:13:51


Chapter 2

The symmetry classes used in seismic exploration were originally developed in crystal acoustics
based on the appearance of crystals and on analysis of their atomic structures (Musgrave, 1970; Auld
1973). Those structures can have two types of symmetries that transform a structure into itself:

(1) a reflection or a mirror symmetry with respect to a plane, such as the symmetry of an ellipsoid
with respect to the coordinate planes , and
(2) a rotation symmetry around an axis by a certain angle, such as the symmetry of the same
ellipsoid with respect to rotation around each coordinate
axis by 180.

Clearly, many geometric shapes (a scalene triangle for example) do not have planes of mirror
symmetry and cannot be transformed into themselves by any rotation except for a trivial one by
360. According to our definition above, they possess no symmetry. There are asymmetric crystals
and rocks too. They are called triclinic. We can say, therefore, that the discussion in the previous
section pertains to triclinic media. Still, for any given triclinic stiffness tensor , there exists such a
coordinate frame , where (Helbig, 1994, p. 116)

(2.31)

As a result, only 18 out of 21 elements of the most general stiffness matrix 2.26 are independent.
Three relationships between 21 stiffness coefficients stem from the arbitrariness of the spatial ori-
entations of the coordinate axes.
While the task of obtaining the stiffnesses of triclinic media from seismic data might seem formi-
dable, sometimes it is feasible. For example, Dewangan and Grechka (2003) present a case study of
estimation of the full triclinic stiffness tensor from multicomponent, multi-azimuth, walkaway VSP
data acquired beneath a laterally homogeneous overburden in the Vacuum Field, New Mexico, USA.
In the rest of the section, we devote our attention to media that have higher symmetries than
triclinic. Here, we limit our discussion to the elastic symmetries important for seismic exploration
and reservoir characterization and refer the reader to the crystal acoustic texts by Fedorov (1968),
Musgrave (1970) and Auld (1973) for a more comprehensive analysis.

2.3.1 Monoclinic media


The lowest symmetry that elastic media can have is termed monoclinic. Such media possess a single
plane of mirror symmetry. To find the structure of a monoclinic stiffness tensor, , let us select a
coordinate system where this symmetry plane is . Mirror symmetry with respect to this plane
implies that rotation 2.30 with the matrix

(2.32)

does not change tensor , that is,

(2.33)

Equations 2.33 do not constrain the stiffness components whose indexes contain any combina-
tion of 1s and 2s because Therefore, the non-trivial equations reduce to

44

12815-OTE III Gretchka.indd 44 28-07-2009 10:13:52


Plane waves

(2.34)

where is the number of occurrences of index 3 in the sequence . Equations 2.34 make it clear
that all , where the index 3 appears one or three times should vanish. Also, in accordance with
equations 2.31, there exists a rotation in the -plane that eliminates the coefficient (Helbig,
1994; Grechka et al., 2000). Consequently, stiffness tensors in monoclinic media with the symmetry
plane have the structure

(2.35)

and monoclinic media are fully characterized by 12 independent stiffnesses and three orientational
parameters. The latter define the direction of normal to the symmetry plane, which is chosen to be
for matrix 2.35, and the rotation within this plane.
Let us also mention another property of media that possess a symmetry plane. As the rotation
matrix appears in the coordinate transform 2.30 four times, the sign of cancels out. This
means that monoclinic media are invariant with respect to rotation

(2.36)

Unlike , the matrix describes a rotation around the vertical axis by 180. Solids invari-
ant with respect to rotation 2.36 are said to have a symmetry axis of the second order. Thus, we
conclude that all elastic media that have a symmetry plane also have a symmetry axis of order 2
and vice versa.
Although monoclinic media are of limited use in the oil industry, rocks containing multiple frac-
ture sets (such as those schematically shown in Figure 2.4) can be effectively monoclinic. Another
plausible reason for monoclinic symmetry is sequences of thin, horizontal anisotropic layers that have
symmetries higher than monoclinic but do not possess a horizontal symmetry plane. Their upscal-
ing is done with the so-called Schoenberg-Muir (1989) calculus that can be viewed as an extension
of Backus (1962) averaging to intrinsically anisotropic fine layers. Figure 2.5 displays sandstones
that exhibit an overall monoclinic symmetry if the formation does not change in the direction
normal to the outcrop. Indeed, each nearly horizontal layer consists of smaller-scale dipping lay-
ers with oblique cracks (shown with arrows) oriented approximately normally to the fine layering.
The tilted orthorhombic symmetry of those fractured, fine dipping layers (see below) yields an
overall monoclinic medium with a symmetry plane parallel to the outcrop.
As we have already mentioned, monoclinic media play a relatively unimportant role in seismic
exploration and exploitation. The main problem with them is similar to that for triclinic media
too many independent stiffness coefficients need to be estimated from seismic measurements.
Nevertheless, Winterstein and Meadows (1991) discuss a VSP data set that suggests monoclinic
symmetry. The model they use to fit their data vertically fractured dipping fine layers is similar
to the outcrop in Figure 2.5.

45

12815-OTE III Gretchka.indd 45 28-07-2009 10:13:52


Chapter 2

Figure 2.4: Two systems of vertical cracks


embedded in azimuthally isotropic host rock
at oblique directions with respect to each
other generally result in monoclinic symme-
try with a horizontal symmetry plane (after Figure 2.5: Fractured sandstones near Paria
Bakulin et al., 2000). Canyon in Arizona, USA.

2.3.2 Orthorhombic media


By definition, orthorhombic (or orthotropic) media have three mutually orthogonal planes of mirror
symmetry. It is natural, therefore, to choose them as the coordinate planes .
Establishing the structure of the orthorhombic stiffness tensor, , then becomes straightforward.
It suffices to note that has to be invariant to reflections, such as that given by equation 2.32,
with respect to all three coordinate planes. Those reflections are described by the matrices

(2.37)

Repeating the logic of section 2.3.1, we conclude that components where any index 1, 2, or 3 appears
an odd number of times in the sequence should vanish. Thus, the tensor has the form

(2.38)

and contains nine independent stiffness elements.

Two microstructures important in practice possess orthorhombic symmetry. In the previous section
we mentioned perhaps the most popular of them. It consists of thin fractured layers with cracks orthogo-

46

12815-OTE III Gretchka.indd 46 28-07-2009 10:13:52


Plane waves

nal to the layering (Schoenberg and Helbig, 1997). This microstructure is shown in Figure 2.6
Clearly, its three symmetry planes are oriented along the cracks, across them and along the layers.
Also, the effective orthotropy can be caused by orthogonal fractures or joints like those displayed
in Figure 2.7. The orientations of planes of the mirror symmetry in rocks with orthogonal fractures
are visually apparent.
Although the number of independent stiffness components nine is still too large for
orthorhombic models to be routinely used in seismic data processing, their applicability is facilitated
by two factors. First, the P-wave kinematic signatures in orthorhombic media are governed by six
combinations of the nine stiffness coefficients (Tsvankin, 1997b). If we are interested in P-wave time
(rather than depth) processing, the number of independent parameters is further reduced to five.
Importantly, all five parameters are the functions of azimuthally-varying P-wave NMO velocities that
can be estimated from wide-azimuth seismic
reflection data (Grechka and Tsvankin, 1999a).
The second useful application of orthorhom-
bic media arises when it is possible to assume
that the anisotropy is crack-induced. Then, the
effective symmetry is close to orthotropy for an
arbitrary distribution of fracture orientations.
Kachanov (1992; 1993) proved this counterin-
tuitive statement theoretically and Grechka and
Kachanov (2006a) supported it with numeri-
cal experiments. The key simplification here
again comes from the reduction in the number
of pertinent parameters. For vertical fractures,
this number is generally five. When the frac-
ture infill is gas, it reduces to four, making Figure 2.6: Orthorhombic media have three
the parameter-estimation problem tractable. mutually orthogonal planes of mirror sym-
Its solution was presented by Vasconcelos and metry. One of the orthotropic models contains
Grechka (2007, also Chapter 8) who relied on vertically fractured fine horizontal layers (after
the notion of crack-induced orthotropy and built Rger, 1998).
an orthorhombic model of the tight-gas Rulison
reservoir from multicomponent, wide-azimuth
surface reflection data.

2.3.3 Transverse isotropy


Transverse isotropy (TI) is arguably the most
commonly used symmetry in seismic data
processing and velocity-model building. There
are several reasons for this:
1. Transverse isotropy is geologically plausible.
Its well understood natural causes are shales
and fine layering.
2. TI media are characterized by relative-
ly few independent stiffness coefficients. Figure 2.7: Orthogonal fracture sets in sand-
Only three combinations of those stiffnesses stones of Cedar Mesa, Utah, USA (source:
are needed to describe P-wave traveltimes Jon Olsons home page, http://www.pe.utexas.
(Tsvankin, 1996). edu/~jolson/Welcome.html).

47

12815-OTE III Gretchka.indd 47 28-07-2009 10:13:53


Chapter 2

3. The stiffness combinations responsible for the key seismic signatures are captured by Thomsen
(1986) parameterization. We will discuss Thomsen parameters in Chapter 4.
4. TI symmetry is sufficient for processing narrow-azimuth P-wave data acquired in structurally
simple lower-symmetry media, for instance, orthorhombic or monoclinic.

By definition, transversely isotropic solids have one axis of symmetry of order . In other words,
they are invariant with respect to any rotation around this axis. Their properties, therefore, are con-
venient to describe in a coordinate frame whose one axis, for instance, the vertical axis coincides
with the axis of rotational symmetry. Media possessing rotational symmetry around the vertical are
called vertically transversely isotropic (VTI). Figure 1.12 shows an example of a VTI medium.
To establish the structure of the VTI stiffness tensor, , it makes sense to start with the
orthorhombic tensor 2.38. Let us note that VTI media, like orthorhombic, possess three orthogonal
planes of mirror symmetry: a horizontal plane normal to the rotational symmetry axis and two
arbitrarily oriented but orthogonal to each other vertical planes. Therefore, the zero elements of
should remain zero in . The fact of arbitrariness of orientations of the symmetry planes
implies that the components do not change if the index 1 is replaced with 2
and vice versa. The interchangeability of indexes entails three equalities,

(2.39)

among the components of the orthorhombic stiffness matrix 2.38. The fourth equality is derived
from the invariance of the VTI stiffness element with respect to rotation by an arbitrary angle
around the vertical -axis. Such a rotation is given by the matrix

(2.40)

Substituting in place of in equation 2.30 and requiring that for any results
in the sought equality

(2.41)

and allows us to write the VTI stiffness tensor as

(2.42)

According equations 2.41 and 2.42, VTI and more generally TI media are characterized by five
independent stiffness coefficients. To these, we need to add two parameters determining the orien-
tation of the symmetry axis.
It is useful to define several terms pertaining to TI media. Since all directions in a plane normal
to the TI symmetry axis are equivalent, this plane is called the isotropy plane. Any plane that contains

48

12815-OTE III Gretchka.indd 48 28-07-2009 10:13:53


Plane waves

Figure 2.8: Sketch of an HTI model. Shear waves polarized parallel and normal to the isotro-
py plane (cracks) propagate with different vertical velocities. The velocity of the -wave is greater than
that of (after Rger, 1997).

the symmetry axis is termed the symmetry-axis plane. If the symmetry-axis direction of a TI medium
deviates from the vertical, such a medium is called tilted transversely isotropic (TTI). The main practi-
cal application of tilted transverse isotropy is the description of anisotropy caused by dipping beds,
for instance, shales.
TI media with a horizontal symmetry axis have a special name. They are called horizontally trans-
versely isotropic (HTI). Their significance is in providing the simplest model for anisotropy caused
by vertical cracks (such as those displayed in photo 1.2). Those cracks are schematically shown with
vertical planes in Figure 2.8. If the properties of the cracks are independent of directions within
their vertical planes and the host rock is isotropic, the overall symmetry is TI with the rotational
axis pointing along the -axis in Figure 2.8. The stiffness tensor, , of such an HTI medium is
obtained from (equation 2.42) by interchanging the indexes 1 and 3:

(2.43)

The same swap of indexes, , applied to relationship 2.41 yields

(2.44)

Applications of transverse isotropy in seismic data processing are too numerous to describe here.
They are discussed in later chapters.

2.3.4 Isotropy
Finally, we briefly touch on isotropy. By definition, all directions in isotropic media are equivalent.
This means that rotation 2.30 with any matrix transforms an isotropic solid into itself. For the

49

12815-OTE III Gretchka.indd 49 28-07-2009 10:13:54


Chapter 2

purpose of deriving the isotropic stiffness tensor, , it suffices to note that isotropic media are
simultaneously VTI and HTI. Based on equations 2.41 2.44, this is only possible when

(2.45)

where and are the Lam constants. Thus, the structure of the isotropic stiffness tensor is

(2.46)

2.4 Christoffel equation


We are now well equipped to return to the equation of motion 2.18. Substituting Hookes law 2.19
and definition 2.5 of the strain tensor, we arrive at the equation

(2.47)

that contains the only unknown the displacement vector . We simplify this equation by assum-
ing the medium to be spatially homogeneous, that is, = const and = const, and ignoring
the influence of body forces (for instance, gravitational or electric forces) on elastic processes.
Equation 2.47 then becomes

(2.48)

Its standard solution is a harmonic plane wave

(2.49)

where is the polarization vector, is the imaginary unity, is the angular frequency, is
the unit wavefront normal, and V is the phase velocity.
Substitution of the plane-wave solution 2.49 into the equation of motion 2.48 leads to the
Christoffel equation

(2.50)

where

(2.51)

50

12815-OTE III Gretchka.indd 50 28-07-2009 10:13:54


Plane waves

is the Christoffel matrix and

(2.52)

is the density-normalized stiffness tensor. Introducing the slowness vector,

(2.53)

we can write the Christoffel equation in an equivalent form,

(2.54)

where

(2.55)

The Christoffel equation 2.50 is arguably the most important equation for analysis of wave phe-
nomena in anisotropic media. In fact, it expresses a classic eigenvalue-eigenvector problem for the
symmetric, positive definite matrix . The positive definiteness of is a direct consequence of
the positive definiteness of tensors , while its symmetry, , follows from definition
2.51 and the symmetries of the stiffness tensor given by equations 2.20 and 2.21. The reducibility of
the Christoffel equation to a textbook problem of linear algebra allows us to state the properties of
its solution.
Three eigenvalues of are the squared phase velocities, , of three body waves
. All three values are strictly positive because the Christoffel
matrix is positive definite.
The naming convention for waves , where P stands for primary and S for sec-
ondary, reflects the fact that velocities and are close to each other in the vast majority
of natural materials, whereas is usually noticeably greater than both and .
The phase velocities generally depend on the direction of wave propagation, , and, hence, exhib-
it anisotropy. For symmetries lower than TI, the wave type is defined according to inequalities

(2.56)

for each wavefront normal .


A different definition is used in VTI media, where one shear-wave is always polarized in the hori-
zontal isotropy plane. Naturally, this wave is called the horizontally polarized shear and denoted SH.
The polarization vector of another shear-wave is confined to the vertical symmetry-axis plane.
Therefore, this wave is named the vertically polarized shear (SV). The phase velocities of the two
shear-waves coincide along the symmetry axis, . Whether or
at oblique directions of the wavefront normal depends on the VTI stiffness
coefficients and the angle between and the vertical.
While the first equality in 2.56, , is theoretically possible, it is an oddity. The
second equality, however, is not. In fact, the wavefront normals, , satisfying the equality

(2.57)

51

12815-OTE III Gretchka.indd 51 28-07-2009 10:13:55


Chapter 2

Figure 2.9: Typical geometry of phase-velocity


surfaces in orthorhombic media. Red arrow
points to the shear-wave singularity (after
Grechka et al., 1999a).

have been found in all anisotropic solids. These directions are called the shear-wave singulari-
ties (e.g., Crampin and Yedlin, 1981; Musgrave, 1985). Commonly (but not always), two shear-
wave phase-velocity surfaces touch each other at isolated points and form the so-called point or
conical singularities. Such a singularity is shown with a red arrow in Figure 2.9.
Except for isotropy, where equation 2.57 is an identity, and for singular directions dif-
fers from in anisotropic media. Inequality is the mathematical reason for
shear-wave splitting.
The three eigenvectors of the Christoffel matrix are mutually orthogonal for any ,

(2.58)

Because of anisotropy, however, the polarization vectors are generally neither parallel nor
orthogonal to the wavefront normal,

(2.59)

Nevertheless, there always exist the so-called longitudinal directions, , where


and, correspondingly, and (e.g., Helbig, 1994).

Varying the wavefront normal in equation 2.50, finding its eigenvalues , and plotting the
vectors from a common origin makes the phase-velocity surfaces. The reciprocal surfaces,
, defined by equation 2.53 are called the slowness
surfaces. Figure 2.10 gives an example of vertical cross-sections of the slowness and phase-velocity
surfaces of SV-wave in a VTI solid. The slowness in Figure 2.10 has both convex and concave parts.
The implications of concavity of the slowness surfaces for the wavefront shapes will be discussed in
the next section. At this point, we only mention the well-known fact that the P-wave slowness sheet,
, is always convex (e.g., Fedorov, 1968; Auld, 1973), whereas the slowness surfaces of shear-waves
can have convex and concave regions.

52

12815-OTE III Gretchka.indd 52 28-07-2009 10:13:55


Plane waves

Figure 2.10: Slowness (dashed blue) and


phase velocity (solid red) of the SV wave in a
VTI medium with the density-normalized stiff-
ness coefficients ,
, and
. The vertical and horizontal lines
indicate the symmetry axis and the isotropy
plane, respectively.

2.4.1 Group velocity


The phase velocity, slowness and polarization of a plane wave discussed in the previous section
were found to be functions of the wavefront normal . Here, we concentrate on the wavefronts
themselves. Specifically, we examine their velocities, which are known as the group or ray veloci-
ties. In general, the group-velocity vector, , characterizing propagation of energy associated with
the wave motion, is defined as the partial derivative of the angular frequency with respect to the
wavenumber vector, ,

(2.60)

or, in components,

(2.61)

The wavenumber vector is parallel to the wavefront normal ,

(2.62)

and its magnitude is

(2.63)

The easiest way to derive the group velocity is to write Christoffel equation 2.50 in the form

, (2.64)

take a dot product of this equation with polarization vector ,

(2.65)

and substitute definitions 2.62, 2.63 into equation 2.65. The result is

53

12815-OTE III Gretchka.indd 53 28-07-2009 10:13:55


Chapter 2

(2.66)

Differentiating equation 2.66 with respect to and applying definitions 2.61 and 2.53 then yields
the expression

(2.67)

This is the most general equation for the group velocity in anisotropic media.

Equation 2.67, however, is not necessarily the most convenient formula for calculation of the
wave-propagation times because it requires computing the polarization vector . Therefore, it
makes sense to explore the options for eliminating from the group-velocity expression. Equation
2.65 suggests how to proceed. We note that it can be rewritten as

(2.68)

Comparing equations 2.67 and 2.68 leads to an important relationship between the group-velocity
and slowness vectors,

(2.69)

Its equivalent for the phase velocity is

(2.70)

Both equations 2.69 and 2.70 serve as the departure points for expressing the group velocity in
terms of the slowness and the phase velocity. We begin with the latter.

Group velocity in terms of phase velocity


To obtain the group velocity from equation 2.70, we note that spatial direction of the unit wavefront
normal can be specified by two directional angles and . These angles are usually taken as
the polar angle of and the azimuth of with respect to the coordinate axis , so the wavefront
normal is given by

(2.71)

As both velocities and V can be considered as functions of and and ,


we can differentiate equation 2.70 with respect to the angles . The result is

(2.72)

The second term on the right-hand side of this equation vanishes because, by definition, is the
normal to the wavefront, whereas the derivatives are tangent to it. Thus,

(2.73)

54

12815-OTE III Gretchka.indd 54 28-07-2009 10:13:56


Plane waves

We now treat equations 2.70 and 2.73 as a linear system for the components of the group-velocity
vector. Its solution is the sought expression

(2.74)

Mutual orthogonality of the vectors , and (except for the polar angle , where
is undefined) makes it possible to relate the polar components of the group-velocity vector
only to the derivative and the azimuthal component of only to . The correspond-
ing expressions are given in Tsvankin (2001, his equations 1.14 1.16).

Equation 2.74 leads to the following observations and definitions.

The magnitude of the group-velocity vector is

(2.75)

Therefore, the phase and group velocities satisfy inequality

(2.76)

for any wavefront normal .

The unit vector, , parallel to the group velocity is called the ray. By definition,

(2.77)

Plotting the group velocities, , from a common origin along the ray directions produces the
group-velocity surface or the wavefront originating from a point source in a homogeneous aniso-
tropic medium.
As follows from equation 2.70,

(2.78)

If , the ray deviates from the wavefront normal towards the phase-velocity increase in
accordance with equation 2.74. Figure 2.11 illustrates this geometrically. In 2D, (black) is the
hypotenuse of the right triangle with sides V (red) and (green). Figure 2.11 also dem-
onstrates that and the phase velocity plotted in the ray direction on the same graph
do not necessarily obey inequality 2.76. Whether or not and satisfy the inequality for
depends on the magnitudes of the phase-velocity derivatives .
According to equation 2.75, the equality = V is reached only when = = 0,
that is, at extrema of . At those extrema, (equation 2.78) and the group- and phase-
velocity surfaces touch each other. Thick gray arrows show such locations in Figure 2.11.

55

12815-OTE III Gretchka.indd 55 28-07-2009 10:13:56


Chapter 2

Figure 2.11: Slowness (dashed blue) and


phase velocity (solid red) shown in Figure 2.10,
and group velocity (solid black). The red,
green and black arrows indicate the vectors
, and , respectively.
The gray arrows point to the phase-velocity
extrema.

Finally, we mention that equation 2.74 breaks down for shear-waves propagating in the direc-
tions of point singularities (equation 2.57). Mathematically, this happens because the shear-
wave phase velocities are nondifferentiable at . Then, instead of equation 2.74, one has to use
a more general formula 2.67. The polarization vectors and needed to calculate
and with equation 2.67, however, are defined nonuniquely. Since the Christoffel equation
has a double root at (see equation 2.57), and can point to any directions orthogonal
to each other in the plane normal to . As a result, the geometry of the shear-wave surfaces
becomes extremely complex in the vicinity of . We, therefore, choose to skip this subject and
reference Fedorov (1968) for a general discussion and Grechka and Obolentseva (1993) for rel-
evant examples.

Group velocity in terms of slowness


We now return to equation 2.69 and use it to obtain a group-velocity formula similar to equation
2.74 but in terms of the slowness vector. To begin, we note that the necessary condition for nonzero
eigenvectors of Christoffel equation 2.54 is

(2.79)

or

(2.80)

Equation 2.80 relates three components of the slowness vector, , and thus makes it possible to
treat, for instance, its vertical component, , as a function of two other components, .
Likewise, the group-velocity vector, , can be considered a function of the same arguments,
.
Essentially, we set the stage for repeating the logic of the previous subsection. Specifically, we
differentiate equation 2.69 with respect to and ,

(2.81)

56

12815-OTE III Gretchka.indd 56 28-07-2009 10:13:57


Plane waves

and notice that the second dot product in this equation is zero because vector , being parallel to
the wavefront normal , is orthogonal to the group-velocity surface, whereas the derivatives
are tangent to it. Hence, it follows from equation 2.81 that

(2.82)

The final step is to solve the system of three linear equations 2.69, 2.82 for the components of the
group-velocity vector. The result is (Grechka et al., 1999b)

(2.83)

Equivalent (except for directions of the point singularities) group-velocity expressions 2.67,
2.74, and 2.83 find their applications in a variety of computations. For example, equation 2.67 will
be used in the next chapter to trace rays in heterogeneous anisotropic media, while formula 2.74
was already helpful in revealing the relations between the wavefront normals and rays and between
the phase- and group-velocity surfaces. Equation 2.83 is useful for building anisotropic velocity
models from seismic reflection data because the horizontal slownesses and can be directly
measured at the Earths surface. Some other expressions for can also be derived if the components
of or are treated as independent quantities (e.g., Helbig, 1994).
To gain more insight into the geometries of the slowness and the group-velocity surfaces, let us
return to equation 2.82 and observe that it implies the orthogonality of the group-velocity vector to
the slowness surface. Thus, we arrive at an important duality:
The slowness vector is normal to the group-velocity surface.
The group-velocity vector is normal to the slowness surface.

Figure 2.12: The same slowness (dashed blue),


phase-velocity (solid red), and group-velocity
(solid black) surfaces as in Figure 2.11. The
dark and light blue arrows are normals to
the slowness surface. Their lengths are equal
to the group velocities . The corresponding
wavefront normals are shown with red and
brown arrows.

57

12815-OTE III Gretchka.indd 57 28-07-2009 10:13:57


Chapter 2

Figure 2.13: Shear-wave slowness (a, b) and group-velocity (c, d) surfaces in an orthorhombic solid
with density-normalized stiffness coefficients (in Voigt notation)
and .
The units of axes are s/km in (a, b) and km/s in (c, d).

This duality makes it possible to understand the reason for triplications on the SV wavefront
shown with a black curve in Figure 2.11. To explain the origin of the triplications, we plot normals
to the slowness surface (light and dark blue arrows in Figure 2.12). When the slowness surface is
convex, the normals spread away from each other (light blue) and, as they are parallel to the group-
velocity vectors , the wavefront remains single-valued. In contrast, the normals to the concave por-
tions of the slowness surface (dark blue) intersect and cause the group-velocity surface to fold onto
itself and become multivalued.

58

12815-OTE III Gretchka.indd 58 28-07-2009 10:13:57


Plane waves

Much more complicated group-velocity surfaces of shear waves can exist for symmetries lower
than TI. Figure 2.13 displays an example of such a complexity. The parameters of the orthorhom-
bic medium in this example, specifically, the low values of the diagonal stiffness elements
, were inspired by a model of the so-called acoustic anisotropy(4). This model was
proposed by Alkhalifah (1998) to simplify wave-equation-based processing of P-waves in VTI media.
Shear waves in acoustically anisotropic media, however, exhibit an extreme degree of anisotropy
(Grechka et al., 2004). This strong anisotropy gets inherited by elastic media whose stiffness coef-
ficients are in some sense close to the acoustic ones. This closeness creates extensive concave regions
on the S-wave slowness surfaces (Figures 2.13a and 2.13b). They, along with contributions from
the shear-wave singularities, lead to the complex, multivalued group-velocity surfaces displayed in
Figures 2.13c and 2.13d.

4 The term acoustic refers to spatial directions where at least one shear-wave phase velocity vanishes.

59

12815-OTE III Gretchka.indd 59 28-07-2009 10:13:58


12815-OTE III Gretchka.indd 60 28-07-2009 10:13:58
RAYS
3 Rays AND
and TRAVELTIMES
traveltimes

Properties of plane waves discussed in Chapter 2 provide the theoretical basis for ray tracing. Here,
we concentrate on kinematic aspects of wave propagation in anisotropic media. This choice is moti-
vated by the fact that the industry currently uses mainly traveltime information to infer anisotropy
of the subsurface and to image seismic data. As we do not cover ray amplitudes, which are undoubt-
edly important for anisotropic AVO applications and true-amplitude imaging, we refer the reader
to a comprehensive ray-tracing volume by Cerven (2001).
We begin this chapter with derivation of the kinematic ray-tracing equations. We then present
an overview of various techniques developed to trace rays and compute traveltimes in anisotropic
media and emphasize the differences of those techniques from more familiar isotropic ray-tracing
methods. We conclude the chapter with an example that illustrates a possible complexity of the ray
trajectories and traveltimes of S-waves in relatively simple, horizontally layered anisotropic struc-
tures.

3.1 High-frequency asymptotics of seismic wavefields


A natural starting point for ray tracing is the equation of motion 2.47, which we rewrite without the
body-force term:

(3.1)

In this equation, is the density, is the displacement vector, t is the time, is the vector of spatial
coordinates and is the stiffness tensor. Unlike in Chapter 2, where homogeneity of both and
was assumed, here they vary arbitrarily in space, that is, .
Supplementing equation 3.1 with the appropriate initial and boundary conditions (e.g., Aki and
Richards, 1980; Ben-Menahem and Singh, 1981) and solving it as such amounts to full wavefield
modeling, which is generally accomplished using the finite-difference or finite-element methods. The
ray theory does not proceed this way. Instead, it introduces the following two key simplifications:

1. The displacement field is sought as a Fourier integral

(3.2)

Here A is the wave-amplitude vector, F is the spectrum of the examined wavefield, i= is


the imaginary unity, is the angular frequency, and is the phase function. Equation =0
describes the wavefront.

61

12815-OTE III Gretchka.indd 61 28-07-2009 10:13:58


Chapter 3

2. The wavefield spectrum F is assumed to contain only high frequencies, that is,

(3.3)

where the frequency is large.

Clearly, equations 3.2 and 3.3 are designed to extract and describe the high-frequency part
of a wavefield. Following the standard approach (see Cerven, 2001, for pertinent references and
details), we substitute the trial solution 3.2 into equation 3.1 and collect the terms containing differ-
ent powers of . This results in three groups of terms proportional to , , and (they are explic-
itly written in Petrashen, 1980). As the frequencies of the analyzed wavefield are large in accordance
with the second assumption above, the terms proportional to dominate. They are given by

, (3.4)

where U is the unit polarization vector. No scale or magnitude is associated with U because equation
3.1 is sourceless.
Next, it is customary to introduce several quantities related to the wavefront =0. As is
the surface moving in space, we can define its normal

(3.5)

the slowness vector

(3.6)

and the phase velocity

(3.7)

Expressing the derivative and the components of gradient from equations 3.5 3.7
and substituting them in equation 3.4 yields the Christoffel equation in form 2.64,

(3.8)

and its equivalent

(3.9)

where is the local density-normalized stiffness tensor.

62

12815-OTE III Gretchka.indd 62 28-07-2009 10:13:58


Rays and traveltimes

There is little surprise that we managed to recover the Christoffel equation derived for plane
waves in homogeneous anisotropic media as a local approximation of the high-frequency waves char-
acterized by curved wavefronts and propagating in heterogeneous solids. The physical reason for the
local validity of the Christoffel equation is the assumed small wavelength of the examined waves.
Such short waves, being insensitive to the medium heterogeneities at distances exceeding a few
wavelengths from their wavefronts, see the medium only locally, where it looks homogeneous.
Also, the wavefront curvature does not appear in equations 3.8 and 3.9. Mathematically, this hap-
pens because equations 3.4 contain only the first-order derivatives of whereas the curvature is
expressed through the second-order derivatives . Physically, the absence of curvature
indicates that a plane is a reasonable local representation of a curved wavefront for the purpose of
traveltime calculation.

3.2 Ray-tracing equations and techniques


The existing approaches for computing traveltimes can be broadly divided into three groups:
wavefront-construction methods,
shooting methods, and
bending methods.

Although each group of methods solves a different equation or system of equations, the resulting
traveltime is known to be exactly the same for any ray geometry in any model. This useful property
allows one to choose the ray-tracing method best suited for a particular application. In this section
we discuss those applications, talk about physical principles underlying different ray-tracing meth-
ods and describe mathematical techniques used to implement the computations.

3.2.1 Wavefront construction


Wavefront construction is perhaps the most computationally intensive ray-tracing technique. To cal-
culate traveltimes with it, we take the dot product of equation 3.9 and the unit polarization vector
U and obtain the so-called Hamiltonian (e.g., Cerven, 2001)

(3.10)

Because (equation 3.6), we arrived at a nonlinear, first-order partial differential equation


for the traveltime t. This equation describes the temporary and spatial evolution of the wavefront
=0 initially located at =0 at time t=t0.
Equation 3.10 is the anisotropic extension of a more familiar isotropic eikonal equation. This
can be directly observed by replacing an arbitrarily anisotropic stiffness tensor in equation 3.10
with , where is the isotropic stiffness tensor given by matrix 2.46. Equation 3.10
then reduces to

(3.11)

where is the isotropic velocity. It is either or


depending on whether the P-wave polarization vector, , or the S-wave polarization vector,
, is used in equation 3.10.
The wavefront evolution expressed by equations 3.10 and 3.11 is a modern mathematical rep-
resentation of the classical Huygens principle. It suggests to treat all points on wavefront =0
as elementary seismic sources (dots in Figure 3.1) that simultaneously excite elementary wavefronts

63

12815-OTE III Gretchka.indd 63 28-07-2009 10:13:58


Chapter 3

Figure 3.1: Seismic wavefront at


two consecutive time moments t
and . The dots indicate ele-
mentary sources, the red arcs are
the elementary wavefronts.

(red arcs) at time t. The position of the wavefront at time , where is a small time interval,
is simply an envelope (blue in Figure 3.1) of the elementary wavefronts.
To construct moving wavefronts based on equations 3.10 and 3.11, various finite-difference
and finite-element techniques can be used. Recently, a particular finite-difference implementa-
tion, called the fast marching method (Sethian, 1999; Sethian and Vladimirsky, 2003), has gained
popularity. It systematically advances the wavefront while maintaining causality and minimizing the
traveltime between any two spatial locations. As a result, the fast marching method cannot produce
multivalued wavefronts such as those shown in Figures 2.13c and 2.13d. Instead, it yields traveltimes
of the fastest arrivals only, sharing this sometimes unfortunate property with numerous other ray-
tracing techniques.
The wavefront-construction methods possess an obvious advantage for computations in which
traveltimes corresponding to many consecutive wavefront positions are of interest. Therefore, their
main application area is generating traveltime tables for Kirchhoff migration. Indeed, to migrate
seismic data, one needs to know the wave-propagation times from seismic sources and receivers to
all potential reflectors that can be anywhere in the subsurface. The wavefront-construction methods
are well-suited for such a task because they allow computing the times without a priori knowledge
of the reflector locations.

3.2.2 Shooting method


Quite a different set of requirements arises when traveltimes at intermediate positions of propa-
gating wavefronts are unnecessary. The goal might then be to calculate the times between some
predetermined source and receiver positions. This is accomplished with the shooting or bending
ray-tracing techniques. We will discuss the shooting method first.
Similarly to the wavefront-construction method, the ray-shooting approaches are based on the
Hamiltonian 3.10. Unlike the wavefront construction, the shooting techniques do not solve equa-
tion 3.10 numerically. Instead, they use the method of characteristics (e.g., Bleistein, 1984, for a
detailed tutorial) to obtain the ray-tracing system. It consists of two groups of ordinary differential
equations

(3.12)

64

12815-OTE III Gretchka.indd 64 28-07-2009 10:13:58


Rays and traveltimes

(3.13)

where is the coordinate along the characteristic or the ray and t is the ray-propagation time.

The first group 3.12 describes how fast rays propagate through a medium. This becomes clear
once we recognize the components of the group velocity introduced by equation 2.67 in the right-
hand side of equations 3.12. The second group 3.13 determines the change in the slowness vector
along a ray. When changes, the ray bends. Therefore, equations 3.13 control the ray bending
and represent Snells law in smoothly varying anisotropic media. If a medium is homogeneous,
and, as equations 3.12, 3.13 indicate, the rays become straight lines.

The ray shooting is usually implemented as follows.


Step 1. Specify the source location, , and the wavefront normal, , or the shooting direction
at .
Step 2. Substitute into the Christoffel equation 2.50, solve it for the phase velocity, V, of a
given wave mode and apply definition 2.53 to calculate the slowness vector, .
Step 3. Solve the system of ordinary differential equations 3.12, 3.13 with the initial conditions
and . In smoothly heterogeneous anisotropic media, this system has
to be integrated numerically, for instance, using Runge-Kutta methods that are available in
many computational packages. For media containing the surfaces of discontinuity (reflec-
tors) across which the derivative becomes infinite, Snells law should be applied
explicitly; we will discuss its application below. The solution or
is obtained when the ray reaches an observation surface. For typical
seismic reflection geometries, this would be the plane .

Figure 3.2: Rays emerging at observation


surfaces typically do not arrive at predeter-
mined receiver locations (squares).

65

12815-OTE III Gretchka.indd 65 28-07-2009 10:13:59


Chapter 3

Regardless of whether the discontinuities of elastic properties are present, the ray shot in the
direction at the source normally will not arrive at receiver located at the observation
surface (Figure 3.2). If we want to find the traveltime , which would be usually the case, we
will shoot a fan of rays and apply one of the following two strategies.

Interpolation. Once the receiver is inside the cloud of the ray-emergence


points , where are the coordinates at the observation surface, we interpolate
the times towards . This is a straightforward 2D interpolation when
, is single-valued. If it is multivalued, the interpolation is more involved and has
to be done in the 4D space because is
always single-valued. We apply such an interpolation, albeit in two-dimensional space ,
to properly describe multivalued traveltimes in the ray-tracing examples presented in the next sec-
tion.

Optimization. We find the closest to and minimize the distance


(Figure 3.2) by changing the initial ray-shooting direction . As the
unit wavefront normal is specified by two directional cosines or two angles and in equation 2.71,
we end up with a relatively simple two-dimensional optimization. Although the minimum value of
the objective function is known, min , the solution is unique only when the traveltime
is single-valued. If it is not, the convergence might be hampered by discontinui-
ties of associated with the presence of different branches of the traveltime surface.

Snells law
Let us examine an important for practice special case of anisotropic media composed of homogene-
ous layers, which are separated by curved interfaces (Figure 3.3). Kinematic ray tracing then consists
of computing the group velocities of straight ray segments (with either equation 2.67, 2.74, or 2.83)
in layers and applying Snells law to transmit a ray through or reflect it from the medium interfaces.
Here, we will discuss implementation of Snells law in anisotropic media.
Snells law is based on an obvious physical requirement of the wavefront continuity across an

Figure 3.3: Ray trajectory


through a sequence of homo-
geneous anisotropic layers.

66

12815-OTE III Gretchka.indd 66 28-07-2009 10:13:59


Rays and traveltimes

interface and has exactly the same mathematical form in both isotropic and anisotropic media.
Snells law says that the cross products of the normal, , to the interface at the reflection or
transmission point and the slowness vectors, , of the incident, reflected and transmitted waves
should be equal,

(3.14)

In other words, equation 3.14 requires the slowness component tangent to the interface to be pre-
served during reflection or transmission. This fact allows us to recast the reflection-transmission
problem in the local coordinate frame, where and . Equation 3.14 in
such local coordinates becomes

(3.15)

To obtain the slowness vectors, , of reflected and transmitted waves, we need to find their nor-
mal to the interface components . This is done in a straightforward fashion by solving for the
Christoffel equation 2.80, which we rewrite here again:

(3.16)

Equation 3.16 is a six-order polynomial for the unknown and thus has six roots that can be either
real or complex-valued. The real roots correspond to body waves while the complex ones describe
propagation of evanescent energy that should decay exponentially from the interface. Exactly three

Figure 3.4: Kinematics of reflection-


transmission problem in the local
-plane. If the incident wave propa-
gates in the upper half-space, the group-
velocity vectors of reflected waves point
up, , while the group-velocity
vectors of transmitted waves point down,
.

67

12815-OTE III Gretchka.indd 67 28-07-2009 10:14:00


Chapter 3

roots out of those six satisfy the so-called radiation conditions. They require
(i) either the group-velocity vector, , of a body wave to point towards the corresponding half-space
so that the body wave propagates away from the interface,
(ii) or the appropriate sign of Im to ensure an exponential decrease of energy of the evanescent
wave into the desired half-space.

Figure 3.4 illustrates the previous paragraph. It depicts the cross-sections of the slowness sur-
faces by the local -plane. The ordinates of intersections of the slowness cross-sections with
the vertical line (dashed green) are the solutions of equation 3.16 for reflected (upper
half-space) and transmitted (lower half-space) waves. The black arrows indicate the group-velocity
vectors satisfying radiation condition (i). The vectors are orthogonal to the slowness surfaces
(equation 2.82) but, in general, are not confined to the plane . In contrast to the black group-
velocity vectors, red -vector in Figure 3.4 disobeys radiation condition (i) because the correspond-
ing transmitted S-wave would propagate towards the interface. The slowness component of the
transmitted P-wave is complex-valued because max ; the sign of Im should be chosen
in accordance with the radiation condition (ii).
Figure 3.4 displays only one possible geometry of the slowness cross-sections. Many other con-
figurations, which give rise to a variety of complicated reflection and transmission phenomena in
anisotropic media, are discussed in Musgrave (1970), Auld (1973), Petrashen (1980), Kashtan et al.
(1984), Carcione (2001) and Goldin (2008).

3.2.3 Bending or two-point ray tracing methods


The third group of ray-tracing techniques aims at finding a ray trajectory that connects the given
source, , and receiver, , locations. For this reason, the methods that belong to this group are
called two-point ray-tracing methods. They are useful when relatively few seismic sources and receiv-
ers are placed inside a medium, such as in typical cross-well tomography geometries.
To apply two-point ray tracing, one would usually begin with connecting the source and receiver
with a curve , which represents an initial guess for the ray. The vector denotes the quan-
tities that govern the shape of the trajectory and, through it, determine the time .
The physical meaning of these quantities depends on the model parameterization. For instance,
might denote the coordinates of intersections of a ray trajectory with the model interfaces for layered
(Figure 3.3) and blocky models or correspond to the coefficients of basis functions describing a curved
ray (e.g., Grechka and McMechan, 1996). In any case, once the initial trajectory is specified, it has
to be bent hence, the origin of the name bending to satisfy Fermats principle

(3.17)

Equation 3.17 looks deceptively simple, however, it is actually a system of nonlinear equations,
where is the number of components of the vector . While the value of varies
depending on the model complexity and the requirements for accuracy of representation of a curved
ray with a chosen set of basis functions, system 3.17 is notoriously difficult to solve even for moderate .
The standard approach, therefore, is to replace solving the nonlinear equations with finding the
minimum of because the resulting multi-dimensional optimization problem is much easier.
The disadvantage of doing optimization, however, is that not all roots of system 3.17 are travel-
time minima. The roots corresponding to parts of the wavefronts that locally have the shape of a
saddle (such parts are apparent in Figure 2.13c) might also be the saddle points of function .

68

12815-OTE III Gretchka.indd 68 28-07-2009 10:14:00


Rays and traveltimes

Table 3.1: Summary of kinematic ray-tracing methods.

Therefore, those legitimate ray trajectories, which satisfy Fermats principle 3.17, cannot be found
by minimizing t, resulting in the loss of some branches of a multivalued traveltime. Whether or not
this is important depends on a particular application.

3.2.4 Summary of ray-tracing techniques


To conclude this section, we emphasize again that all three discussed ray-tracing techniques
wavefront construction, shooting, and bending are guaranteed to yield exactly the same travel-
times, even though they are based on different physical principles and require solving different (but
related) equations or sets of equations. While the traveltimes should not depend on how they were
computed, their computational cost does. Therefore, it makes sense to apply each technique where
it performs best. Such possible applications are listed in Table 3.1.
The differences in computational efficiency of the ray-tracing methods imply that ideally, one
might wish to have codes implementing all three techniques. If this is not feasible, perhaps priority
should be given to the shooting method as the most versatile. Good-quality anisotropic ray-tracing
codes are collected in package ANRAY (http://seis.karlov.mff.cuni.cz/software/sw3dcd8/anray/anray.
htm), which was originally written by Gajewski and Penk (1987) and has been significantly expand-
ed and improved in the last 20 years.

3.3 Examples
We finish this chapter with several 2D examples of rays and traveltimes of reflected SV-waves in
simple, horizontally layered VTI structures. The goal of those examples is to illustrate that triplica-
tions on the SV-wave group-velocity surfaces might cause complex reflection traveltimes and create
unusual ray geometries from the isotropy standpoint.
Figure 3.5a shows traveltimes and rays in our first two-layer VTI model. The reflection time is
multivalued at offsets (source-receiver distances) smaller than 1.25 km because the SV-wave group-
velocity surface in the bottom layer triplicates near the vertical(5). This triplication gives raise to ray
trajectories that cannot exist in isotropic horizontally layered media. For example, the ray shown
with green arrows in the bottom panel of Figure 3.5a begins at the source at km

5 The SV wavefront in the bottom layer has a triplication near the horizontal too but it does not show up
in the reflection traveltimes displayed in Figure 3.5a.

69

12815-OTE III Gretchka.indd 69 28-07-2009 10:14:01


Chapter 3

Figure 3.5: Reflection traveltimes (top row) and ray trajectories (bottom row) of SV-waves in two-layer
VTI models. The relevant density-normalized stiffness coefficients in Voigt notation are:
in the top layer and in the
bottom layer. Two models differ in the values of stiffness coefficient in the top layer. It is equal to
=4.50 in (a) and =2.36 in (b). The units of all are km2/s2. Red horizontal lines in the bot-
tom panels are the model interfaces.

and arrives at the receiver located at


km. This ray travels from the
source to the first interface and transmits
through it at km. It then
turns back and reflects at km
exactly below the midpoint in accordance
with the model symmetry. The reflected path
is symmetric to the inci-
dent trajectory with respect to the vertical,
again, as prescribed by the symmetry of the
model. Clearly, this ray trajectory would violate
Fermats principle in isotropic layered media;
for anisotropy, however, it obeys it.
Figure 3.5b displays traveltimes and rays Figure 3.6: Detail of the multivalued
similar to those in Figure 3.5a. The two models traveltime in Figure 3.5b.
differ only in the stiffness coefficient in the

70

12815-OTE III Gretchka Chap3.indd 70 28-07-2009 11:02:38


Rays and traveltimes

top layer. Its value is important for the appear-


ance of the reflection times because a smaller
causes the SV group-velocity surface in the
shallow layer to triplicate in the vicinity of 45
from the vertical (such triplications are shown in
Figures 2.11 and 2.12). Rays passing through SV
triplications in both the shallow and deep layers
create a multivalued reflection traveltime that
apparently has several branches near a time of 2 s
in Figure 3.5b. The zoom-in of this portion of
Figure 3.5b displayed in Figure 3.6 revels a com-
plex traveltime geometry. Figures 3.5b and 3.6
show, for example, that the reflected SV-wave has
seven distinct arrivals at a half-offset of 0.1 km.
To compute their times with such precision, we
used the shooting method followed by interpola-
tion in the 2D space-slowness ; -domain,
where the traveltime function ; is single-
valued and smooth.
Finally, we decrease further and present
the result in Figure 3.7. A smaller compared
to that in Figure 3.6 increases the size of the SV
triplication in the top layer, causing more rays
Figure 3.7: Same as Figure 3.5 but for to propagate through both triplications. As a
result, the region occupied by complex, multi-
valued times grows.

71

12815-OTE III Gretchka.indd 71 28-07-2009 10:14:01


12815-OTE III Gretchka.indd 72 28-07-2009 10:14:01
4 THOMSEN
4 Thomsen PARAMETERS
parameters

Let us begin this chapter with an instructive lesson from the history of seismic anisotropy. As we
already know, the phase and group velocities of plane waves in anisotropic crystals had been thor-
oughly studied and well understood by the mid-1960s and early-1970s, when the books of Fedorov
(1965), Musgrave (1970) and Auld (1973) appeared. It took another ten or so years to extend
isotropic ray-tracing techniques, which were introduced in the late-1950s and early-1960s (e.g.,
Alekseev and Gelchinsky, 1959; Alekseev et al., 1961), to anisotropic media and develop software
capable of handling realistic subsurface structures. Thus, it seems that the theoretical basis and
computational machinery for practical applications of seismic anisotropy were in place by the mid-
1980s. Yet, the first anisotropy case studies in the oil and gas industry (excluding a few pertaining
to shear-wave splitting) were not reported until the early and mid-1990s. What was the reason for
such a delay?
To answer this question we need to realize that the ability to trace rays in anisotropic struc-
tures is not necessarily useful for seismic exploration or exploitation. What is actually needed is
to model seismic data acquired in a given geologic environment. This task is quite different from
forward traveltime computations because it involves estimation or inversion of the relevant subsur-
face parameters. Such an inversion is especially difficult in anisotropic media due to the inherently
multi-parameter nature of the inverse problems there (remember, the quantity to be estimated is
a tensor rather than a scalar). The vastness of the parameter space translates into nonuniqueness
of the estimated anisotropy when we lack a clear understanding of what is constrained by the data
at hand. Therefore, there is no surprise that it took some time to figure out which combinations of
the stiffnesses coefficients in common anisotropic symmetries can be inverted from typical seismic
data. The key contribution to this critical for the exploration seismology issue was made by Leon
Thomsen.
This chapter is devoted to Thomsens (1986) anisotropy coefficients. We explain their impor-
tance for understanding seismic signatures in VTI media and emphasize the insights gained by
reparameterizing seismic velocities in terms of those coefficients. In addition, we take a close look
at the values of Thomsens anisotropy coefficients for various lithologies and discuss extensions of
Thomsens notation to symmetries lower than transverse isotropy.

4.1 Phase and group velocities in VTI media


To set the stage for Thomsen coefficients, we need to analyze the phase velocities in VTI media
because, by design, Thomsen parameters capture the most essential features of the velocity behav-
ior. A natural place to begin this analysis is the Christoffel equation 2.50, where we now substitute
VTI stiffness tensor, , given in Voigt notation by matrix 2.42. To simplify the forthcoming equa-
tions, we take advantage of the rotational symmetry of VTI media around the vertical and place
the wavefront normal, , in the vertical symmetry-axis plane . In such a coordinate frame,
and the Christoffel equation 2.50 is given by

73

12815-OTE III Gretchka.indd 73 28-07-2009 10:14:02


Chapter 4

(4.1)

where is the density, V is the p phase velocity, is the unit polarization vector, and the superscript
(vti) in stiffness coefficients is dropped for brevity.
The system of three linear equations 4.1 for the components of splits into two systems. The
first contains just one equation,

(4.2)

while the second comprises the remaining equations,

(4.3)

Equation 4.2 is simpler than 4.3 and we discuss it first. The polarization component in this equa-
tion has to be equal to 1 because according to equation 2.58. This means that a wave described
by equation 4.2 is polarized horizontally in the direction. Its polarization vector, , is
always orthogonal to the wavefront normal, . Therefore, this wave is a pure shear wave.
As it is polarized horizontally, it is called SH the horizontally polarized shear mode.
The right-hand side of 4.2 is zero. Because , this equality is only possible when the term
in parentheses is zero, which implies that the squared phase velocity of the SH wave is

(4.4)

where we used the expression of the wavefront normal,

(4.5)

through its angle with the vertical symmetry axis. The phase velocity is an oval (not exactly
an ellipse but close to it) in the vertical -plane. It can be shown that the inverse quantity, the
slowness , is exactly an ellipse and so is the group velocity. The latter can be obtained from
equation 2.74,

(4.6)

which contains only two terms in VTI media because the phase velocities of all modes are azimuth-
ally independent. We mentioned in Chapter 2 that the group-velocity equation 2.74 breaks down
in the directions of conical (or point) shear-wave singularities. Such singularities are absent in TI
media, therefore, this limitation is lifted and equation 4.6 is always valid.
Let us now turn our attention to system 4.3. It governs waves whose nonzero polarization com-
ponents are and . Clearly, such waves are polarized in the vertical symmetry-axis plane .
One of those waves has to be P while the other is called the vertically polarized shear and denoted
SV. As prescribed by equation 2.58, the polarization vector, , has unit magnitude for either of these

74

12815-OTE III Gretchka.indd 74 28-07-2009 10:14:02


Thomsen parameters

waves; hence, their phase velocities, and , are obtained by setting the determinant of matrix
in 4.3 to zero. This results in a quadratic equation for whose solution is

(4.7)

where the plus sign in front of the radical corresponds to the P-wave, the minus to SV, and equa-
tion 4.5 for the wavefront normal was used.
The phase velocities of P- and SV-waves can be calculated in a straightforward manner from
equation 4.7. A bit more complicated computation, which involves differentiation of equation 4.7 to
find followed by applying equation 4.6, leads to analytic (but rather lengthy) expressions for
the group-velocity vectors and . To the best of the authors knowledge, both calculations were
first carried out by Rudzki (1911), who was also the first to demonstrate that the SV group-velocity
surfaces can have triplications or cusps such as the ones shown in Figure 2.11.
While equations 4.6 and 4.7 are easy to use for computing the group and phase velocities, they
are complex enough to preclude us from gaining insights into how the four stiffness coefficients that
enter equation 4.7 control the shapes of the group- and phase-velocity surfaces of P- and SV-waves.
An obvious step towards this understanding is to analyze the velocities along the vertical symmetry
axis and in the horizontal isotropy plane. The phase and group velocities in these directions are
obtained from equations 4.4, 4.6, and 4.7 by substituting the polar angles for
vertical and horizontal wave propagation, respectively. Once we make these substitutions, we dis-
cover that and for each wave mode , SV, and SH. This result
should not surprise us because due to the symmetry around the vertical and the
velocities are supposed to behave isotropically, meaning , in the horizontal isotropy plane.
The wave-propagation velocities calculated at are listed in Table 4.1. Let us
observe the following:

The vertical and horizontal velocities are expressed in terms of four stiffness coefficients
. The fifth coefficient , therefore, influences the velocities at oblique propaga-
tion directions only.
The shear-wave velocities coincide along the symmetry axis,

(4.8)

The direction , however, is not the shear-wave point singularity because the two
phase-velocity sheets touch each other tangentially at . Such a geometry is termed the kiss
singularity by Crampin and Yedlin (1981).

The SV-wave velocities in the symmetry-axis direction and the isotropy plane coincide,

(4.9)

75

12815-OTE III Gretchka.indd 75 28-07-2009 10:14:02


Chapter 4

Table 4.1: Phase and group velocities of P-,


SV- and SH-waves in the vertical symmetry-
axis direction, , and in the horizontal
isotropy plane, .

As follows from equation 4.9, the influence of anisotropy on SV-waves takes place at oblique
directions of propagation.
Both P- and SH-wave velocities differ in the vertical and horizontal directions (Table 4.1). Their
differences might be used to quantify the magnitude of velocity anisotropy associated with P- and
SH-waves.

4.2 Thomsen notation


The analysis presented in the previous section suggests that understanding of seismic signatures in
VTI media could be facilitated if we capture the four stiffness coefficients that appear in Table 4.1
and are responsible for the vertical and horizontal velocities. Thomsens (1986) choice for this is to
define the P- and S-wave vertical velocities

(4.10)

(4.11)

and characterize anisotropy by the dimensionless coefficients

(4.12)

and

(4.13)

The intuitive appeal of coefficients and is clear: they vanish in isotropic media where
and (see equations 2.45). Hence, the values of and can be used
to measure the magnitudes of the P- and SH-wave anisotropy.
Several other options for quantifying anisotropy of P- and SH-waves have been proposed in the
literature. For instance, Nevsky (1974) uses the coefficients and .
While the transitions from to and from to are obvious, and are somewhat less
convenient than Thomsen coefficients perhaps because when anisotropy is absent. A
different choice was made by Schoenberg and de Hoop (2000). They suggest to replace Thomsen
with a more symmetric coefficient based on the finding that improves
the accuracy of approximations of the P- and SV-wave phase velocities and slownesses in the vicinity
of the horizontal isotropy plane. Coefficient , however, has not been widely used, in part, because

76

12815-OTE III Gretchka.indd 76 28-07-2009 10:14:03


Thomsen parameters

seismic reflection data mainly contain wave-propagation directions close to the vertical and accurate
approximations for the nearly horizontal propagation velocities are seldom needed.
Equations 4.10 4.13 establish the one-to-one correspondence between four density normalized
stiffness coefficients and four Thomsen parameters . To complete
the notation, the fifth anisotropy parameter incorporating is needed. This parameter is the
least intuitive of all because does not influence the vertical or horizontal velocities (Table 4.1)
and thus has to be related to some other seismic signature. Thomsens definition for the stiffness
combination containing is

(4.14)

This expression might seem unexpected. Why would one choose a significantly more compli-
cated combination 4.14 than those for other anisotropy parameters? The significance of becomes
apparent once we notice that according to the exact phase-velocity equation 4.7,

(4.15)

Therefore, is not just an arbitrary combination of the stiffnesses containing . Instead, Thomsen
is responsible for the curvature of the P-wave velocity function at the vertical. It has key importance
for seismic reflection data because it governs the P-wave normal-moveout velocities from horizontal
reflectors. We will discuss the NMO velocities in much more detail in the next chapter.
Since the definition of is not as transparent as those of and , other alternatives exist. Nevsky
(1974) proposes two combinations that incorporate . The first, , reduces to the
expression containing the Poissons ratio for isotropy, . The significance of
parameter is unclear because its value is not necessarily related to the strength of velocity anisot-
ropy. Besides, becomes imaginary when , which can happen in VTI media (e.g., Helbig
and Schoenberg, 1987). If it does, one would replace real stiffness coefficients with complex-valued
anisotropic parameters, which hardly appears to be a simplification. Another combination sug-
gested by Nevsky (1974) is , where in the maximum value of the SV-wave
phase velocity given by equation 4.7. The combination is difficult to use because in order to
calculate , we need to find the corresponding phase angle as a solution a relatively
complicated equation (see Berryman, 2008, for detail). In addition, if Thomsen coef-
ficients and satisfy inequality , the maxima of are located at and .
As , parameter and contains no information about .

Schoenberg and de Hoop (2000) suggest to use the combination

(4.16)

in place of . Although is a valid choice for anisotropy parameter, it does not relate to the P-wave
NMO velocity in the way Thomsen does and, for this reason, is not as useful for seismic data
processing.

77

12815-OTE III Gretchka.indd 77 28-07-2009 10:14:03


Chapter 4

4.2.1 Physical meaning of Thomsen coefficients


We now express from equations 4.10 4.14 and substitute them into the phase-velocity equa-
tion 4.7(6). The result,

(4.17)

where

(4.18)

looks only marginally more concise than the original equation 4.7, so it seems we have not achieved
much.
To show that in fact, a substantial progress has been made, let us assume a weak anisotropy, that
is, , and linearize equation 4.17 with respect to and . Such a linearization,
called the weak anisotropy approximation (WAA), for the P-waves reads

(4.19)

or, equivalently,

(4.20)

where

(4.21)

is the Alkhalifah-Tsvankin (1995) anellipticity coefficient. Hereafter, we use the sign to indicate
the weak anisotropy approximation. Linearizing equation 4.17 with the minus sign in front of the
radical yields

, (4.22)

where the coefficient

6 Note that does not constrain the sign of . Although this sign does not influence the phase veloci-
ties because equation 4.7 contains , it should be taken into account when calculating the polariza-
tion vectors of P- and SV-waves (see Tsvankin, 2001, for a discussion and Helbig and Schoenberg, 1987, for
examples).

78

12815-OTE III Gretchka.indd 78 28-07-2009 10:14:04


Thomsen parameters

(4.23)

was introduced by Tsvankin and Thomsen (1994). For completeness, we also present the lineariza-
tion of equation 4.4 in terms of ,

(4.24)

Equations 4.19, 4.20, 4.22, and 4.24 are well suited for understanding the physical meaning of
Thomsen coefficients and for explaining the behavior of the phase velocities.
Let us begin with the P-waves. Equation 4.19 clearly indicates that and are responsible for
different angular ranges of the phase velocities. Thomsen coefficient governs at near-vertical
propagation (this becomes obvious if we recall equation 4.15) while controls at approaching the
horizontal. When is negative, at small and bends inward as shown in Figure 4.1;
for positive , the phase velocity would bulge outward. Thomsen coefficient offers an even simpler
interpretation. The relationship between the vertical and horizontal velocities,

(4.25)

implies that if is positive (Figure 4.1) and otherwise.

Comparison of equations 4.20 and 4.24 explains why is named the anellipticity coefficient. We can
obtain the approximate SH-wave phase velocity from the P-wave velocity given by equation 4.20
with the following formal substitutions:

(4.26)

Figure 4.1: P-wave phase velocity function Figure 4.2: SV-wave phase velocity function
(solid) for positive and negative . The (solid) for a positive . The dashed circular
dashed line is a circular arc whose radius is arc has a radius equal to .
equal to the vertical velocity .

79

12815-OTE III Gretchka.indd 79 28-07-2009 10:14:04


Chapter 4

Since the SH-wave anisotropy is always ellipti-


cal, the correspondence 4.26 suggests that the
P-wave anisotropy also becomes elliptical when

(4.27)

Table 4.2: Average values of anisotropic While the ellipticity conditions 4.27 are obtained
coefficients measured on core samples by in the WAA, they turn out to be exact. To prove
Wang (2002). this, we just need to substitute in equation
4.17 and verify that

(4.28)

has exactly the form of the first equality 4.24. The phase velocity given by equation 4.28
does not have an elliptical shape but the slowness, , and the group velocity,
, do. Equation 4.28 also explains why is sometimes called the P-wave ellipticity coefficient.
The phase velocity of SV-waves (equation 4.22) is controlled by the Tsvankin-Thomsen (1994)
coefficient . Similarly to substitutions 4.26, we can derive from by making the
replacements

(4.29)

Therefore, one can say that plays the role of for SV-waves and its influence on is equiva-
lent to that of on . Specifically, when is nonnegative (Figure 4.2) and vice
versa, .
If according to the WAA 4.22 and, hence, behaves isotropically. Like
conditions 4.27 for the elliptical anisotropy of P-waves, this weak-anisotropy result is exact. It allows
us to draw an important conclusion:

elliptical anisotropy for P-waves implies isotropy for SV-waves and vice versa.

4.2.2 Values of the anisotropy coefficients


The weak-anisotropy approximations discussed in the previous section are strictly valid only when
the magnitudes of all relevant coefficients are significantly smaller than unity. Therefore, a reason-
able practical question might be what values of those coefficients should be expected in rocks. They
can be found in Thomsen (1986), who compiled the laboratory measurements available by the mid-
1980s. Wang (2002) presented a more comprehensive set of core measurements made with modern
techniques. We list the mean values of anisotropic coefficients published by Wang
(2002) in Table 4.2.
The conclusion that emerges from Table 4.2 is obvious. Shales are the most anisotropic sedimen-
tary rocks. Sandstones are weakly anisotropic (the magnitude of their anisotropy usually increases
with the clay content), while carbonates are virtually isotropic. The coefficients for
shales are positive; can be both positive and negative. This observation from Wangs (2002) results
is consistent with earlier measurements done on a different collection of shale cores by Vernik and
Liu (1997).

80

12815-OTE III Gretchka.indd 80 28-07-2009 10:14:05


Thomsen parameters

The magnitudes of anisotropy coefficients in Table 4.2 are not large, except for in shales,
implying that the accuracy of approximations 4.19, 4.20 and 4.24 is probably sufficient for practical
needs. However, there is no reason to apply WAAs for actual computations because once the influ-
ence of anisotropy parameters on a signature in question, say, on the phase velocity, is understood,
switching to the exact equations would prevent us from making any numerical errors.
This point becomes especially important when anisotropy is inferred from measurements and
the magnitudes of anisotropic coefficients are unknown a priori. For example, the maximum of is
1.05 in Wangs (2002) table. Clearly, such a value has to be computed from the measured velocities
based on exact equations. The same pertains to calculations involving . It is usually greater than
other anisotropic coefficients because in equation 4.23 is multiplied by , whose
typical values lie between 3 and 10. They make commonly measured in shales to be between
0.5 and unity and sometimes exceed 1 [for instance, Wang (2002) measured max , clearly
violating the requirement needed to ensure the numerical accuracy of equation 4.22. Thus,
the preferred approach that both takes advantage of the simplicity of the WAAs and avoids any
approximation errors would be to

understand the influence of anisotropy from the WAAs and then use the exact equations for actual
computations.

Tsvankin (2001) demonstrates the usefulness of this recipe for a variety of seismic signatures.
While different rocks exhibit different magnitudes of seismic anisotropy, it would be useful to
know whether any bounds on the anisotropy coefficients can be derived theoretically. Such low
bounds are available. Perhaps unexpectedly, they are the same for all anisotropic coefficients,

(4.30)

The bounds for and follow directly from their definitions 4.12 and 4.13, if we take into account
the fact that all stiffnesses in those equations are nonnegative to satisfy the stability conditions
for anisotropic solids and liquids(7). Derivations of bounds 4.30 for the other anisotropy coefficients
are more involved and we skip them here.
The stability conditions do not lead to any upper bounds on the anisotropic coefficients. The
absence of the upper bounds can be directly verified from equations 4.12 4.14, 4.21, and 4.23.
They indicate that

(4.31)

In fact, the properties of VTI media where and, consequently, and


are well known (e.g., Alkhalifah, 1998; 2000; Grechka et al., 2004). These media,
termed acoustically anisotropic, are considered helpful for imaging P-wave reflection data. Although it
is unclear whether any literature on anisotropic media characterized by vanishingly small exists,
this is unimportant for us because rocks definitely have nonzero .

7 Excluding anisotropic liquids would result in a strong inequality in 4.30.

81

12815-OTE III Gretchka.indd 81 28-07-2009 10:14:06


Chapter 4

Finally, we mention that a suite of upper bounds for can be derived if we restrict the
ratio to some realistic values. For instance, Tsvankin (2001) finds that for a typical
velocity ratio .

4.2.3 Parameter reduction for P-waves


The WAAs are useful not only for capturing and explaining various features of the velocity behavior.
Perhaps more importantly, they identify the parameters that control the velocities. To be specific,
let us observe that there is no in the weak-anisotropy expressions 4.19 and 4.20 for . Does
this mean that is independent of for arbitrary anisotropy?
The answer is given in Figure 4.3. Even though anisotropy is uncommonly strong in the presented
example and the entire range of plausible ratios is sampled, the maximum separa-
tion between the curves is only about 1%. Clearly, such an accuracy is sufficient for practical
applications.
It is critical to note that the P-wave phase velocity does depend on in the stiffness notation
(equation 4.7). It depends on in Thomsen notation too. However, this dependence is now largely
absorbed by . As a result, once we switch to Thomsen parameterization, the dependence of
on virtually disappears and we reduce the number of parameters governing the P-wave phase
velocity by one:

(4.32)

Figure 4.3 illustrates the accuracy of this parameter reduction.


Correspondence 4.32 carries another important message. As is fully controlled by
and depends on the combination (equation 4.14), the stiffness coefficient cannot be
uniquely determined from the P-wave velocities or traveltimes. This result is known for P-wave VSP
data and P-wave cross-well tomography from studies of Michelena (1994). Here, we drew a more
general conclusion much simpler just by inspecting the P-wave phase velocity parameterized in
terms of Thomsen coefficients.

Figure 4.3: Dependence of


the exact P-wave phase veloc-
ity (calculated with equation
4.17) on in VTI media with
. The
legend indicates the
ratios. The corresponding ver-
tical shear-wave velocities are
, 1.6 km/s, and
2.4 km/s.

82

12815-OTE III Gretchka.indd 82 28-07-2009 10:14:06


Thomsen parameters

4.2.4 Advantages of Thomsen notation


Let us summarize the advantages brought by Thomsen reparameterization of the stiffness coefficients.
1. Because Thomsen parameters vanish for isotropy, their nonzero values explicitly quantify the
magnitude of seismic anisotropy.
2. Thomsen parameters are extremely well suited for developing the weak-anisotropy approxima-
tions of seismic signatures. Those WAAs have been derived for the phase and group velocities,
the polarization vectors (e.g., Tsvankin, 2001), the NMO velocities, the quartic moveout term
(Chapters 5, 6 and references therein), and the reflection and transmission coefficients (Rger,
1997; 2002; Vavrycuk and Penck, 1998; Jlek, 2002).
3. Typically, the WAAs allow us to get an analytic insight into the combinations of anisotropy
parameters that control a given seismic signature and, hence, should be targeted in its inver-
sion.
4. Some WAAs, such as the P-wave ellipticity condition 4.27, turn out to be exact whereas many
others remain useful approximations. Consequently, their accuracy needs to be verified with
numerical computations based on the exact formulae. Whatever magnitudes of errors this veri-
fication reveals, it is always advisable to use the exact equations in actual calculations to avoid
any inaccuracies in the results.
5. While Thomsen coefficients have been originally introduced to simplify the treatment of weak
anisotropy, the strength of anisotropy does not invalidate them. In either case, they capture the
most essential, first-order parameter combinations important for a signature in question. The
previous section illustrates this point.

4.3 Extensions of Thomsen parameterization


4.3.1 Orthotropy
Following the success of using Thomsen notation for vertical transverse isotropy, it has been
generalized to lower symmetries. The first and perhaps most logical extension to orthotropy was
proposed by Tsvankin (1997b). He noticed that the Christoffel equation 2.50 has the VTI form 4.1
in the symmetry planes of orthorhombic media. Specifically, substituting (equation 2.38) in
equation 2.50 yields

(4.33)

in the -plane, where the wavefront normal , and

(4.34)

in the -plane, where the wavefront normal . For brevity, the superscript (ort)
in the stiffness coefficients is dropped in both of the equations above.
The limited equivalence of VTI and orthorhombic media expressed by equations 4.33 and 4.34
makes generalization of Thomsen notation straightforward. Tsvankin (1997b) chooses to define
Thomsens vertical velocities,

83

12815-OTE III Gretchka.indd 83 28-07-2009 10:14:07


Chapter 4

(4.35)

(4.36)

and anisotropy parameters in the -plane,

(4.37)

(4.38)

(4.39)

The superscript (2) in equations 4.37 4.39 indicates that the parameters , and belong
to the plane normal to the coordinate axis .
A simple permutation of subscripts leads to anisotropy parameters in the symmetry
plane,

(4.40)

(4.41)

(4.42)

Equations 4.35 4.42 introduce eight anisotropy parameters and relate them to eight stiffness
coefficients. To complete the notation, Tsvankin (1997b) replaces the remaining independent stiff-
ness element with the parameter , which plays the role of Thomsen in the horizontal sym-
metry plane and is defined as

(4.43)

By design, Tsvankins notation 4.35 4.43 inherits the advantages of Thomsens notation 4.10
4.14. In addition, it makes VTI and HTI media easily treatable special cases of orthotropy. Indeed,
Thomsens VTI parameterization follows from Tsvankins by setting

84

12815-OTE III Gretchka.indd 84 28-07-2009 10:14:07


Thomsen parameters

(4.44)

Likewise, specifying, for instance, the -plane to be the isotropy plane for an HTI medium
results in equalities

(4.45)

and becomes a function of the other parameters. The nonzero anisotropy coefficients ,
and given by equations 4.37 4.39 are usually denoted , , and , respectively, in HTI
media with the symmetry axis (Rger, 1997; Tsvankin, 1997a).
The parameters introduced by equations 4.35 4.43 are useful for understanding seismic sig-
natures not only within the vertical symmetry planes but also outside them. This can be shown by
linearizing the P-wave phase velocity in an arbitrary direction characterized by the polar angle
and azimuth , that is, when the wavefront normal in given by equation 2.71. The WAA for the
P-wave phase velocity reads (Tsvankin, 1997b)

(4.46)

where the azimuthally dependent and are

(4.47)

and

(4.48)

Equation 4.46 closely resembles Thomsens WAA 4.19. The only difference between the two is
that now and vary azimuthally, as one would expect in the presence of azimuthal anisotropy. If
we fix the azimuth , equations 4.19 and 4.46 become identical. This observation has an interesting
implication for non-true-amplitude imaging of 2D P-wave seismic data acquired in the dip direction
of subsurface structures. Such narrow-azimuth data can be processed in a VTI mode, that is, ignor-
ing weak azimuthal anisotropy even when it presents. Again, we gained this important insight just
by switching from the stiffness to Thomsen-type notation.

4.3.2 Monoclinic and triclinic symmetries


Grechka et al. (2000) extended Tsvankins orthorhombic parameterization to monoclinic media.
We skip the analysis of their notation because we will not specifically discuss monoclinic symmetry
in the forthcoming chapters. It suffices to say that Grechka et al. (2000) preserved the advantages
of Tsvankins parameters and incorporated the monoclinic stiffness elements that
vanish for orthotropy in such a way that the key seismic signatures the normal-moveout velocities
of P-, -waves from a horizontal reflector are given in terms of the introduced anisotropy
coefficients exactly.
Several authors (Jech and Penck, 1989; Sayers, 1994; Mensch and Rasolofosaon, 1997; Vavrycuk
and Penck; 1998) have proposed a number of different parameterizations for orthorhombic and

85

12815-OTE III Gretchka.indd 85 28-07-2009 10:14:07


Chapter 4

lower-symmetry media that contain the -coefficients linearized in terms of small stiffness perturba-
tions from isotropy. Those sets of anisotropy parameters are also suitable for developing weak-ani-
sotropy approximations and have been successfully used to this end. Unlike Thomsen coefficients,
however, they lack the ability to handle strong anisotropy, which limits their application.

86

12815-OTE III Gretchka.indd 86 28-07-2009 10:14:08


NORMAL
5 Normal MOVEOUT
moveout

The hyperbolic portion of the pure-mode reflection moveout is undoubtedly the most stable and
reliable part of conventional seismic data. Arguably, it is the most significant part too because it
provides the key quantity, known as the normal-moveout (NMO) velocity, which is routinely used
to build velocity models for seismic imaging. If the subsurface happens to be anisotropic, which is
often the case, isotropic velocity-model building techniques are bound to result in an inaccurate
velocity field and consequently, in blurred and misplaced seismic images.
In this chapter we analyze the influence of anisotropy on NMO velocities. We begin with 2D VTI
models and find out which anisotropic parameters control the NMO velocities from horizontal and
dipping reflectors. Next, we discuss an extension of the Dix (1955) formula to laterally homogene-
ous anisotropic media and examine the opportunities provided by this extension for anisotropic
parameter estimation.
The second part of this chapter is devoted to wide-azimuth seismic data and azimuthal anisot-
ropy. We introduce the concept of the NMO ellipse, which has become a standard tool for azimuthal
velocity analysis, and describe further generalizations of the Dix formula that can be used to build
azimuthally anisotropic velocity models of heterogeneous subsurface.

5.1 NMO velocity in a single VTI layer


Reflection traveltimes, t, of pure (non-converted) modes are usually approximated by the Taylor
series with respect to the offset, (e.g., Bolshikh, 1956; Taner and Koehler, 1969):

(5.1)

where the coefficients and are given by

(5.2)

and is the two-way zero-offset traveltime. The series 5.1 does not contain odd powers of because
of the reciprocity: interchanging the source and receiver locations changes the sign of the offset
but does not alter the pure-mode reflection times. Although the same is true for converted waves
in laterally homogeneous anisotropic media with a horizontal symmetry plane, our discussion will
be limited to pure modes.
The number of terms in equation 5.1 needed to accurately represent traveltimes usually depends
on the ratio of the maximum offset to the reflector depth(8). For conventional seismic acquisition,
this ratio is close to unity and accuracy sufficient for practice is usually achieved with only two
terms,

8 Elliptically anisotropic media (Grechka, 2009) and homogeneous isotropic media are an exception.

87

12815-OTE III Gretchka.indd 87 28-07-2009 10:14:08


Chapter 5

(5.3)

where

(5.4)

is the normal-moveout velocity. Equation 5.3 describes a hyperbola in the -plane and is called
the hyperbolic moveout approximation. As follows from the Pythagorean theorem, this approximation
is exact if a medium above a horizontal reflector is homogeneous and isotropic. Then, the NMO
velocity, , is simply equal to the medium velocity, , pointing to a possibility of using reflection
traveltimes to estimate the velocities. When the reflector beneath a homogeneous isotropic medium
is dipping, on the dip line relates to as (Levin, 1971)

(5.5)

where is the reflector dip.

The presence of anisotropy considerably complicates things. Under the assumption of 2D wave
propagation (Figure 5.1), the NMO velocity from a dipping interface beneath a homogeneous ani-
sotropic layer measured in the reflector dip plane is given by (Tsvankin, 1995)

(5.6)

Here, is the phase velocity and is the phase angle. The derivatives in equation 5.6 are evalu-
ated at the phase angle of the zero-offset ray because, according to Snells law, the slowness
vector, , of the zero-offset ray is orthogonal to the reflector (Figure 5.1) whereas its group-velocity

Figure 5.1: Ray trajectories in


the reflector dip plane at zero
and nonzero offsets. The vertical
-plane is assumed to be a
symmetry plane so that the strike
components of vectors and
are equal to zero (modified from
Tsvankin, 1995).

88

12815-OTE III Gretchka.indd 88 28-07-2009 10:14:08


Normal moveout

vector, , generally deviates from the reflector normal. Although equation 5.6 is valid for any ani-
sotropic medium that has the symmetry plane (for example, orthorhombic or monoclinic),
we analyze this equation for vertical transverse isotropy.

5.1.1 Horizontal layer


Let us begin with a horizontal VTI layer. When the reflector dip , equation 5.6 reduces to

(5.7)

If we substitute the exact phase-velocity equation 4.17 in formula 5.7, we obtain

(5.8)

(5.9)

Both NMO velocities 5.8 and 5.9 are exact and valid for arbitrary strength of anisotropy provided
that and satisfy inequality 4.30. Similarly, the exact zero-dip NMO velocity of SH-waves is

(5.10)

Now we can appreciate Thomsen parameters even more: they yield exact and concise expressions
for perhaps the most important quantities in seismic exploration the NMO velocities from hori-
zontal reflectors.
Equations 5.8 5.10 show that the zero-dip NMO velocities are not equal to the corresponding
vertical velocities in the presence of anisotropy. The discrepancy between the two is the main physi-
cal reason for depth misties between well logs and seismic sections when the latter are isotropically
depth-migrated with the stacking velocities inferred from the data. As those misties are ubiquitously
observed throughout the world, the industry came up with a simple fix that places seismic hori-
zons at the correct depths: one needs to squeeze or stretch the isotropic images in depth to match
the well-log or check-shot data. For P-waves, the stretch factor is in accordance with
equation 5.8. Although this stretching procedure removes the depth misties, it does not solve other
imaging problems caused by the presence of anisotropy. We discuss these problems next.

5.1.2 Dipping layer


To derive the NMO velocity in a dipping VTI layer, we have to return to equation 5.6. Its significant
simplification is achieved if we assume that anisotropy is weak. The WAA of equation 5.6 for P-waves
reads (Tsvankin, 1995)

(5.11)

Let us make two important observations:


1. The isotropic term in equation 5.11, , exhibits the familiar cosine-of-dip depend-
ence 5.5. Unlike isotropic media, where , the presence of anisotropy explicitly
requires the use of the zero-dip NMO velocity.

89

12815-OTE III Gretchka.indd 89 28-07-2009 10:14:08


Chapter 5

2. The influence of anisotropy in equation 5.11 is dominated by the second term in brackets. This
becomes evident if we compare the multipliers of the and terms. The former is greater
than the latter by a factor that changes from 3 to 6, depending on the reflector dip. As
is the weak-anisotropy approximation for the anellipticity coefficient (equation 4.21), we con-
clude that the dip moveout of P-waves in VTI media is largely controlled by .

The second point implies that a simple depth stretch discussed in the previous section is not
going to help us to image relatively steeply dipping features of the subsurface because the stretch
compensates for the influence of whereas the dip moveout is mainly governed by . We illustrate
this anisotropy-related imaging problem with the case study from offshore Africa described by
Alkhalifah and Tsvankin (1995).
Figure 5.2 shows a time-migrated seismic section processed without taking anisotropy into
account. While subhorizontal reflectors are imaged well, dipping faults, such as the one at a time
close to 2 s at common midpoint (CMP) 900, are hardly visible. To demonstrate that the inability to
image the faults is caused by anisotropy, Alkhalifah and Tsvankin (1995) generated constant-velocity
stacks by applying a conventional isotropic dip-moveout (DMO) correction, which is designed to
focus both horizontal and dipping events on the same velocity panel (e.g., Yilmaz, 2001). Figure 5.3,
however, shows that the subhorizontal reflectors are imaged best at a stacking velocity of 2200 m/s,
whereas the dipping reflector comes into focus at a higher velocity of approximately 2450 m/s.
As a result, isotropic DMO fails to image both dips simultaneously. Clearly, to bring them to the
same velocity panel, the NMO velocity has to increase with dip faster than predicted by the isotropic
cosine-of-dip dependence 5.5.
One way to properly account for the influence of dip on the NMO velocity is to treat the entire
subsurface as a homogeneous VTI medium, apply the anisotropic DMO equation 5.11 or its exact
counterpart 5.6, and find the anellipticity coefficient , which provides the required increase of
the NMO velocity. Alkhalifah and Tsvankin (1995) obtained such and Figure 5.4 displays
the stacks that result from VTI DMO processing and exhibit both dips simultaneously. This value
of can be regarded as an effective parameter because it incorporates the influence of both anisot-
ropy and heterogeneity of the subsurface. To separate the two, more elaborate data processing is
needed.

Figure 5.2: Time-migrated seismic


line. The gray bar indicates the
locations of CMP gathers exam-
ined in Figures 5.3 and 5.4 (after
Alkhalifah and Tsvankin, 1995).

90

12815-OTE III Gretchka.indd 90 28-07-2009 10:14:09


Normal moveout

Figure 5.3: Constant-velocity stacks after the


conventional isotropic NMO-DMO sequence.
The velocity values correspond to the NMO
velocities of the subhorizontal reflectors (after
Alkhalifah and Tsvankin, 1995).

Figure 5.4: Anisotropic constant-velocity


stacks. The velocities correspond to the stack-
ing velocities of the subhorizontal reflectors
(after Alkhalifah and Tsvankin, 1995).

5.2 Anisotropic Dix equation


Here, we discuss such processing under the assumption of lateral homogeneity. We begin with horizon-
tally layered media.

5.2.1 Stack of horizontal VTI layers


If the layers were isotropic, we could apply the conventional Dix (1955) formula to express the effective
NMO velocity from the bottom of layer number (Figure 5.5),

(5.12)

in terms of the interval NMO velocities, , the interval zero-offset times, , and the effective
zero-offset time

91

12815-OTE III Gretchka.indd 91 28-07-2009 10:14:09


Chapter 5

Figure 5.5: Stack of


horizontal layers.

(5.13)

Equation 5.12 can be rewritten in the so-called differentiation mode,

(5.14)

which allows one to calculate the interval NMO velocities from the effective quantities.

Remarkably, equations 5.12 5.14 remain valid in horizontally layered VTI media. Unlike isot-
ropy, however, where the interval NMO and isotropic layer velocities can be used interchangeably,
the presence of anisotropy explicitly requires equations 5.12 and 5.14 to be applied to the NMO
velocities. While the effective is given by equation 5.12, the vertical velocities, , obey quite a
different averaging in either isotropic or VTI layered media,

(5.15)

The latter can also be cast in the differentiation form,

(5.16)

In equations 5.15 and 5.16, are the layer thicknesses and is the depth of the bottom of the
layer.

92

12815-OTE III Gretchka.indd 92 28-07-2009 10:14:10


Normal moveout

In isotropic stratified media, the difference between the root-mean-square (RMS) average 5.12
and arithmetic average 5.15 leads to inequality

(5.17)

It originates from treating the vertically inhomogeneous medium as a homogeneous layer and
indicates the presence of the so-called processing-induced anisotropy (Grechka and Tsvankin, 2002b).
For the P-waves, the difference expressed by inequality 5.17 translates into a nonnegative effective
, which is defined as

(5.18)

in accordance with equation 5.8. Although the bias towards higher is always present due to
the inability of relatively low-frequency seismic reflection data to fully account for fine layering
observed, for instance, in well logs, processing-induced anisotropy seldom increases by more than
0.03. This bias also tends to decrease for narrower averaging intervals.
In VTI media, processing-induced anisotropy and the intrinsic anisotropy of layers are super-
imposed. While equation 5.18 can still be used to calculate the effective , the interval vertical and
NMO velocities, and , that enter expressions for the effective and are
no longer equal. Instead, their relationship given by equation 5.8 leads to the expression of the
interval ,

Figure 5.6: (a) Average P-wave vertical velocity from a check shot (gray), the effective NMO velocity
(black), and (b) the interval Thomsen coefficient (after Alkhalifah et al., 1996).

93

12815-OTE III Gretchka.indd 93 28-07-2009 10:14:10


Chapter 5

(5.19)

The inherent difference between the NMO and


vertical velocities in VTI media implies that the
latter cannot be inferred from seismic reflection
data and have to be measured independently.
Conventionally, vertical velocities are obtained
from check shots. Once both the effective NMO
velocities and the check-shot data are differenti-
ated (with equation 5.14 and the first equation
5.16, respectively), the interval can be calcu-
lated using equation 5.19.
Such a calculation was presented by Alkhalifah
et al. (1996) for check-shot and NMO velocities
measured near CMP 200 in Figure 5.2. The
check-shot data and the effective NMO veloc-
ity are displayed in Figure 5.6a. They yield
the interval -curve shown in Figure 5.6b. The
Thomsen coefficient increases with depth and
reaches a fairly large value of about 0.35 at a
vertical time of 2.2 s. This behavior, however, is
corroborated by the available geologic informa-
Figure 5.7: Zero-offset ray (blue) reflected tion the section above 2.3 s consists of strongly
from a dipping interface beneath a layered anisotropic shales (Alkhalifah et al., 1996).
anisotropic medium. The dashed lines indicate
nonexisting interfaces that would generate zero- 5.2.2 Dip moveout and its inversion in
VTI media
offset reflected rays with exactly the same trajec-
tories as the one shown with the solid blue line Let us now examine NMO velocities in more
(modified from Alkhalifah and Tsvankin, 1995). complicated structures, where a laterally homo-
geneous overburden overlays a dipping reflector
(Figure 5.7). If we keep our assumption that the
vertical dip plane in Figure 5.7 is a symmetry plane of the entire model, the NMO velocity in this plane
is given by a Dix-type formula that closely resembles equation 5.12 (Alkhalifah and Tsvankin, 1995),

(5.20)

where

(5.21)

and are the interval NMO velocities and zero-offset times along oblique segments of
the zero-offset ray in Figure 5.7. The quantity is the horizontal slowness component in the
-plane. It is preserved along rays in a stratified overburden according to Snells law.

94

12815-OTE III Gretchka.indd 94 28-07-2009 10:14:10


Normal moveout

Equations 5.20 and 5.21 have been known for isotropic media since the work of Shah (1973).
They explicitly require the use of the dip-dependent interval NMO velocities, which are given by the
isotropic Levins (1971) equation 5.5. This equation, rewritten in terms of the horizontal slowness
rather the reflector dip, has the form

(5.22)

which allows us to make the following instructive observation. Let us note that the reflector dips cor-
responding to the NMO velocities for (dashed lines in Figure 5.7) are absent in our
horizontally layered model. Therefore, to apply equation 5.20 in practice, we have to measure the
effective NMO velocities from horizontal interfaces, use the Dix-differentiation formula 5.14 to obtain
the interval , and only then can the interval be calculated from equation 5.22. To
perform this calculation we also need to know the reflector slope , which is conventionally estimated
from zero-offset time sections. Although the procedure outlined above is definitely more involved
than that for horizontal reflectors,, its implementation
p is nevertheless straightforward because both
and are expressed via the directly measurable quantities
and .
Recalling our discussion in the previous section, we might expect that the presence of vertical trans-
verse isotropy ought to complicate things. This extra level of complexity originates from equations 5.8
and 5.11. Specifically, anisotropy influences the P-wave NMO velocity, , only through for hori-
zontal reflectors whereas the reflector dip introduces an additional dependence of on the anel-
lipticity coefficient . Therefore, it is unclear whether knowing would be helpful for obtaining
when both and are unknown. This issue was resolved by Alkhalifah and Tsvankin
(1995), who showed that is virtually independent of . They demonstrated this in the WAA,

(5.23)

where

(5.24)

the superscript indicating the layer number was dropped for brevity. Equations 5.23 and 5.24
indicate that as a function of the reflection slope depends on just two combinations of
the VTI parameters: the zero-dip NMO velocity and the anellipticity coefficient .
Independence of of the individual values of Thomsen coefficients and , which follows
from the WAA 5.23, can be verified by exact numerical calculations. They are presented in Figure 5.8.
Variations in and have almost no influence on as long as is fixed (Figure 5.8a). In con-
trast, changes in the anellipticity coefficient strongly influence . Figure 5.8b demonstrates
that can be reasonably well resolved from the dip-dependence of the NMO velocity at dips exceeding
approximately 25-30.
As is controlled by just two quantities, and , they both can be estimated
from the NMO velocities measured from two sufficiently different dips. One of those dips is typi-
cally close to zero, leading to the following procedure for obtaining the interval and as
functions of the vertical time :

95

12815-OTE III Gretchka.indd 95 28-07-2009 10:14:10


Chapter 5

Figure 5.8: Influence of transverse isotropy on the P-wave NMO velocity. Calculations are done with
equation 5.6 normalized by (after Alkhalifah and Tsvankin, 1995). The horizontal
axes approximately correspond to sin ; the dip angles range from 0 to 70. (a) Models with the
same but different and , (solid black); (gray); and
(dashed). The three curves in (a) nearly overlap. (b) Models with different values:
(solid black); (gray); and (dashed).

estimate from subhorizontal reflections, for example, from conventional semblance


analysis;
apply Dix differentiation [equation 5.14 or its appropriately regularized version (e.g., Han et al.,
2001)] to calculate the interval NMO velocities;
measure the slopes, , of dipping horizons on the zero-offset time section (e.g., Fomel, 2002);
obtain the NMO velocities corresponding to the measured slopes ; and
find the interval by inverting the exact function .

Han et al. (2001) applied the outlined processing sequence to 2D synthetic data generated by
J. Leveille and F. Qin (pers. comm.) for the VTI model displayed in Figure 5.9. In Chapter 1 we dis-
cussed a suite of migrations of this data set. The images of the dipping fault at midpoints ranging from
14 km to 18 km are shown in Figure 1.3. We now describe the results of inverting the intervals
and . There are two places in the model in Figure 5.9, where sufficiently steeply dipping reflectors
are available and our parameter-estimation procedure can be implemented: a flank of the salt dome
at midpoints around 5.5 km and the dipping fault.
The solid lines in Figures 5.10 and 5.11 mark the inversion results. As Han et al. (2001) chose to
regularize the problem by smoothing the estimated interval NMO velocity and values, the obtained
-curve (solid) in Figure 5.10b represents an accurate but smooth version of the actual piecewise-constant
-function (dashed). The imposed smoothing did not allow the inversion algorithm to resolve a relatively
thin low- layer at a vertical time of around 3 s in Figure 5.11b. The insensitivity of the inversion to high-
frequency variations of can be also understood on physical grounds: those variations do not produce
noticeable changes in the measured effective NMO velocities. Apart from the problem with vertical reso-
lution, Figure 5.11b adequately reproduces the low-frequency trend of interval .

To summarize this section we state that the extension of the Dix formula (equation 5.20) to
laterally homogeneous anisotropic media leads to development of practical techniques for aniso-
tropic parameter estimation. When applied to P-waves propagating through a stack of horizontal

96

12815-OTE III Gretchka.indd 96 28-07-2009 10:14:11


Normal moveout

Figure 5.9: Parameters of a 2D VTI model: (a) the P-wave vertical velocity , the Thomsen coefficients
(b) and (c) and (d) the anellipticity coefficient (after Han et al., 2001).

Figure 5.10: Interval (a) P-wave NMO


velocity and (b) anellipticity coefficient
(solid lines) as functions of the two-
way vertical time estimated from sub-
horizontal and dipping reflections at
midpoints between 4.9 km and 6.7 km
in Figure 5.9. The dashed line in (b)
indicates the true values of (after
Han et al., 2001).

VTI layers above a dipping reflector, these techniques result in the interval zero-dip NMO velocity
and the anellipticity coefficient as functions of the vertical time. The horizon depths in such VTI
media, however, cannot be obtained from reflection data and have to be constrained, for example,
by check shots.

5.3 NMO ellipse


The previous two sections were devoted to 2D problems. While we allowed a reflector to be dipping, we
had to assume that both the seismic acquisition line and the ray trajectories are confined to the reflector
dip plane. Here, we relax those assumptions and analyze NMO velocities measured in wide-azimuth or
multi-azimuth surveys.

97

12815-OTE III Gretchka.indd 97 28-07-2009 10:14:12


Chapter 5

Figure 5.11: Same as Figure 5.10 but


for midpoints in the range 11.0 km
16.8 km (after Han et al., 2001).

Let us suppose that pure-mode reflection data are acquired along CMP lines that make different
azimuths with respect to the coordinate direction (Figure 5.12). We are interested in the reflec-
tion traveltime, t, at the common midpoint, , where those CMP lines intersect. An approxima-
tion of this traveltime is given by the Taylor series (Grechka and Tsvankin, 1998b),

(5.25)

where are the source and receiver coordinates with respect to , is the two-way
zero-offset time, and are the second-order partial derivatives
of the one-way traveltime, , from the zero-offset reflection point. The dots in equation 5.25 denote
terms higher than quadratic in and that are ignored throughout this chapter. The contribution
of those terms will be examined in Chapter 6, where we discuss nonhyperbolic moveout.
If we express the source and receiver coordinates via the full offset, , and the azimuth, , of the
CMP line (Figure 5.12),

(5.26)

and substitute expressions 5.26 in equation 5.25, it takes a familiar form

(5.27)

where

(5.28)

is the azimuthally varying NMO velocity. The elements of the 22 symmetric matrix are

(5.29)

98

12815-OTE III Gretchka.indd 98 28-07-2009 10:14:12


Normal moveout

Figure 5.13: NMO ellipse that has the


azimuth of the major axis and the semi-
axes = and = .

Here, are the components of the slowness vec-


tor corresponding to the ray propagating from
the zero-offset reflection point (Figure 5.12)
to the point at the horizontal observation
plane. The zero-offset traveltimes appear in
equation 5.29 because reflection-point disper-
sal has no influence on NMO velocity of pure
Figure 5.12: Ray trajectories and pure-mode modes (e.g., Hubral and Krey, 1980).
reflection traveltime in wide-azimuth CMP It is instructive to use the eigenvectors of
geometry (after Grechka et al., 1999b). the matrix as auxiliary horizontal axes and
rotate by the angle

(5.30)

This rotation reduces equation 5.28 to

(5.31)

where and are the eigenvalues of and is the angle between the eigenvector correspond-
ing to and the -axis. As the reflection traveltimes expressed by equation 5.27 typically
increase with offset in CMP geometry, the squared NMO velocity, , has to be positive. Hence,
the matrix is positive definite and its both eigenvalues and are positive. Consequently,
equation 5.31 represents an ellipse. Grechka and Tsvankin (1998b) termed it the NMO ellipse
(Figure 5.13).
A negative eigenvalue implies a negative in certain azimuthal directions and a
decrease of the CMP traveltime with the offset. Although such a reverse moveout sometimes exists
(e.g., for turning waves, as described by Hale et al., 1992), both and are usually positive and the

99

12815-OTE III Gretchka.indd 99 28-07-2009 10:14:12


Chapter 5

azimuthal dependence of NMO velocity is indeed elliptical. This conclusion is valid for arbitrarily het-
erogeneous, anisotropic media provided that the traveltime field is sufficiently smooth to be adequately
approximated by the Taylor series 5.25. Thus, we can state that

as long as the hyperbolic moveout approximation is satisfactory at all azimuths, the azimuthal
variation of the NMO velocity is an ellipse.

The elliptical azimuthal dependence of the NMO velocity entails the following practical implications:
1. Because an ellipse is fully described by three parameters the lengths of its two semi-axes and
the orientation or the elements of the symmetric matrix at least three well-separated (in
the azimuthal sense) CMP lines are required for its reconstruction. This requirement is impor-
tant for marine seismic acquisition. Wide-azimuth geometries on land often look like that shown
in Figure 5.17 below, not only satisfying the three-line requirement but also providing a useful
redundancy for the estimation of NMO ellipses.
2. The three-parametric nature of an ellipse implies that wide-azimuth reflection data recorded at
conventional offsets constrain up to three combinations of the medium parameters per common mid-
point and per reflection event.

Next, we determine those parameter combinations for common anisotropic symmetries.

5.3.1 Homogeneous anisotropic layer


It is logical to start our analysis with a homogeneous anisotropic layer above a dipping and possibly
curved reflector. For such a model, an exact analytic expression for the matrix exists (Grechka
et al., 1999b):

(5.32)

Here and are the components of the slowness vector, , of the zero-offset ray
and the vertical slowness, , is treated as a function of the horizontal components, and
. The partial derivatives and , defined as ,
are evaluated at the and values of the zero-offset ray.
In general, the slowness vector is available once the zero-offset reflected ray is traced (see
Chapter 3 for detail). Therefore, equation 5.32 is fully analytic because the derivatives and
can be obtained by implicit differentiation of the Christoffel equation 2.80. Consequently,

computation of NMO ellipses requires tracing of the zero-offset rays only.

This conclusion is extremely important for developing efficient anisotropic velocity-model


building methods as it hints at a possibility of replacing multi-offset, multi-azimuth ray-tracing,
which would be required to compute NMO ellipses, with tracing of just zero-offset rays followed
by straightforward analytic computations. To proceed with anisotropic parameter estimation,
first we need to find out what quantities could be obtained from NMO ellipses.

Horizontal orthorhombic layer


In the special case of a horizontal layer that has a vertical symmetry plane, equation 5.32 becomes
simpler. As the reflector is horizontal, the slowness components and vanish, .

100

12815-OTE III Gretchka.indd 100 28-07-2009 10:14:13


Normal moveout

We can also choose the symmetry plane to be the coordinate plane . This makes =0 and
reduces equation 5.32 to

(5.33)

Substituting this matrix into equation 5.28 yields

(5.34)

Clearly, the semi-axes of the NMO ellipse are

(5.35)

When a horizontal layer has orthorhombic symmetry (e.g., in Figure 2.6), the principal (symme-
try-plane) P-wave NMO velocities, , can be concisely expressed in terms of the vertical P-wave
velocity and Tsvankins (1997b) anisotropy coefficients and (equations 4.41 and 4.38).
The exact result reads

(5.36)

Equations 5.36 represent an obvious extension of Thomsens (1986) VTI equation 5.8 to orthot-
ropy. Similarly to the P-wave NMO velocities in VTI media, the P-wave NMO ellipses in orthorhom-
bic media do not allow us to estimate the vertical velocities and, hence, cannot be used to constrain
the reflector depths. There is one model, however, where obtaining the true depths from P-wave
reflection data is feasible. This model is the horizontal transverse isotropy. As follows from equa-
tions 4.45, = 0 in HTI media with the symmetry axis and semi-axes 5.36 of the NMO ellipse
become (Tsvankin, 1997a)

(5.37)

It is no coincidence that the P-wave NMO velocity in the vertical -plane is equal to the
vertical velocity this is the isotropy plane of the horizontal HTI layer. The result expressed by
the second equation 5.37 is certainly useful for building HTI velocity models from P-wave reflec-
tion data. Unfortunately, it seldom leads to practical applications because HTI symmetry itself is
deemed too restrictive for sedimentary formations of exploration interest.

Dipping VTI layer


Next, we apply the same formalism to the NMO velocities from a plane dipping reflector beneath a
homogeneous VTI layer. We already know that: (i) the azimuthal variation of the NMO velocity is an
ellipse given by equations 5.28 and 5.32, and (ii) the ellipse has to be aligned with the reflector dip
plane in Figure 5.14 because this plane is the symmetry plane of the entire model. Hence,
the NMO velocity at an arbitrary azimuth is given by

101

12815-OTE III Gretchka.indd 101 28-07-2009 10:14:13


Chapter 5

Figure 5.14: NMO ellipse in a dipping VTI layer.

Figure 5.15: The strike-line P-wave NMO


velocity (normalized by the
zero-dip NMO velocity) as a function of the Figure 5.16: Dipping reflector beneath
ray parameter for dips ranging from 0 a horizontally layered overburden. The
to 90 (after Grechka and Tsvankin, 1998b). effective NMO ellipse in this model can
(a) Different models with the same be obtained from the generalized Dix
but different and equation 5.40 (after Grechka et al., 1999b).
(solid); (dashed);
, (dotted). (b) Models with
different values: (solid);
(dashed); (dotted).

102

12815-OTE III Gretchka.indd 102 28-07-2009 10:14:13


Normal moveout

(5.38)

where is the azimuth from the reflector dip plane, is the reflection slope, and
and are the dip- and strike-line NMO velocities, respectively.
The WAA 5.23 and Figure 5.8 indicate that for the P-waves, is a function of the zero-
dip NMO velocity and the anellipticity coefficient . The same remains true for the strike-
component of the P-wave NMO-velocity. Figure 5.15 demonstrates this and allows us to draw an
important conclusion:

in VTI media, the P-wave NMO ellipse as a whole is governed by and .

For typical VTI models with a positive anellipticity coefficient (Table 4.2), both the dip- and
strike-line NMO velocities increase with dip faster than they do for isotropy. However, as follows
from comparison of the vertical scales in Figures 5.8b and 5.15b, the NMO velocity in the strike
direction is less sensitive to anisotropy than that in the dip direction.
The dependencies presented in Figures 5.8 and 5.15 can be summarized in the following
form:

(5.39)

where and are different functions. Therefore, it is possible to invert two equations 5.39 for
and . Grechka and Tsvankin (1998b) showed that such an inversion is stable provided
that the reflector dip exceeds approximately 40. In practice, however, one would typically combine
information contained in the dip dependence of the NMO ellipse with that in the NMO velocities
from subhorizontal reflectors to improve the reliability of the estimated quantities.

5.4 3D Dix equation


Similarly to effective NMO velocities in 2D anisotropic media, effective NMO ellipses can be construct-
ed from the interval ones. This result, which can be viewed as an extension of the 2D Dix-averaging
equation 5.20 to laterally homogeneous azimuthally anisotropic media, was obtained by Grechka et al.
(1999b). The effective NMO ellipse from a dipping reflector (Figure 5.16) is given by

(5.40)

where

(5.41)

103

12815-OTE III Gretchka.indd 103 28-07-2009 10:14:14


Chapter 5

All quantities in equations 5.40 and 5.41 are computed for the horizontal slowness components of
the zero-offset ray, which are preserved along the ray trajectories in accordance with Snells law.
Unlike equation 5.20, which gives the RMS averaging procedure of the dip components of the
NMO velocities, the generalized Dix formula 5.40 averages the NMO ellipses as a whole. Clearly,
equations 5.20 and 5.40 describe different averages. If the reflector dip plane in Figure 5.16 is also
a symmetry plane of the entire model, equation 5.40 reduces to 5.20 in that plane. Under those
conditions, the RMS averaging 5.20 is also applicable to the strike-components of the NMO veloc-
ity (Grechka et al., 1999b). In all other azimuthal directions, however, equation 5.20 is inexact. Its
accuracy depends on the eccentricities (or elongations) of the interval NMO ellipses, which are
controlled by both the azimuthal anisotropy and the reflector dip. The approximation provided
by equation 5.20 is likely to be sufficient for practical needs when the magnitude of interval azi-
muthal anisotropy is moderate (say, up to 20% 30%) and the reflector dip is less than about 45
(Grechka et al., 1999b). At steeper dips, equation 5.20 rapidly loses its accuracy and averaging of
the full NMO ellipses is required even for a purely isotropic overburden. This interesting find-
ing exposes the inherent limitation of narrow-azimuth acquisition geometries in building velocity
models in structurally complex areas, where steep dips (associated, for instance, with salt tectonics)
are present.
To the best of the authors knowledge, the first practical application of the generalized Dix for-
mula 5.40 was presented by Grechka and Tsvankin (1999b). They used it to process a wide-azimuth
data set acquired in the Powder River Basin, Wyoming, USA. To enhance the signal-to-noise ratio,
the data were collected into the so-called superbins, each with a somewhat random distribution
of azimuths and offsets (Figure 5.17). The effective NMO ellipses were then obtained from 3D sem-
blance analysis for the whole superbin. This global semblance analysis is based on the hyperbolic
moveout equation 5.27 and requires simultaneous scanning over all three elements of the matrix
responsible for the NMO ellipse (equation 5.28)(9).
Figure 5.18 shows typical semblance curves obtained from conventional and azimuthal velocity anal-
yses. While the two curves do not differ over most of the interval because the NMO ellipses are close
to circles, the azimuthal velocity analysis noticeably improves the fit to moveout and provides higher
semblance values at a vertical time of 2.57 s. According to Withers and Corrigan (1997), the event at
2.57 s is the basement reflection beneath the presumably fractured Frontier/Niobrara formation.
In general, the measured moveout ellipticity can be attributed to any combination of the follow-
ing three factors:
(1) reflector dip,
(2) lateral velocity variation in the overburden and
(3) azimuthal anisotropy.
While the distortions due to dips were negligible in the discussed data set, Grechka and Tsvankin
(1999b) had to correct the NMO ellipses for the presence of lateral heterogeneity. The eccentricities,

(5.42)

9 Recently, Lynn (2007) showed that this method is more robust in the presence of noise than the conven-
tional velocity analysis of data divided into narrow azimuthal sectors followed by an elliptical fit of the
obtained moveout velocities. Another approach for extracting the NMO ellipses was proposed by Jenner
(2001) and Jenner et al. (2001). They cross-correlate moveout-corrected traces to retrieve the azimuthally-
dependent relative time shifts.

104

12815-OTE III Gretchka.indd 104 28-07-2009 10:14:14


Normal moveout

of the NMO ellipses, smoothed and corrected for the lateral velocity heterogeneity, are displayed in
Figure 5.19. Each tick in this figure corresponds to the effective NMO ellipse at a certain superbin.
The direction of a tick indicates the azimuth of the semi-major axis of the NMO ellipse and the
ticks length is proportional to the eccentricity E. We observe little eccentricity, that is, virtually no
effective azimuthal anisotropy for the shallow reflection at s (Figure 5.19a). In contrast,
the NMO eccentricity is evident for the deeper event (Figure 5.19b); the NMO ellipses are extended
predominantly in the N-S direction.
To find the NMO ellipses in the interval s, the so-called differentiation form of
the generalized Dix equation 5.40,

(5.43)

Figure 5.17: Plan view of


the source and receiver posi-
tions (small circles) in a single
superbin. The superbin contains
approximately 400 source-receiver
pairs with a common-midpoint
scatter of about 80 m or 2% of the
maximum offset (after Grechka
and Tsvankin, 1999b).

Figure 5.18: Semblance curves


obtained by the conventional
velocity analysis, which ignores
the azimuthal dependence of
moveout velocity (dashed) and
by the azimuthal velocity analy-
sis (solid). Arrows indicate the
major reflection evens (after
Grechka and Tsvankin, 1999b).

105

12815-OTE III Gretchka.indd 105 28-07-2009 10:14:14


Chapter 5

was used. Applying this equation to the effective NMO ellipses in Figure 5.19 results in the interval
ellipses, whose eccentricities are shown in Figure 5.20. The interval ellipses inherit the N-S orienta-
tion of the effective ones at s because the ellipses of the shallow events are close to circles.
Also, comparing the E scales in Figures 5.19 and 5.20 reveals that the interval ellipses have higher
eccentricities, as one would expect. The maximum E value reaches 0.08 in Figure 5.20.

5.4.1 Further developments


To conclude this chapter, we mention a useful extension of the 3D Dix formula that resulted in a
suite of anisotropic parameter-estimation techniques collectively known as stacking-velocity tomogra-
phy. Grechka and Tsvankin (2002a) introduced the concept of the NMO-velocity surfaces that allowed

Figure 5.19: Effective NMO eccentricities at (a) s and (b) s (after Grechka and
Tsvankin, 1999b).

Figure 5.20: Interval NMO eccentricities for


the horizon between 2.14 s and 2.57 s (after
Grechka and Tsvankin, 1999b).

106

12815-OTE III Gretchka.indd 106 28-07-2009 10:14:14


Normal moveout

them to remove the requirement of lateral homogeneity of the overburden and generalized for-
mula 5.40 to more realistic, laterally heterogeneous subsurface structures with dipping and curved
intermediate interfaces. The key feature of this extension is that it models the entire wide-azimuth
moveout at conventional offsets by tracing a single zero-offset ray per common midpoint and per
reflector, thus, resulting in a computationally efficient implementation of anisotropic inversion. In
addition, analytic insights gained from analysis of the NMO-velocity surfaces make it possible to
understand which anisotropic parameters are constrained for various subsurface geometries and
therefore, should be targeted in the inversion. Here, we list the most significant results obtained
with this technique.
Le Stunff et al. (2001) demonstrate that under certain circumstances the P-wave reflection data
can be inverted for all three relevant VTI parameters and , making depth processing
in VTI media at least theoretically feasible. Grechka et al. (2002a) extend the finding of Le Stunff
et al. (2001) to a broader class of VTI media composed of homogeneous layers and explain when
and what kind of a priori information is needed to ensure the uniqueness of anisotropic param-
eter estimation.
Grechka et al. (2002b) apply stacking-velocity tomography to multicomponent, wide-azimuth
data in VTI and tilted TI media. They prove that supplementing reflected P-waves with shear-
waves or mode conversions (PS) yields more stable estimates of the anisotropy coefficients and,
in many cases, helps to constrain the reflector depth.
Grechka et al. (2005) examine the feasibility of building orthorhombic velocity models from mul-
ticomponent reflection data. The uniqueness of parameter estimation strongly depends on the
reflector dips and orientations. Numerical tests employing multicomponent stacking-velocity
tomography for layered orthorhombic media show that a priori constraints on the vertical veloci-
ties are required to avoid instabilities in the inversion. Those constraints, however, are no longer
needed for a special case of orthotropy induced by vertical fractures in otherwise isotropic host
rock (Chapter 8; Vasconcelos and Grechka, 2007).

107

12815-OTE III Gretchka.indd 107 28-07-2009 10:14:15


12815-OTE III Gretchka.indd 108 28-07-2009 10:14:15
NONHYPERBOLIC
6 Nonhyperbolic moveoutMOVEOUT

Although the results presented in Chapter 5 might seem impressive, their area of applicability
is inherently limited to conventional offsets, which are defined to be approximately equal to the
reflector depth. On the other hand, the industry has long recognized the value of acquiring seismic
reflection data at longer spread lengths. Motivations for recording such data include, for instance,
an increased signal-to-noise ratio because of a greater fold, an improved illumination of the subsur-
face targets due to a larger angular aperture of the reflected rays and a more reliable characteriza-
tion of lithologies and fluids through application of large-angle AVO. Each of those tasks requires
an accurate moveout correction. However, if the subsurface is anisotropic and/or heterogeneous,
the hyperbolic moveout equation loses its accuracy with increasing offsets. This fact necessitates the
development of more sophisticated moveout equations that could be successfully used at offset-to-
depth ratios significantly greater than unity.
This chapter is devoted to such nonhyperbolic (NH) moveout corrections. We explicitly focus
on seismic anisotropy as the underlying cause of the moveout nonhyperbolicity and refer readers
to Fomel and Grechka (2001), for a review of other possible causes. We also restrict our analysis
in this chapter to P-waves because they are the subject of the vast majority of practical nonhyper-
bolic-moveout applications. Our discussion will concentrate on two issues: (i) whether a given NH
moveout equation flattens reflection moveout at long offsets and (ii) whether information contained
in long-spread data allows robust estimation of the relevant anisotropy parameters. We will explain
why virtually all available NH moveout approximations, for example, those proposed by Dellinger
et al. (1993), Tsvankin and Thomsen (1994), Alkhalifah and Tsvankin (1995), Siliqi and Bousqu
(2000), Zhang and Uren (2001), Fomel (2004), Ursin and Stovas (2006) and Douma and Calvert
(2006) successfully deal with the first issue but none of them provides a fully satisfactory solution
to the second.

6.1 NH moveout of P-waves in a single VTI layer


Our discussion in this chapter naturally starts from the Taylor series expansion 5.1 that expresses the
pure-mode reflection traveltime, , as a function of the offset, . We write the series in the form

(6.1)

where is the zero-offset time, is the normal-moveout velocity,

(6.2)

is the quartic moveout coefficient, which was already introduced by the third equation 5.2, and dots
represent the neglected higher-order terms in offset.
Hake et al. (1984) observed that contains contributions of both layering and anisotropy in
stratified VTI media. Equation 6.1 itself, however, becomes inaccurate at large offsets even in a
single anisotropic layer if . This is clear from the fact that given by 6.1 behaves as

109

12815-OTE III Gretchka.indd 109 28-07-2009 10:14:15


Chapter 6

(6.3)

at large , whereas the correct limit should obviously be

(6.4)

where is the horizontal velocity in a single VTI layer. Tsvankin and Thomsen (1994) noticed this
discrepancy and modified equation 6.1 to

(6.5)

where the quantity

(6.6)

ensures the correct asymptotic behavior of .

The next essential step was made by Alkhalifah and Tsvankin (1995), who suggested to ignore
the contribution of the shear-wave vertical velocity, , to the exact quartic moveout coefficient, ,
of P-waves. A justification for this approximation is presented in Figure 4.3. The final result of
Alkhalifah and Tsvankin (1995) reads

(6.7)

Its key significance for processing of long-spread seismic data is that only a single parameter the
anellipticity coefficient in addition to needs to be estimated to flatten the reflection travel-
times at arbitrarily large offsets.
Equation 6.7 carries another important message. As NH moveout of P-waves is governed solely
by the anellipticity coefficient, the

P-wave reflection moveout is purely hyperbolic in elliptically anisotropic VTI layer,

where (equation 4.27). This conclusion turns out to be exact even though we inferred it from
approximation 6.7.

The relative traveltime errors incurred by applying approximation 6.7 in a single VTI layer
usually amount to a few percent. The presence of those errors was the main motivation for numer-
ous efforts to improve this approximation (e.g., Siliqi and Bousqu, 2000; Zhang and Uren, 2001;
Fomel, 2004; Ursin and Stovas, 2006). Aleixo and Schleicher (2009) show, however, that some of
those approximations are, in fact, worse than equation 6.7. Besides, according to E. Blias (pers.
comm.), it is possible to find such pairs of and that make any of the above referenced
approximations slightly more accurate than all its competitors.

110

12815-OTE III Gretchka.indd 110 28-07-2009 10:14:15


Nonhyperbolic moveout

In fact, the issue of making the NH moveout approximations exact or nearly exact is far less
important than one might think. There are two reasons for this.
1. In practice, these approximations are applied with fitted rather than the (unknown) actual
parameters. Fitting a NH moveout approximation to P-wave reflection traveltimes typically
produces traveltime errors of less than 1%, which are definitely acceptable for practical needs.
Although such a fit may result in somewhat biased estimates of and in a single VTI layer,
this bias is often much smaller than other uncertainties that will be discussed below.
2. Small-scale vertical and lateral heterogeneity, which cannot adequately be sampled with seis-
mic data, makes the issue of sub-percent accuracy of moveout approximations unimportant in
practice. Usually, a geophysicist is satisfied when no observable residual moveout is present. All
aforementioned NH moveout approximations achieve this goal.

6.1.1 Inversion of nonhyperbolic moveout


The simplicity of equation 6.7 prompted a number of geophysicists to believe that can reliably
be estimated from P-wave nonhyperbolic moveout. Indeed, one might mute long-offset data and
estimate . Adding large offsets would then yield because

(6.8)

and, according to equation 6.4,

(6.9)

Unfortunately, this straightforward logic fails to deliver satisfactory results. Its failure ultimately
relates to the absence of ideal, laterally homogeneous plane layers in the subsurface and infinite
offsets in real reflection data. The latter is reflected in the usage of the term long-spread in the
industry: it usually refers to the maximum offset to reflector-depth ratios (ODR) of about two or
a bit greater. Although substantially larger ODR are available for shallow horizons, reflections at
those ODR are often post-critical and interfere with either head or diving waves. In addition, lateral
heterogeneity or a slight dip of the layers reduces the accuracy of equation 6.7 and all other similar
approximations. As a result, typical ODR for the NH moveout analysis lie between two and three.
It is clear that both and have to be obtained simultaneously at such offset-to-depth ratios. In
practice, this is done via a two-dimensional semblance scan (e.g., Taner and Koehler, 1969) in vari-
ables and or and (Alkhalifah, 1997; Grechka and Tsvankin, 1998a).
Another issue then gains significance. Estimation of two unknowns, and , from one equation
6.7 is only possible because of the difference in the offset dependencies of the hyperbolic and non-
hyperbolic moveout terms. This difference happens to be relatively mild, though. Figure 6.1 illus-
trates this. The two dependencies are semi-parallel rather than orthogonal, as would be preferable
for a stable parameter estimation. Other existing NH moveout equations exhibit the same feature.
Consequently, we might expect a strong tradeoff between and . The analysis of equation 6.7
reveals that a small velocity variation can be compensated by a variation that has a
larger magnitude and the opposite sign. The differences in the signs of changes in and and
relationship 6.8 imply that more robust results of the NH moveout analysis could be obtained in
terms of the variables rather than . The corresponding moveout equation is
derived by substituting from equation 6.8 into 6.7:

111

12815-OTE III Gretchka.indd 111 28-07-2009 10:14:15


Chapter 6

Figure 6.1: Hyperbolic (solid) and nonhyper-


bolic (dashed) moveout terms normalized by
their values at the offset-to-depth ratio equal
to two. Equation 6.7 with the parameters
, was used
for this computation.

(6.10)

Equations 6.7 and 6.10 are fully equivalent to each other. As such, they are both just approxima-
tions to the exact P-wave reflection traveltimes. Grechka and Tsvankin (1998a) proposed to improve
the accuracy of these approximations at ODR between 2 and 3 by sacrificing the correct asymptotic
behavior of equations 6.7 and 6.10 at infinity. They changed equation 6.10 to

(6.11)

where the coefficient for typical anellipticities . Douma and van der Baan (2008)
showed that the value of needs to be modified when and, predictably, when ODR .
The performance of equation 6.11 for nonhyperbolic moveout analysis in a single-layer VTI
model is displayed in Figure 6.2. The coordinates of the semblance maximum are close to the
actual parameters and ; therefore, a sufficiently accurate estimate of is obtained as well.
The shape of the semblance contours in Figures 6.2b and 6.2c corroborates our earlier analysis of
the tradeoffs between the scanned variables. We observe that and can vary simultaneously
along the diagonal ridge in the semblance contours without producing significant changes in the
values of semblance (Figure 6.2b). This ridge narrows and becomes longer in the coordi-
nates (Figure 6.2c), indicating that is more poorly constrained than either or . The higher
values of are compensated by lower values of (Figure 6.2b) and the absolute percentage
inaccuracies in and sum up and propagate into in accordance with equation 6.8.
Adding mild correlated noise to the data supports the established instability of estimation
from NH moveout. For instance, Figure 6.3 is generated for a linear error function that is equal to
(+4 ms) at zero offset and to ( 4 ms) at the end of the spread length. Such an error mitigates the
traveltime increase with offset and makes the effective NMO velocity greater. To provide a good fit
at large offsets, the horizontal velocity has to be smaller than its model value and the semblance

112

12815-OTE III Gretchka.indd 112 28-07-2009 10:14:16


Nonhyperbolic moveout

Figure 6.2: (a) Synthetic seismogram of the P-wave reflected from the bottom of a VTI layer described
by parameters , and . The spread length is equal to two
reflector depths; the source pulse is a Ricker wavelet with central frequency of 40 Hz. (b) Semblance
contours in the coordinates calculated with equation 6.11 for . (c) Same as (b) but
plotted in the coordinates (after Grechka and Tsvankin, 1998a).

analysis yields and . Those velocities correspond to , which


is about one-half of its actual value. The results of semblance analysis in Figure 6.3 confirm the
conclusion that follows from Figure 6.2: the horizontal velocity is as tightly constrained as by
reflection moveout, provided that the spread length is about twice the reflector depth. However,
long-period traveltime errors, which can be considered insignificant in the practice of seismic data
processing, may cause inaccuracies in the inverted reaching . These errors are entirely due to
the tradeoff between and on long-spread gathers.
We illustrate the presented error study with a walkaway VSP data set acquired by the Reservoir
Characterization Project of the Colorado School of Mines at Vacuum Field, located in New Mexico,
USA. The data, recorded by a downhole receiver placed at a depth of about 1 km, were excited at
616 surface source locations with offsets ranging from 60 m to 1,370 m from the well. The dots in
Figure 6.4 display the picked traveltimes of the first P-wave arrivals that can be used as substitutes
for the reflection traveltimes from a horizontal interface at the receiver depth. The data points plot-
ted in the squared coordinates in Figure 6.4 cluster around a curve rather than a straight line,
indicating the presence of substantial nonhyperbolic moveout. Using equation 6.11, we can find a
set of moveout parameters that gives the best least-squares fit to the observed traveltimes. The cor-

113

12815-OTE III Gretchka.indd 113 28-07-2009 10:14:16


Chapter 6

Figure 6.3: Semblance contours for the model


from Figure 6.2 after addition of linear travel-
time error that changes from +4 ms at zero
offset to at the offset . The
semblance maximum corresponds to
(after Grechka and Tsvankin, 1998a).

Figure 6.4: Traveltimes for different source


positions (dots) recorded by a single downhole
receiver and the best-fit nonhyperbolic moveout
curve (solid) computed using equation 6.11.
The offsets and traveltimes are doubled to simu-
late a reflection experiment (after Grechka and
Tsvankin, 1998a).

Figure 6.5: RMS time residuals (in ms)


calculated with equation 6.11 for different pairs
. The residual for the best-fit model
at the center of the contours is 6.7 ms (after
Grechka and Tsvankin, 1998a).

114

12815-OTE III Gretchka.indd 114 28-07-2009 10:14:17


Nonhyperbolic moveout

responding moveout is plotted as the solid curve in Figure 6.4. Note that the largest offset-to-depth
ratio in this data set is over 2.7 (the offset should be doubled for comparison with reflection data),
which is quite favorable for nonhyperbolic moveout analysis. The inverted values of the NMO and
horizontal velocities are
Next, we analyze the root-mean-square (RMS) time residuals for different pairs of and
in Figure 6.5. These residuals look very similar to the semblance contours displayed in Figures 6.2b
and 6.3. The center of the contours in Figure 6.5 corresponds to a residual of 6.7 ms. For all models
with and lying inside the nearest contour line, the RMS residuals are equal to 8 ms or
less. We might view this contour as the boundary of a family of anisotropic models that deviate
from the best-fit model by less than 1.3 ms. The existence of this family of models is in agree-
ment with the modeling results discussed above, while the large magnitudes of the time residuals
are caused by a sizeable scatter in the traveltime measurements. If we estimate the error bars on
by picking its extreme values corresponding to the 8 ms contour line in Figure 6.5 (again, 8 ms
is close to the minimum RMS residual plus one time sample), we obtain a wide range of values:
.
The known receiver depth makes it possible to calculate the average vertical velocity
from the zero-offset traveltime and, using equation 5.8, find Thomsens .
Finally, combining the obtained values of and yields (of course, the error bars on
depend on the uncertainties in and ). Those anisotropy coefficients should be regarded
as the effective values that absorb the contribution of the subsurface heterogeneity. Still, they are
indicative of non-negligible anisotropy in the predominantly carbonate section above the reservoir
in the Vacuum Field.
Perhaps the most important conclusion that we can draw from our discussion is the existence
of a family of VTI models with very similar P-wave nonhyperbolic moveout functions measured at
the offset-to-depth ratios between two and three. The presence of those models manifests itself in
the tradeoff between and . The structure of the NH moveout equation 6.11 implies that
variations in and within the model family have comparable magnitudes but the opposite
signs. As a result, the corresponding absolute error in the parameter is about as large as the sum
of the relative changes in and . The discussed walkaway VSP, which might be perceived as
an ideal testing ground for the nonhyperbolic moveout inversion, produced values spanning the
interval . Clearly, such a broad range of the estimated precludes one from using
those results for lithology discrimination.

6.2 NH moveout in stratified VTI media


Despite significant errors in effective values estimated from nonhyperbolic moveout, a Dix-
type differentiation procedure resulting in the interval has been developed (e.g., Tsvankin and
Thomsen, 1994; Alkhalifah, 1997; Tsvankin, 2001). We refrain from reproducing the correspond-
ing math here because the equations are lengthy and the expected uncertainties in the interval
coefficients are generally greater than those in the effective ones. Nevertheless, it is instructive to
discuss a case study that illustrates the error amplification caused by the transition from the effec-
tive to interval quantities.
Figure 6.6 shows a section from offshore Africa dominated by subhorizontal reflections. While
their large number facilitates estimation of the interval and , Figure 6.7 indicates that the
accuracy of velocity picking at times exceeding 2 s suffers from a low resolution of the semblance
maxima. The values of and estimated from the semblance peaks in Figure 6.7, as well as
those for other reflections (not shown), yield the interval NMO velocity and curves (black) dis-
played in Figure 6.8. The gray curves in Figure 6.8 correspond to the upper and lower limits of

115

12815-OTE III Gretchka.indd 115 28-07-2009 10:14:17


Chapter 6

Figure 6.6: Time-migrated seis-


mic line (after Alkhalifah, 1997).

Figure 6.7: Semblance analysis for


CMP 300 in Figure 6.6 at different
zero-offset times . Gray lines cor-
respond to contours of the effective .
The maximum offset in the data is
4.3 km (after Alkhalifah, 1997).

the parameter values inferred from the uncertainties in picking of and within the black
regions in Figure 6.7.
The resolution of in Figure 6.7 noticeably decreases with the zero-offset time because
of the reduction in the offset-to-depth ratio. Correspondingly, the errors in the interval quickly
grow with , and the estimated values become useless at times greater than . In addition,
as with any other Dix-type layer-stripping procedure, the interval parameters in a deeper section
are compromised by errors incurred in shallower horizons.

116

12815-OTE III Gretchka.indd 116 28-07-2009 10:14:18


Nonhyperbolic moveout

Figure 6.8: Interval and (black) as func-


tions of at CMP 300 in Figure 6.6. The gray
curves show the error bars caused by the picking
uncertainties (after Alkhalifah, 1997).

6.3 Azimuthal variation of quartic moveout


The analysis presented above is likely to break down for long-spread, wide-azimuth reflection data
because either lateral heterogeneity or azimuthal anisotropy in the overburden, reflector dip and
curvature, or any combination of these factors make nonhyperbolic moveout azimuthally variable.
A general analytic treatment of the NH moveout in 3D was presented by Pech et al. (2003). Here,
we mention only its most important points.
Pech et al. (2003) base their development on the NH moveout equation 6.5. They realize that
maintaining the exact asymptotic behavior of the reflection traveltimes at infinite offsets is not criti-
cal for obtaining a good-quality, practically useful moveout approximation (compare equations 6.10
and 6.11) and concentrate their efforts on deriving the exact azimuthal dependence of the quartic
moveout coefficient . The result of Pech et al. (2003) can be written in the form

(6.12)

where is the azimuth and the five coefficients absorb the velocity heterogeneity and azimuthal
anisotropy of the overburden as well as the reflector dip and curvature.
It is clear from equation 6.12 that unlike NMO velocity, whose azimuthal variation is typically
elliptical (Figure 5.13), the dependence of the quartic moveout coefficient on azimuth is much more
complicated. We illustrate the behavior of with several examples.

6.3.1 Dipping VTI layer


The weak-anisotropy approximation for the P-wave quartic coefficient in a homogeneous dipping
VTI layer reads (Pech et al., 2003)

(6.13)

117

12815-OTE III Gretchka.indd 117 28-07-2009 10:14:18


Chapter 6

where is the reflector dip and is the azimuth from the dip plane. For a vertical reflector
, regardless of the azimuth of the CMP line because reflected rays are confined
to the horizontal isotropy plane, where the velocity is constant and the moveout is purely hyper-
bolic. If the reflector is horizontal , the model as a whole is azimuthally isotropic and
, which is consistent with equation 6.7.

For a dipping reflector, the coefficient vanishes in azimuthal directions satisfying

(6.14)

If the dip is smaller than 30, equation 6.14 has no solution and the sign of is opposite to that
of in all azimuthal directions (Figure 6.9a). For a dip of 30, goes to zero only at a single
azimuth that corresponds to the dip plane (Figure 6.9b). This analytic result is supported
by the numerical study of Tsvankin (1995, 2001), who showed that the P-wave dip-line moveout
approaches a hyperbola for reflector dips close to 30.

Figure 6.9: Magnitude of the azimuthally-varying quartic coefficient computed using


equation 6.13 for a VTI layer above a planar reflector with the dip , and
. The arrows mark the azimuth of the reflector dip plane (after Pech et al., 2003).

118

12815-OTE III Gretchka.indd 118 28-07-2009 10:14:19


Nonhyperbolic moveout

For dips in the range , equation 6.14 predicts two azimuths for which
. If the dip is equal to 45, the quartic coefficient vanishes at (Figure 6.9c). The
sign of coincides with sign( ) near the dip plane and is opposite to sign( ) in the vicinity of
the strike direction.

6.3.2 Dipping HTI layer


In a homogeneous dipping HTI layer with the symmetry axis oriented in the reflector strike direc-
tion, the WAA for the P-wave quartic moveout coefficient is given by (Grechka and Pech, 2006)

(6.15)

where is the azimuth from the reflector strike plane. Equation 6.15 indicates that
(i) nonhyperbolic moveout vanishes in the reflector dip plane and
(ii) is independent of the dip at any azimuth .
The first result is exact. It can readily be understood if we recall that the dip plane coincides with
the isotropy plane in this HTI model. Equation 6.15 correctly states that, in accord-
ance with the known fact of the absence of NH moveout in homogeneous isotropic media above a
planar reflector.
Observation (ii) holds only in the weak anisotropy limit. Nevertheless, it predicts the exact azi-
muthal dependence for a horizontal reflector, where (Al-Dajani and Tsvankin,
1998).

6.3.3 Horizontal layer of orthorhombic and lower symmetries


Comparing the WAAs for the P-wave quartic moveout coefficients in horizontal layers of monoclinic
and triclinic symmetries, Grechka and Pech (2006) found an unexpected equality:

(6.16)

It indicates that

asymmetry with respect to the horizontal does not influence the NH moveout of P-waves reflected from
horizontal interfaces.

The conclusion expressed by equation 6.16 is similar to that drawn by and


(1998) for the P-wave reflection coefficients from small-contrast interfaces in azimuthally anisotropic
media characterized by weak anisotropy. and (1998) also discovered that only the
monoclinic components of the stiffness tensor are important for the P-wave reflectivity; deviations
of from those in monoclinic media with a horizontal symmetry plane (equation 2.35) do not
matter.
For a horizontal orthorhombic layer, the P-wave quartic moveout coefficient reads (Al-Dajani et
al., 1998; Pech and Tsvankin, 2004; Grechka and Pech, 2006)

(6.17)

where is the azimuth from the symmetry plane and the coefficients

119

12815-OTE III Gretchka.indd 119 28-07-2009 10:14:19


Chapter 6

(6.18)

(6.19)

and

(6.20)

generalize the Alkhalifah-Tsvankin (1995) anellipticity coefficient (equation 4.21) for orthorhombic
media. The anisotropy coefficients and are given by equations 4.37, 4.40 and 4.38, 4.41, 4.43,
respectively. As follows from equation 6.17, it is possible for to change sign and vanish
along up to four distinct azimuthal directions.

6.4 Layered orthorhombic media


Although extension of nonhyperbolic moveout analysis to stratified azimuthally anisotropic media
faces the same stability problem as the one discussed in section 6.2, the azimuthal variation of
NH moveout might be detected more reliably than the anellipticity coefficients averaged over all
azimuthal directions. The feasibility of such
an approach was demonstrated by Vasconcelos
and Tsvankin (2006). They applied azimuthal
NH moveout inversion to wide-azimuth P-wave
data acquired at Weyburn Field located in
Saskatchewan, Canada, where the presence of
azimuthal anisotropy at the reservoir level has
been established from P-wave NMO ellipses,
azimuthal AVO analysis ( Jenner, 2001; Jenner
et al., 2001) and shear-wave splitting (Cardona,
2002). A contribution of the nonhyperbolic
moveout inversion, therefore, could be in iden-
tifying the vertical variations of the anellipticity
parameters (equations 6.18
6.20) obtained under the assumption of effective
orthotropy.
Vasconcelos and Tsvankin (2006) applied 3D
Figure 6.10: Source-receiver geometry for a nonhyperbolic moveout inversion to the P-wave
superbin used for azimuthal NH moveout analy- data collected into superbins, one of which
sis. Each blue dot marks a source-receiver pair; is shown in Figure 6.10. The NMO-corrected
the polar angle corresponds to the source-receiv- gather in Figure 6.11 confirms the presence of
er azimuth from the north (see the numbers on NH moveout. The apparent hockey sticks are
the perimeter), whereas the radius is the offset. indicative of a positive effective , whereas their
The total fold for this superbin is 2,491 (after jitter implies non-negligible azimuthal varia-
Vasconcelos and Tsvankin, 2006). tions of the reflection traveltimes. The inver-
sion was applied to four reflected events in the

120

12815-OTE III Gretchka.indd 120 28-07-2009 10:14:20


Nonhyperbolic moveout

Figure 6.12: Results of nonhyperbolic


moveout inversion for the reflections from
Figure 6.11: Seismic data at superbin in the Viking horizon (the maximum offset-
Figure 6.10 after a conventional hyperbolic to-depth ratio is 2.5), the Blairmore
NMO correction. The gather is sorted by (ODR = 2.0), the Lower Vanguard (ODR
increasing offset and contains all azimuths, = 1.9) and the Mississippian Unconformity
which are irregularly sampled (Figure 6.10). (ODR = 1.8). The arrows mark the esti-
Nonhyperbolic moveout manifests itself mated direction of the semi-major axis
through the curvature of the NMO-corrected of the NMO ellipse. The number by each
events at long offsets (so-called hockey arrow is the azimuth of the axis with
sticks). The jittery character of the reflections respect to the north. All parameters are
suggests the presence of traveltime varia- the effective values for a given reflection
tions with azimuth (after Vasconcelos and event (after Vasconcelos and Tsvankin,
Tsvankin, 2006). 2006).

overburden above the Weyburn reservoir (Figure 6.12) because the recorded spread lengths were
insufficient for accurate estimation of the quartic moveout term for deeper reflections.
The values of the anellipticity coefficients in Figure 6.12 indicate that anisotropy is quite sub-
stantial through the entire overburden. Although Vasconcelos and Tsvankin (2006) have not derived
the interval anisotropic parameters from the effective ones (possibly because of error-amplification
issues), a general increase in the magnitude of the effective coefficients with depth suggests that the
deeper portion of the section is more anisotropic than the overburden.
To verify whether azimuthal anisotropy could be ignored, Vasconcelos and Tsvankin (2006) ran
the inversion algorithm for VTI media (section 6.2) and found that the semblance values for the

121

12815-OTE III Gretchka.indd 121 28-07-2009 10:14:20


Chapter 6

best-fit VTI models were on average 10% smaller than those for orthorhombic media. Since a 10%
gain in semblance is not negligible (see Figure 5.18 and the related discussion), accounting for the
azimuthal variation of reflection moveout is justified by the data.

To conclude this section, we mention two alternative approaches that circumvent the amplification of
errors in the interval quantities by exploiting (the presumed) lateral homogeneity of the overburden.

Hake (1986) and van der Baan and Kendall (2002) propose to carry out the parameter estima-
tion in the domain. The fact that the horizontal slowness components are preserved along
rays propagating through a stack of horizontal layers allows isolating the influence of each layer
and leads to an inversion scheme capable of retrieving the transforms of the local slowness
surfaces in 3D or the slowness curves in 2D. Either can be inverted individually, as if no overbur-
den existed.
Another alternative is the so-called velocity-independent layer-stripping method published by
Dewangan and Tsvankin (2006) and Wang and Tsvankin (2009). It is based on matching the
slopes of reflections from the top and the bottom (which does not have to be horizontal) of
a target horizon and results in a straightforward procedure that removes the influence of the
overburden, leaving only the propagation times in the target layer itself. The latter can be used
as the input to any appropriate single-layer parameter-estimation technique.

122

12815-OTE III Gretchka.indd 122 28-07-2009 10:14:21


7 ANISOTROPY
7 Anisotropy ESTIMATION
estimation from VSP data FROM VSP DATA

So far, our discussion has concentrated on the issues of estimating seismic anisotropy from surface
reflection data. We found Thomsen notation and its extensions to be especially useful for this pur-
pose because they allowed us to capture the combinations of the stiffness coefficients responsible
for such commonly measured seismic signatures as the NMO velocities and NH moveout. Although
substantial progress has been made, we noticed that the estimated parameters typically pertained to
relatively thick subsurface intervals and instabilities associated with traveltime inversion precluded
us from deriving anisotropy on a finer scale. In this chapter, we examine another type of seismic
acquisition geometry the vertical seismic profiling (VSP) and demonstrate its capability of pro-
viding more localized estimates of the in-situ anisotropy.
We begin the chapter with an overview of various approaches for inverting seismic anisotropy
from VSP measurements. Our analysis will show that a sufficiently wide angular data coverage at
downhole geophone locations is the most important requirement for any successful VSP inversion.
However, the details of which parameters can be estimated and how accurately depend on the over-
burden complexity. For a laterally homogeneous overburden, obtaining a complete (that is, triclinic)
local stiffness tensor from wide-azimuth, multicomponent VSP might be feasible. This is no longer
the case in the presence of strong lateral heterogeneity, which usually makes shear-wave arrivals too
noisy to be used in the inversion. Consequently, anisotropy has to be inferred from P-waves only,
leading to the introduction of Thomsen-style parameters that govern the P-wave VSP signatures.

7.1 Laterally homogeneous overburden


The term vertical seismic profiling refers to recordings of elastic waves excited at the surface by
geophones located in a borehole. VSP as a tool for seismic exploration and development origi-
nated in the Soviet Union in the late 1950s. The fundamentals of the VSP method, as well as its
early applications, are covered in the book by Galperin (1971). After the English translation of
Galperins book appeared in 1974, VSP studies took off in the west.

Figure 7.1: First-break P-wave times


(in ms) recorded by a geophone placed
at a depth of 4,509 ft (1,374 m) in a
vertical well at Rulison Field, Colorado,
USA. The data were acquired by the
Reservoir Characterization Project,
Colorado School of Mines (after
Grechka et al., 2007).

123

12815-OTE III Gretchka.indd 123 28-07-2009 10:14:21


Chapter 7

The simplest option for estimating the effective anisotropy is to use the first-break times
measured by a downhole geophone. Although those times might result in reasonable estimates of
effective anisotropy (see Figures 6.4, 6.5 and the corresponding discussion), the overburden het-
erogeneity often creates distortions in the measured traveltimes that cannot be explained by any
homogeneous anisotropic model. An example of such distortions is presented in Figure 7.1. We
observe not only a shift of the traveltime minimum to the south-east from the geophone location
but also a sharp traveltime change at some 2,000 ft (610 m) to the west from the well. The reason
for this traveltime step is the presence of a fault, which is also recognizable in seismic reflection data
(Franco et al., 2007; Grechka et al., 2007).
While a homogeneous anisotropic model that does not possess a horizontal symmetry plane
could be used to fit the shift of the traveltime minimum (which is likely caused by lateral hetero-
geneity rather than by anisotropy), no homogeneous model yields a sufficiently rapid change of
the P-wave velocity to explain the traveltime behavior due to the velocity contrast across the fault.
Thus, Figure 7.1 demonstrates the importance of taking into account subsurface heterogeneity
when estimating anisotropy. Failure to do so would produce unrealistic or even implausible anisot-
ropy parameters, whose values are expected to be significantly compromised by the unaccounted
heterogeneity. On the other hand, if our goal it just to image seismic reflection data, the traveltime
surface in Figure 7.1 might be helpful because it provides the exact depth-migration operator for
the geophone location. Whether or not this operator is applicable to adjacent locations depends on
the strength of lateral and vertical velocity variations.

7.1.1 Slowness method


The issue of velocity heterogeneity as it pertains to inferring anisotropy from VSP data has been
recognized since the earliest attempts to estimate anisotropy. Historically, the first assumption made
about the overburden was its stratification or lateral homogeneity. It led to the technique called the
slowness method (Gaiser, 1990; Miller and Spencer, 1994; Jlek et al., 2003).
Figures 7.2 and 7.3 explain how the slowness method works. Let us suppose that we measure
the traveltimes, t, of waves excited at a number of surface shot locations by an array of geophones
placed in a vertical well with the depth increment (Figure 7.2). If we sort the VSP data in com-
mon-shot gathers, the vertical slowness, , at the geophone locations is according to the
definition given by equation 3.6.

Figure 7.2: Measurements carried out for estimating


anisotropy in a typical VSP geometry. The traveltime
difference, , between geophones (dots) located at
a distance along a wellbore defines the appar-
ent slowness, . Three-component traces
recorded by each downhole geophone yield the
direction of particle motion, , defined by the polar
polarization angle, , and azimuth, . The azimuths
of the horizontal components and of the
geophones usually vary along the tool string and
should be treated as unknown unless independently
measured (after Grechka et al., 2007).

124

12815-OTE III Gretchka.indd 124 28-07-2009 10:14:21


Anisotropy estimation from VSP data

Next, we resort the recorded traces into com-


mon-receiver gathers and find the derivatives
of the same traveltimes, t, with respect to the
source coordinates . In accordance
with definition 3.6, are the hori-
zontal slowness components measured at the
Earths surface. To bring these down to the
geophone depths, we rely on the assumption
of lateral homogeneity and apply Snells law. It
states that and remain unchanged along
each ray. Hence, we obtain the slowness surface
for 3D source geometries (or the slow-
ness curve in 2D) at the downhole geophone loca-
tions. Once the surface is constructed, Figure 7.3: P-wave slowness surface (in s/km)
the density-normalized stiffness tensor, , for a constructed from 3D VSP data (after
chosen anisotropic symmetry can be computed Dewangan and Grechka, 2003).
by solving the Christoffel equation 2.80. Its
solution leads to a nonlinear inverse problem
for the stiffness components .
Figure 7.3 displays such a slowness surface
that was obtained from a 3D walkaway VSP data
set acquired by the Reservoir Characterization
Project at Vacuum Field, New Mexico, USA.
The data were recorded by a string of 10 three-
component (3C) geophones placed in a verti-
cal well between the depths of 304.8 m and
439.8 m with increments . Vibrators
at 250 shot points uniformly distributed around
the borehole with the offsets reaching 510 m
from the well were used as seismic sources. We
note that the slowness contours in Figure 7.3
deviate from circles, implying that an azimuth- Figure 7.4: Horizontal projections of the
ally anisotropic model is required to fit the P-wave polarization vectors (tick lengths
measurements. are proportional to the magnitudes of the
projections) plotted at the source locations
7.1.2 Slowness-polarization method with respect to the well, which is placed at
The particle motions of waves recorded by geo- the coordinate origin (after Dewangan and
phones in a well comprise an inherent part of Grechka, 2003).
any VSP data set. If individual body waves can
be extracted from the obtained waveforms, we
should be able to measure the body-wave polarization vectors, (inset in Figure 7.2). Experience
shows that this is feasible, at least for first-break P-waves, when heterogeneity in the vicinity of the
geophones is not extreme. For example, Figure 7.4 displays the vectors corresponding to the
P-wave slowness surface in Figure 7.3. As both the slowness vectors, , and the polarization vectors,
, pertain to the same downhole geophone locations, they can be used together to estimate the
anisotropy of a rock volume around the geophones that has a linear size approximately equal to
the seismic wavelength.

125

12815-OTE III Gretchka.indd 125 28-07-2009 10:14:21


Chapter 7

Combining the vectors and leads to the so-called slowness-polarization technique for measuring
anisotropy. Using this method, we can find the density-normalized stiffness tensor, , directly from the
Christoffel equation 2.54. It results in a linear inverse problem of minimizing the functions

(7.1)

for all source-receiver pairs in terms of the stiffness coefficients . If all three body waves (P, fast
shear and slow shear ) are available and the data coverage at geophone locations approaches
a full hemisphere, such as that shown in Figure 7.3, no a priori assumption about the symmetry of
tensor is needed, and a local triclinic stiffness tensor can be obtained (Dewangan and Grechka,
2003). If we operate with P-waves only, which might be the case in many applications, equations 7.1
constrain up to 15 combinations of the 21 triclinic stiffness coefficients (Zheng and Penck, 2002).

7.2 Laterally heterogeneous overburden


In a laterally heterogeneous subsurface, the horizontal slowness components and calculated
at the surface are no longer preserved along rays and should be treated as unknown at the geo-
phone locations. While, in principle, they can be recovered along with the stiffnesses from equation
7.1, resulting in a nonlinear inverse problem, Dewangan and Grechka (2003) showed that such an
increase in the number of unknowns causes a significant deterioration in the accuracy of the esti-
mated stiffness coefficients and renders their values almost useless.
An alternative approach to inferring local VTI parameters beneath a laterally heterogeneous
overburden was proposed by White et al. (1983) and further developed by de Parscau (1991) and
Hsu et al. (1991). These authors combine the apparent slownesses, , obtained by differentiating the
arrival times along a vertical borehole and the polarization vectors, , recorded by 3C geophones
in a wellbore. As a result, the horizontal slowness components, which might vary laterally, never
enter the inversion, making the slowness-polarization technique theoretically applicable to subsur-
face of any complexity. In fact, this implementation of the slowness-polarization method rests on
the only assumption that the recorded body waves are locally plane.
Both de Parscau (1991) and Hsu et al. (1991) analyzed which parameters of VTI media can be
estimated from the vertical slownesses and polarizations. They conclude that four density-normal-
ized stiffness coefficients and or, equivalently, two vertical velocities , and
two Thomsen anisotropy coefficients , can accurately be determined from a combination of
P- and SV-wave VSP data when the propagation angles of the P- and SV-waves range from 0 to 80
from the vertical. In addition, de Parscau (1991) and Hsu et al. (1991) note that neither P nor SV
data alone constrain the four quantities in a unique fashion. This theoretical conclu-
sion was later confirmed by Horne and Leaney (2000).
The necessity of relying on both P- and SV-waves to ensure the uniqueness of VTI parameter esti-
mation is unfortunate because, in general, it requires identifying SV arrivals in the data and separating
them from SH-waves. Such an event identification and separation is especially problematic in a typical
scenario when the azimuths of geophones in a vertical well are not measured independently (Figure 7.2).
Another issue arises if the local symmetry happens to be lower than TI and, strictly speaking, there are
no SV- and SH-waves (Chapter 2). Enforcing the SV-wave properties on either as would
be required by the inversion algorithm, will then bias the estimated quantities. Thus, it appears that
estimation of anisotropy from VSPs acquired beneath complicated velocity structures (e.g., subsalt) would
benefit from avoiding shear waves. If this can be done successfully, we might expect a higher accuracy
of the inversion because (i) shear data are usually much noisier than P and (ii) there will be no need to
deal with the uncertainty of determining the type of shear mode. There is an important operational

126

12815-OTE III Gretchka.indd 126 28-07-2009 10:14:22


Anisotropy estimation from VSP data

benefit too implementation of the inversion becomes more straightforward because only direct,
easy-to-pick P-wave arrivals are used.

7.2.1 Inversion of P-wave VSP data for VTI parameters


To perform the inversion based on P-waves alone, we need to determine the combinations of VTI
parameters constrained by the slowness-of-polarization dependence of P-wave VSP data. The first
attempt to address this question was made by Williamson and Maocec (2001). They noticed that
the ratio of the vertical velocities, , had to be known a priori to resolve the Thomsen coef-
ficients and individually. Acknowledging this conclusion, Grechka and Mateeva (2007) proposed
to reparameterize the problem in order to absorb into the quantities uniquely constrained
by the data. Their result is best understood from the weak-anisotropy approximation of the P-wave
vertical slowness, , as a function of the P-wave polar polarization angle, (Figure 7.2). The WAA
derived by Grechka and Mateeva (2007) reads

(7.2)

The newly introduced anisotropy coefficients and are

(7.3)

(7.4)

and

(7.5)

where is defined by equation 4.18.

Equation 7.2 is similar to the WAA 4.20 for the P-wave phase velocity. Based on this similarity, we
might expect that the pairs and play comparable roles for P-wave VSPs acquired
along vertical boreholes in VTI media and for processing of P-wave surface reflection data, respec-
tively. Indeed, a closer look at equation 7.2 reveals that is responsible for the near-vertical
behavior of , while governs the vertical slowness at larger angles. Importantly, and
incorporate the shear-wave velocity, , rendering its value unnecessary for fitting the P-wave
slowness-of-polarization functions.
The relative sensitivities of to and at different polarization angles are also appar-
ent from equations 7.2 7.5. To gain this analytic insight, let us introduce the ratio of the squared
vertical velocities

(7.6)

127

12815-OTE III Gretchka.indd 127 28-07-2009 10:14:22


Chapter 7

For a typical velocity ratio , equation 7.6 yields . Consequently,

(7.7)

is a reasonable approximation. Substituting it into equations 7.2 7.5 yields

(7.8)

and demonstrates that and have the prefactors and , respectively.


Therefore, the -term dominates the -term in equation 7.8 for polarization angles 25 with
the vertical. At 25, the influence of on is still relatively minor because the quantity
and is mostly determined by the vertical velocity .
Like approximation 4.20, which allowed us to eliminate from the parameters governing
the P-wave kinematics in strongly anisotropic VTI media, WAA 7.2 suggests that a reduced set of
quantities might be sufficient to fit P-wave VSP data. Figure 7.5 confirms this.
It demonstrates that the exact curves (shown with symbols) nearly overlap for a family of VTI
models with fixed , , and whose Thomsen anisotropy coefficients vary from moderate
and to uncommonly large and (not typos). Clearly, these vastly
different models cannot be distinguished from the function. Even though the WAA (dashed
line in Figure 7.5) slightly deviates from the exact , it has accomplished its most important goal
to identify the parameter combinations governing the influence of anisotropy on .
Grechka and Mateeva (2007) applied the described technique to estimate anisotropy from a
P-wave VSP data set acquired in the deepwater Gulf of Mexico. Figure 7.6 displays the geometry of
the 2D walkaway VSP survey. The data comprise 677 shot locations spaced at 100 ft (30.48 m) along
a line, with a maximum offset of 36,660 ft (11,174 m) from the wellhead. A receiver tool with 24
3C geophones, 100 ft (30.48 m) apart, was used to cover the 17,500 22,250 ft (5,334 6,782 m)
depth interval. As the geophone azimuths in the borehole were not measured and the sources were
located along a single line, the local azimuthal anisotropy could not be resolved and the medium
had to be treated as VTI.

Figure 7.5: Exact func-


tions for various ratios
but fixed parameters

and (symbols). The


dashed line shows the WAA of
given by equation 7.2 (after
Grechka and Mateeva, 2007).

128

12815-OTE III Gretchka.indd 128 28-07-2009 10:14:22


Anisotropy estimation from VSP data

Figure 7.6: Geometry of walkaway VSP (after Grechka and Mateeva, 2007). Surface shot locations are
shown in red. Positions of 3C receivers in the borehole are marked in cyan. Depth-migrated surface
seismic data are displayed on the background of the isotropic P-wave depth-velocity model (white
color corresponds to the P-wave velocity in water, magenta to the velocity in salt).

The slowness and polarization picks made from the data were subjected to the criteria below to
ensure that the picks correspond to locally plane body waves. To enter the inversion, each slowness-
polarization pair had to possess the following:

1. Linearity of particle motion


Body waves are polarized linearly in homogeneous anisotropic media at sufficiently large dis-
tances from seismic sources. The linearity of their polarization can be verified directly from the
recorded waveforms. For instance, Figure 7.7 displays particle motion for a data point accepted
as an input for the inversion.
2. Locally linear moveout
A plane wave is characterized by a constant slowness vector. Consequently, the traveltimes of
a wave arriving at geophones located in a small depth interval are supposed to be linear in a
common-shot gather. This property was used to both calculate the apparent slowness, , and to
reject the data exhibiting too much scatter from a straight line. Figure 7.8a shows the traveltimes
(dots) picked for seven geophones on a common shot gather and their linear fit (solid line). The
dashed lines indicate one standard deviation from the best-fit straight line.
3. Small scatter of polarization angles
Not only the slowness but also the polarization vector is constant for a plane body wave.
Therefore, the polarization angles, , should not vary significantly in a common-shot gather for
a selected geophone interval. Figure 7.8b displays the measured polarization angles. In practice,
they are never constant because of noise; for instance, the standard deviation shown with dashed
lines in Figure 7.8b is about 1.

129

12815-OTE III Gretchka.indd 129 28-07-2009 10:14:23


Chapter 7

Figure 7.9 gives an example of typical data and estimated parameters both in the salt and below
it. Clearly, the salt turns out to be virtually isotropic: all data points (black dots with crosses) in Figure
7.9a tightly cluster around the isotropic slowness curve (red solid line). Perhaps
more importantly, the salt isotropy is estimated with high confidence, which comes primarily from
a nearly ideal data coverage. Figure 7.9a demonstrates that P-wave polarization angles reach 75
from the vertical leading to precise estimates of both and . We observe that the projection of
the model uncertainties (open circles) onto the data space overlaps with that of the best-fit VTI model
(solid circles) and with the isotropic dependence.

Figure 7.7: Particle-motion hodo-


gram for a typical source-receiver
pair used in the inversion. The
geophone axis is vertical; the
axes and are horizontal
but their azimuths are unknown.
Open circles indicate the picked
first-break time, dots mark the
particle motions at 2 ms time
increments. The hodogram cor-
responds to approximately one
quarter of the dominant period
of P-waves (after Grechka and
Mateeva, 2007).

Figure 7.8: P-wave traveltimes (a)


and polarization angles (b). The
dots indicate the quantities picked
from several traces in a common-
shot gather. The solid lines are the
linear traveltime fit (a) and the
mean polarization angle (b). They
give the slowness-polarization pair,
, which is one data point for
the inversion. The dashed lines
correspond to one standard
deviation from the best-fit lines.
These standard deviations are
0.25 ms for the times and 1 for
the polarization angles (after
Grechka and Mateeva, 2007).

130

12815-OTE III Gretchka.indd 130 28-07-2009 10:14:23


Anisotropy estimation from VSP data

Figure 7.9b illustrating parameter estimation beneath the salt tells quite a different story. First,
the subsalt sediments are noticeably anisotropic. The presence of anisotropy, especially that of nega-
tive (the corresponding ), is undeniable in Figure 7.9b. We clearly see that
all data points fall below the isotropic curve (red) at polarization angles . A negative
is needed to fit these data. Second, the
angular aperture of data in Figure 7.9b is not as
wide as that in Figure 7.9a. The reduction in the
aperture is caused by the strong velocity contrast
at the base of salt (compare the values in
Figures 7.9a and 7.9b), which bends seismic rays
toward the vertical beneath the salt. As implied
by this ray behavior, the P-wave slowness-polar-
ization inversion yields larger standard devia-
tions of both anisotropy coefficients and
compared to those in the salt.
Figure 7.10 presents the parameter-estima-
tion results for the entire data set. The anisot-
ropy is displayed in terms of the conventional
coefficients and computed by supplement-
ing the VSP coefficients and with the
S-to-P-wave velocity ratios. Those ratios were
derived from sonic logs averaged over depth
intervals of 600 ft (183 m), which were used to
infer anisotropy from the VSP. The parameters
near the salt bottom could not be obtained
because the interference of various reflected
and transmitted waves there violates the key
assumption of a single, locally plane body wave.
The velocity and anisotropy estimates displayed
in Figure 7.10 fall into two distinct groups cor-
responding to salt and sediments.
The salt turns out to be nearly isotropic by
seismic standards (Figure 7.10b). This conclu- Figure 7.9: Slowness-polarization data and
sion, which follows from the inversion, is note- estimated parameters (a) in the salt at a
worthy because individual salt (NaCl) crystals depth of 18,500 ft (5,639 m) and (b) beneath
are known to be anisotropic (they have cubic the salt at a depth of 21,750 ft (6,629 m).
symmetry, e.g., Sun, 1994). Perhaps the simplest Crosses associated with each data point
way to reconcile this apparent discrepancy is (black dot) indicate the standard deviations
to suppose that the orientations of salt crystals in the picked and values. The solid
are random at the examined VSP location and circles correspond to the best-fit VTI
recognize that the presented estimates pertain functions, open circles to those functions
to the seismic wavelength, which is about 300 m one standard deviation. The obtained
in the salt. P-wave vertical velocities in SI units are (a)
The anellipticity coefficient is mostly and (b)
negative for the subsalt sediments The solid red lines show the
(Figure 7.10b). Although negative have been isotropic dependencies implied by those
measured in a lab (for instance, Thomsen, 1986, (after Grechka and Mateeva, 2007).
cites more than a dozen of shale and sandstone

131

12815-OTE III Gretchka.indd 131 28-07-2009 10:14:24


Chapter 7

Figure 7.10: Vertical velocities (a) and anisotropy coefficients (b). The thin solid lines in (b) are the
standard deviations of and estimated from the uncertainties in the and picks (after Grechka
and Mateeva, 2007).

Figure 7.11: Anisotropy coefficient (a) and gamma-ray log for subsalt sediments (b). Thin lines
in (a) show the standard deviation of (after Grechka and Mateeva, 2007).

132

12815-OTE III Gretchka.indd 132 28-07-2009 10:14:24


Anisotropy estimation from VSP data

cores that exhibit ), such values are seldom observed in seismic data. Perhaps the known
anomalous stress regime beneath the salt might provide a possible explanation for the negative .
The depth dependence of in Figure 7.10b inherits that of in Figure 7.11a. Both dependencies
clearly correlate with lithology marked on the gamma-ray log in Figure 7.11b.

7.2.2 Inversion for parameters of azimuthal anisotropy


The VTI anisotropy-estimation procedure can be extended to lower symmetries. While a general
solution for weak anisotropy can be found in Grechka and Mateeva (2007), here we only discuss
their results for orthotropy. The parameters that control the P-wave slowness-of-polarization
dependence are similar to Tsvankins (1997b) parameters for orthorhombic media (see section 4.3).
This similarity can be observed by comparing the P-wave phase-velocity equation 4.46 with WAA for
the vertical P-wave slowness, , expressed in terms of the polar and azimuthal P-wave polarization
angles, and , respectively,

(7.9)

Here

(7.10)

(7.11)

are Tsvankins s introduced by equations 4.38 and 4.41, are the anellipticity coefficients
defined by equations 6.18 6.20 and quantity is given by equation 7.5. The -term in parenthe-
ses on the first line of 7.9 is easily recognizable as the azimuthally varying (equation 4.47) scaled
by ( 1). Similarly, the -term in parentheses on the second line of 7.9 is the azimuthally varying
[multiplied by (2 1)] that governs NH moveout of P-waves (equation 6.17). Hence, essentially
the same coefficients are responsible for both kinematic signatures and the slowness-of-polarization
dependence of P-waves in orthorhombic media.

Inspection of equations 7.9 7.11 reveals the following:


1. Function constrains up to five anisotropy coefficients in
orthorhombic media. Knowledge of the ratio of the vertical velocities, , is not needed to
uniquely determine those coefficients from wide-azimuth P-wave VSP data that have sufficiently
good polar coverage.
2. The velocity ratio is required only to make a transition from
to the set of more conventional parameters .

We illustrate these findings with the Rulison data set that was used at the beginning of this
chapter (Figure 7.1). The vertical slowness for the best-fit orthorhombic model corrected for
isotropy, that is, , is displayed in Figure 7.12. We have chosen to subtract the
isotropic slowness-of-polarization dependence, , to enhance the observed
azimuthal variation. The azimuthal dependence of the vertical slowness is apparent at near-
vertical propagation (or at small horizontal components of polarization) and becomes less obvi-
ous away from the vertical. The parameters of the best-fit orthorhombic model and their stand-

133

12815-OTE III Gretchka.indd 133 28-07-2009 10:14:25


Chapter 7

Figure 7.12: Azimuthal varia-


tion of the fitted vertical slow-
ness (in s/kft) corrected for
isotropy as a function of the
P-wave horizontal polarization
components for the best-fit
orthorhombic model in the
depth range 4,510 ft 4,910 ft
(1,375 m 1,497 m). The white
circles indicate the slowness var-
iations expected in the absence
of azimuthal anisotropy (after
Grechka et al., 2007).

ard deviations are:


It is instructive to note a sizable differ-
ence between and , which stands out despite the significant uncertainties in both coeffi-
cients and the approximate equalities and for the estimated s. These values of
the anisotropy parameters are consistent with the presence of gas-filled vertical fractures in a VTI
host rock (Grechka, 2007), which are plausible at Rulison.
We conclude this chapter by saying that while shear waves recorded in VSP geometries are
certainly sensitive to anisotropy, the structural complexity of the overburden might make them
too noisy for adding useful information to the anisotropy inversion. Even when this is the case,
Thomsen-type anisotropy coefficients specifically tailored to VSP applications can be still obtained
from generally less noise-prone first-break P-wave data. Working with P-waves is also attractive
because they make it possible to estimate local velocity and anisotropy without knowledge of the
properties of the overburden.

134

12815-OTE III Gretchka.indd 134 28-07-2009 10:14:25


8 FRACTURE
8 Fracture CHARACTERIZATION
characterization

The final chapter of the book is devoted to seismic signatures of naturally fractured rocks. Unlike the
previous chapters where we discussed estimation of seismic anisotropy without going into details of its
underlying physical causes, here we explicitly assume that the anisotropy is crack-induced. The moti-
vation for emphasizing this particular reason for anisotropy among a number of others (see section
1.4) is purely practical. Fractured formations are known to contain substantial hydrocarbon resources.
Production of hydrocarbons from these formations, whose matrix permeability often lies in the low
microdarcy range, is facilitated by the presence of networks of natural cracks that might provide high-
ly permeable pathways for the fluid flow. Given the importance of fractures for production, reservoir
development is greatly aided by any information about the cracks obtained from seismic data.
Perhaps this is why the 2004 Summer Research Workshop of the Society of Exploration Geophysics
(SEG) and the 2007 Workshop of the European Association of Geoscientists and Engineers (EAGE)(10)
were held to determine the state of the art of seismic characterization of fractured reservoirs. These
workshops demonstrated the following:
1. Current understanding of the influence of small-scale fractures on seismic is based on the so-
called effective media theories that aim at replacing a microheterogeneous rock containing the
cracks with a homogeneous one that has exactly the same overall (or effective) elastic properties
in the regime of static deformation.
2. The existing industry methodologies pertaining to fracture characterization from seismic data
are limited to a single set of penny-shaped, vertical cracks embedded in an otherwise isotropic
host rock. This model is described by effective horizontal transverse isotropy, whose anisotropy
coefficients can be estimated from various seismic signatures (e.g., NMO ellipses, azimuthal
AVO, shear-wave splitting) and related back to the fractures.
While the choice of static effective media schemes for the theoretical framework reflects the fact of
significant difference in size between the seismic wavelengths and the lengths of
fractures that contribute to production, assumption of the presence of only a sin-
gle set of cracks seems inadequate because it contradicts numerous observations of multiple fracture
sets at outcrops around the globe (e.g., Figure 1.11) and in borehole televiewer data acquired in a
variety of geologic environments (Laubach et al., 2004; Narr et al., 2006).
In this chapter, we will address this issue. To this end, we begin the chapter with an overview of
effective media theories for cracked solids containing multiple, differently-oriented fracture sets.
It will naturally lead us to:
(i) analysis of the accuracy of two popular effective media schemes proposed by Schoenberg
(1980) and Hudson (1980);
(ii) identification of the geometric features of fractures that cannot be inferred from the effective
properties; and
(iii) determining which fracture properties are constrained by seismic data.
We conclude the chapter with a case study of seismic characterization of multiple fracture sets.

10 Several papers presented at the Workshop are published in the special issue of Geophysical Prospecting
(2009, 57, No. 2).

135

12815-OTE III Gretchka.indd 135 28-07-2009 10:14:26


Chapter 8

8.1 Effective media theories


8.1.1 Basic concepts
A standard departure point for discussion of the mechanics of microheterogeneous media (e.g.,
Nemat-Nasser and Hori, 1999; Milton, 2002) is introduction of the representative volume element
(RVE) that we denote . On an intuitive level, the RVE can be understood as follows. Suppose we
take a piece of rock that has a linear size and volume . It usually contains small-scale heteroge-
neities, such as pores, cracks and individual grains, whose characteristic lengths are (e.g., Figure
1.10) and whose elastic properties can be described by the spatially variable stiffness tensor ,
where is the vector of Cartesian coordinates. As there are many heterogeneities in our piece of
rock, it is reasonable to assume that . The rock itself, however, appears statistically homogeneous
(see Ostoja-Starzewski, 2008, for a mathematical definition of this term and a detailed discussion),
in the sense that the displacement field of a long seismic wave that has the wavelength
is nearly constant in . Furthermore, this large-scale field remains unchanged if we replace our
particular rock sample with a different one that has the same average elastic properties or, more
precisely, the same effective stiffness tensor, . Thus, we can say that this homogenizes the original
stiffness tensor .
Consequently, the first goal of the homogenization or effective media theories is to derive from
. The second, more practical goal is to determine what information about can be inferred
from the knowledge of the full tensor or certain combinations of its components measured, for
example, from seismic data. We will discuss both issues here.
The effective stiffness tensor, , of a heterogeneous (for instance, fractured) solid relates the
stress, , and the strain, , tensors averaged over the RVE via Hookes law (equation 2.19)

(8.1)

Although Hookes law 8.1 is applicable to both static and dynamic elastic processes (Chapter 2), we
use it in the static regime in this chapter. This corresponds to probing the RVE with infinitely long
waves or to assuming that .

Equation 8.1 can be represented in an equivalent form,

(8.2)

in terms of the compliance tensor defined as the inverse of the stiffness tensor,

(8.3)

Formulation 8.2 turns out to be more appropriate for fractured solids because cracks are sources
of extra strains. This can be made explicit by splitting into the background, , and the fracture-
related, , compliances,

(8.4)

and rewriting equation 8.2 as

(8.5)

136

12815-OTE III Gretchka.indd 136 28-07-2009 10:14:26


Fracture characterization

where

(8.6)

is the extra strain due to cracks.

The strain can be expressed through physically transparent quantities the displacement
discontinuities across crack surfaces, . For flat cracks, this results in (Vavakin and Salganik, 1975)

(8.7)

where is the displacement-discontinuity vector averaged over the crack area, is


the unit normal to the fracture face, and the sum is taken over all cracks in volume . Obtaining
the effective compliance is thus reduced to finding the vectors .

8.1.2 Non-interaction approximation


Let us observe that our discussion in the previous section was fairly general. All we did was to define
a crack as a flat inhomogeneity and rely on the linearity of Hookes law. At this point, however,
we need to make certain assumptions to proceed further. Their necessity stems from the fact that
the displacement-discontinuity vector at the -th fracture is a function of the local tractions
applied to its faces that depend on the positions and orientations of all adjacent cracks because
their presence might create complicated local stress fields (see, for example, Figures 8.3 and 8.4
below). Therefore, strictly speaking, complete microstructural information is required to account
for the contribution of each individual fracture to the effective elasticity. Such detailed informa-
tion, however, is unavailable in the majority of applications. To circumvent this difficulty, we make
the important assumption that interactions in the stress fields of different fractures can be ignored
and each crack senses the far-field stress . This assumption leads to the so-called non-interaction
approximation (NIA). Bristow (1960) was the first to apply it to cracked media.
The NIA is not the only available option and other approximations have been proposed to
account for fracture interactions. They typically assume non-interacting cracks to be placed into
either the effective elastic matrix or effective stress field. The self-consistent (OConnell and
Budiansky, 1974) and differential (Vavakin and Salganik, 1975) schemes put the cracks into the
effective matrix; the latter does this in increments. The Mori and Tanaka (1973) scheme, reformu-
lated for cracks by Benveniste (1986), places non-interacting fractures into the average stress field
that remains unchanged by the cracks. Perhaps the most advanced extension of the Mori-Tanakas
ideas is called the method of effective field (Levin et al., 2004; Levin and Markov, 2005). This meth-
od can account for the statistics of the crack positions but does not always yield explicit results.
We prefer the NIA to all other effective media theories because, being the simplest scheme, it
accurately describes the influence of thin fractures (but not round pores) on the overall elasticity
(e.g., Figure 8.5 below). In the NIA, the displacement-discontinuity vector is given in terms of
the symmetric, second-rank excess fracture-compliance tensor (Schoenberg, 1980; Kachanov, 1992).
The latter relates to the uniform traction (equation 2.17) induced at the crack face by
the remotely applied stress ,

(8.8)

137

12815-OTE III Gretchka.indd 137 28-07-2009 10:14:26


Chapter 8

The eigenvectors of the symmetric, positive definite tensor are the principal compliance direc-
tions of a flat crack that has an arbitrary shape. For a purely isotropic host material, one of the eigen-
vectors coincides with the crack normal, , while the other two lie in the crack plane (Sevostianov
and Kachanov, 2002). The eigenvalue, , of the tensor corresponding to the eigenvector is
called the normal crack compliance; the other two eigenvalues, , are known as the shear or
tangential crack compliances.

8.1.3 Scalar cracks


The simplest and yet most important special case is when the -tensor is proportional to the unit tensor,

(8.9)

where is a scalar and is the Kronecker delta. Schoenberg and Sayers (1995) call such cracks
scalar. All excess compliances of scalar cracks are equal, .
The scalar nature of the excess fracture compliance tensor has far-reaching consequences. To
understand them, we substitute equations 8.8 and 8.9 into equation 8.7 to obtain

(8.10)

Comparison of equations 8.6 and 8.10 helps to identify the quantity

as the proper crack-density tensor. It is symmetric and has the second rank; therefore, the symmetry of an
originally isotropic material containing scalar fractures cannot be lower than that of the crack-density
tensor itself. This observation allowed Kachanov (1980) to draw an important conclusion:

a purely isotropic solid with any orientation distribution of scalar cracks is orthorhombic.

Also, the crack-induced orthotropy is always elliptical and characterized by fewer than nine inde-
pendent constants that determine general orthorhombic media (see the stiffness matrix 2.38).

8.1.4 Dry penny-shaped cracks


For dry circular (or penny-shaped) cracks, the proportionality 8.9 is satisfied only approximately
with the accuracy dependent on the background Poissons ratio . This is clear from the eigenvalues
of the tensor (Kachanov, 1992; 1993),

(8.11)

138

12815-OTE III Gretchka.indd 138 28-07-2009 10:14:26


Fracture characterization

where is the Youngs modulus of the host and is the crack radius. The normal, , and tangen-
tial, , excess fracture compliances are relatively close to each other because the (usually positive)
Poissons ratio satisfies the inequality . The difference between leads to two
fracture-related tensors (Kachanov, 1980)

(8.12)

and

(8.13)

The second-rank crack-density tensor and the fourth-rank tensor contain all information
about the crack distribution over orientations and sizes relevant for the effective properties in the
non-interaction approximation. Tensor can be viewed as a natural tensorial extension of the scalar
crack density,

(8.14)

introduced by Bristow (1960). Remarkably, the aspect ratios of cracks (defined as the ratios of
the fracture widths and their diameters) do not enter equations 8.12 and 8.13 implying that the
effective properties of solids with dry fractures are almost independent of (provided that .
Consequently, for thin cracks,

the crack-related porosity has virtually no effect on the effective elasticity.

According to equations 8.11, dry circular cracks become scalar (in the terminology of Schoenberg
and Sayers, 1995) only when the background has a zero Poissons ratio . Although in reality ,
the influence of the tensor on the effective properties is still relatively minor because its mag-
nitude is significantly smaller than that of due to the presence of the multiplier (equation
8.13) that cannot be greater than 1/4. Hence, neglecting and retaining as the sole crack-density
parameter constitutes a reasonable approximation. Thus, we are back to the conclusion drawn for
the (unrealistic) scalar fractures:

a solid with arbitrarily oriented circular cracks is nearly orthorhombic.

The numerical studies of Grechka and Kachanov (2006a; 2006b) confirm this statement.
The possibility of describing the effective elasticity in terms of just results in the following
important properties of the crack-induced anisotropy:
The overall influence of multiple, differently oriented dry fracture sets is indistinguishable from
that of three orthogonal or principal sets.
The normals to those principal sets coincide with the eigenvectors of the tensor (equation
8.12). The corresponding principal crack densities are the eigenvalues of .

139

12815-OTE III Gretchka.indd 139 28-07-2009 10:14:27


Chapter 8

The crack-induced orthotropy is elliptical. Furthermore, it is controlled by only four independent


quantities [the combinations of , and three eigenvalues of s; see Kachanov (1980; 1993) for
details] rather then nine needed for general orthotropy (equation 2.38).

The eigenvalues 8.11 of the excess fracture-compliance tensor lead to the following expression
for (Kachanov, 1980; Sayers and Kachanov, 1995; Schoenberg and Sayers, 1995):

(8.15)

Either ignoring tensor in this equation or approximating its components as


recovers the effective elliptical orthotropy.

8.1.5 Liquid-filled fractures


The influence of liquid infill on the overall compliance was first examined by OConnell and
Budiansky (1974) and Budiansky and OConnell (1976) for cracks with identical aspect ratios.
Shafiro and Kachanov (1997) extended their analysis to fractures that have variable aspect ratios
and generalized equation 8.15 to

(8.16)

where

(8.17)

The dimensionless parameters,

(8.18)

are called the fluid factors. Their magnitudes are governed by the bulk modulus of the fluid, , and
the crack aspect ratios, .
The structure of equations 8.17 and 8.18 provides important clues for understanding the
influence of fluids on the effective elasticity. Let us note the minus sign in front of the second
term in equation 8.17. It implies that the presence of fluids stiffens the fractures compared to
dry ones, exactly as our intuition would tell us. Next, the fluid factors given by equation 8.18
are bounded by . Near-zero mean dry fractures , while indi-
cate either thin cracks or a relatively stiff fluid infill, or both. Increasing the aspect
ratios of cracks with such a stiff infill reduces the corresponding fluid factors. Those fat frac-
tures soften the rock and make the liquid-filled cracks look similar to dry ones, thus, pointing
to the importance of the aspect ratios for the effective elasticity. As a direct consequence, the
crack-density tensor alone is no longer sufficient for describing the effective properties; the

140

12815-OTE III Gretchka.indd 140 28-07-2009 10:14:27


Fracture characterization

additional (and essential) microstructural parameters are captured by the fourth-rank tensor
(equation 8.17).
Analysis of the magnitude of the tensor reveals another apparent complication brought by the
presence of fluids in cracks. Equations 8.12 and 8.17 indicate that the norms of tensors and are
comparable when the fluid factors are not small. Therefore, tensor cannot be ignored similarly to
tensor , and, in principle, the effective symmetry might be lower than orthorhombic. Nevertheless,
the deviations from orthotropy remain small because of the stiffening effect of fluids: the tensor ,
acting in some sense opposite to , reduces the overall influence of fractures. The net result is that
the magnitude of crack-induced anisotropy decreases and orthotropy still holds because of the
proximity of the effective elasticity to isotropy (Grechka and Kachanov, 2006a).

8.1.6 Pore-like cracks


For completeness, we mention an extension of the non-interaction approximation to ellipsoidal
pore-like fractures whose aspect ratios are not necessarily small. Their contribution, , to the effec-
tive compliance (equation 8.4) is given in terms of Eshelbys (1957) tensor, , as (Walpole, 1969;
Kachanov et al., 2003)

(8.19)

where is the crack porosity (the fraction of volume occupied by the crack), is the compli-
ance tensor of (generally anisotropic) infill material, is the background stiffness tensor and
is the fourth-rank identity tensor. The colon in equation 8.19 denotes
a double dot product (a contraction over two indexes) defined as .
While we use expression 8.19 below to illustrate the influence of nonzero crack aspect ratios
on the effective properties, we caution the reader that the NIA loses its accuracy for fat cracks.
Hence, we do not recommend relying on equation 8.19 for cracks that have aspect ratios .

8.1.7 Linear-slip theory


The non-interaction approximation and the linear-slip theory of Schoenberg (1980) have identical
forms in terms of the -tensors for dry fractures. The difference is that the former yields the effec-
tive elastic constants in terms of geometric crack-density parameters (at least, for circular cracks),
whereas the latter lacks a direct link to the microstructure (Schoenberg, 1980; Schoenberg and
Sayers, 1995).
For liquid-filled fractures, the linear-slip differs from that given by equation 8.16. Schoenberg
and Douma (1988) obtained the linear-slip results by inverting Hudsons (1981) effective stiffnesses.
We do not reproduce Schoenberg and Doumas (1988) equations here because the quantitative dif-
ference between them and equation 8.16 is small due to the overall weak influence of liquid-filled
fractures on the effective properties.

8.1.8 Hudsons theory


In contrast to the non-interaction approximation, which yields compliances as linear functions of
the crack density, , Hudson (1980) focuses on the effective stiffnesses, . He constructs them as a
power series with respect to , truncating the series after either the linear (the first-order theory) or
quadratic term (the second-order approximation).

141

12815-OTE III Gretchka.indd 141 28-07-2009 10:14:27


Chapter 8

The first-order theory of Hudson (1980; 1981) has the form

(8.20)

It represents the so-called dilute limit, which is essentially the NIA for compliances being inverted
and linearized with respect to the crack density. For a single set of penny-shaped fractures that has
the crack density and the normal in equation 8.20 is given by

(8.21)

where are the Lam coefficients of the host rock. The quantities are (Hudson,
1980; 1981; Peacock and Hudson, 1990)

(8.22)

(8.23)

(8.24)

Here are the P- and S-wave velocities of the background, are the Lam
parameters of the isotropic infill and all cracks are assumed to have the same aspect ratio . If
several differently oriented fracture sets are present, their stiffness contributions, , are simply
summed (Hudson, 1981),

(8.25)

The substitution 8.25 is insensitive to the spatial distribution of fractures, as it should be in the non-
interaction approximation.
The second-order Hudsons theory (1980; 1991) extends the linear approximation 8.20 by add-
ing the term quadratic in the crack density:

(8.26)

142

12815-OTE III Gretchka.indd 142 28-07-2009 10:14:28


Fracture characterization

where

(8.27)

(8.28)

and is given by equation 8.24. Note that the second-order term in equation 8.27 is constructed
from the first-order term (equation 8.21) without bringing in any additional information about the
fractures.

8.2 Comparison of theoretical predictions


It is easy to realize that the linear-slip and Hudsons first- and second-order theories predict differ-
ent effective properties. Before we compare them quantitatively, it is instructive to gain a qualita-
tive insight into which theory is expected to be more accurate. The easiest way to achieve this is to
assume the presence of a single fracture set with the normal along the coordinate axis and use
Hudsons (1980) power series expansion, for instance, of the effective stiffness coefficient ,

(8.29)

where dots denote the higher-order terms in .

Since the presence of cracks reduces the stiffness, has to decrease. As the linear term
dominates at sufficiently small ,

(8.30)

Thus, truncating series 8.29 after the linear term in accordance with the first-order Hudsons (1980)
approximation inevitably results in an incorrect negative at some crack density . This crack
density, where the first-order theory of Hudson starts violating the stability conditions and its pre-
dictions become physically implausible, can be obtained from equations 8.20 8.24 [see also equation
21 in Hudson (1981) or equations 20a and 24b in Liu et al. (2000)]. For dry fractures, these equa-
tions result in the inequalities

(8.31)

Similar analysis utilizing matrix 8.21 yields for :

(8.32)

Both inequalities indicate that Hudsons first-order scheme encounters problems for small
ratios or equivalently, for large Poissons ratios . Hence, equations 8.31 and 8.32 demonstrate that

143

12815-OTE III Gretchka.indd 143 28-07-2009 10:14:28


Chapter 8

Figure 8.1: Effective stiffness coefficients and for a single set of dry cracks. The
background velocities are , and density is ; they
yield the Lam coefficients . Symbols indicate different theoretical
predictions: the first-order Hudsons (equations 8.20 and 8.21), the second-order Hudsons
(equations 8.21 and 8.26 8.28), Schoenbergs (equations 8.4, 8.12, 8.13, and 8.15), and
the NIA (equations 8.4 and 8.19), which takes into account nonzero crack aspect ratios ( for
all fractures). The bars correspond to the 95% confidence intervals (the mean values two standard
deviations) of the numerically computed stiffness coefficients obtained for 100 random realizations of
the fracture locations (after Grechka and Kachanov, 2006c).

Figure 8.2: Anisotropy coefficients of effective HTI media. The symbols


are the same as those in Figure 8.1 (after Grechka and Kachanov, 2006c).

144

12815-OTE III Gretchka.indd 144 28-07-2009 10:14:28


Fracture characterization

the first-order Hudsons theory breaks down for any nonzero crack density of dry fractures in the limit
.

On the other hand, the first-order Hudsons predictions for the shear moduli , and
are known to be better.
Hudsons (1980) second-order theory yields a positive coefficient of the quadratic term in expansion
8.29. Therefore, begins to increase at some also exhibiting an unphysical behavior. In contrast,
the NIA (e.g., Bristow, 1960; Schoenberg, 1980) which ignores elastic interactions between the cracks
and sums their contributions to the effective compliance, predicts the effective in the form

(8.33)

where is a positive coefficient. Therefore, is always a positive, monotonically decreasing func-


tion of , as it should be.

Figure 8.1 confirms the presented qualitative analysis. We see that the first-order Hudsons
theory (marked with ) yields the unphysical negative and at the crack densities
greater than 0.07 and 0.12, respectively. The second-order theory of Hudson (marked with in
Figure 8.1) results in obviously incorrect, monotonically growing .
Such behavior implies that adding fractures stiffens rather than softens the rock. This theory leads
to another equally unphysical prediction: and exceed their background values
, thus indicating that a solid containing fractures is stiffer
than the uncracked matrix. In fact, the tendency of the second-order Hudsons theory to produce the
unreasonably high effective stiffnesses displayed in Figure 8.1 has been known for quite some time,
both for a single fracture set and for randomly-oriented cracks (Sayers and Kachanov, 1991; Cheng,
1993). In contrast, the predictions of the linear slip theory ( ) and the NIA ( ) shown in Figure 8.1
are plausible. They almost coincide with each other, indicating a weak influence of the crack aspect
ratio on the effective properties, and are close to the results of numerical modeling (bars;
see the next section).
Additional insights into the behavior of effective elastic properties can be gained by examining
the HTI anisotropy coefficients in Figure 8.2. These coefficients were introduced
in section 4.3, where we examined horizontal transverse isotropy as a special case of orthotropy.
Figures 8.2a and 8.2b do not display the first-order Hudsons predictions for because
they fall out of the range of values on the plots. Let us make the following observations:
The linear slip theory ( ) indicates that , therefore, an approximate elliptical anisot-
ropy is expected.
It has been pointed out in many papers (e.g., see Bakulin et al., 2000, for review) that the shear-
wave splitting coefficient is close to the crack density, . Figure 8.2c makes it clear that
this conclusion is mainly based on Hudsons theory . While equality is not
supported by the linear slip theory ( ) at large crack densities, it might be viewed as a reason-
able approximation when .

Overall, we conclude that the linear slip theory (or the non-interaction approximation in com-
pliances) is superior to either first- or second-order Hudsons schemes for a single set of dry penny-
shaped cracks.

145

12815-OTE III Gretchka.indd 145 28-07-2009 10:14:29


Chapter 8

8.3 Numerical modeling of effective elasticity


We now discuss the computations that produced the bars shown in Figures 8.1 and 8.2. They were
obtained by performing static, finite-element modeling on so-called digital rocks. The modeling
methodology is described in Zohdi and Wriggers (2005). It amounts to specifying a desired micro-
structure (such as that in Figure 8.3) and numerically computing the local stress, , and strain,
, tensors that correspond to six linearly independent boundary conditions. Then Hookes law
8.1, written for the components of the volume-averaged and , is treated as a system of
linear equations for the unknown effective stiffness components . There is no need to know or
assume the effective symmetry a priori because it can be inferred from the solution.

8.3.1 Fracture interactions


The numerical solution, however, always depends on details of the mutual positions of individual
fractures that cause scatter in the effective parameters for any nonzero crack density. The reason
for the scatter is the interaction of the stress fields of the adjacent cracks. Figure 8.3 shows the local
behavior of the stress component . When the locations of the crack centers (which are supposed to
be random and uncorrelated for the effective parameters to make sense) vary in , the patterns of
interactions change, introducing variations in the numerically computed effective stiffness tensors ,
shown as bars in Figures 8.1 and 8.2. Such variations are inevitable for any finite number of cracks (e.g.,
Zohdi and Wriggers, 2001). The magnitudes of these variations and their dependence on the applied
boundary conditions allow one to determine the proximity of a computational volume to the RVE.
Huet (1990) and Ostoja-Starzewski (2008) explain how this proximity and the bounds on the effective
elasticity for a given crack array can be derived based on the results of numerical modeling.

Figure 8.3: Horizontal cross-section of the stress component through a model containing dry
fractures (wire spheroids). The crack density is . The arrow indicates the direction of applied
remote load, whose magnitude is 1 MPa (after Grechka and Kachanov, 2006c).

146

12815-OTE III Gretchka.indd 146 28-07-2009 10:14:29


Fracture characterization

Let us observe that bars in Figures 8.1 and 8.2 are located above the predictions of the NIA
( and ). To understand why this is so, we need to examine Figure 8.3 more closely. It reveals
two types of stress disturbances: relatively extended areas of low stress called the shielding (blue)
and smaller areas of elevated stress called the amplification (yellow and red). The physical reason for
shielding is easy to grasp. The crack faces are traction-free because the cracks are dry. This means
that at the faces of all fractures regardless of the applied load. As the stress is continuous in
the background, slowly changes from zero to its far-field value as the distance from a crack face
increases. On the other hand, the static stress in linear elasticity obeys an analog of Gauss diver-
gence theorem (e.g., Markov, 1999). In particular, the theorem says that the volume integral of is
equal to the surface integral of the applied normal traction . Therefore, areas of higher stress that
compensate for the presence of the low-stress regions in the vicinity of fracture faces are required to
yield the proper volume average. These stress-amplification areas always form at the fracture tips
that primarily support a load applied to a fracture.
Figure 8.3 demonstrates that the shielding occupies a major portion of volume , and hence,
dominates the amplification. As a result, the numerically computed effective media come out to be
stiffer than those predicted by the non-interaction approximation. Figure 8.1 illustrates this directly.
The stiffening naturally translates into a reduction of the magnitude of the crack-induced anisotro-
py; this is why bars in Figure 8.2 correspond to smaller absolute values of the anisotropy coefficients
than those marked by the symbols and .
We note that Saenger et al. (2004) observed quite an opposite, softening effect of the crack
interactions in their finite-difference wave-propagation experiments. They concluded that the dif-
ferential effective media scheme (Vavakin and Salganik, 1975) is more appropriate than the NIA for
explaining the results of their modeling. We suggest that further numerical studies and perhaps a
step change in computing power are necessary to clarify this discrepancy.

8.3.2 Intersecting cracks


Next, we analyze whether the presence of fracture intersections can be inferred from effective
elasticity. To this end, we follow Grechka and Kachanov (2006b; 2006d), who examined more than
a hundred fracture arrays that contained both non-intersecting and intersecting cracks. Some of
those arrays are displayed in Figure 8.4. Note that intersecting fractures create intricate geometries,
ranging from relatively simple X-, 8- and V-shapes (Figures 8.4b, 8.4c), to more complicated ones
shown in Figure 8.4d. Clearly, once we begin dealing with these geometries, the cracks are no longer
penny-shaped; they even cease to be planar.
Figure 8.4 demonstrates that fracture intersections are not important for the effective prop-
erties. Indeed, we observe that literally nothing happens to the stress fields in the vicinity of
crack intersections. Therefore, the latter have little influence on the effective elasticity and can
be safely ignored. Figure 8.5 substantiates this conclusion. It shows the anisotropy coefficients of
the effective media for models containing non-intersecting and intersecting fractures. The bars
in Figure 8.5 correspond to the 95% confidence intervals for the anisotropy coefficients obtained
from finite-element simulations for 40 random realizations of locations of the non-intersecting
fractures at four different values of the total crack density . The trian-
gles indicate the same coefficients but for models where the fractures intersect (such as those in
Figures 8.4b 8.4d). We observe the following:
The bars and triangles overlap, implying the absence of influence of crack intersections of the
effective anisotropy.
Both the linear-slip theory and the NIA tend to slightly overestimate the magnitude of the
anisotropy coefficients. Similarly to models with a single fracture set (Figure 8.2), this is a conse-
quence of stiffening due to fracture interactions.

147

12815-OTE III Gretchka.indd 147 28-07-2009 10:14:30


Chapter 8

Figure 8.4: Horizontal cross-sections of the stress component for arrays of (a) non-intersecting and
(b, c, d) intersecting fractures. The aspect ratios of the fractures lie in the range .
The arrows indicate the directions of applied uniaxial remote load whose magnitude is 1 MPa (after
Grechka and Kachanov, 2006d).

The linear-slip (stars) and NIA predictions (circles) in Figure 8.5 do not significantly deviate
from each other, again confirming the insensitivity of the effective properties to the aspect ratios
of dry cracks.

We conclude that geometric intersections of fractures have virtually no influence on the effective
elasticity. Consequently, the interconnectedness of real fracture networks can be hardly established
from seismic data.

8.3.3 Non-circular fractures


Natural fractures in rocks are notoriously irregular. As their shapes resemble neither circles nor
ellipses (which can be modeled with equation 8.19), it is unclear to what extent our discussion
above is applicable to real cracks. Here, following Grechka et al. (2006), we illustrate a somewhat
unexpected result:

circular, penny-shaped fractures can be legitimately used to represent flat irregular cracks with random
shape irregularities.

148

12815-OTE III Gretchka.indd 148 28-07-2009 10:14:30


Fracture characterization

Figure 8.5: Effective anisotropy coefficients of fractured media. The bars correspond to the 95% con-
fidence intervals (the mean two standard deviations) of the numerically computed coefficients. The
triangles indicate their values for models with intersecting cracks. The predictions of the linear-slip
theory, which ignores the nonzero crack aspect ratios (equations 8.4, 8.12, 8.13, and 8.15) and the
non-interaction approximation, which accounts for them (equations 8.4 and 8.19), are shown with
, respectively. All numerical effective models are triclinic but only Tsvankins orthorhombic
coefficients are displayed (after Grechka and Kachanov, 2006d).

To do so, we examine six fracture shapes shown in Figure 8.6. Let us use each of them to build
a single vertical fracture set and allow the cracks in this set to have not only random locations but
also random orientations (or angles) in the vertical -plane. We would like to know whether or
not these irregular fractures can be replaced with penny-shaped ones for the purpose of obtaining
the effective stiffnesses. To answer this question, we compute the effective stiffnesses numerically,
average those in the -plane (this yields ), and calculate the crack contribution to the
effective stiffness,

149

12815-OTE III Gretchka.indd 149 28-07-2009 10:14:30


Chapter 8

Figure 8.6: Fracture geometries created to study the influence of crack shape on the effective properties.
All fractures are vertical and planar; their normals are parallel to the -axis. Geometries 4, 5, and 6
contain rock islands inside the cracks and model partially closed fractures (after Grechka et al., 2006).

(8.34)

We then use equations 8.4 and 8.19 to calculate

(8.35)

for penny-shaped cracks. Their crack densities, , and aspect ratios, , are the fitting parameters.
They are obtained from the nonlinear optimization

(8.36)

The fit quality is measured by the stiffness misfits

(8.37)

We intentionally compare the stiffness contributions, , rather than the effective stiffness ten-
sors themselves because the former are much more sensitive to the fractures. Figure 8.7 shows the
magnitudes of the misfits calculated as the norm

(8.38)

The misfits for all our fracture geometries. Clearly, the irregular fracture shapes in
Figure 8.6 can accurately be represented by circular cracks.

150

12815-OTE III Gretchka.indd 150 28-07-2009 10:14:31


Fracture characterization

Figure 8.7: Misfits (equation 8.38)


for the six fracture shapes in Figure 8.6
(after Grechka et al., 2006).

Figure 8.8: Models containing three sets of vertical rectangular cracks. Fractures that intersect their
neighbors are shaded, isolated cracks are transparent. The background Poissons ratio is
(after Grechka et al., 2006).

As a side observation, let us note that geometries 24, which are built from ellipsoids, are some-
what rough. They contain sharp edges, where normals to the crack faces are undefined. Berg et al.
(1991) term such fractures microcorrugated. The values for those crack models do not stand out
in Figure 8.7, implying that the circular-crack approximation performs equally well for all fracture
shapes, regardless of the smoothness of their faces.

8.3.4 Multiple sets of non-circular cracks


Finally, we combine all previously discussed features of cracks and directly examine the effective
elasticity due to multiple, differently-oriented, non-circular, intersecting fractures. Three repre-
sentative crack arrays out of a total of 20 built for this purpose are displayed in Figure 8.8. The
arrays contain three sets of vertical, dry, rectangular cracks oriented at azimuths

151

12815-OTE III Gretchka.indd 151 28-07-2009 10:14:31


Chapter 8

with respect to the coordinate axis . Each set has five cracks rotated around their normals, with a
increment to remove any preferential in-plane fracture orientation. The locations of the fracture
centers are random and uncorrelated to the in-plane crack rotations. The crack densities of the frac-
ture sets (computed from the optimization equation 8.36) are .
They yield the cumulative crack density .
It is virtually impossible to place the above described fractures in the volume randomly and
avoid their intersections, so all created models contain intersecting cracks. Sometimes relatively
few fractures intersect (Figure 8.8a), sometimes many (Figure 8.8b), sometimes all, forming a single
interconnected fracture network (Figure 8.8c). Fracture arrays shown in Figure 8.8 exhibit a high
level of 3D geometrical complexity. Specifically, the fracture shapes are non-circular, their faces are
not smooth because the cracks often protrude through each other, and there are irregular pieces of
host rock between the cracks, owing to the complicated geometry of their intersections.
As the models in Figure 8.8 possess no geometric symmetry, the corresponding effective stiffness
tensors, , computed with the finite-element method, are generally triclinic. To find how close
they are to orthotropy, we approximate them with the orthorhombic stiffness tensors (Arts et
al., 1991; Dewangan and Grechka, 2003) and calculate the relative misfits

(8.39)

Figure 8.9 shows that for all our 20 crack arrays are smaller than 1%. Hence, the effective crack-
induced anisotropy is virtually orthorhombic, even though the fractures have non-circular shapes
and intersect each other.

8.4 Governing parameters for vertical cracks


The comparison of various effective media schemes with finite element modeling presented in the
previous sections of this chapter leads to an important conclusion:

Figure 8.9: Relative deviations


(equation 8.39) from orthotropy of
the numerically computed effective
stiffness tensors for intersecting crack
arrays such as those shown in Figure 8.8
(after Grechka et al., 2006).

152

12815-OTE III Gretchka.indd 152 28-07-2009 10:14:32


Fracture characterization

originally isotropic rocks containing multiple fracture sets can be treated as orthotropic for
practical purposes.

Also, for dry cracks, effective orthotropy has a simplified type. If the fractures are vertical, the effec-
tive medium is fully characterized by only four independent parameters instead of nine needed
for the description of general orthorhombic media. These parameters are the density-normalized
Lam constants of the isotropic background (or, equivalently, the isotropic velocities
) and two principal crack densities . When all cracks have the same infill, which
is likely to be the case for interconnected fracture networks, the average fluid factor might be used
to describe the presence of fluids, bringing the total number of governing parameters to five. Thus,
the parameter vector, , quantifying anisotropy induced by multiple sets of vertical fractures in an
otherwise isotropic host rock, has the form

(8.40)

As the fracture sets are vertical, one of the symmetry planes of the effective orthorhombic medium
is always horizontal.
While the parameters 8.40 can potentially be estimated from a variety of seismic signatures,
Grechka and Kachanov (2006a) showed that can unambiguously be inverted from the following
data set:

(8.41)

where are the velocities of vertically propagating P- and two split shear-waves (fast
). The velocity ratios in can be computed from the zero-offset times after establish-
ing the P-to-S event correspondence. The symmetric matrices in equation 8.41 are the
NMO ellipses of pure modes reflected from a horizontal interface (see section 5.3). Clearly, wide-
azimuth, multicomponent reflection data are needed to characterize multiple fracture sets.

8.5 Seismic characterization of multiple fracture sets


Vasconcelos and Grechka (2007) tested this methodology of fracture characterization on a 3D, 9C
seismic data set acquired by the Reservoir Characterization Project (Colorado School of Mines) at
the tight-gas Rulison Field located in Colorado, USA. The reflector dips at Rulison are small (Figure
8.10) and allow application of the theory outlined in section 8.4.
The NMO ellipses were obtained over the entire survey area. The azimuthal velocity variations
for reflections from the bottom of the producing interval (Figure 8.10) are strong and visible on
both P- and S-wave data (Figure 1.5). In contrast, the eccentricity of the P-wave NMO ellipses in the
overburden is consistently smaller than 3%, suggesting that fractures in the producing interval are
the main source of the observed azimuthal anisotropy. The azimuthal velocity analysis, followed
by the generalized Dix differentiation (equation 5.43), provided the interval ellipses (equation
8.41) for estimating the fracture parameters.
Figure 8.11 displays the results of fracture characterization the inverted background
velocity fields and the principal crack densities (by definition, ).
The crack densities in Figure 8.11c are considerably greater than those in Figure 8.11d (the
color scale is the same), implying that the fracturing is dominated by the cracks trending in the
WNW-ESE direction. The results indicate that the western part of the area is mostly controlled

153

12815-OTE III Gretchka.indd 153 28-07-2009 10:14:32


Chapter 8

Figure 8.10: P-wave seismic section at


Rulison Field. The arrows mark the
reflection events used for azimuthal
velocity analysis (after Vasconcelos and
Grechka, 2007).

Figure 8.11: Output of the fracture characterization: the background velocities and of P- (a)
and S-waves (b), and the principal crack densities . The directions of the principal
fracture sets are shown with ticks; their lengths are proportional to the eccentricities of the interval
P-wave NMO ellipses (Figure 8.14b). The star indicates the well location from where the FMI log
shown in Figure 8.12 was acquired (after Vasconcelos and Grechka, 2007).

154

12815-OTE III Gretchka.indd 154 28-07-2009 10:14:32


Fracture characterization

by a single fracture set that has the density ,


while the eastern part has a non-negligible
contribution of other, differently-oriented frac-
tures that exhibit themselves as the set with
crack density (Figures 8.11c and 8.11d). The
fluid factors were also estimated but they are
not displayed because all the obtained values
were smaller than 0.01.
The survey area contains a well (star in Figure
8.11), where an FMI (Formation MicroImager)
log was run and fractures in the entire reservoir
were counted. Figure 8.12 compares the borehole Figure 8.12: Fracture count (blue) in well
fracture count with the fracture orientation esti- shown with star in Figure 93 and the 90%
mated from reflection seismic. As a dominant set confidence interval (dashed red) correspond-
of cracks has been identified at the well location ing to the azimuth of the fracture set with the
(Figures 8.11c and 8.11d), the subsidiary set is density estimated from seismic data (after
not shown in Figure 8.12 because it has nearly Vasconcelos and Grechka, 2007).
zero crack density. The fractures observed in FMI
form two sets oriented at approximately N70W
and N73E (blue in Figure 8.12). Clearly, these sets are not orthogonal to each other; yet, their influ-
ence on propagation of long (compared to the fracture sizes) seismic waves is equivalent to that of two
orthogonal sets. The dashed red line in Figure 8.12 shows the azimuth of the dominant equivalent set
estimated from seismic data.
Figure 8.12 can be regarded as an illustration of the resolution achievable from seismic data:
while long seismic waves cannot resolve each individual fracture set, they are sensitive to all fractures
simultaneously. This cumulative influence of all cracks on elastic wave propagation is described by the
crack-density tensor , which can be equivalently represented in terms of the contributions of mutually
orthogonal (or principal) fracture sets. It is these principal sets that control the seismic signatures and,
therefore, can be estimated from them. Thus, we can state that the fracture-characterization results in
Figure 8.11 are consistent with the FMI log in Figure 8.12.
The estimated background velocities and crack densities (along with the fluid factors ) are
sufficient for building an orthorhombic depth model of the reservoir. Figure 8.13 shows the vertical
velocities and some of Tsvankins anisotropy coefficients. It is instructive to point out that
the model in Figures 8.13 was obtained from reflection data only without using any borehole infor-
mation to constrain the vertical depth scale. Although well-log or check-shot data are usually neces-
sary for building orthorhombic subsurface models in the depth domain, such data are not needed
for the discussed data set. The reason is that the crack-induced rather than general orthotropy is
targeted. The former is significantly simpler because it is governed by fewer (five instead of nine)
independent parameters. This reduction in the number of unknowns makes it possible to rely solely
on surface seismic for building the unique orthorhombic model of Rulison reservoir. As Figures 8.11
and 8.13 indicate, the reservoir is noticeably anisotropic: the magnitude of the anisotropy coefficient
reaches 0.2 at the highest total crack density (Figures 8.11c and 8.11d).
Having estimated the reservoir parameters from both P- and S-wave data, we can predict what
would happen if we used either S- or P-waves alone for fracture characterization. In a typical pure
shear-wave survey, one would measure the shear-wave splitting coefficient, , and
interpret it as the crack density of a single fracture set. The result of this interpretation (Figure
8.14a) suggests that the western part of the study area is more fractured than the eastern one, quite
the opposite to the conclusion drawn from Figures 8.11c and 8.11d.

155

12815-OTE III Gretchka.indd 155 28-07-2009 10:14:33


Chapter 8

Figure 8.13: Vertical velocities and anisotropy coefficients ,


at Rulison reservoir (after Vasconcelos and Grechka, 2007).

Figure 8.14: The shear-wave splitting coefficient and the eccentricity of the P-wave NMO
ellipses at Rulison reservoir (after Vasconcelos and Grechka, 2007).

156

12815-OTE III Gretchka.indd 156 28-07-2009 10:14:33


Fracture characterization

If only P-wave data were used, one would estimate the eccentricities, (equation 5.42), of the
P-wave NMO ellipses as we did in the Powder River Basin case study discussed in section 5.4. The
eccentricity is quantified by the difference between the two Tsvankins coefficients, .
This difference, shown in Figure 8.14b, yields a similar result that the western part of the area is
apparently more fractured than the eastern part.
The origin of these contradictory conclusions can be understood from the weak-anisotropy
approximations (Vasconcelos and Grechka, 2007)

(8.42)

and

(8.43)

Clearly, both quantities are proportional to the difference between the


crack densities of the two principal fracture sets. As a consequence, the shear-wave splitting coef-
ficients and the eccentricities of the P-wave NMO ellipses are useful for fracture characterization
only if one fracture set dominates, that is, when . If multiple sets of cracks resulting in com-
parable are present in the subsurface, the differences become
ambiguous for fracture characterization. In particular, if the two principal crack densities coincide,
, both differences vanish, , and one would arrive at an obviously
incorrect result that the fractures are absent.
In summary, we saw that 3D, wide-azimuth, multicomponent seismic data make it possible to
establish the presence of multiple vertical fracture sets in an otherwise isotropic host rock. The
obtained principal crack densities indicate the presence of an interconnected fracture network in
the eastern part of the study area. This follows from a straightforward geometric consideration: it is
extremely difficult to place multiple sets of fractures that have random locations and the principal
crack densities and into a rock volume without letting those fractures intersect
each other. Figures 8.4 and 8.8 illustrate this point.
Despite the success in inferring intersecting fractures from seismic data, comparison of the
obtained crack densities with the estimated ultimate recovery (EUR) of wells in the study area shows
little correlation. Its absence might be explained by the physics of fluid flow through fractured
media. For low-porosity rocks, the crack-related permeability is essentially controlled by the fracture
widths (e.g., Kachanov and Sevostianov, 2005; Jaeger et al., 2007), whereas seismic signatures are
governed by the crack densities or the fracture lengths. If a relationship between the crack widths
and lengths is absent, as seems to be the case at Rulison Field, the effective elasticity and permeabil-
ity are uncorrelated. A similar lack of correlation between seismic fracture-characterization results
and gas production has been observed in a number of other tight-gas fields in the continental U.S.
and Canada.

157

12815-OTE III Gretchka.indd 157 28-07-2009 10:14:34


12815-OTE III Gretchka.indd 158 28-07-2009 10:14:34
BIBLIOGRAPHY
9 Bibliography

Aki, K. and Richards, P. G. (1980). Quantitative Seismology. W. N. Freeman & Co. (second edition,
2002, University Science Books, Sausalito).

Al-Dajani, A. and Tsvankin, I. (1998). Nonhyperbolic reflection moveout for horizontal transverse
isotropy. Geophysics, 63, 17381753.

Al-Dajani, A., Tsvankin, I. and Toksz, M. N. (1998). Nonhyperbolic reflection moveout for azi-
muthally anisotropic media. 68th Annual Meeting, Society of Exploration Geophysics, Expanded
Abstracts, 14791482.

Aleixo, R. and Schleicher, J. (2009). Traveltime approximations for qP waves in VTI media.
Geophysical Prospecting, 57, submitted.

Alekseev, A. S. and Gelchinsky, B. Y. (1959). On the ray method of computation of wavefields for
inhomogeneous media with curved interfaces. Petrashen, G. I. (editor), Problems of the dynamic the-
ory of propagation of seismic waves, 3, 5472, Leningrad, Leningrad University Press (in Russian).

Alekseev, A. S., Babich, V. M. and Gelchinsky, B. Y. (1961). Ray method for the computation of
intensity of wavefronts. Petrashen, G. I. (editor), Problems of the dynamic theory of propagation of
seismic waves, 5, 324, Leningrad, Leningrad University Press (in Russian).

Alexandrov, K. S. and Ryzhova, T. V. (1961). Elastic properties of rock-forming minerals. II. Layered
silicates. Bulletin of the USSR Academy of Sciences, Geophysics Series, 9, 11651168.

Alford, R. M. (1986). Shear data in the presence of azimuthal anisotropy. 56th Annual Meeting, Society
of Exploration Geophysics, Expanded Abstracts, 476479.

Alkhalifah, T. (1997). Velocity analysis using nonhyperbolic moveout in transversely isotropic


media. Geophysics, 62, 18391854.

Alkhalifah, T. (1998). Acoustic approximations for processing in transversely isotropic media.


Geophysics, 63, 623631.

Alkhalifah, T. (2000). An acoustic wave equation for anisotropic media. Geophysics, 65, 12391250.

Alkhalifah, T. and Tsvankin, I. (1995). Velocity analysis for transversely isotropic media. Geophysics,
60, 15501566.

Alkhalifah, T., Tsvankin, I., Larner, K. and Toldi, J. (1996). Velocity analysis and imaging in trans-
versely isotropic media: Methodology and a case study. The Leading Edge, 15, No. 5, 371378.

159

12815-OTE III Gretchka.indd 159 28-07-2009 10:14:34


Chapter 9

Arts, R. J., Helbig, K. and Rasolofosaon, N. J. P. (1991). General anisotropic elastic tensor in rocks
approximation, invariants, and particular direction. 61st Annual Meeting, Society of Exploration
Geophysics, Expanded Abstracts, ST 2.4.

Auld, B. A. (1973). Acoustic Fields and Waves in Solids, Volumes I and 2. John Wiley & Sons (second
edition, 1990, Robert E. Krieger Publishing Company).

Backus, G. E. (1962). Long-wave elastic anisotropy produced by horizontal layering. Journal of


Geophysical Research, 67, No. 11, 44274440.

Bakulin, A. and Grechka, V. (2003). Effective anisotropy of layered media. Geophysics, 68, 1817
1821.

Bakulin, A., Grechka V. and Tsvankin, I. (2000). Estimation of fracture parameters from reflection
seismic data. Parts I, II, III. Geophysics, 65, 17881830.

Banik N. C. (1984). Velocity anisotropy in shales and depth estimation in the North Sea basin.
Geophysics, 49, 14111419.

Belikov, B. P., Alexandrov, K. S. and Ryzhova, T. V. (1970). Elastic Properties of Rock-Forming Minerals
and Rocks. Nauka, Moscow (in Russian).

Ben-Menahem, A. and Singh, S. J. (1981). Seismic Waves and Sources. Springer-Verlag, New York
(second edition, 2000, Dover Publications, Mineola).

Benveniste, Y. (1986). On the Mori-Tanaka method in cracked solids. Mechanics Research


Communications, 13(4), 193201.

Berg, E., Hood, J. and Fryer, G. (1991). Reduction of the general fracture compliance matrix Z to
only five independent elements. Geophysical Journal International, 107, 703707.

Berryman, J. G. (2008). Exact seismic velocities for transversely isotropic media and extended
Thomsen formulas for stronger anisotropies. Geophysics, 73, D1D10.

Bleistein, N. (1984). Mathematical Methods for Wave Phenomena. Academic Press.

Bolshikh, S. F. (1956). On the approximate representation of the traveltime of reflected waves in the
case of a multilayered medium. Applied Geophysics, 15, 314 (in Russian).

Bristow, J. R. (1960). Microcracks and the static and dynamic elastic constants of annealed and
heavily cold-worked metals. British Journal of Applied Physics, 11, 8185.

Budiansky, B. and OConnell, R. J. (1976). Elastic moduli of a cracked solid. International Journal of
Solids and Structures, 12, 8197.

Carcione, J. M. (2001). Wave Fields in Real Media: Wave Propagation in Anisotropic, Anelastic and
Porous Media. Pergamon, Amsterdam.

160

12815-OTE III Gretchka.indd 160 28-07-2009 10:14:34


Bibliography

Cardona, R. (2002). Fluid Substitution Theories and Multicomponent Seismic Characterization of Fractured
Reservoirs. PhD. Thesis, Colorado School of Mines.

Cerven, V. (2001). Seismic Ray Theory. Cambridge University Press, Cambridge.

Cheng, C. H. (1993). Crack models for a transversely isotropic medium. Journal of Geophysical
Research, 98, 675684.

Chesnokov, E. M. (1977). Seismic Anisotropy of the Upper Mantle of the Earth. Nauka, Moscow (in
Russian).

Crampin, S. and Yedlin, M. (1981). Shear-wave singularities of wave propagation in anisotropic


media. Journal of Geophysics, 49, 4346.

Crampin, S., Evans, R. and Atkinson, B. K. (1984). Earthquake prediction: A new physical basis.
Geophysical Journal of the Royal Astronomical Society, 76, 147156.

de Parscau, J. (1991). P- and SV-wave transversely isotropic phase velocity analysis from VSP data.
Geophysical Journal International, 107, 629638.

Dellinger, J., Muir, F. and Karrenbach, M. (1993). Anelliptic approximations for TI media. Journal
of Seismic Exploration, 2, 2340.

Dewangan, P. and Grechka, V. (2003). Inversion of multicomponent, multiazimuth, walkaway VSP


data for the stiffness tensor. Geophysics, 68, 10221031.

Dewangan, P. and Tsvankin, I. (2006). Velocity-independent layer stripping of PP and PS reflection


traveltimes. Geophysics, 71, U59U65.

Dix, C. H. (1955). Seismic velocities from surface measurements. Geophysics, 20, 6886.

Douma, H. and Calvert, A. (2006). Nonhyperbolic moveout analysis in VTI media using rational
interpolation. Geophysics, 71, D59D71.

Douma, H. and van der Baan, M. (2008). Rational interpolation of qP-traveltimes for semblance-
based anisotropy estimation in layered VTI media. Geophysics, 73, D53D62.

Eshelby, J. D. (1957). The determination of the elastic field of an ellipsoidal inclusion and related
problems. Proceedings of the Royal Society, A241, 376396.

Fedorov, F. I. (1968). Theory of Elastic Waves in Crystals. Plenum Press, New York (originally in
Russian, 1965, Nauka, Moscow).

First Break Special Topic on Marine Seismic. (2008). 26, No. 12, 60108.

Fomel, S. (2002). Applications of plane-wave destruction filters. Geophysics, 67, 19461960.

161

12815-OTE III Gretchka.indd 161 28-07-2009 10:14:34


Chapter 9

Fomel, S. (2004). On anelliptic approximations for qP velocities in VTI media. Geophysical Prospecting,
52, 247259.

Fomel, S. and Grechka, V. (2001). Nonhyperbolic reflection moveout of P-waves: An overview


and comparison of reasons. Colorado School of Mines, Center for Wave Phenomena, Project Review,
127138.

Franco, G., Davis, T. L. and Grechka, V. (2007). Seismic anisotropy of tight-gas sandstones, Rulison
Field, Piceance Basin, Colorado. 77th Annual Meeting, Society of Exploration Geophysics, Expanded
Abstracts, 14611464, RC 3.3.

Gaiser, J. E. (1990). Transversely isotropic phase velocity analysis from slowness estimates. Journal
of Geophysical Research, 95, 241254.

Gajewski, D. and Penk, I. (1987). Computation of high-frequency seismic wavefields in 3D later-


ally inhomogeneous anisotropic media. Geophysical Journal of the Royal Astronomical Society, 91,
383411.

Galperin, E. I. (1971). Vertical Seismic Profiling. Nedra, Moscow (in Russian; English translation by
Hermont, A. J. and White, J. E. (editor), Society of Exploration Geophysics, 1974).

Goldin, S. V. (2008). Seismic Waves in Anisotropic Media. Siberian Branch of the Russian Academy
of Sciences Press, Novosibirsk (in Russian).

Grechka, V. (2007). Multiple cracks in VTI rocks: Effective properties and fracture characterization.
Geophysics, 72, D81D91.

Grechka, V. (2009). On the nonuniqueness of traveltime inversion in elliptically anisotropic media.


Geophysics, 74, in print.

Grechka, V. and Kachanov, M. (2006a). Seismic characterization of multiple fracture sets: Does
orthotropy suffice? Geophysics, 71, D93D105.

Grechka, V. and Kachanov, M. (2006b). Effective elasticity of rocks with closely spaced and intersect-
ing cracks. Geophysics, 71, D85D91.

Grechka, V. and Kachanov, M. (2006c). Effective elasticity of fractured rocks: A snapshot of the work
in progress. Geophysics, 71, W45W58.

Grechka, V. and Kachanov, M. (2006d). Effective elasticity of fractured rocks. The Leading Edge, 25,
No. 2, 152155.

Grechka, V. and Mateeva, A. (2007). Inversion of P-wave VSP data for local anisotropy: Theory and
a case study. Geophysics, 72, D69D79.

Grechka, V. and McMechan, G. A. (1996). 3-D two-point ray tracing for heterogeneous, weakly
transversely-isotropic media. Geophysics, 61, 18831894.

162

12815-OTE III Gretchka.indd 162 28-07-2009 10:14:34


Bibliography

Grechka, V. and Obolentseva, I. R. (1993). Geometrical structure of shear-wave surfaces near singu-
larity directions in anisotropic media. Geophysical Journal International, 115, 609616.

Grechka, V. and Pech, A. (2006). Quartic reflection moveout in a weakly anisotropic dipping layer.
Geophysics, 71, D1D14.

Grechka, V. and Rojas, M. A. (2007). On the ambiguity of elasticity measurements in layered rocks.
Geophysics, 72, D51D59.

Grechka, V. and Tsvankin, I. (1998a). Feasibility of nonhyperbolic moveout inversion in transversely


isotropic media. Geophysics, 63, 957969.

Grechka, V. and Tsvankin, I. (1998b). 3-D description of normal moveout in anisotropic media.
Geophysics, 63, 10791092.

Grechka, V. and Tsvankin, I. (1999a). 3-D moveout velocity analysis and parameter estimation for
orthorhombic media. Geophysics, 64, 820837.

Grechka, V. and Tsvankin, I. (1999b). 3-D moveout inversion in azimuthally anisotropic media with
lateral velocity variation: Theory and a case study. Geophysics, 64, 12021218.

Grechka, V. and Tsvankin, I. (2002a). NMO-velocity surfaces and Dix-type formulas in anisotropic
heterogeneous media: Geophysics, 67, 939951.

Grechka, V. and Tsvankin, I. (2002b). Processing-induced anisotropy. Geophysics, 67, 19201928.

Grechka, V., Theophanis, S. and Tsvankin, I. (1999a). Joint inversion of P- and PS-waves in
orthorhombic media: Theory and a physical-modeling study. Geophysics, 64, 146161.

Grechka, V., Tsvankin, I. and Cohen, J. K. (1999b). Generalized Dix equation and analytic treatment
of normal-moveout velocity for anisotropic media. Geophysical Prospecting, 47, 117148.

Grechka, V., Contreras, P. and Tsvankin, I. (2000). Inversion of normal moveout for monoclinic
media. Geophysical Prospecting, 48, 577602.

Grechka, V., Pech, A. and Tsvankin, I. (2002a). P-wave stacking-velocity tomography for VTI media.
Geophysical Prospecting, 50, 151168.

Grechka, V., Pech, A. and Tsvankin, I. (2002b). Multicomponent stacking-velocity tomography for
transversely isotropic media. Geophysics, 67, 15641574.

Grechka, V., Tsvankin, I., Bakulin, A., Hansen, J. O. and Signer, C. (2002c). Anisotropic inver-
sion and imaging of PP and PS reflection data in the North Sea. The Leading Edge, 21, No. 1,
9097.

Grechka, V., Zhang, L., and Rector, J. W. (2004). Shear waves in acoustic anisotropic media.
Geophysics, 69, 576582.

163

12815-OTE III Gretchka.indd 163 28-07-2009 10:14:34


Chapter 9

Grechka, V., Pech, A. and Tsvankin, I. (2005). Parameter estimation in orthorhombic media using
multicomponent wide-azimuth reflection data. Geophysics, 70, D1D8.

Grechka, V., Vasconcelos, I. and Kachanov, M. (2006). The influence of crack shape on the effective
elasticity of fractured rocks. Geophysics, 71, D153D160.

Grechka, V., Mateeva, A., Franco, G., Gentry, C., Jorgensen, P. and Lopez, J. (2007). Estimation of
seismic anisotropy from P-wave VSP data. The Leading Edge, 26, No. 6, 756759.

Hake, H. (1986). Slant stacking and its significance for anisotropy. Geophysical Prospecting, 34,
595608.

Hake, H., Helbig, K. and Mesdag, C. S. (1984). Three-term Taylor series for curves of P- and
S-waves over layered transversely isotropic ground. Geophysical Prospecting, 32, 828850.

Hale, D., Hill, N. R. and Stefani, J. (1992). Imaging salt with turning wave seismic waves. Geophysics,
57, 14531462.

Han, B., Galikeev, T., Grechka, V., Le Rousseau, J. and Tsvankin, I. (2001). A synthetic example of
anisotropic P-wave processing for a model from the Gulf of Mexico. Ikelle, L. and Gangi, A. (edi-
tors), Anisotropy 2000: Fractures, converted waves and case studies. Proceedings of the 9th International
Workshop on Seismic Anisotropy, Society of Exploration Geophysics, 311325.

Helbig, K. (1984). Anisotropy and dispersion in periodically layered media. Geophysics, 49, 364
373.

Helbig, K. (1994). Foundations of Anisotropy for Exploration Seismics. Elsevier Science Publishing.

Helbig, K. and Schoenberg, M. (1987). Anomalous polarization of elastic waves in transversely iso-
tropic media. Journal of Acoustical Society of America, 81, 12351245.

Hornby, B. E., Schwartz, L. M. and Hudson, J. A. (1994). Anisotropic effective-medium modeling


of the elastic properties of shales. Geophysics, 59, 15701583.

Horne, S. and Leaney, S. (2000). Short note: Polarization and slowness component inversion for TI
anisotropy. Geophysical Prospecting, 48, 779788.

Hsu, K., Schoenberg, M. and Walsh, J. (1991). Anisotropy from polarization and moveout. 61st
Annual Meeting, Society of Exploration Geophysics, Expanded Abstracts, 15261529.

Hubral, P. and Krey, T. (1980). Interval Velocities From Seismic Reflection Measurements. Society of
Exploration Geophysicists.

Hudson, J. A. (1980). Overall properties of a cracked solid. Mathematical Proceedings of Cambridge


Philosophical Society, 88, 371384.

Hudson, J. A. (1981). Wave speeds and attenuation of elastic waves in material containing cracks.
Geophysical Journal of the Royal Astronomical Society, 64, 133150.

164

12815-OTE III Gretchka.indd 164 28-07-2009 10:14:34


Bibliography

Hudson, J. A. (1991). Overall properties of heterogeneous material. Geophysical Journal International,


107, 505511.

Huet, C. (1990). Application of variational concepts to size effects in elastic heterogeneous bodies.
Journal of the Mechanics and Physics of Solids, 38, 813841.

Jaeger, J. C., Cook, N. G. W. and Zimmerman, R. W. (2007). Fundamentals of Rock Mechanics, fourth
edition, Blackwell Publishing.

Jech, J. and Penk, I. (1989). First-order perturbation method for anisotropic media. Geophysical
Journal of the Royal Astronomical Society, 99, 367376.

Jenner, E. (2001). Azimuthal Anisotropy of 3-D Compressional Wave Seismic Data, Weyburn Field,
Saskatchewan, Canada. PhD. Thesis, Colorado School of Mines.

Jenner, E., Williams, M. and Davis, T. (2001). A new method for azimuthal velocity analysis and
application to a 3D survey, Weyburn field, Saskatchewan, Canada. 71st Annual Meeting, Society of
Exploration Geophysics, Expanded Abstracts, 102105.

Jlek, P. (2002). Converted PS-wave reflection coefficients in weakly anisotropic media. Pure and
Applied Geophysics, 159, 15271562.

Jlek, P., Hornby, B. and Ray, A. (2003). Inversion of 3D VSP P-wave data for local anisotropy: A case
study. 73rd Annual Meeting, Society of Exploration Geophysics, Expanded Abstracts, 13221325.

Kachanov, M. (1980). Continuum model of medium with cracks. Journal of the Engineering Mechanics
Division, ASCE, 106 (EM5), 10391051.

Kachanov, M. (1992). Effective elastic properties of cracked solids: Critical review of some basic
concepts. Applied Mechanics Review, 45, 304335.

Kachanov, M. (1993). Elastic solids with many cracks and related problems. In Hutchinson, J. W.
and Wu, T. (editors), Advances in Applied Mechanics, 30, 259445.

Kachanov, M., and Sevostianov, I. (2005). On quantitative characterization of microstructures and


effective properties. International Journal of Solids and Structures, 42, 309336.

Kachanov, M., Shafiro, B. and Tsukrov, I. (2003). Handbook of Elasticity Solutions, Kluwer Academic
Publishers.

Kashtan, B. M., Kovtun, A. A. and Petrashen, G. I. (1984). Algorithms and techniques for computa-
tion of the fields of body waves in arbitrary anisotropic media, 108248. In Propagation of Body
Waves and The Methods for Computing of Wavefields in Anisotropic Elastic Media, Leningrad, Nauka
(in Russian).

Katahara, K. W. (1996). Clay mineral elastic properties. 66th Annual Meeting, Society of Exploration
Geophysics, Expanded Abstracts, 16911694.

165

12815-OTE III Gretchka.indd 165 28-07-2009 10:14:34


Chapter 9

Landau, L. D. and Lifshitz E. M. (1959). Theory of Elasticity. Course of Theoretical Physics, Vol. 7.
Pergamon (third edition, Butterworth-Heinemann, 1998; translated from Russian).

Laubach, S. E., Olson, J. E. and Gale, J. F. W. (2004). Are open fractures necessarily aligned with
maximum horizontal stress? Earth and Planetary Science Letters, 222, 191195.

Le Stunff, Y., Grechka, V. and Tsvankin, I. (2001). Depth-domain velocity analysis in VTI media
using surface P-wave data: Is it feasible? Geophysics, 66, 897903.

Levin, F. K. (1971). Apparent velocity from dipping interface reflections. Geophysics, 36, 510516.

Levin, V. and Markov, M. (2005). Elastic properties of inhomogeneous transversely isotropic rocks:
International Journal of Solids and Structures, 42, 393408.

Levin, V., Markov, M. and Kanaun, S. (2004). Effective field method for seismic properties of
cracked rocks: Journal of Geophysical Research, 104, B08202.

Liner, C. L. and Fei, T. W. (2006). Layer-induced seismic anisotropy from full wave sonic logs:
Theory, application, and validation. Geophysics, 71, D183D190.

Liu, E., Hudson, J. A. and Pointer, T. (2000). Equivalent medium representation of fractured rock.
Journal of Geophysical Research, 105, No. B2, 29813000.

Love, A. E. H. (1944). A Treatise on The Mathematical Theory of Elasticity. Dover Publications, New York
(fourth edition, 2001).

Lynn, H. B. and Thomsen, L. A. (1986). Reflection of shear-wave data along the principal axes of
anisotropy. 56th Annual Meeting, Society of Exploration Geophysics, Expanded Abstracts, 473476.

Lynn, W. (2007). Uncertainty implications in azimuthal velocity analysis. 77th Annual Meeting, Society
of Exploration Geophysics, Expanded Abstracts, 8488.

MacBeth, C. (2002). Multi-Component VSP Analysis for Applied Seismic Anisotropy. Pergamon,
Amsterdam.

Markov, K. (1999). Elementary Mechanics of Heterogeneous Media. Markov, K. and Preziosi, L. (editors),
Heterogeneous Media: Micromechanics Modeling Methods and Simulations, Birkhuser.

Michelena, R. J. (1994). Elastic constants of transversely isotropic media from constrained aperture
traveltimes. Geophysics, 49, 658667.

Miller, D. E. and Spencer, C. (1994). An exact inversion for anisotropic moduli from phase slowness
data. Journal of Geophysical Research, 99, 651657.

Milton, G. M. (2002). The Theory of Composites. Cambridge University Press, Cambridge.

166

12815-OTE III Gretchka.indd 166 28-07-2009 10:14:34


Bibliography

Mori, T. and Tanaka, K. (1973). Average stress in matrix and average energy of materials with misfit-
ting inclusions: Acta Metallurgica, 21, 571574.

Mueller, M. C. (1992). Using shear waves to predict lateral variability in vertical fracture intensity.
The Leading Edge, 11, No. 2, 2935.

Mensch, T. and Rasolofosaon, P. (1997). Elastic-wave velocities in anisotropic media of arbitrary


symmetry generalization of Thomsens parameters , , and . Geophysical Journal International,
128, 4364.

Murnaghan, F. D. (1951). Finite Deformation of an Elastic Solid. John Wiley & Sons, New York.

Musgrave, M. J. P. (1970). Crystal Acoustics. Holden-Day (second edition, 2003, Acoustical Society of
America).

Musgrave, M. J. P. (1985). Acoustic axes in orthorhombic media. Proceedings of Royal Society of


London, A 401, 131143.

Narr, W., Schechter, D. W. and Thompson, L. B. (2006). Naturally Fractured Reservoir Characterization.
Society of Petroleum Engineers.

Nemat-Nasser, S. and Hori, M. (1999). Micromechanics: Overall Properties of Heterogeneous Materials.


Elsevier, Amsterdam.

Nevsky, M. V. (1974). Quasi-Anisotropy of The Velocities of Seismic Waves. Nauka, Moscow (in Russian).

OConnell, R. J. and Budiansky, B. (1974). Seismic velocities in dry and saturated cracked solids.
Journal of Geophysical Research, 79, 54125426.

Ostoja-Starzewski, M. (2008). Microstructural Randomness and Scaling in Mechanics of Materials.


Chapman & Hall, New York.

Peacock, S. and Hudson, J. A. (1990). Seismic properties of rocks with distributions of small cracks.
Geophysical Journal International, 102, 471484.

Pech, A. and Tsvankin, I. (2004). Quartic moveout coefficient for a dipping azimuthally anisotropic
layer. Geophysics, 69, 699707.

Pech, A., Tsvankin, I. and Grechka, V. (2003). Quartic moveout coefficient: 3-D description and
application to tilted TI media. Geophysics, 68, 16001610.

Petrashen, G. I. (1980). Wave Propagation in Anisotropic Elastic Media. Nauka, Leningrad (in
Russian).

Postma, G. W. (1955). Wave propagation in a stratified medium. Geophysics, 20, 780806.

Riznichenko, Yu. V. (1949). On seismic quasi-anisotropy. Izvestia of the Academy of Sciences, Geography
and Geology Series, 13, No. 6, 518543 (in Russian).

167

12815-OTE III Gretchka.indd 167 28-07-2009 10:14:34


Chapter 9

Rudzki, M. J. P. (1911). Parametric representation of elastic wave in anisotropic media. Bulletin of


the Academy of Sciences, Cracow (A), 503536 (in German).

Rger, A. (1997). P-wave reflection coefficients for transversely isotropic models with vertical and
horizontal axis of symmetry: Geophysics, 62, 713722.

Rger, A. (1998). P-wave reflectivity with offset and azimuth in anisotropic media. Geophysics, 63,
935947.

Rger, A. (2002). Reflection Coefficients and Azimuthal AVO Analysis in Anisotropic Media. Society of
Exploration Geophysics, Tulsa.

Rytov, S. M. (1956). Acoustic properties of a thinly laminated medium. Acoustics Journal, 2, 6780
(in Russian).

Saenger, E. H., Krger, O. S. and Shapiro, S. A. (2004). Effective elastic properties of randomly
fractured soils: 3D numerical experiments. Geophysical Prospecting, 52, 183195.

Sarkar, D. and Tsvankin, I. (2004). Anisotropic migration velocity analysis: Application to a data
set from West Africa. 74th Annual Meeting, Society of Exploration Geophysics, Expanded Abstracts,
23992402.

Sayers, C. M. (1994). P-wave propagation in weakly anisotropic media. Geophysical Journal


International, 116, 799805.

Sayers, C. M. and Kachanov, M. (1991). A simple technique for finding effective elastic constants of
cracked solids for arbitrary crack orientation statistics. International Journal of Solids and Structures,
27, 671680.

Sayers, C. M. and Kachanov, M. (1995). Microcrack-induced elastic wave anisotropy of brittle rocks,
Journal of Geophysical Research, 100, B3, 41494156.

Schoenberg, M. (1980). Elastic wave behavior across linear slip interfaces. Journal of Acoustical Society
of America, 68, 15161521.

Schoenberg, M. and de Hoop, M. (2000). Approximate dispersion relations for qPqSV-waves in


transversely isotropic media. Geophysics, 65, 919933.

Schoenberg, M. and Douma, J. (1988). Elastic wave propagation in media with parallel fractures
and aligned cracks. Geophysical Prospecting, 36, 571590.

Schoenberg, M. and Helbig, K. (1997). Orthorhombic media: Modeling elastic wave behavior in a
vertically fractured earth. Geophysics, 62, 19541974.

Schoenberg, M. and Muir, F. (1989). A calculus for finely layered anisotropic media. Geophysics, 54,
581589.

168

12815-OTE III Gretchka.indd 168 28-07-2009 10:14:34


Bibliography

Schoenberg, M. and Sayers, C. (1995). Seismic anisotropy of fractured rock. Geophysics, 60, 204
211.

Sethian, J. A. (1999). Level Set Methods and Fast Marching Methods. Cambridge University Press,
Cambridge.

Sethian, J. A. and Vladimirsky, A. (2003). Ordered upwind methods for static Hamilton-Jacobi equa-
tions: theory and algorithms. SIAM Journal of Numerical Analysis 41, 235263.

Sevostianov, I. and Kachanov, M. (2002). On elastic compliances of irregularly shaped cracks.


International Journal of Fracture, 114, 245257.

Shafiro, B. and Kachanov, M. (1997). Materials with fluid-filled pores of various shapes: effective
moduli and fluid pressure polarization. International Journal of Solids and Structures, 34, 3517
3540.

Shah, P. M. (1973). Use of wavefront curvature to relate seismic data with subsurface parameters.
Geophysics, 38, 812825.

Shapiro, S. A. and Kaselow, A. (2005). Porosity and elastic anisotropy of rocks under tectonic stress
and pore-pressure changes. Geophysics, 70, N27N38.

Siliqi, R. and Bousqu, N. (2000). Anelliptic time processing based on a shifted hyperbola approach.
70th Annual Meeting, Society of Exploration Geophysics, Expanded Abstracts, 22452248.

Sirotin, Yu. I. and Shaskolskaya, M. P. (1978). The Foundations of Physics of Crystals. Nauka, Moscow
(in Russian).

Sun, Z. (1994). Seismic Anisotropy of Salt from Theoretical Study, Modeling, and Field Experiments. MSc.
thesis, University of Calgary.

Taner, M. T. and Koehler, F. (1969). Velocity spectra digital computer derivation and applications
of velocity functions. Geophysics, 34, 859881.

The Leading Edge Special Section on Offshore Technology Conference and Wide-Azimuth Seismic Acquisition.
(2007). 26, No. 4, 448529.

Thomsen, L. (1986). Weak elastic anisotropy. Geophysics, 51, 19541966.

Thomsen, L. (2002). Understanding Seismic Anisotropy in Exploration and Exploitation. Society of


Exploration Geophysics, Tulsa.

Thurston, R. N. (1965). Effective elastic coefficients for wave propagation in crystals under stress. Journal of
Acoustical Society of America, 37, No. 2, 348356.

Thurston, R. N. (1974). Waves in Solids. Encyclopedia of Physics, Mechanics of Solids, vol. VIa/4,
Truesdell, C. (editor), Springer-Verlag, 109302.

169

12815-OTE III Gretchka.indd 169 28-07-2009 10:14:34


Chapter 9

Tsvankin, I. (1995). Normal moveout from dipping reflectors in anisotropic media. Geophysics, 60,
268284.

Tsvankin, I. (1996). P-wave signatures and notation for transversely isotropic media: An overview.
Geophysics, 61, 467483.

Tsvankin, I., (1997a). Reflection moveout and parameter estimation for horizontal transverse isot-
ropy. Geophysics, 62, 614629.

Tsvankin, I. (1997b). Anisotropic parameters and P-wave velocity for orthorhombic media.
Geophysics, 62, 12921309.

Tsvankin, I. (2001). Seismic Signatures and Analysis of Reflection Data in Anisotropic Media. Pergamon,
Amsterdam (second edition, 2005).

Tsvankin, I. and Thomsen, L. (1994). Nonhyperbolic reflection moveout in anisotropic media.


Geophysics, 59, 12901304.

Ursin, B. and Stovas, A. (2006). Traveltime approximations for a layered transversely isotropic
medium. Geophysics, 71, D23D33.

van der Baan, M. and Kendall, J.-M. (2002). Estimating anisotropy parameters and traveltimes in
the - domain. Geophysics, 67, 10761086.

Vasconcelos, I. and Grechka, V. (2007). Seismic characterization of multiple fracture sets at Rulison
Field, Colorado. Geophysics, 72, B19B30.

Vasconcelos, I. and Tsvankin, I. (2006). Non-hyperbolic moveout inversion of wide-azimuth P-wave


data for orthorhombic media. Geophysical Prospecting, 54, 535552.

Vavakin, A. S. and Salganik, R. L. (1975). Effective characteristics of nonhomogeneous media


with isolated inhomogeneities. Mechanics of Solids, Allerton Press, 5866 (English translation of
Izvestia AN SSSR, Mekhanika Tverdogo Tela, 10, 6575.)

Vavrycuk, V. and Penck, I. (1998). PP-wave reflection coefficients in weakly anisotropic elastic
media. Geophysics, 63, 21292141.

Vernik, L. and Liu, X. (1997). Velocity anisotropy in shales: A petrophysical study. Geophysics, 62,
521532.

Walpole, L. J. (1969). On the overall elastic moduli of composite materials. Journal of the Mechanics
and Physics of Solids, 17, 235251.

Wang, X. and Tsvankin, I. (2009). Interval anisotropic parameter estimation using velocity-inde-
pendent layer stripping. Geophysics, 74, in print.

Wang, Z. (2002). Seismic anisotropy in sedimentary rocks, part 2: Laboratory data. Geophysics, 67,
14231440.

170

12815-OTE III Gretchka.indd 170 28-07-2009 10:14:34


Bibliography

White, J. E., Martineau-Nicoletis, L. and Monash, C. (1983). Measured anisotropy in Pierre shale.
Geophysical Prospecting, 31, 709725.

Williamson, P. and Maocec E. (2001). Estimation of local anisotropy using polarizations and travel
times from the Oseberg 3D VSP. Ikelle, L. and Gangi, A. (editors), Anisotropy 2000: Fractures, con-
verted waves and case studies. Proceedings of the 9th International Workshop on Seismic Anisotropy, Society
of Exploration Geophysics, 339348.

Winterstein, D. F. (1990). Velocity anisotropy terminology for geophysicists. Geophysics, 55, 1070
1088.

Winterstein, D. F. and Meadows, M. A. (1991). Shear-wave polarizations and subsurface stress direc-
tions at Lost Hills field. Geophysics, 56, 13311348.

Winterstein, D. F., De, G. S. and Meadows, M. A. (2001). Twelve years of vertical birefringence in
nine-component VSP data. Geophysics, 66, 582597.

Withers, R. and Corrigan, D. (1997). Fracture detection using wide azimuth 3D seismic surveys. 59th
EAGE Conference, Geneva, Extended Abstracts, paper E003.

Yilmaz, . (2001). Seismic Data Analysis. Society of Exploration Geophysics, Tulsa.

Zhang, F. and Uren, N. (2001). Approximate explicit ray velocity functions and traveltimes for
P-waves in TI media. 71st Annual International Meeting, Society of Exploration Geophysics, Expanded
Abstracts, 106109.

Zheng, X. and Penck, I. (2002). Local determination of weak anisotropy parameters from
qP-wave slowness and particle motion measurements. Pure and Applied Geophysics, 159, 1881
1905.

Zohdi, T. I. and Wriggers, P. (2001). Aspects of the computational testing of the mechanical prop-
erties of microheterogeneous material samples. International Journal for Numerical Methods in
Engineering, 50, 25732599.

Zohdi, T. I. and Wriggers, P. (2005). Introduction to Computational Micromechanics, Springer.

171

12815-OTE III Gretchka.indd 171 28-07-2009 10:14:34


12815-OTE III Gretchka.indd 172 28-07-2009 10:14:34

You might also like