You are on page 1of 87

Computations of the complex phase-,

group-, ray-velocity and attenuation for


seismic ray tracing in viscoelastic
anisotropic media

Tayammum Mohamed Ali Abdulqader Alkatheeri

MSc. Thesis

May 2020

A thesis submitted to Khalifa University of Science and Technology in accordance with the
requirements of the degree of MSc. in Petroleum Geosciences in the Department of Earth
Sciences.

i
Computations of the complex phase-, group-,
ray-velocity and attenuation for seismic ray
tracing in viscoelastic anisotropic media
by

Tayammum Mohamed Ali Abdulqader Alkatheeri

A thesis submitted in partial fulfillment of the


requirements for the degree of

MSc. in Petroleum Geosciences

at

Khalifa University

Thesis Committee

Dr. Bing Zhou (Supervisor), Dr. Youcef Bouzidi (Committee Member),


Khalifa University Khalifa University
Dr. Mohammad AlKhaleel (Co-Supervisor), Dr. Mohamed Kamel Riahi (Committee Member),
Khalifa University Khalifa University
Dr. Fateh Bouchaala (Thesis Defense
Coordinator),
Khalifa University

May 2020

ii
Abstract
Tayammum Mohamed Ali Abdulqader Alkatheeri, Computations of the complex phase-,
group-, ray-velocity and attenuation for seismic ray tracing in viscoelastic anisotropic
media, M.Sc. Thesis, MSc in Petroleum Geosciences, Department of (Earth Sciences), Khalifa
University of Science and Technology, United Arab Emirates, May 2020.

Seismic imaging is still the sole reconnaissance means to provide subsurface information with
no direct tangible contact. The resulting outcome of the seismic exercise greatly depends on
the understanding of the sent energy, as well as the medium that the energy waves are travelling
through. Too many assumptions and approximations would render the seismic depreciating in
value in direct proportion to the degree of violation of reality by those assumptions. The earth
is often assumed elastic and isotropic to simplify calculations while it is known that the earth
is in fact attenuating and does exhibit directional trends. To accurately model Earth (through
seismic surveying), it should be regarded as more complex than is being done currently. The
preliminary stages of ray tracing include the calculation of complex-valued wave quantities,
such as the complex phase slowness vector, the phase-, group- and ray-velocity and the ray
attenuation. These parameters become complex-valued because of the viscoelasticity of the
rock media which has to be taken into account because it causes amplitude decay of the wave,
hence the imaginary part, which is directly linked with attenuation of the propagating wave
energy. This complicates the problem. We present here three different methods which enable
the computation of the wave quantities: the iterative solver method, the perturbation method,
and the RSD method. The first proved satisfactory but computationally expensive due to
iterations with fine steps. The second proved successful in most of the cases but failed with
some highly complex media. The third and last proved efficient and fitting even at the
challenging triplication of wave-fronts in vertical shear waves. Further development of such
methods is needed especially for more complex media with strong anisotropy and strong
attenuation, because they mimic the earth more truthfully therefore they need to be
investigated. Subsequently, after obtaining such parameters, seismic ray tracing for accurate
complex travel-times can be conducted.

Indexing Terms: viscoelastic anisotropic media (VEAM); amplitude attenuation; ray and
phase velocities; transverse isotropy; seismic ray tracing

iii
Acknowledgement
I would like to acknowledge and thank all those who contributed greatly to this study. First and

foremost, I would like to express my sincerest appreciation and gratitude to my advisor, Dr.

Bing Zhou for his unrelenting support and constant observation throughout this study. I would

also like to extend my appreciation to my co-advisor Dr. Mohammad AlKhaleel for his

constructive advice and insightful suggestions and continuous concerns. I gratefully

acknowledge the support of the committee members, Dr. Youcef Bouzidi and Dr. Mohamed

Riahi, and the pleasure it was interacting with them. Special thanks to the team members, Dr.

Jianlu Wu, Dr. Qijnjie Yang, Mr. Shangbei Yang, Mr. Moosoo Won and Mr. Weining Liu for

their helpful contributions. Many thanks go to my family and close friends for their

unconditional support. Completing this would not have been possible without those incredible

people.

iv
Declaration and Copyright

Declaration

I declare that the work in this thesis was carried out in accordance with the regulations of
Khalifa University of Science and Technology. The work is entirely my own except where
indicated by special reference in the text. Any views expressed in the thesis are those of the
author and in no way represent those of Khalifa University of Science and Technology. No part
of the thesis has been presented to any other university for any degree.

Tayammum Mohamed Alkatheeri


Author Name: _____________________________________________________

Author Signature: __________________________________________________

07/05/2020
Date: _________________________________

Copyright ©

No part of this thesis may be reproduced, stored in a retrieval system, or transmitted, in any
form or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise,
without prior written permission of the author. The thesis may be made available for
consultation in Khalifa University of Science and Technology Library and for inter-library
lending for use in another library and may be copied in full or in part for any bona fide library
or research worker, on the understanding that users are made aware of their obligations under
copyright, i.e. that no quotation and no information derived from it may be published without
the author's prior consent.

v
Contents
Abstract .................................................................................................................................................. iii
Acknowledgement ................................................................................................................................. iv
Declaration and Copyright ...................................................................................................................... v
List of Figures ....................................................................................................................................... vii
List of Tables .......................................................................................................................................... x
1. Introduction ......................................................................................................................................... 1
1.1 Viscoelastic anisotropic rocks ....................................................................................................... 1
1.2 Seismic ray tracing in viscoelastic anisotropic media................................................................... 4
1.3 Goals of this study ........................................................................................................................ 5
2. Plane-wave theory in VEAM .............................................................................................................. 7
2.1 Inhomogeneous complex phase slowness vector .......................................................................... 7
2.2 Complex travel-time in the phase velocity form ......................................................................... 11
2.3 Phase attenuation ........................................................................................................................ 13
2.4 The Christoffel Equation ............................................................................................................. 16
3. Group and ray velocity in VEAM ..................................................................................................... 23
3.1 Group velocity vector ................................................................................................................. 23
3.2 Ray velocity vector ..................................................................................................................... 27
3.3 Complex travel-time in the ray velocity form ............................................................................. 29
3.4 Ray attenuation ........................................................................................................................... 31
4. Computations of complex phase and ray velocity ............................................................................ 33
4.1 Iterative algorithm ....................................................................................................................... 33
4.2 Perturbation method .................................................................................................................... 40
4.3 Real Slowness Direction approximation ..................................................................................... 48
5. Numerical Results ............................................................................................................................. 52
5.1 VTI models ................................................................................................................................. 52
5.2 TTI model ................................................................................................................................... 60
5.3 ORT models ................................................................................................................................ 62
6. Conclusions ....................................................................................................................................... 65
References ............................................................................................................................................. 69
Appendices............................................................................................................................................ 73
Appendix A ....................................................................................................................................... 73
Appendix B ....................................................................................................................................... 75
Appendix C ....................................................................................................................................... 76

vi
List of Figures
Figure 1: A very simplistic illustration showing the absorption/attenuation effect on seismic

waves as they travel through Earth and how wave energy decays with time. Wave energy and

shape are altered (vertical axis is energy, horizontal axis is travel-time) .................................. 1

Figure 2: Simplistic models of different types of transversely isotropic media: (a) vertical

symmetry axis (vertical transverse isotropy VTI), (b) horizontal symmetry axis (horizontal

transverse isotropy HTI), and (c) tiled symmetry axis (tilted transverse isotropy TTI) (Zhang,

2017) .......................................................................................................................................... 2

Figure 3: The stress and strain (cause & effect) impacts relationship on different materials at

different points in time. Note the different recovery modes which are associated with the

different medium responses to stress (inflicted energy) ............................................................ 3

Figure 4: Illustrating the wave-field models which arise from the nature of the source wave as

well as the geological model ...................................................................................................... 5

Figure 5: Illustration behind the concept of plane waves. The source is far away enough from

the receiver that until reaching there, the wave arrives in the form of its asymptote (planar).

Where τ is the travel-time of the wave and d is distance ........................................................... 8

Figure 6: A very concise illustration of a source generating a wave-front tracked by a ray-path

ȓ, with two distinct slowness directions associated with the complex slowness vectors p, and τ

is the travel-time; the real parts indicated in red and imaginary parts in blue ......................... 10

Figure 7: Multiple models of VTI media which are derived from geological features ........... 19

Figure 8: Multiple models of TTI media which are derived from geological features and are a

result of tiltation of VTI media ................................................................................................ 19

Figure 9: Simple illustration of the polarization direction in 3D of the different wave modes

qP, qSV, qSH. The q prefix added to any mode denotes quasi- which is a more

vii
realistic/representative visualization of particle motion with respect to the overall direction of

the wave travel; where the angle between wave propagation and particle vibration is not exactly

900 ............................................................................................................................................ 23

Figure 10: A simplistic illustration of a ray-path going from a source S to a receiver R through

time τ with slowness vector p (implicitly incorporating phase velocity c) and group velocity U

in an elastic anisotropic medium.............................................................................................. 24

Figure 11: A flowchart summarizing the first branch of the iterative solver algorithm steps,

focusing on the iterative element, to obtain the ray velocity, attenuation and quality factor .. 38

Figure 12: A flowchart summarizing the second branch of the iterative solver algorithm steps,

focusing on the iterative element, to obtain the ray velocity, attenuation and quality factor .. 39

Figure 13: Simplistic illustration of the concept of perturbation. Properties of the background

are summed with properties of a distinct disturbance deviating from the background, resulting

in an aggregate of the two properties ....................................................................................... 40

Figure 14: A flowchart summarizing the perturbation method steps to obtain the ray velocity,

attenuation and quality factor................................................................................................... 47

Figure 15: A flowchart summarizing the RSD approach steps to obtain the ray velocity,

attenuation and quality factor................................................................................................... 51

Figure 16: Ray and phase velocities (first row), ray and phase attenuations (second row), and

ray and phase quality factors for the three wave modes: qP (first column), qSV (second column),

qSH (third column) for a shale model of a viscoelastic VTI medium ...................................... 53

Figure 17: The computed ray velocity, attenuation, and quality factor of the qP-wave in Model

A1. The absolute relative error is also plotted with respect to each method in comparison with

the exact solution form the literature ....................................................................................... 55

Figure 18: The computed ray velocity, attenuation, and quality factor of the qP-wave in Model

B1. The absolute relative error is also plotted with respect to each method ........................... 56

viii
Figure 19: The computed ray velocity, attenuation, and quality factor of the qSV-wave in Model

A1. The absolute relative error is also plotted with respect to each method ........................... 57

Figure 20: The computed ray velocity, attenuation, and quality factor of the qSV-wave in Model

B1. The absolute relative error is also plotted with respect to each method ........................... 57

Figure 21: The computed ray velocity, attenuation, and quality factor of the qSH-wave in Model

A1. The absolute relative error is also plotted with respect to each method ........................... 58

Figure 22: The computed ray velocity, attenuation, and quality factor of the qSH-wave in Model

B1. The absolute relative error is also plotted with respect to each method ........................... 59

Figure 23: Ray and phase velocities (first row), ray and phase attenuations (second row), and

ray and phase quality factors for the three wave modes: qP (first column), qSV (second column),

qSH (third column) for a tilted shale model of a viscoelastic TTI medium based on the Bond

transformation .......................................................................................................................... 61

Figure 24: The absolute relative errors of the ray velocity, ray attenuation, and ray quality factor

of the two methods for the qP-wave in Model C1 ................................................................... 63

Figure 25: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qSV-wave in Model C3 ....................................................... 64

Figure 26: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qSH-wave in Model C3 ....................................................... 64

ix
List of Tables

Table 1: Elastic moduli and quality matrix of a viscoelastic VTI shale .................................. 52

Table 2: Elastic moduli and quality matrix of 8 viscoelastic VTI models............................... 54

Table 3: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qP-wave .............................................................................. 56

Table 4: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qSV-wave ............................................................................ 58

Table 5: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qSH-wave. ........................................................................... 59

Table 6: Elastic moduli and quality matrix of 3 viscoelastic ORT media ............................... 62

Table 7: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality

factor of the two methods for the qP-wave .............................................................................. 63

x
1. Introduction

1.1 Viscoelastic anisotropic rocks

Through seismic wave analysis, it is possible to visualize the subsurface and image it.

This is highly dependent on how the waves travel through the ground and are received back on

the surface. The better the understanding gets for how these waves propagate through the

multiple rock layers, the better and the closer to reality the imaging is going to be. This is

especially needed for complex geological features such as heavy thrusting and cumulative

folding. Traditionally, an assumption is followed to simplify the calculations and the

processing of the seismic data; and it is that the rock media which the seismic waves travel

through are elastic and isotropic. This is a violation of the real situation because rock media

are almost always not elastic nor isotropic; they facilitate energy decay, otherwise known as

attenuation/absorption (Figure 1) as waves travel through them, this is also associated with the

medium properties like pore fluids and the presence of fractures or joints (Carcione, 1995). Not

only that, but these changes vary in magnitude along different directions, which is otherwise

known as anisotropy (Figure 2). Waves’ velocities vary as they cross different rock types, and

on top of that, different layers of the same rock.

Figure 1: A very simplistic illustration showing the absorption/attenuation effect on seismic waves as they travel through Earth
and how wave energy decays with time. Wave energy and shape are altered (vertical axis is energy, horizontal axis is travel-
time)

1
Figure 2: Simplistic models of different types of transversely isotropic media: (a) vertical symmetry axis (vertical transverse
isotropy VTI), (b) horizontal symmetry axis (horizontal transverse isotropy HTI), and (c) tiled symmetry axis (tilted transverse
isotropy TTI) (Zhang, 2017)

It is known throughout the literature that rock layers are typically anisotropic and

attenuating (Tao and King, 1990; Burton 2007; Behura et al., 2012; Shekar and Tsvankin, 2012;

Zhubayev et al., 2016; Vavryčuk et al., 2016). Anisotropy means that the medium is not

homogeneous throughout; its properties change from sample to sample across the medium

along certain axes (directional dependency). In the real world, there are several examples of

naturally anisotropic geologic formations. Many factors result in forming such properties in

rocks; it might come down to the mineralization of the rock’s constituents where they have

preferred orientations reflected in the grain deposition like in the example of shales or olivine-

rich rocks, or in other cases where the depositional environment plays a role like in the cases

of mud cracks in clays or thin beds in limestones (Thomsen, 1986). This intrinsic or inherent

anisotropy can be divided into four broad categories: crystalline anisotropy, constraint induced

anisotropy (fractures due to confining pressure), lithologically induced anisotropy (favored

depositional orientation), and paleomagnetically induced anisotropy (related to magnetic

minerals directed by Earth’s magnetic field).

Attenuation practically represents the gradual loss of energy as that energy travels

through a medium due to the various interactions with the various components of the

surroundings. A medium that is perfectly elastic does not exhibit any attenuation, and that is

why attenuation is associated with inelastic or viscoelastic media; media which get

2
permanently deformed when subjected to stress (Figure 3). These types of media get

permanently deformed because they absorb portions of the energy travelling across them.

Figure 3: The stress and strain (cause & effect) impacts relationship on different materials at different points in time. Note the
different recovery modes which are associated with the different medium responses to stress (inflicted energy)

In this particular context, the energy is the seismic waves and the medium is Earth’s

subsurface. The combined effects of both of the mentioned phenomena result in the directional

variation of the seismic wave velocities and amplitudes as they travel underground. Velocity is

the main parameter for seismic studies since only through it can the subsurface be modelled

and visualized. There have been some theoretical and practical approaches to investigate the

anisotropic attenuation of rocks but they are considered rough estimates. The theory has

succeeded in the modelling but as for experimental results, they have been based on somewhat

invalid/limiting assumptions. The difficulty arises from the association of viscoelasticity with

complex numbers, wherein the real part of the quantities describes the propagation of wave

energy while the imaginary part represents the attenuation of the wave energy (Corcione,

2007). The challenges are significantly less apparent when the medium is assumed to be elastic;

no attenuation of energy is exhibited and hence all parameters are real-valued. This approach

3
have already been studied and explored (Thomsen, 1986). Reality also has another layer of

complexity; anisotropy, where a medium’s attributes vary across it with different directions.

1.2 Seismic ray tracing in viscoelastic anisotropic media

One of the numerical modelling techniques of seismic waves is seismic ray tracing. It

involves simulating the receiver response and predicting the ray-path which the waves follow

as they cross Earth’s media from the starting point, the source, to the end point, the receiver.

In seismic wave modelling, most studies assume that the wave-field is a 2D model, i.e. a line

source (planar waves) and a 2D geological structures. However, the actual seismic sources are

always point sources (spherical/curving waves). Thus, the assumption of 2D wave-field will

undoubtedly cause some discrepancies in data processing and interpretation and ultimately

yield a slightly different image than what the subsurface actually looks like. In addition, most

studies typically regard acoustic media or elastic isotropic media in seismic wave modelling,

but most geological structures in reality are more likely to be anisotropic due to the usually

naturally thin horizontally deposited layers, vertical fractures or parallel cracks. To simplify

the mathematical computations, as abovementioned, most previous research assumes the wave-

field to be 2D, i.e. both the seismic source and the geological model are 2D; however, in

practice, the seismic sources generated are always 3D (called point sources), even though the

geological model is considered 2D, the resultant wave-field caused by the point source become

2.5D (Figure 4). Attention rose to such conditions in order to more realistically approach the

modelling.

Reconstructing the subsurface structures is the main value of seismic data acquisition

and processing. If the seismic data failed to do so, then it becomes irrelevant. Additionally, the

more accurate the subsurface representation by the seismic is, the higher its value becomes and

the closer we are to studying and understanding what we cannot physically make direct contact

4
with. Therefore, it is also critical to study reality as it. The current ray tracing methods mainly

focus on elastic media, far fewer ones deal with viscoelastic anisotropic media (VEAM). On

top of that, most of the seismic ray tomography procedures use the assumption that equates the

response of ray tracing in VEAM and in elastic media, but the fact remains that this is not a

wholly valid assumption in reality. Hence, there is a need to address VEAM differently

corresponding to how different they are from any basic media.

Figure 4: Illustrating the wave-field models which arise from the nature of the source wave as well as the geological model

1.3 Goals of this study

We can consider this study a pioneer in its field as nobody has undertaken the seismic

ray tracing in viscoelastic 2D and 3D heterogeneous anisotropic media for its complexity and

involvement in computation. However, doing this will tremendously increase the accuracy in

seismic data processing and interpretation, as well as the overall quality and perhaps resolution,

because the image that is going to be examined is essentially closer to the reality of Earth’s

natural layering. And this practical application is advantageous over most of the research

ignoring the viscoelastic anisotropic rocks in carbonate reservoirs for example, which make up

roughly half of the world’s hydrocarbon reserves. Moreover, this study aids in eventually

carrying out seismic anisotropy ray tomography to improve the imaging of the aforementioned

carbonate reservoirs by helping to simultaneously locate the subsurface reflective interfaces

and reconstruct the heterogeneity of viscoelastic anisotropic layers using the complex travel-

times of the multi-arrivals. The stress is on carbonates here because they are well known for

5
their complexity and inherent heterogeneity. On top of that, studying and developing in this

domain could prove to be useful and applicable for research involving micro-seismicity, e.g.

relocation of fracturing or cracking in rocks which is critical in fracturing (fracking)

technology.

Several significant objectives can be achieved by conducting this study. Ultimately,

computing the ray-paths and travel-times in both isotopic and anisotropic media, in other

words, seismic ray tracing in arbitrary 2D & 3D VEAM. What is applicable for anisotropic

media is surely applicable for isotropic media. Since VEAM properties are closer to the actual

Earth properties than the traditional assumptions discussed above, understanding and defining

them allow for much more accurate reconstruction of Earth’s layers with a higher resolution of

subsurface interfaces. Finding the complex ray velocity gives the seismic ray speed and

attenuation in arbitrary VEAM, then integrating several factors aids in determining the ray-

path and the calculation of the real and imaginary travel-time (wave energy attenuation), and

therefore the accomplishment of seismic complex ray tracing in heterogeneous viscoelastic

anisotropic media. Having a method to accurately trace ray-paths would surely create

opportunities for a more meaningful interpretation of seismic images and deduce more realistic

subsurface structures. Having a robust ray tracing technique, is a critical part of seismic ray

tomography or seismic depth migration for reconstructing the subsurface image of the

viscoelastic anisotropic Earth. This research is to find a more effective and efficient algorithm

to compute the phase and ray velocities, as well as the ray attenuation for the final goal of ray

tracing in any given arbitrary VEAM, which is the natural next step and continuation of the

study. Computing these quantities is fundamental for seismic ray tracing as it cannot be

conducted without having known these variables.

6
2. Plane-wave theory in VEAM

2.1 Inhomogeneous complex phase slowness vector

According to the complex ray theory, the elastic moduli which describe the properties

of a medium; the phase velocity and ray velocity, the ray travel-time and the ray polarization

vector all become complex-valued with real parts to represent the actual energy propagation

aspect of the waves, and imaginary parts to represent the attenuation the wave energy goes

through as it travels an inelastic medium (Hearn and Krebes, 1990a, b; Chapman et al., 1999;

Hanyga and Seredyńska, 2000). Moreover, all of these viscoelastic quantities in our case are,

obviously, complex in value but they are also proved to be frequency dependent (Auld, 1973;

Carcione, 2014). Involving the aforementioned theory in order to address the problems in ray

tracing in viscoelastic media is a solution; but it is often overlooked due to the unavoidable

incorporation of imaginary numbers in the calculations, which renders dealing with the

problem computationally expensive as well as time consuming.

This is a study involving seismic waves in VEAM, the basis principle is understandably

governed by the equation of motion, or the wave equation, the derived subset. For the sake of

simplifying the problem, the waves are considered planar, meaning that the source is assumed

to be far enough that the wave-front’s bend is negligible and the wave is abstracted into its

tangent resulting in a plane wave (Figure 5). In addition, according to the Fourier Transform,

any seismic signal can be decomposed into its constituents of different combinations of plane

waves.

7
Figure 5: Illustration behind the concept of plane waves. The source is far away enough from the receiver that until reaching
there, the wave arrives in the form of its asymptote (planar). Where τ is the travel-time of the wave and d is distance

The equation describing a plane wave is given by

u  Aei ( px) , (2-1)

where u is the displacement vector, A is the amplitude vector of the wave, i is the imaginary

unit,  is the angular frequency of the wave,  is the travel-time of the wave at a certain point

in elapsed time, p is the slowness vector and its magnitude can be thought of as an inverse to

the phase velocity with a certain direction, x is the position vector denoting the wave travelling

location in space. This representation of the wave can be further extended to this special case

of viscoelasticity to incorporate the complex numbers induced by the attenuation property;

namely changing the travel-time and the phase slowness to be complex-valued. Equation (2-1)

becomes:

u  Aei [(  i ( I ) )  ( p( R )  ip( I ) )x ]


(R)

 Aei [( p( R ) x )  i ( ( I ) p( I ) x )]


(R)

(2-2)
=Ae  ( p x ) i ( p x )
(I ) (I ) (R) (R)
e
where any superscript R denotes the real component of the complex parameter, and any

superscript I denotes the imaginary component of the complex parameter. The final form of (2-

2) is rearranging the variables in the equation aiming at separating the real components form

the imaginary components. This is with the intent to distinguish between the portion describing

8
the propagation of the wave (the real part) and the portion describing the attenuation of the

wave (the imaginary part). Carefully examining equation (2-2), it is possible to deduce these

physical meanings from the mathematical forms. First, by setting up two obvious conditions

from equation (2-2):

 ( I )  p ( I )  x  C1  0
(2-3)
 ( R )  p ( R )  x  C2  0

The imaginary portion in the exponential function should and must represent amplitude decay.

And since there is a negative sign in the exponent term, the article within must yield a positive

value, hence  ( I )  p( I )  x  C1  0 , where C1 is a constant which much satisfy this condition

and be a value above zero, otherwise the attenuation of energy and the amplitude decay will

not happen. It is worth to note that the same conditions are also prevalent with the real term of

the wave equation. Although it is true that the real portion could be represented in a similar

form to the imaginary portion:  ( R)  p( R)  x  C2  0 where C2 is a constant which much

satisfy this condition and be a value above zero, but looking carefully in the wave equation’s

real part in the exponential function there is not a negative sign; meaning that inversely, the

article within must satisfy a positive value too as to reflect the propagation of the energy of the

wave.

The travel-time and the slowness are complex-valued while the wave amplitude,

frequency and position are real-valued. Subsequently, taking the gradient of equation (2-3)

yields

 ( I )
p( I )    ( I ) ,
x
(2-4)
 ( R )
p( R )    ( R ) .
x

As aforementioned, the slowness p is thought of as the inverse of the phase velocity vector;

time over distance, and at the same time it can be interpreted as the gradient of the travel-time,

9
as seen above in (2-4). This complex slowness vector is in general considered inhomogeneous

in nature. Vector inhomogeneity in this context is defined as any complex vector where the

directions of its real and imaginary components are not the same (Vavryčuk, 2008a).

Accordingly, and with having these two components, the slowness could be represented in

another form,

p  p( R )  ip( I ) = p( R )n( R )  ip( I )n( I ) , (2-5)

whereas established previously, p is the complex slowness vector with the two parts indicated

by the superscripts, p is the magnitude of p, n ( R ) and n ( I ) are the directional unit vectors shown

in Figure 6. A homogeneous complex vector has only one directional vector to represent it, as

the real and imaginary parts’ directions coincide, but generally this is not the case of the

slowness vector, meaning n ( R )  n( I ) .

Figure 6: A very concise illustration of a source generating a wave-front tracked by a ray-path ȓ, with two distinct slowness
directions associated with the complex slowness vectors p, and τ is the travel-time; the real parts indicated in red and imaginary
parts in blue

Finally, iterating on the definition of the slowness as an inverse of velocity, it is worth to note

that that velocity is the phase velocity symbolized by c. Throughout, there is a distinction

between phase quantities and ray quantities. Many of the wave quantities have those two sides

or versions; where the first describes the propagation of a plane wave while the latter describes

the propagation of wave energy following a ray-path and is more often associated with curved

10
wave-fronts related to point sources (Vavryčuk, 2016). Concerning the slowness vector, its

form with respect to the phase velocity is simply p  n / c , and decomposing that into its

complex components yields:

n( R ) n( I )
p  p ( R )  ip ( I )  i (2-6)
cR cI

Therefore, to determine the slowness vector, one must find the unit vectors n ( R ) and n ( I ) , and

the real and imaginary components of the phase velocity c .

1 1
p( R)  , p( I )  (2-7)
cR cI

2.2 Complex travel-time in the phase velocity form

Kinematic ray tracing mainly involves travel-times and ray-path calculations (Zhang,

2017). Essentially, to trace the rays as they are transmitted through the earth is the main

objective of this field of seismic modelling. Hence, the quantity representing the travel-time of

a wave is such a critical element as it is involved in multiple aspects of wave modelling and is

a characteristic by which a wave could be defined. Simply put, a travelling wave can be

comprehended as energy propagating through either of the domains; the spatial domain where

the position of the wave and its associated properties at a certain point is characterized by its

location in the 3D space coordinates, or the temporal domain where the position of the wave

and its associated properties at a certain point is characterized by the time elapsed since the

wave is triggered. The link between those two domains is the speed or velocity at which the

wave travels across a certain distance in the medium in a period of time. Great care should be

taken as in this case of a VEAM, the problem differs on two levels; there is an additional

imaginary part which must be considered, linked to the attenuating aspects of the media, as

11
well as the anisotropic aspects which results in the splitting and the heterogeneity of parameters

at different positions in the media (real and imaginary components of the slowness vector and

travel-times, etc., and different velocities of phase and group quantities, etc.).

Referring back to the definition of the phase slowness in terms of the travel-time, i.e.

equation (2-4), we have

    dx  p (n  rˆ ) ds (2-8)

where the gradient of the travel-time  is equal to the slowness vector p  pn , and dx is an

infinitesimal ray-path in the direction r̂ and is found by dx  rˆ ds . Here ds is the segment of

the ray-path. Furthermore, the real and imaginary travel-times could be obtained by

decomposing equation (2-8) to give

 ( R )   ( R )  dx  p ( R ) (n( R )  rˆ )ds
(2-9)
 ( I )   ( I )  dx  p ( I ) (n( I )  rˆ )ds

Moreover on that, it is possible to discover:

p ( I ) (n ( I )  rˆ )  ( I )
= . (2-10)
p ( R ) (n ( R )  rˆ )  ( R )

Having proven equation (2-10), which translates to show that a minute change in the travel-

time of a wave or a very fine step in time is dependent on the ray-path direction of where the

wave is going and its phase slowness vector, or conversely the inverse of the velocity by which

the wave is propagating. This suggests that in order to obtain a function for the travel-time of

a wave, an integration along the ray is needed. And it could simply read

 (R)  
Ray
p ( R ) (n( R )  rˆ )ds,
(2-11)
   p ( I ) (n ( I )  rˆ )ds.
(I )

Ray

Similarly, this integral equation could be manipulated and presented in the form of the phase

velocity. Recalling equation (2-6), which prompts that the phase velocity is the inverse of the

slowness with along its direction. The combination of equations (2-6) and (2-11) results in,

12
(n  rˆ )
 
Ray
c
ds . (2-12)

When the travel-time is complex (    ( R )  i ( I ) ), and the slowness direction vector n is taken

to have an equal direction with its components n  n ( R )  n ( I ) , this gives

c* c  ic
   ( R )  i ( I ) 
Ray
 (n  rˆ )
cc *
ds   (n  rˆ ) R * I ds
Ray
cc
(2-13)
c c
  (n  rˆ ) R 2 ds  i  (n  rˆ ) I2 ds
Ray
|c| Ray
|c|

where the asterisks represent complex conjugate quantities, and the modulus signifies the

magnitude of a vector. Comparing equation (2-11) and equation (2-13) to split the travel-time

into its components, we have:

cR
 (R)   (n  rˆ ) | c |
Ray
2
ds,
(2-14)
c
 (I )
  (n  rˆ ) I2 ds.
Ra y
|c|

And as such, obtaining the complex phase velocity grants the arrival at the function of travel-

time, among other quantities which will be discussed hereafter.

2.3 Phase attenuation

Attenuation is a core subject in this field of study. As heavily emphasized previously,

attenuation is an inseparable parameter in wavelet characterization, although it is considered to

be a property of the medium. A strongly attenuating medium has its Q factor lower than 5, and

weakly attenuating media’s Q factor is higher than 30 (Vavryčuk, 2008b). This quality factor

is unitless as it is a ratio, more on that below. Affinity to attenuating energy through it, the

quality factor of the medium is inversely proportional to that: the higher the attenuation

inflicted on waves travelling through the medium, the lower the quality factor of said medium

is. Attenuation is only an issue if the medium is inelastic, elastic media do not exhibit

13
attenuation. Waves’ energy gradually decay holding other factors constant as they travel across

viscoelastic media; this prompts a relationship between the travel-time of the wave, and the

amount of attenuation it undergoes as it progressively propagate through the particles of the

media.

Like the phase slowness, there are other wave quantities. Additionally, many of these

quantities could either be defined as phase quantities or ray quantities as mentioned previously.

Being one of the wave quantities, attenuation could be obtained through calculations using

phase quantities where it then becomes phase attenuation, while on the other hand, if it is

obtained through calculations using ray quantities, it becomes ray attenuation. However,

measuring phase quantities and obtaining them is theoretically more desirable as then doing

the inversion with them is much easier than with the ray quantities (Vavryčuk, 2016). Phase

velocity and phase attenuation both depend on the complex phase velocity c , and as just

illustrated, the travel-time of the wave could be formulated in terms of this complex phase

velocity through substitution of the phase slowness vector (equation 2-12 and assuming

n  n ( R ) ). In addition, using the identity dx  rˆ ds , it is possible to rewrite equation (2-13) as

c* ( R ) c  ic ( R )
   ( R )  i ( I )  Ray cc* (n  dx)  Ray Rcc* I (n  dx)
cR c
 
Ray
c  cI
2
R
2
( n ( R )  dx )  i  2 I 2 ( n ( R )  d x )
c  cI
Ray R
(2-15)

(n ( R )  dx)
  phase
 i  A phase (n ( R )  dx)
Ray
V Ray

where i is the imaginary unit, and c* is the complex conjugate of c . The real portion of the

travel-time is the inverse of one of the phase quantities mentioned earlier, the phase velocity:

14
cR2  cI2
V phase  . (2-16)
cR

All the while the imaginary portion is equal to another phase quantity and it is the subject of

discussion; the phase attenuation,

 cI
A phase  . (2-17)
c  cI2
2
R

To honor the theory, one must highlight that an approximation/assumption have been made

here. The unit vector n used in equation (2-12) is in reality complex-valued and is generally

unknown, therefore it is common to make the supposition where the vector is real

n  n ( R )  n ( I ) for equations (2-13) and (2-14) for constraining the problem. Following behind

this idea, it is frequently considered better to resort to the opposite side of the coin and choose

to utilize the laterally adjacent workflow of ray quantities over phase quantities in determining

the velocities and attenuations. Lastly, the third interesting phase quantity which is oftentimes

incorporated in such studies is the phase quality factor Q phase . It is a scalar which directly

represents a medium’s attenuating magnitude: the higher the Q factor of a medium, the less

attenuating it is. It mathematically reads (Carcione, 2000; Chichinina et al., 2006; Vavryčuk,

2007),

(c 2 ) ( R )
Q phase   . (2-18)
(c 2 ) ( I )

Moreover, Vavryčuk (2016) argues that ray quantities are preferable for another reason

involving practical laboratory experiments complications. Plane waves, which are associated

with phase quantities, are generally difficult to generate and control in lab settings, the

alternative is generating point source waves which naturally have a curved wave-front in 3D

space and are also indeed linked with ray quantities.

15
2.4 The Christoffel Equation

All VEAM may be described by a density-normalized (in order to eliminate the effects

of varying medium densities) four-ordered complex tensor aijkl (i, j, k, l = 1, 2, 3) (Vavryčuk,

2008b),

aijkl  aijkl
( R)
 iaijkl
(I )
(2-19)

where the superscripts R and I respectively represent the real component of the elastic moduli

and the imaginary component of the elastic moduli. The combination of the two effectively

describe the viscoelastic medium since each part denotes properties pertaining to a certain

behavior; the real part is the elastic behavior and the imaginary part is the viscous behavior

mirroring the attenuation of energy. Taking a step back from this, and to find where this elastic

moduli came from, it is inevitable to examine the plane wave (equation 2-1) and recognize it

presents a trial solution for plane waves in a homogeneous VEAM (Zhang, 2017) which satisfy

the differential equation based on Hook’s Law and the motion equation (Aki and Richards,

1980; Pereyra et al., 1980),

  2u j   i (cijkl  k ul )  f j
(2-20)
or   2u j   i (aijkl  k ul )  f j

where  is density of the medium,  is the frequency of the travelling wave, u j is the jth

component of the displacement vector u ,  i is the partial derivative of u with respect to the ith

coordinate, cijkl is the four-ordered elastic moduli tensor of the medium, and finally, f j is the

jth component of the body force. It is worth to note here that most commonly, the density-

normalized moduli aijkl is the one used in formulation instead of cijkl in order to eliminate the

density out of the equation and disregard its individual effects, and have it incorporated from

within the elastic moduli. Moreover, this moduli is not dealt with the way it is in its higher

order tensor form, but it is rather simplified in order to be able to manipulate it and its

16
calculations more conveniently. Hence, this four-ordered tensor aijkl is reduced into a two-

ordered tensor a pq (p, q = 1, 2, 3, 4, 5, 6) by means of the Voigt recipe transformation

(Vavryčuk, 2008b):

p  i ij  (9  i  j )(1   ij ),
(2-21)
q  k kl  (9  k  l )(1   kl ).

Kronecker delta  jk is the operator used here for carrying out the transformation. Applying

equation (2-21) to the four order moduli aijkl , we obtain the two order moduli a pq , which

replaces its counterpart in equation (2-20). For example, to describe the parameter a2312 , it is

referred to by its transformation complement which is corresponding to a46 as the equation

dictates. Additionally, the tensor also exhibits symmetry when considered in its matrix form;

the subscripts can be addressed interchangeably as the two resulting moduli would have the

same value; a12 equals a21 ( a1122  a2211 ). Originally, and since there are four subscripts in the

elastic tensor (i, j, k, l = 1, 2, 3), and each of these subscripts can take up to three values (3D

space), the total number of independent moduli resulting out of that will be

34  3  3  3  3  81 . After the Voigt recipe however, that number is reduced as there is now

two variables only (p, q = 1, 2, 3, 4, 5, 6), and both of them can take up to six values this time

62  6  6  36 . On the other hand, the assumption of the matrix symmetry equates all the

moduli values diagonally, and this leads to having the amount of independent moduli reach 21

independent parameters to completely describe any arbitrary VEAM. Equation (2-22) shows

the matrix form of the density-normalized elastic moduli reduced in two dimensions by the

Voigt recipe exhibiting symmetry, hence showing only the 21 independent variables needed to

describe any general medium,

17
 a11 a12 a13 a14 a15 a16 
 
 a22 a23 a24 a25 a26 
 a33 a34 a35 a36 
apq  . (2-22)
 a44 a45 a46 
 a55 a56 
 
 a66 

This is considered to be as the most general case for defining a VEAM, or any arbitrary media.

However, in specifically oriented media, the number of describing variables decreases as the

medium becomes simpler and simpler and more homogeneous in its properties. Ten cases have

been showed by Crampin (1984) as he applied the symmetric properties of rocks and deduced

the elastic moduli of different anisotropy cases. For example, to define an isotropic medium, it

is only necessary to define a11 and a44 as they are the sole independent variables out of the 21

(equation 2-23). Similarly, to define a transversely isotropic medium (VTI, Figure 7), the

moduli a11 , a44 , a13 , a33 , a66 are the ones needed (equation 2-24). In the case of a tilted

transversely isotropic media (TTI, Figure 8), the number of parameters increases to seven as

the additional two represent the tilt angles (inclination  and azimuth  ) of the symmetry axis

as given by equation (2-25). And as the medium becomes more and more complex, more

moduli need to be incorporated; as in the orthorhombic anisotropy case (ORT), 9 independent

moduli are needed to fully describe media with this orientation (equation 2-26).

18
 a11 a11  2a44 a11  2a44 0 0 0
 
 a11 a11  2a44 0 0 0 
 a44 0 0 0 
ISO
a pq  . (2-23)
 a44 0 0 
 
 a44 0 
 a44 

Figure 7: Multiple models of VTI media which are derived from geological features

 a11 a11  2a66 a13 0 0 0 


 
 a11 a13 0 0 0 
 a33 0 0 0 
aVTI  . (2-24)
 0 
pq
a44 0
 
 a44 0 
 a66 

Figure 8: Multiple models of TTI media which are derived from geological features and are a result of tiltation of VTI media

19
TTI
aijkl  ai ' j ' k 'l 'ei 'i e j ' j ek ' k el 'l (2-25a)

where ai ' j ' k 'l ' are given by equation (2-24) and

e  ( cos 0 , 0, sin 0 )
x' (2-25b)
e ( 0 , 1, 0 )
y'
e  ( sin 0 , 0, cos 0 )
z'

 a11 a12 a13 0 0 


0
 
 a22 a23 0 0 0
 a33 0 0 0 
a ORT
pq   (2-26)
 a44 0 0 
 a55 0 
 
 a66 

An additional property that is often also associated with viscoelasticity and the study of

viscoelastic media is the matrix of quality factor qijkl . This quality factor is the quotient of the

two previously mentioned parts of the elastic moduli and it represents the attenuation

quantitatively. The higher the quality factor value, the less absorption of energy the medium

applies to the travelling wave energy,

qijkl  aijkl
( R) (I )
/ aijkl (2-27)

It is essentially a descriptive parameter of the medium’s affinity to attenuate energy.

Having established all of that, the obvious follow up is to define a methodology to

obtain some of the aforementioned viscoelastic parameters like the phase slowness and the

phase velocity. The famous Christoffel Equations is the integral relationship which is

cornerstone to almost all discussions about seismic waves in anisotropic media (Vavryčuk,

2007; 2008a; 2008b; Vavryčuk et al., 2016; Zhang, 2017). The Christoffel equation could be

obtained by the motion equation (2-20) in a VEAM, which leads to the following form,

20
 2u j  aijkl ,i uk ,l  aijkl uk ,li (2-28)

where i u  u,i and so on and so forth with respect to any partial derivative of any variable.

The trial solution is u j  Aj (x)ei ( x ) (see equation 2-1), and then we calculate the other

quantities:

uk ,l  [ Ak (x)ei ( x ) ],l  [ Ak ,l (x)  i ,l Ak ]ei ( x )

uk ,li  {[ Ak ,l (x)  i ,l Ak ],i  [ Ak ,l (x)  i ,l Ak ]i ,i }ei ( x ) (2-29)

Substituting these equation in, equation (2-28) is expanded like so to obtain:

  2 Aj (x)ei ( x )  cijkl ,i [ Ak ,l (x)  i ,l Ak ]ei ( x )  cijkl {[ Ak ,l (x)  i ,l Ak ],i  [ Ak ,l ( x)  i ,l Ak ]i ,i }ei ( x )

  2 Aj (x)  cijkl ,i [ Ak ,l (x)  i ,l Ak ]  cijkl {[ Ak ,li (x)  i ( ,li Ak   ,l Ak ,i )]  [ Ak ,l (x)  i ,l Ak ]i ,i }

  2 Aj (x)  cijkl ,i Ak ,l (x)  cijkl Ak ,li (x)  i[cijkl ,i ,l Ak  cijkl ( ,li Ak   ,l Ak ,i )  cijkl Ak ,l (x) ,i ]   2cijkl ,l ,i Ak

(2-30)

thus, there are two parts to the last equation, a real one and an imaginary one. The latter is

concerned with the amplitude and is not the focus. The real part is the main concern and it

gives,

  2 Aj  (cijkl Ak ,l ),i   2cijkl ,l ,i Ak


(2-31)

and with applying the density normalization to the elastic moduli, as well as making the

approximation of very high frequencies (    ), we obtain one form of what is referred to as

the Eikonal Equation:

Aj  aijkl ,i ,l Ak
(2-32)

or

 jk Ak  aijkl ,i ,l Ak
(2-33)

and

21
(aijkl ,i ,l   jk ) Ak  0
, (2-34)

where the matrix aijkl ,i ,l is often written as aijkl pi pl recalling equation (2-4) relating the travel-

time to the slowness vector. This matrix defines the so called Christoffel tensor:

 jk (p)  aijkl pi pl , (2-35)

which is a 3  3 symmetric matrix dependent on the density-normalized elastic moduli and the

slowness, or slowness direction in the more common form:

 jk (n)  aijkl ni nl . (2-36)

We express the amplitude vector by A  Agˆ , where A is the amplitude of the wave, and

gˆ  ( gˆ1 , gˆ 2 , gˆ 3 ) is the polarization vector of the wave. Plugging this form into equation (2-34)

yields the Christoffel equation:

( jk  c 2 jk ) gˆ k  0 , (2-37)

where  jk is the Christoffel tensor given by equation (2-36), c is the phase velocity, and  jk

is the Kronecker delta. Subsequently, as gˆ k is a unit vector which represents the polarization

direction of the wave, and cannot be equal to zero, hence for the equation to be true, the

following is implicit,

det( jk  c 2 jk )  0 , (2-38)

where in this determinant, the squared phase velocity is sometimes denoted as G because it

represents the three eigenvalues of the Christoffel equation, which in general has three

solutions of eigenvalues and three eigenvectors. From equation (2-37), we have:

G  n   aijkl ni nl g j g k  c 2 (2-39)

the eigenvalues’ matrix in terms of the direction vector. It can also be reformed in terms of the

slowness vector naturally,

G  p   aijkl pi pl g j g k  1 . (2-40)

22
The three eigenvalues give three phase velocities of wave modes, and three eigenvectors ĝ

describing the three different polarization directions of the three wave modes. When addressing

vector ĝ , it is defined as the vector denoting the directions of the eigenvector. And in other

words, the polarization direction of the particle motion, which is characteristic of each wave

mode (qP, qSV, qSH) (Figure 9).

All of these values are dependent either directly or indirectly on the complex elastic

moduli, which in turn makes them complex in nature as well. And this is what is distinct about

VEAM; a layer of complexity is there taking the form of anisotropy, but another layer is also

added from the viscoelastic aspect of the medium reflecting the attenuating features.

Figure 9: Simple illustration of the polarization direction in 3D of the different wave modes qP, qSV, qSH. The q prefix added
to any mode denotes quasi- which is a more realistic/representative visualization of particle motion with respect to the overall
direction of the wave travel; where the angle between wave propagation and particle vibration is not exactly 90 0

3. Group and ray velocity in VEAM

3.1 Group velocity vector

The group velocity is defined by the travelling of the wave amplitude. As opposed to

all the previous discussion about the phase velocity, the physical meaning of the group velocity

is the wave speed of the seismic energy along a ray-path. However, in the isotropic case, the

distinction between the phase velocity and the group velocity becomes meaningless as they

23
both are equivalent, but that is not the case in other types of media, namely, elastic anisotropic

media (Zhou and Greenhalgh, 2004) (Figure 10). In addition, and as emphasized previously, in

VEAM, all of the quantities become complex-valued; therefore illustrating them would be

complicated as the different values have different vectors, and those vectors have different

directions, and on top of that, real and imaginary quantities also must be considered and their

directions taken into account.

Figure 10: A simplistic illustration of a ray-path going from a source S to a receiver R through time τ with slowness vector p
(implicitly incorporating phase velocity c) and group velocity U in an elastic anisotropic medium

Mathematically, and since the group velocity denotes the energy flux alongside the

propagating wave, it can be defined as:


U (3-1)
κ

where U is the group velocity, and κ is the wave number (spatial frequency) vector

κ  (1 , 1 , 1 )   n . The wave number is also related to the frequency in another form; the

magnitude of the wave number vector is,


  12   22   32 (3-2)
c

where c is the phase velocity. Having established the above relationships, we can further

expand on the group velocity as follows (Zhang, 2017):

24
  ( c)  c
Ui   c 
 i  i  i  i
(3-3)
c c  c 
 cni    cni   (  )
 i   i   i

where c  cm ( ,  ) the phase velocity is a function of two angles reflecting the inclination (  )

and the azimuth (  ) of the phase slowness direction. Moreover, the two derivatives of the

 
angles with respect to the wave number read  sin 1 ( 12   22 /  ) ,  tan 1 ( 2  1 ).
 i  i

Then, with this, the derivatives can be replaced by their expansions in the equation of the group

velocity,

cm sin  cm


U x( m )  (cm sin   cos  ) cos   ,
 sin  
c cos  cm
U y( m )  (cm cos   cos  m ) sin   , (3-4)
 sin  
c
U x( m )  cm cos   sin  m ,


where the superscripts or subscripts ‘m’ takes the integers 1, 2 ad 3 and stand for the wave

modes (qP, qSV, qSH). Thus, with obtaining the group velocity in the three directions in the 3D

space, it is possible to also find the amplitude,

cm 2 1 c
U ( m)  cm2  ( )  2 ( m )2 . (3-5)
 sin  

Like so, the group velocity could be obtained, and viewing it from equation (3-5) suggests that

it is dependent on the phase velocity and the inclination and azimuth angles. This confirms the

concept and supports the abovementioned idea about the association between the phase

velocity and the group velocity in isotropic media; in isotropic media where the two angles are

zero (thus eliminating the derivative terms in (3-5)), the velocities end up being equivalent.

Whereas, in anisotropic media, the two velocities are indeed proven to be distinct. As for

25
calculating these derivatives of the phase velocity with respect to the angles, analytic

calculations could be carried out by several formulations (Zhou and Greenhalgh, 2005; Zhang,

2017).

In another way, and recalling equation (2-39) describing the eigenvalues, the phase

velocity could be defined differently by also employing the Christoffel tensor equation (2-39),

cm2   jk gˆ (jm ) gˆ k( m ) . (3-6)

In addition to that, the definition of the Christoffel tensor above could be put to use further by

i
recognizing a property of the wave number vector with its directional descriptors where ni 

l
and nl  (due to   n ), so we have:

 i l ( m ) ( m )
cm2  aijkl ( ) gˆ gˆ . (3-7)
2 j k

Afterwards, to manipulate the above equation aiming to come up with an alternative expression

of the group velocity, we proceed with deriving equation (3-7) with respect to the wave number

 y and calculate,

cm 1   i l 
 [aijkl gˆ (jm ) gˆ k ( 2 )  aijkl pi pl ( gˆ (jm ) gˆ k( m ) )]
 2cm   
1 il 2 i l  gˆ (jm ) gˆ k( m )
 [ 2 ( l i   i l )   il  (  ˆ
g (m)
 ˆ
g ( m)
 )]
2cm  3   
jk k j jk

1 gˆ (jm ) ( m ) ( m ) gˆ k( m )


 [ l nl  i ni  2cm n  (
2
gˆ j  gˆ j )]
2cm  
1
 ( i pi  cm n )

(3-8)

26
gˆ (jm) gˆ k( m)
where il  aijkl gˆ j gˆ k , and the identity  jk gˆ k( m)  gˆ (jm)  jk  0 have been applied in
 

these calculations. Substituting equation (3-8) into equation (3-3) for the derivative in the group

velocity, we obtain another form of the group velocity:

cm aijkl nl ( m ) ( m )
U i( m )  cm ni    gˆ j gˆ k
 i cm (3-9)
= aijkl p (m)
l gˆ ( m)
j gˆ ( m)
k

Thus far, presented here are two methods by which the group velocity could be obtained, which

in turn demonstrated why the phase velocity and the group velocity generally have significant

difference when it comes to anisotropic media as opposed to isotropic media in which they are

both interchangeable. The first way of finding the group velocity utilized defining it in terms

of the phase velocity and additional terms associating with the wave normal inclination and

azimuth angles, while the second approach required finding the eigenvectors along with the

slowness (or exchangeably the phase velocity). Any parameter based on a complex parameter

(since the main subject is the viscoelastic realm) is rendered by default to be complex-valued

as well.

3.2 Ray velocity vector

The ray velocity vector is a member of the ray quantities, the category of parameters

briefly cited earlier opposing phase quantities. And in contrast to the latter, the former as its

name suggests is directly linked to the ray-path of a wave. It can be explained as the propagation

speed of a wave along its ray-path. Therefore, it can be defined simply in a mathematical form

by a basic relationship with travel-time and wave displacement,

27
dx ds
v  rˆ , (3-10)
d d

where x is the coordinate vector of the ray-path, and it is exactly the same if defined by a

segment ds in the direction of the ray-path r̂ . Based on the definition of the slowness as the

gradient of the travel-time ( p   , see equation (2-4)), it is possible to derive a well-known

identity by applying the a dot product of the phase slowness vector and the ray velocity (Helbig,

1994; Vavryčuk, 2007; 2008a; 2008b; Vavryčuk et al., 2016; Zhang, 2017):

ds d ds
p  v    rˆ   1, (3-11)
d ds d

c
n r  . (3-12)
v

where v is the -scalar- ray velocity coming from v   rˆ . By this, a direct link is found between

the slowness vector and the ray velocity vector. The dot product could suggest the nature of

the relationship between the two vector variables, as it encloses a trigonometric function of the

angle between the two directions of the two vectors. If a dot product of two vectors is zero then

they are perpendicular to each other, if it is equal to 1 then they are parallel to each other (given

they are both normalized to length equal to 1), any other number indicates any other angle.

In an attempt to further develop the expression of the ray velocity, the step of

manipulating equation (3-6) by dividing both sides of the equation by cm2 could be taken, and

also by accompanying that with the definitions of the Christoffel tensor and the slowness

vector, we get:

1  aijkl pi( m ) pl( m ) gˆ (jm ) gˆ k( m ) . (3-13)

and with modifying this equation to isolate pi( m) , and recalling the famous identity equation (3-

11), the ray velocity is found to also have the form:

28
vi( m )  aijkl pl( m ) gˆ (jm ) gˆ k( m ) . (3-14)

This equation will look familiar as it is exactly identical to equation (3-9). This quickly prompts

to deduce that the group velocity is in fact equivalent to the ray velocity in anisotropic media

as it is in isotropic media. But note that both are not equal to the phase velocity in anisotropic

media as they are in isotropic media. One will also note that both the group and ray velocities

have similar physical meanings where both describe the propagation of the wave energy along

a ray-path. This might prove important for some specific applications or particular

methodologies for when for example measuring and calculating ray velocities is not reasonable,

hence one would resort to measuring the easier quantity and then that would be equivalent to

the actual quantity sought after. Again, it is worthy to reiterate on the fact that most of these

parameters are complex-valued in the subject case of viscoelasticity and attenuating media.

Therefore, all equations shown here are applicable to VEAM so long as the parameters are

taken to have an imaginary portion.

3.3 Complex travel-time in the ray velocity form

Having defined the ray velocity (or the group velocity for that matter, it makes no

difference in isotropic as well as anisotropic media), which corresponds to the phase velocity

when mirroring ray and phase quantities, subsequent parameters can easily be derived and

calculated. Most of the relationships between these viscoelastic quantities have been shown

above where linking the phase with ray calculations is unavoidable and measuring or handling

either one or the other purely depends on the nature of the anticipated application. Therefore,

some view both domains as being quite alike where they only differ in physical meanings or

concept of study; “the only difference is that the ray quantities are defined along a ray, whereas

the phase quantities are defined along a wave normal” (Vavryčuk, 2007).

29
The complex ray velocity relates to the complex travel-time through the ray coordinate

vector as equation (3-11) indicates. But additionally, and to make things clearer, the equation

could be further developed to show the two components of the complex ray velocity; the real

and the imaginary one,

dx
v  vrˆ
d
ds
v  rˆ  v  rˆ  ( v ( R )  iv ( I ) ) (3-15)
d
vR  rˆ  v ( R ) , vI  rˆ  v ( I )

where bold characters indicate vectors while others indicate components or magnitudes of

vectors. The ray velocity is simple and straightforward to deduce, it is intuitive when the term

‘velocity’ is expressed; it relates to the travel-time  and the ray-path/direction r̂ , so it

encompasses the basic definition of speed which is the distance over the time. Consequently,

in order to obtain the travel-time it need be isolated and have the equation rewritten in terms of

the travel-time, which results in:

ds
 
Ray
v
(3-16)

If we were to continue to make analogy to the phase realm, equation (3-16) should be

equivalent to equation (2-12) in phase terms. This form of the travel-time is then inserted to

account for its two components –being a complex parameter–,

v* v  iv
 (R)
 i (I )
  * ds   R * I ds
Ray
vv Ray
vv
(3-17)
v  vI
  R 2 ds  i  ds
Ray
|v| Ray
| v |2

Correspondingly, equation (3-17) correlates to equation (2-13) of the phase travel-times. The

formulation is identical, however note that the dot product between the slowness direction

30
vector and the ray direction vector is not explicitly required in the ray realm, which is the reason

why oftentimes calculating the travel-time by equation (3-17) rather than by equation (2-13) is

more preferable and reliable as it does not necessitate making any approximation which could

very well impose inaccuracies and errors on the results of the calculations.

3.4 Ray attenuation

The exact same correlation applies when realizing the ray attenuation and ray velocity

parameters as with their counterparts as phase quantities. In an almost identical fashion to the

travel-times in phase form, the real and imaginary components of equation (3-17) demonstrate

the two sought after quantities. We rewrite (3-17) to highlight that as such,

v* v  iv
 ( R )  i ( I )  Ray vv* ds  Ray Rvv* I ds
vR v
 
Ray
v  vI
2
R
2
ds  i  2 I 2 ds
v  vI
Ray R
(3-18)

ds
 V
Ray
Ray
i 
Ray
ARay ds

The real portion is the inverse of one of the ray quantities mentioned earlier, the ray velocity:

vR2  vI2
V ray  . (3-19)
vR

All the while, the imaginary portion is equal to another ray quantity and it is the subject of

discussion; the ray attenuation, which guides the amplitude decay of a wave along a ray-path:

 vI
Aray  . (3-20)
v  vI2
2
R

Lastly, the third interesting ray quantity which is always incorporated in such studies is the ray

quality factor Q ray . It is a scalar which directly represents a medium’s attenuating capacity; the

31
higher the Q factor of a medium, the less attenuating it is. It mathematically reads (Carcione,

2000; Chichinina et al., 2006; Vavryčuk, 2007),

(v 2 ) ( R )
Q ray   (3-21)
(v 2 ) ( I )

All of these quantities are associated with v , the complex ray velocity parameter. Similarly,

and as have been demonstrated previously, it is simple to obtain the same quantities but for the

counterparts; the real phase velocity V phase , the real phase attenuation A phase , and the real phase

quality factor Q phase shown in equations (2-16) ~ (2-18), simply by replacing the complex ray

velocity with the aforementioned complex phase velocity c in the formulas.

An obvious deduction can be made clearer here. If it is assumed that the imaginary part

of the elastic moduli aijkl  aijkl


( R)
 iaijkl
(I )
, is equal to zero hence leaving only the real part to the

parameter, the abovementioned six quantities (expressed in equations (2-16) ~ (2-18) and (3-

19) ~ (3-21)) change drastically to accommodate the absence of any imaginary component,

leading to:

vR2  vI2 vR2  0


V ray   v
vR vR

cR2  cI2 cR2  0


V phase   c
cR cR

 vI 0
Aray   2 0
v  vI vR  0
2
R
2

cI 0
A phase   2 0 (3-22)
c  cI c R  0
2
R
2

(v 2 ) ( R ) (v 2 ) ( R )
Q ray     
(v 2 ) ( I ) 0

32
(c 2 ) ( R ) (c 2 ) ( R )
Q phase     
(c 2 ) ( I ) 0

This perfectly represents elastic media; the ray velocity and phase velocity are exactly equal to

their real parts, the attenuation is brought to zero because there is zero energy loss in perfect

elastic media, which is also translated in an infinitely rising Q factor for both of the ray and

phase quantities.

To reiterate, it is preferable to follow equations (3-19) ~ (3-21) (ray quantities) over

equations (2-16) ~ (2-18) (phase quantities) for computing the travel-time and attenuation

because the former does not incorporate any conditions, but the latter implies the assumption

n  n( R ) .

4. Computations of complex phase and ray velocity

4.1 Iterative algorithm

For the ultimate goal of conducting ray tracing in VEAM, the determination of the

phase slowness vector and hence the ray velocity vector is inevitable. Having stated that, the

dimensions of the problem differ as the conditions of the settings vary. It is an undeniable fact

that dealing with elastic media is much more convenient in terms of calculations/computations

and the limitations which have be to be taken into account. This is all due to the simplicity of

the medium’s properties when compared with the more tortuous viscous medium or even the

viscoelastic medium. The complexity sprouts from the absorbent properties of such media

which theoretically translate into complex-valued numbers with imaginary parts which must

be considered in the calculations. Conversely, in elastic media, there are only real numbers to

be involved in calculations; the elastic moduli aijkl and the slowness vector p only have to be

dealt with as their real numbers values. However, in the subject case (the viscoelastic case),

33
there is a distinction between p ( R ) and p ( I ) , and one way to calculate this imaginary part is by

iteratively varying it through a range of values (for a constant p ( R ) ) with the purpose of

minimizing the ray direction vector’s imaginary part (Vavryčuk, 2008b). Commonly, the

imaginary part of the slowness pales in value with respect to the real part, and this makes for

not so many iterations suffice for an acceptable result.

According to Vavryčuk (2007), the stationary slowness vector, denoted by p 0 ,

determination can be done through two procedures; by an iterative inversion method or by

solving a system of polynomial equations. The latter though exhibits some hindering

limitations seeing that it was initially developed for dealing with elastic media, however this is

not an alarming issue as it is possible to adaptively convert the solution into the viscoelastic

domain by incorporating the imaginary parts of the complex quantities. The real issue with the

polynomial equations approach, even after adopting complex numbers, is that it might still

produce highly invalid results when the assumption that the complex ray velocity vector for

the given slowness vector being homogenous is violated (Vavryčuk, 2006). This leaves the

iterative procedure. On top of that, this procedure also has limitations where it does not provide

accurate results; “the iterative procedure is fast and works well, provided that a wave-front is

free of triplications” (Vavryčuk, 2007). The wave-front triplication only occurs with the qSV

wave, which will be shown later.

The iteration is searching for a complex slowness direction vector n generating a

homogeneous complex ray velocity v , whose real and imaginary parts have a real ray direction

N . The initialization is by calculating the Christoffel tensor  jk (n)  aijkl ni nl with the complex-

valued moduli aijkl  aijkl


( R)
 iaijkl
(I )
and a complex slowness direction:

34
 sin( R  i I ) cos( R  i I ) 
 
n   sin( R  i I )sin( R  i I )  . (4-1)
 cos( R  i I ) 
 

Subsequently, the Christoffel tensor has three eigenvalues and three eigenvectors each to be

calculated through equations (2-38) and (2-37), which implicitly suggests:

cm2  aijkl ni nl gˆ (jm) gˆ k( m )


. (4-2)

The eigenvalues ( cm ) and equation (2-6) may then be used to calculate the complex slowness

( pi( m)  ni / cm ). Having done that, it is a direct matter of simply computaion by applying

equation (3-14). The ray velocity vector calculated by equation (3-14) is generally

inhomogeneous. Additionally, it can be normalized as it is along the ray-path to obtain N

through ( v  Nv ). The method is referred to as an iterative inversion because it governs the

inverse approach of the above stated procedure. There are two very similarly-working branches

which can be followed to inversely iterate to find the directional vector n , and the misfit

function explained by Vavryčuk (2008b). One, is when a ray direction vector N is fixed and

then the actual search is for two parameters: the two angles (  ( I ) ,  ( I ) ) describing the

imaginary part of the slowness direction n , with given angles (  ( R ) ,  ( R ) ) the other two

describing the real part’s. Consequently, the slowness direction n with its complex

components ( n ( R ) , n ( I ) ) is obtained. Henceforth, the rest of the parameters are orderly obtained

(the Christoffel tensor, the stationary slowness and the ray velocity vectors, etc.), and then

through the several iterations, the to-be optimized misfit function could be constructed by

setting a condition relating the initially fixed ray vector N and the derived ray vector and the

difference between them (see the flowchart Figure 11).

35
The other branch is to define the slowness direction as,

n  zˆ1n( R )  zˆ2n( R,) , (4-3)

where n ( R ,  ) is the unit vector perpendicular to n ( R ) , and ẑ1 , ẑ2 are complex values defined

by:

zˆ1  cos   i sin 


(4-4)
zˆ2    i 

Considering n  n  1 , it is possible to determine the variables  and  , i.e.

1  cos 2 1  cos 2
  ,
2 2
 sin 2 (4-5)
 .
1  cos 2 1  cos 2
2 
2 2

Accordingly, given any angle  , one is capable of obtaining ẑ1 and ẑ2 , hence knowing n ( R )

and n ( R ,  ) . Therefore, one only needs to search a single angle  for the homogeneous ray

velocity vector given by equation (3-14).

At this point, the two branches analogously converge as the coming step is to calculate

the remaining parameters (the Christoffel tensor, the stationary slowness and the velocity

vectors, etc.). This time around however for the second approach, the aim of the to-be

optimized, through the several iterations, misfit function is to minimize the imaginary part of

the ray direction vector N in order to end up with the real-valued component alone –after

searching and varying the angle–. And when that condition is satisfied, the iteration halts and

one last run through the calculations is done with the optimum values: starting from the

36
searched angles up to calculating the real ray velocity V ray , the real ray attenuation Aray , and

the real ray quality factor Q ray using equations (3-19) ~ (3-21).

These methods are thought to be more accurate and more dependable when compared

with one of the predecessors, the aforementioned method utilizing the polynomial equations,

which is considered to have a high possibility of producing some unreliable results (Vavryčuk,

2006; Vavryčuk, 2008b). So, this was the alternative concept; to deal with wave energy in its

ray-path carried by higher frequency ranges of a stationary slowness direction knowing the ray

vector direction with an iterative inversion. Figures 11 and 12 show and summarize the general

workflows of these two branches, and where the iteration element actually is.

Utilizing this method for the ultimate objective of ray tracing in viscoelastic anisotropic

media is reasonable. The iterative method incorporates less assumptions; there is only one

assumption, which dictates that the complex ray velocity vector is homogeneous i.e. there is

only the real ray direction to be taken into account as both the real and imaginary directions

are parallel ( n  n ( R ) ). This is highly advantageous for this method. In fact, this is the particular

reason why this certain method is used to generate the solution by which we benchmark

modelling results in order to compare any other methods’ against. Additionally, the results

proved to show high accuracy for the qP-wave even in cases of strong attenuation, as well as

at the interfaces of two media (Vavryčuk, 2008a real ray; Vavryčuk, 2010). Having stated that,

the iterative process is considered to be a time-consuming process due to the fine step required

for searching for the solution of the complex slowness direction for VTI media for example at

the cusps of qSV, and fails for the qSV- and qSH-waves in a highly complex VEAM, for instance

the ORT (orthorhombically anisotropic) medium. For this reason among others, other methods

are explored to tackle and overcome these limitations.

37
Figure 11: A flowchart summarizing the first branch of the iterative solver algorithm steps, focusing on the iterative element,
to obtain the ray velocity, attenuation and quality factor

38
Figure 12: A flowchart summarizing the second branch of the iterative solver algorithm steps, focusing on the iterative element,
to obtain the ray velocity, attenuation and quality factor

39
4.2 Perturbation method

This method is, as the name suggests, based on the concept of an anomalous

perturbation happening in the vicinity of normal conditions (Figure 13). The assumption here

is basing on a hugely complex ray vector based on considering the normal/default conditions

being the elasticity of the media and the disturbance (hence, perturbation) from that is a

component that is behaving in a viscoelastic manner. There is a key assumption here that is

made and is essential in order for this concept to yield valid results; the assumption is that the

complex ray velocity vector is homogenous, in other words, for this complex vector quantity,

the real part’s vector’s direction is identical to the imaginary part’s vector’s direction. This

might not always be the case but it makes for a decent approximation.

Figure 13: Simplistic illustration of the concept of perturbation. Properties of the background are summed with properties of
a distinct disturbance deviating from the background, resulting in an aggregate of the two properties

The following methodology is based on detailed calculations and formulations from

which the thorough work is a work-in-progress lead by Dr. Jianlu Wu. Concisely, the

Hamiltonian function is employed alongside the phase slowness vector –in conjunction with a

few approximations and conditions– to derive the eigenvalues representing the phase velocity

vector, and eventually the final form of the ray velocity vector. Moving on with that, the

magnitudes of the ray velocity vector, ray attenuation, and ray quality factor can be calculated.

The Hamiltonian function is a function of the position vector x , the slowness vector p , and

the elastic moduli aijkl , and it is equal to zero in a viscoelastic medium (Červený, 2001):

40
H  x, p, aijkl   0 (4-6)

The point here then, as mentioned above, is to split the response of the viscoelastic medium

into the background elastic behavior and the perturbing viscoelastic behavior as the figure

above demonstrated; the total elastic moduli becomes: aijkl  aijkl  iaijkl , and the total
(0)

slowness becomes: p  p(0)  ip . Note that both the background and the perturbation are dealt

with as anisotropic media. Hereafter, the Hamiltonian function and its components in turn are

considered as a summation of the two responses as such:

H  x, p, aijkl   H ( 0)  x, p ( 0) , aijkl
(0 )
  iH  x, p, aijkl  (4-7)

(0)
Accordingly, the first part denotes the elastic parameters ( aijkl , p (0) ) and the latter part refers

to the complex parameters ( aijkl , p ). The elastic slowness is clearly a real number which

can be calculated simply because of all the real quantities of the H function (Zhou and

Greenhalgh, 2004). Another note to add relating to the slowness, is that the resulting

perturbation portion is minor in magnitude in comparison with its real complement. As

Vavryčuk (2008a) demonstrated, the Hamiltonian equates to zero as it is plugged and

substituted in the Eikonal equation cited implicitly in equation (2-38),

H (0)  x, p (0) , aijkl


(0)
  det(aijkl(0) pi(0) pl (0)   ij )  0 (4-8)

H  x, p, aijkl   H  x, p (0) , p, aijkl


(0)
, aijkl   0 (4-9)

The same paper, and others (Červený, 2001) also proves how the Hamilton equation is directly

linked to the ray velocity vector which is sought after. The link is found by partially deriving

the Hamiltonian function with respect to the slowness from the Eikonal equation. Applying

41
that will yield a ray velocity vector of two parts, the real part arising from the background

response ( vi ) and an imaginary part arising from the perturbation response ( vi ),
(0)

H  x, p, aijkl  H (0)  x, p(0) , aijkl


(0)
 H  x, p, aijkl 
vi   
pi pi pi (4-10)
 vi(0)  p(0) , aijkl
(0)
  ivi p(0) , p, aijkl(0) , aijkl  .

Both parts of the ray velocity are considered to have the same direction as mentioned earlier to

preserve the assumption of the homogeneity of the complex vector. By this, it is possible to

calculate all the parameters and finally obtain the ray velocity vector ( vi  aijkl pl g j gk ).

Subsequently, and as long as that the two components, the real and the imaginary, of the ray

velocity are known, the vector magnitude V ray can be calculated, as well as the attenuation

magnitude Aray , along with the associated quality factor Q ray as demonstrated previously

(equations (3-19) ~ (3-21)) but not before calculating from the here mentioned two velocity

components the magnitude of the complex ray velocity vector as,

N v/v 
 vR N R  iv I N I 
v (4-11)
v  vR  ivI  vi vi  v  v  2ivR vI ( N R  N I )
2
R
2
I

The main advantage of this method is that it enables the direct calculation of the

imaginary parts of the slowness vector, ray velocity, and quality factor of all of the three wave

modes, the qP-, qSV- and qSH- waves. Up to here, is the broad demonstration of the method

and its procedure for any general case. This could be further developed and extended to be

applied for specific media, like VTI and other types of transversely isotropic/anisotropic media

such as HTI, TTI, and ORT, all of which is in the viscoelastic domain. Here, VTI and ORT are

discussed.

42
For the modes qP and qSV in the VTI medium, they are dealt with separately from qSH

because the former two depend on the elastic moduli a11 , a33 , and a44 , while the latter only

depends on a44 and a66 , hence the equations differ. Accordingly, the Hamiltonian function for

the former two modes (qP, qSV) becomes

H  x, p    a44  a11   px2    a33  a44  pz2  a11a44 px4  a33a44 pz4  Dpz2 px2  1  0

D  a33  a11  a33  2a44    a13  a44    a33  a44 


2 2

(4-12)

where D is the coefficient of the Hamiltonian function which is entirely dependent on the

elastic moduli alone. Referring back to the totals of the elastic moduli and slowness vector, the

two Hamiltonian function components, the background and the perturbation of equation (4-9)

become

H (0)  x, p, aijkl   det(aijk


(0)
l pi
(0)
pl (0)   ij )  0 (4-13)

H  x, p   W0  W1px  W2 pz  0 (4-14)

where,

W0   a11  a44   px(0)    a33  a44   pz(0)    a44 a44 


2 2
(0)
a33  a33
(0)

 p    a a  a a  p   D  p   p 
(0) 4
z
(0)
44 11
(0)
11 44
(0) 4
x
(0) 2
x
(0) 2
z ,

W1  2  a  a  p  4a a  p   2 D p  p  ,
(0)
11
(0)
44
(0)
x
(0) (0)
11 44
(0) 3
x
(0) (0)
x
(0) 2
z

W2  2  a  a  p  4a a  p   2 D  p  p .
(0)
33
(0)
44
(0)
z
(0) ( 0)
33 44
(0) 3
z
(0) (0) 2
x
(0)
z

(4-15)

In the same way, referring to the ray velocity in terms of the Hamiltonian function derivative,

the complex ray velocity components corresponding to equation (4-10) become,

43
vx  H  x , p  px  vx(0)  ivx  vx(0)  i  A1  A2 px  A3pz 
(4-16)
vz  H  x , p  pz  vz(0)  ivz  vz(0)  i  B1  B2 px  B3pz 

where,

vx(0)  2  a11(0)  a44  px(0)  4a11(0) a44(0)  px(0)   2D(0) px(0)  pz(0)  ,
(0) 3 2

A1  2 px(0)  a11  a44   4  px(0)   a44 a11  a11(0) a44   2 p x(0)  p z(0)  D;
(0)3 2

A2  2  a11(0)  a44   2 D(0)  pz(0)   12a11(0) a44(0)  px(0)  ,


(0) 2 2

A3  4 D (0) px(0) pz(0) ;

(4-17)

and

vz(0)  2  a33  pz(0)  4a33(0) a44(0)  pz(0)   2D(0)  px(0)  pz(0) ,


3 2
(0)
 a44
(0)

B1  2 pz(0)  a33  a44   4  a44 a44  pz(0)   2  p x(0)  p z(0) D,


3 2
(0)
a33  a33
(0)

B2  4 D (0) px(0) pz(0) ,


B3  2  a33   2D(0)  px(0)   12a33(0) a44(0)  pz(0)  .
2 2
(0)
 a44
(0)

(4-18)

Employing the Hamiltonian function perturbation and the ray velocity definitions above, along

with the identity vx vz  vx vz , which arises from the condition of the complex ray
(0) (0)

velocity vector homogeneity (the real and the imaginary parts of the direction vectors are

parallel), the perturbation slowness vector p  (px , pz ) , and the ray velocity vector

v  (vx , vz ) are finally obtained. This covers the details associated with the qP and qSV wave

modes. For qSH, the case is similar, however with different parameters. The Hamiltonian

function perturbation and the complex ray velocity components in terms of the Hamiltonian

function are similarly governed by the equations (4-13) and (4-14), where this time, the

variables are defined differently as follows,

44
W0   px(0)  a66   pz(0)  a44 ,
2 2

W1  2a66
(0) (0)
px , W2  2a44
(0) (0)
pz ,
vx(0)  2a66
(0) (0)
px , A1  2 px(0) a66 , (4-19)
A2  2a66
(0)
, vz(0)  2a44
(0) (0)
pz ,
B1  2 pz(0) a44 , B2  2a44
(0)
.

And once again, similarly like it was for the other two wave modes’ case, and using the same

identity from the homogeneity condition of the ray velocity, it is a matter of plugging in the

equations to obtain the perturbation slowness vector, and the ray velocity vector, associated

with the qSH-wave.

As for the ORT media, the subject is a bit more involved and complicated as the elastic

moduli needed for describing this category of anisotropic media are nine independent variables

( a11 , a12 , a13 , a22 , a23 , a33 , a44 , a55 , a66 ) out of the total of twenty one independent variables

(equation 2-26) which could describe any media in general. The Hamiltonian function in this

case reads,

H (x, p)  D1 px6  D2 p 6y  D3 pz6  D4 px4 p y2  D5 px4 pz2  D6 p x2 p y4


+D7 p y4 pz2  D8 px2 pz4  D9 p y2 pz4  D10 px2 p y2 pz2  D11 px4
+D12 p y4  D13 pz4  D14 px2 p y2  D15 px2 pz2  D16 p y2 pz2
 D17 px2  D18 p y2  D19 pz2  1

(4-20)

where the vector of the coefficients of the Hamiltonian function Di is constituted of the nine

elastic moduli and the can be found in details in Appendix A. Henceforward, the perturbation

of the Hamiltonian function becomes,

H  x, p   W0  W1px  W2 p y  W3pz  0 , (4-21)

45
where the vector of the coefficients of the perturbation of the Hamiltonian Wi is dependent on

the slowness and the coefficients of the Hamiltonian Di , and can be found in details in

Appendix B. Finally, the complex ray velocity and in its components form show as,

vx  H px  vx(0)  ivx  vx(0)  i  E1  E2 px  E3p y  E4 pz  ,

y  iv y  v y  i  F1  F2 p x  F3 p y  F4 p z  ,
v y  H p y  v (0) (0)

vz  H pz  vz(0)  ivz  vz(0)  i  G1  G2 px  G3p y  G4 pz  ,

(4-22)

where these ray velocity coefficients are dependent on the slowness and the coefficients Di ,

and can be found in details in Appendix C. Figure 14 below summarizes the procedure above

broadly. In short, it starts from the elastic moduli inserted in to the Hamiltonian function. Then,

the derivative of that function with respect to the slowness gives the ray velocity as the literature

have proven. Finally, the total slowness (real and imaginary parts) is found along with the ray

velocity, which is employed to eventually obtain the real ray velocity V ray , the real ray

attenuation Aray , and the real ray quality factor Q ray as equations (3-19) ~ (3-21) demonstrate.

To ensure the validity of this method’s results, two essential assumptions must be

considered. The first assumption is that the complex ray velocity vector is homogenous. The second

is that the Hamiltonian function perturbation is linear, which limits the applicability of the method

to weak anisotropy (as the aijkl is relatively very small), and that will be investigated and verified.

The main advantage of this method is that, on top of having significantly reduced computational

expenses due to the lack of an iterative search for the slowness angle, it promotes the direct

calculation of the imaginary parts of the slowness vector, ray velocity, and quality factor of all of

the three wave modes, the qP-, qSV- and qSH- waves through the Hamiltonian perturbation function.

In addition, this method is expected to more accurately solve for the real ray velocity and also better

tackle less complicated media and wave-fronts and shape the optimum solution for such settings.

46
Figure 14: A flowchart summarizing the perturbation method steps to obtain the ray velocity, attenuation and quality factor

47
4.3 Real Slowness Direction approximation

This approach could be thought of as similar to its previous counterpart when examined

from a certain point of view. A comparable assumption is put in place in order to simplify the

problem, so long as the results are valid and the approximation does not lead to the violations

of base principles. As stated previously, and generally but specially in viscoelastic anisotropic

media, all the complex vector parameters could be thought of as being inhomogeneous as both

of the two components –the real and the imaginary– could very well have different directions.

However, and for the purposes of this method, the slowness vector is chosen to abide by the

homogeneous complex vector condition. This facilitates neglecting dealing with the

inhomogeneous state of the slowness vector, and on top of that, enables the direct

approximation of the complex ray velocity which is similarly going to be bound under the

condition of the homogeneous complex ray velocity vector. The team dubbed this approach the

Real Slowness Direction (RSD) approximation, in reference to the fact that the direction of the

slowness vector is only one instead of two, and it is by default denoted by the real direction,

the imaginary direction is substituted by the real one (but it is established that the real direction

is identical to the imaginary direction).

Generally, the slowness vector is defined by a real component, an imaginary

component, a direction for each component, and a complex phase velocity. Decomposing it

from equations (2-5) ~ (2-6) and forcing homogeneity on the slowness vector ( n ( R )  n ( I ) )

yields:

n( R )
p  p ( R )n ( R )  ip ( I )n ( I )  , (4-23)
c

1
p  p ( R )  ip ( I )  , (4-24)
c

and

48
cR
p( R)  n( R ) ,
c  cI
2
R
2
(4-15)
c
p (I )
 2 I 2 n( R) .
cR  cI

Hence, applying the homogeneous condition ( n ( I )  n ( R ) ), both components of the slowness

vector are defined by the same direction (which is taken to be the real one), where generally

the two direction could very well be unequal. Thus, by means of substitution, a simple

relationship can be established between p ( R ) and p ( I ) :

cI ( R )
p( I )   p (4-16)
cR

The slowness vector could also be defined in another way; associated with the wave travel-

time ( p   ). And the travel-time itself is in turn constitutional in the definition of ray

velocity by simple physics by means of a partial derivative involving the ray coordinates (recall

equations (3-10) and (3-15)). The ray velocity vector is a homogeneous complex vector too,

forcibly. The above leads to the deduction that the slowness vector and the velocity vector are

parallel following the famous identity of equation (3-11) ( p  v  1 ) (Vavryčuk, 2008a). This is

the main distinction between this method and the perturbation method. This identity is not used

as a condition constraining the solution. Having established this identity formula, expanding in

terms of direction vectors results in equation (3-12) but only in the real domain as we forced

the homogeneity condition,

cR
(n ( R )  rˆ )  . (4-27)
c
vR [1  ( I ) 2 ]
cR

This suggests that the direction of the slowness n and the direction of velocity/ray-path r̂ are

not aligned, in proportion to the difference of the phase velocity cR and the ray velocity vR ,

which is normally the case in VEAM. Through the above identity as well, it is possible to

purposefully derive further formulations as such,

49
p  v  1  (p ( R )  ip ( I ) )  (vR  ivI )rˆ  1
cI ( R )
 (p ( R )  i p )  (vR  ivI )rˆ  1
cR
cI
 (1  i )(vR  ivI )(p ( R )  rˆ )  1
cR
(4-28)
c c n( R)
 (vR  I vI )  i (vI  I vR )(  rˆ )  1
cR cR cR
cI cI n ( R )  rˆ
( vR  vR )( ) 1
cR cR cR
cI 2 c
[1  ( ) ](n ( R )  rˆ )  R ,
cR vR

which enables the discovery of yet another relationship used amidst expanding the terms:

cI
vI  vR
cR (4-29)

relating the real and imaginary parts of both the phase and the ray velocities. The advantage of

this equation is that it imposes and satisfies the homogeneity condition when calculating the

real and imaginary ray velocities. Having known these quantities of the ray velocity

components, the velocity vector magnitude V ray , the attenuation magnitude Aray , along with

the associated quality factor Q ray can straight away be calculated as well, as priorly

demonstrated through equations (3-19) ~ (3-21). The procedure is summarized in the following

flowchart (Figure 15).

As the name of the method indicates, we regard only the real slowness direction and

enforce homogeneity on this vector. This limits the overall reach of the method however,

enables the direct approximation of the complex ray velocity, and also enforces its

homogeneity. This distinguishes RSD as constraining condition p  v  1 is observed

particularly, and it ensures the parallel relationship between the two vectors. This method

optimizes computation time as well as it is suitable for targeting more complex media with

stronger anisotropy and higher attenuation, even for the problematic tripling of q-SV wave-

front having cusps with several spatial values for a single ray direction.

50
Figure 15: A flowchart summarizing the RSD approach steps to obtain the ray velocity, attenuation and quality factor

51
5. Numerical Results

5.1 VTI models

The easiest natural VTI media that one would think of as fitting in a geological sense

complying with vertical symmetry is the shale lithology. Shale is naturally layered with

horizontal planes overlying each other (principle of original horizontality), so long as no further

deformation caused any inclination after sedimentation. There are many natural geological

phenomena which could result in encountering VTI media in the environment because there

are many sources of transverse isotropy. We assume that we have a viscoelastic perfectly VTI

shale with the following properties (Table 1):

Table 1: Elastic moduli and quality matrix of a viscoelastic VTI shale

Moduli a11 a13 a33 a44 a66

15.064 1.639 10.837 3.126 4.251


R km 2
a pq ( )
s
q pq 20 25 30 20 15

Using the iterative method workflow (Figure 11), the above information is used to construct,

for the three wave modes (qP, qSV, qSH) at different ray directions, their real ray velocity, real

ray attenuation, and ray quality factor based on the given elastic moduli. The phase quantities

are also obtainable easily as long as the phase velocity is found. Since the medium is symmetric,

the results are displayed in half a 2D plane due to the reflection of the results on the other half

(Figure 16).

52
Figure 16: Ray and phase velocities (first row), ray and phase attenuations (second row), and ray and phase quality factors
for the three wave modes: qP (first column), qSV (second column), qSH (third column) for a shale model of a viscoelastic VTI
medium

The three wave modes slightly behave similarly. The qP-wave’s behavior in terms of each pair

of quantities in both realms is quite regular, as it is the case for the qSH-wave whose phase and

ray quantities are almost inseparable. Both of the velocities certainly do change with the

changing directions, however, they both converge the more we move away from the center and

get closer towards the horizontal and the vertical axes where they meet and become equal,

which is the similar case reported in elastic VTI media by Thomsen (1986). The same behavior

is also noticed in the case of the attenuation and the quality factor. As for the qSV-wave, it

53
always exhibits bizarre shapes as that is due to the aforementioned triplication of wave-front

the wave undergoes from its natural structure. In this instance and concluding from the

supporting plot, for single ray directions there is more than one value for these wave quantities

and that explain the cusps and twists of the shape that is a signature of the qSV-wave alone.

With this, there is proof that the iterative algorithm provides satisfactory results when it comes

to finding the complex ray velocity and the slowness vectors to deduce the three wave modes’

phase and ray quantities in a viscoelastic purely VTI medium such as shale.

In addition, eight other models are used to verify the efficiency of the perturbation

approach and the RSD method. The Thomsen models’ parameters’ can be found below (Table

2) (Hao and Alkhalifah, 2017):

Table 2: Elastic moduli and quality matrix of 8 viscoelastic VTI models

Upon examining the values, models belonging to the same category (either A or B) have the

same elastic moduli but different quality factors, with Model A1 being the most complex with

the highest anisotropy and attenuation. Vavryčuk (2008b) derived the analytic solutions which

our numerical solutions will compare against. The deviation or the difference between the two

will be measured by an absolute value of the relative error between the numerical and the

54
analytical results ( V ray ,  Aray ,  Qray ). Figure 17 below shows the results of the 3 ray quantities

using the two methods against the exact solution with the error plotted as well, and that is for

the qP-wave in Model A1. These two methods achieved solutions that are in excellent

agreement with the exact solutions, even though this model is the most complex out of the 8.

The maximum errors of the ray velocity and the quality factor are 0.01% and 0.22%

respectively for the perturbation approach. As for the attenuation, the errors are comparable to

the quality factor (0.21%). The RSD slightly falters behind for the velocity but is on par for the

other two quantities in terms of maximum errors. In Figure 18, however, the results are even

brighter with much reduced maximum errors due to the weaker anisotropy of Model B1,

although the same level of attenuation. The error reduction is more than the double for the

perturbation method, while it is minimal for RSD. Thus far, the perturbation method yielded

slightly better results than the RSD approximation, and by examining Table 3, one could

observe that RSD only performs better than its peer in the four A Models (having stronger

anisotropy) only with respect to the quality factor. Overall, the perturbation method is superior

almost always, and the RSD is only faintly sensitive to anisotropy.

Figure 17: The computed ray velocity, attenuation, and quality factor of the qP-wave in Model A1. The absolute relative error
is also plotted with respect to each method in comparison with the exact solution form the literature

55
Figure 18: The computed ray velocity, attenuation, and quality factor of the qP-wave in Model B1. The absolute relative error
is also plotted with respect to each method

Table 3: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qP-wave

As for the other wave modes, Figures 19 and 20 show the same as Figures 17 and 18 (Models

A1 and B1) but with respect to the qSV-wave. A remarkable observation needs to be made here

as the RSD method succeeds in better estimating the complicated tripling wave-front of the

qSV-wave, where other methods fall behind. Table 4 explicitly shows how the maximum errors

of RSD are significantly less, by orders of x10, than those of the perturbation method when it

comes to the qSV-wave. We continue to observe the insensitivity of RSD to the degree of

anisotropy.

56
Figure 19: The computed ray velocity, attenuation, and quality factor of the qSV-wave in Model A1. The absolute relative error
is also plotted with respect to each method

Figure 20: The computed ray velocity, attenuation, and quality factor of the qSV-wave in Model B1. The absolute relative error
is also plotted with respect to each method

57
Table 4: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qSV-wave

The remaining wave mode is the horizontal shear wave and its results for Model A1 are given

in Figure 21, and for Model B1 given in Figure 22. The accuracy of both methods proved to be

generally higher and better for the qSH-waves when compared with the other two modes. All

the while, the RSD performed generally much better than the perturbation approach with

respect to the ray attenuation and ray quality factor, however, using the perturbation approach

is far more accurate for estimating the ray velocity. The maximum absolute errors of the two

methods for all models particularly for qSH are reported in Table 5.

Figure 21: The computed ray velocity, attenuation, and quality factor of the qSH-wave in Model A1. The absolute relative
error is also plotted with respect to each method

58
Figure 22: The computed ray velocity, attenuation, and quality factor of the qSH-wave in Model B1. The absolute relative
error is also plotted with respect to each method

Table 5: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qSH-wave.

59
5.2 TTI model

A tiltation of rock units is very common and can be observed widely, obviously in

tectonically active areas, but even in passive margins and other inactive regions. For example,

clinoforms highly mimic tilted transverse isotropy conditions, assuming that each single layer

is homogeneous; that yields a TTI model which simplifies reality but is at least the best

approximation. Other than that, having a TTI medium could be as simple as applying an

inclination, and/or azimuthal, dip to a VTI medium. Which is the case here as we attempt

employing the iterative solver on a more complicated model, we use the same VTI medium

model in Table 1 to estimate a TTI medium by applying a dip in 3D space through the Bond

transformation (Bond, 1943). TTI is considered a more anisotropic case than VTI which

naturally makes the problem more complicated as more variables enter the equation and

therefore have to be taken into account, on top of the trigonometric relations which have to be

made to work with the two reference coordinates in space. But for simplicity sake, the

inclination angle 0 is arbitrarily chosen to be 30, while the azimuthal shift is considered to

be zero ( 0  0 ); the image to be visualized is much like Figure 8 describes. Figure 23

subsequently shows the results of the TTI model in the same fashion as Figure 16 does with

the VTI model, for the 3 wave modes and wave quantities plots. There are no major differences

in the graphs when comparing them to the VTI case, similar trends can be readily observed but

with a tilt corresponding to the axis tilt; the cusps in the qSV-wave-front, and the convergence

at the vertical and horizontal axes of most of the quantity pairs. However, slight changes have

to be noted too where the pairs of quantities of different modes coincide in less space and the

contrast between them becomes more apparent. This is probably due to the higher level of

complexity of the medium which unsurprisingly causes irregularities and highlights the

accuracy of ray quantities over phase quantities, this is not the main aim but worth to not. The

qSH-wave though seems to always have the most regular behavior in all cases with all

60
quantities. Once again, obtaining such results validates the iterative solver as a satisfactory

method which works reasonably well with somewhat complex conditions of anisotropy.

Figure 23: Ray and phase velocities (first row), ray and phase attenuations (second row), and ray and phase quality factors
for the three wave modes: qP (first column), qSV (second column), qSH (third column) for a tilted shale model of a viscoelastic
TTI medium based on the Bond transformation

61
5.3 ORT models

Orthorhombic anisotropy is not often discussed or studied as it is considered highly

complex, and is in need of 9 independent elastic moduli to fully characterize it. On top of that,

adding the viscoelasticity complicates the subject further. However, as new techniques arise

and computing power enhances, the more complicated a model is, the more probable it is closer

to reality than what we simplify it as. Hence, it would be a good challenge to try and tackle this

problem with a new approach. From Hao and Alkhalifah (2017), the model parameters are

given in the Table 6. There are three models which will be used to validate the results of the

perturbation and the RSD methods as the analytic solution is available and proved by Vavryčuk

(2008b). As previously, the aim is to calculate the ray quantities and then compute the error

between the numerical and the analytical solutions.

Table 6: Elastic moduli and quality matrix of 3 viscoelastic ORT media

Figure 24 shows the errors of the computed ray velocity, attenuation, and quality factor using

the two proposed methods for the qP-wave mode for Model C1, which is anisotropic with the

strongest attenuation. Table 7 summarizes the maximum errors for these ray quantities of the

qP- wave mode for the all of three models. In light of these results, one may conclude that the

results obtained by the RSD method have higher accuracy where the error is less than half a

percent than those obtained by the perturbation approach in almost all cases, particularly for

62
ray attenuation and quality factor for all of the three models. For the ray velocity however the

two methods’ results are quite similar with the perturbation method performing slightly better.

Figure 24: The absolute relative errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qP-wave in Model C1

Table 7: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qP-wave

For the two shear waves, the qSV- and qSH-wave modes' results for Models C1 and C2, which

exhibit stronger attenuation, it is found that the iteration algorithm proposed by Vavryčuk

(2008b) is time consuming because of the choice of a fine step of the complex slowness angle

for searching for an acceptable slowness vector for a given ray direction. And on the other

hand, the perturbation method results in unacceptable large errors of more than ten percent for

63
the two shear modes for all the models, in contrast with the tolerable order of errors. Having

made that statement, the displayed results are only of the RSD approach which offered much

more appropriate outcomes at least for the weakest attenuation model C3. From these results

one can see that highest errors are commonly around 0.1% except for some which reach about

1~5% (especially for the complex qSV), and that is unavoidable and occurs as a consequence

of the triplication of the qSV-wave-fronts (Figures 25 and 26). Overall, these results confirm

that the RSD approach is an effective and efficient technique to follow for dealing with

arbitrary ORT media.

Figure 25: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qSV-wave in Model C3

Figure 26: The maximum absolute errors of the ray velocity, ray attenuation, and ray quality factor of the two methods for the
qSH-wave in Model C3

64
6. Conclusions

In the grand-scheme of things, the focus of the study is but a pathway for the next step.

The main topic is seismic ray tracing, and the ultimate objective in this case is seismic ray

tracing in viscoelastic anisotropic media. Many studies have delved into studying elastic

isotropic media, which might be fine for certain purposes but not all. It superimposes

oversimplifications which inevitably lead to approximations and therefore definite levels of

errors. Thus, there is a need to develop more and more efficient ways to achieve and deal with

the complexity of reality.

Here, the problem is the accuracy of the ray tracing through the subsurface, which

reflects back as the image of the underlying structures. If the directed waves are not actually

behaving as expected, then the resulting modelling and estimation could vary and might very

well be misleading. When we make the assumption that the earth is elastic and does not

attenuate wave energy as it is travelling through it, then the received energy is going to tell a

whole different story than anticipated. So, in order to do accurate and correct ray tracing,

several parameters need to be studied and obtained; the ray velocity, ray attenuation and ray

quality factor. Not to mention that the calculations become held in the complex domain to

account for the attenuating wave energy. Naturally, these quantities are complex-valued due to

viscoelasticity and the associated attenuation. The ray velocity describes the speed of a wave’s

amplitude propagation, while the ray attenuation describes that amplitude’s decay as the wave

propagates through a viscoelastic medium. The ray quality factor is essentially a measure of

the strength of the attenuation; the lower the quality factor is, the more the attenuation to be

expected is.

What I presented are three methods by which a step closer to more accurate ray tracing

takes place. In order to conduct ray tracing, several of the waves’ as well as the mediums’

properties should be known, but most importantly, three ray quantities which are addressed

65
extensively above: the real ray velocity, the real ray attenuation, and the real ray quality factor.

Phase quantities are also important and could be as important but due to some aspects related

to applicability, some choose not to involve them.

To validate the results of the presented methods, which are numerical approximation

attempts at being comparable with the existing analytic solutions, we compare the outcomes

and measure them against the exact solution and examine the relative errors. The methods are

the iterative algorithm, the perturbation method, and the real slowness direction approach.

Some methods work for some media better than others, and some methods work for some wave

modes better than other.

We demonstrate the success in the application of the iterative method to compute the

complex ray velocity which gives the real ray velocity and ray attenuation in a homogeneous

VEAM, namely VTI and TTI. The comparison of the computational expenses for different

media shows that the computing time consumption is increased due to the process of iteratively

running to determine the homogeneous ray velocity vector in a specific direction. It would be

more efficient for a direct solver or a parallel computing scheme to be implemented which will

improve the computational efficiency, or even utilizing preconditioners. In addition, and

naturally, the more complex the medium is, the more time it takes to arrive at a convergent

solution. However, and overall, the iterative method is quite accurate in producing reliable

results –for simple and more complex anisotropic media– as it does not incorporate too many

constraining assumptions. This makes it the scale which we can compare the results of other

methods against to assess their validity; as emphasized, the main pitfall of iteration is that it is

time consuming.

As for the perturbation method and the RSD method, they seem to provide highly

satisfactory results, which approach the exact solution with minimal errors in most of the cases.

Even for the different wave modes, and even for moderate and strong viscoelastic VTI media.

66
There are two dimensions which each method should be assessed by; the performance in

handling the degree of complexness of the media (denoted by the strength of anisotropy and

attenuation), and the performance in handling the different wave modes (qP, qSV, qSH) with

the vertical shear wave being the most problematic as it always exhibits wave-front triplication.

The perturbation method excelled in resolving the 3 ray quantities, especially the ray velocity,

for qP and qSH wave modes in the most complex VTI medium model, yet it disappoints for

qSV as the wave-front triplication kicks in and challenges this method. Additionally, for the

more complex ORT media, the results suggest that using the perturbation method is only

satisfactory for the qP-wave, and that is also only for the ray velocity, while it is notably less

effective for highly complex media for both shear waves. Therefore, this method is only

optimal for simpler media and specifically qP-waves, and also particularly for the ray velocity

and not ray attenuation nor ray quality factor.

The RSD approach is quite peculiar, it performs better for more complex cases than it

does for simpler cases. For example, the results indicate that for VTI’s qP wave mode, RSD

yields just reasonable solutions for ray quality and sometimes ray attenuation. On the other

hand, specifically for qSV, RSD always proves to be the best approach to follow to obtain more

trustworthy results. However, it is worthy to note that these most of the results strongly suggest

that this method is insensitive to anisotropy, as several models with varying anisotropy degrees

share very similar accuracies. What is more promising is that although the qSV-wave is

generally problematic due to its wave-front triplications, that was handled well especially by

the RSD approach. However, for the most complicated media tackled here, the viscoelastic

orthorhombically anisotropic media, it was discovered that the perturbation method fails in

computations of both of the shear waves –as stressed above– due to the wave-front triplication.

Moreover, the iterations make for extensive computational time, which is disadvantageous.

Fortunately enough, in these cases, the RSD approach becomes available to efficiently obtain

67
highly acceptable results for all three wave modes (qP, qSV, qSH). Mainly, this method is the

optimal choice when aiming to obtain accurate solutions for higher complexity media, for qSV

for all anisotropy and attenuation cases, and specifically for ray attenuation and ray quality for

less complex media for qP and qSH. However, it still requires further development to be able

to tackle strongly anisotropic and especially, strongly attenuating media.

Overall, the perturbation approach and the RSD method both show significant results

and exhibit notable success, each for certain settings and specific targets . They surely work

purposefully and yield better results at lower expense (as they cost less computing efforts to

run when compared with the iterative algorithm for example). Both can be applied to seismic

ray tracing in VEAM and yield reliable results depending on the user’s desired application and

goals.

68
References

Aki, K., Richards, P.G., 1980. Quantitative Seismology: Theory and Methods. W.H. Freeman

and Co., New York.

Auld, B. A., 1973, Acoustic fields and waves in solids. JohnWiley & Sons, Inc.

Behura, J., I. Tsvankin, E. Jenner, and A. Calvert, 2012, Estimation of interval velocity and

attenuation anisotropy from reflection data at Coronation Field, The Leading Edge, 31, 580–587.

Bond, W. L., 1943, The mathematics of the physical properties of crystals, The Bell System

Technical Journal, 22, 1–72.

Burton, N., 2007, Rock quality, seismic velocity, attenuation and anisotropy: Taylor & Francis.

Carcione, J.M. & Cavallini, F., 1995. Forbidden directions for inhomogeneous pure shear

waves in dissipative anisotropic media, Geophysics, 60, 522–530.

Carcione, J. M., 2000, A model for seismic velocity and attenuation in petroleum source rocks,

Geophysics, 65, 1080–1092.

Carcione, J.M., 2007. Wave Fields in Real Media: Wave Propagation in Anisotropic, Anelastic,

Porous and Electromagnetic Media, 38(2), Elsevier.

Carcione, J.M., 2014. Wave Fields in Real Media: Theory and Numerical Simulation of Wave

Propagation in Anisotropic, Anelastic, Porous and Electromagnetic Media, (3), Elsevier.

Červený, V., 1972. Seismic rays and ray intensities in inhomogenous anisotropic media

Geophys. J. R. Astron. Soc. 29 1–13.

Červený, V., 2001. Seismic Ray Theory, Cambridge University Press, Cambridge, U.K.

Červený, V., 2004, Inhomogeneous harmonic plane waves in viscoelastic anisotropic media:

69
Stud. Geophys. Geod., 48, 167–186.

Chapman, S.J., Lawry, J.H., Ockendon, J.R. & Tew, R.H., 1999. On the theory of complex

Rays, SIAM Review, 41, 417–509.

Chichinina, T., Sabinin, V., & Ronquillo-Jarillo, G. 2006, QVOA analysis: P-wave attenuation

anisotropy for fracture characterization: Geophysics, 71, (3), C37–C48.

Crampin, S., E. M. Chesnokov, and R. G. Hipkin, 1984, Seismic anisotropy—the state of the

art: II: Geophysical Journal International, 76, 1–16.

Hanyga, A. & Seredyńska, M., 2000. Ray tracing in elastic and viscoelastic media, Pure Appl.

Geophys., 157, 679–717.

Hao, Q., and T. Alkhalifah, 2017, An acoustic eikonal equation for attenuating orthorhombic media,

Geophysics, 82, WA67-WA81.

Hearn, D.J. & Krebes, E.S., 1990a. On computing ray‐ synthetic seismograms for anelastic

media using complex rays, Geophysics, 55, 422–432.

Hearn, D.J. & Krebes, E.S., 1990b. Complex rays applied to wave propagation in a viscoelastic

medium, Pure Appl. Geophys., 132, 401–415.

Helbig, K., 1994, Foundations of Anisotropy for Exploration Seismics, Elsevier, NY.

Musgrave, M. J. P., 1970, Crystal Acoustics: Introduction to the Study of Elastic Waves and

Vibrations in Crystals, 72 pp., Holden-Day, Inc., San Francisco, CA.

Pereyra, V., W. Lee, and H. Keller, 1980, Solving two-point seismic-ray tracing problems in a

heterogeneous medium Part 1. A general adaptive finite difference method: Bulletin of the

Seismological Society of America, 70, 79–99.

70
Shekar, B., and I. Tsvankin, 2012, Anisotropic attenuation analysis of crosshole data generated

during hydraulic fracturing, The Leading Edge, 31, 588–593.

Tao, G., and M. S. King, 1990, Shear-wave velocity and Q anisotropy in rocks, A laboratory

study: International Journal of Rock Mechanics and Mining Sciences and Geomechanics Abstracts,

27, 353–361.

Thomsen, L., 1986, Weak Elastic Anisotropy, Geophysics, 51, (10): 1954–1966.

Vavryčuk, V, 2008b, Velocity, attenuation and quality factor in anisotropic viscoelastic media:

A perturbation approach, Geophysics, 73(5), D63–D73.

Vavryčuk, V. 2008b. Real ray tracing in anisotropic viscoelastic media, Geophys. J. Int., 175,

617–626.

Vavryčuk, V., Svitek, T., & Lokajíček, T., 2016. Anisotropic attenuation in rocks: theory,

modelling and lab measurements, Geophys. J. Int., 208, 1724–1739.

Vavryčuk, V., 2007b. Ray velocity and ray attenuation in homogeneous anisotropic

viscoelastic media, Geophysics, 72, D119–D127.

Vavryčuk, V, 2006, Calculation of the slowness vector from the ray vector in anisotropic

media, Proceedings of the Royal Society A, 462, 883–896.

Zhou, B., & Greenhalgh, S., 2004. On the computation of elastic wave group velocities for a

general anisotropic medium, J. Geophys. Eng., 1, 205–215.

Zhou, B., & Greenhalgh, S. A., 2005, ‘Shortest path’ ray tracing for most general 2D/3D

anisotropic media, J. Geophys. Eng., 2, 54–63.

Zhubayev, A., M. E. Houben, D. M. J. Smeulders, and A. Barnhoorn, 2016, Ultrasonic velocity

71
and attenuation anisotropy of shales, Whitby, United Kingdom, Geophysics, 81(1), D45–D56.

Zhang, L., 2017. Seismic Ray Tracing and Traveltime Inversion Using Multi-arrivals in

Anisotropic Media, Master Thesis, The Petroleum Institute, UAE.

72
Appendices

Appendix A

The coefficients of the Hamiltonian function for viscoelastic ORT media:

D1  a11a55 a66 ,

D2  a22 a44 a66 ,

D3  a33a44 a55 ,

D4   a11a22  a66 a66  a55  a11a44 a66   a12  a66  a55  ,


2
 

D5   a55 a66  a11a44  a55  a11a33 a66   a13  a55  a66  ,


2
 

D6   a11a22  a66 a66  a44  a22 a55 a66   a12  a66  a44  ,
2
 

D7   a22 a55  a44 a66  a44  a22 a33a66   a23  a44  a66  ,
2
 

D8   a55 a66  a11a44  a33  a44 a55 a55   a13  a55  a44  ,
2
 

D9   a22 a55  a44 a66  a33  a44 a44 a55   a23  a44  a55  ,
2
 

D10   a22 a55  a44 a66  a55   a55 a66  a11a44  a44   a11a22  a66a66  a33
  a13  a55  a22  2  a12  a66  a23  a44  a13  a55    a23  a44  a11
2 2

  a12  a66  a33 ,


2

D11    a11  a66  a55  a11a66  ,

D12    a66  a22  a44  a22 a66  ,

D13    a55  a44  a33  a44 a55  ,

73
D14   a12  a66    a66  a22  a55   a11  a66  a44   a11a22  a66 a66   ,
2
 

D15   a13  a55    a55  a44  a55   a11  a66  a33   a55 a66  a11a44   ,
2
 

D16   a23  a44    a55  a44  a44   a66  a22  a33   a22 a55  a44 a66   ,
2
 

D17   a11  a66  a55  ,

D18   a66  a22  a44  ,

D19   a55  a44  a33  .

74
Appendix B

The coefficients of perturbation of Hamiltonian function for viscoelastic ORT media:

W0  D1  px(0)   D2  p (0)


y   D3  p z   D4  p x 
p 
6 (0) 6 (0) 6 4 (0) 2
y

 D  p   p   D  p   p   D  p   p 
5
(0) 4
x
(0) 2
z 6
(0) 2
x
(0) 4
y 7
(0) 4
y
(0) 2
z

 D  p   p   D  p   p   D  p p p 
8
(0) 2
x
(0) 4
z 9
(0) 2
y
(0) 4
z 10
(0)
x
(0)
y
(0) 2
z

 D  p   D  p   D  p   D  p p 
11
(0) 4
x 12
(0) 4
y 13
(0) 4
z 14
(0)
x
(0) 2
y

 D  p p   D  p p   D  p   D  p 
15
(0)
x
(0) 2
z 16
(0)
y
(0) 2
z 17
(0) 2
x 18
(0) 2
y

 D  p  ,
19
(0) 2
z

W1  2[ D6  p (0)
y   D8  p z   D10  p y  p  D p 
4
(0) (0) 4 2 (0) 2 (0) 2
z 14 y

 D15  pz(0)   D17 ] px(0)  6 D1  px(0)   4[ D  p 


2 5 (0) 2
4 y

 D5  pz(0)   D11 ]  px(0)  ,


2 3

W2  2[ D4  px(0)   D9  pz(0)   D10  px(0)  p  D p 


4 4 2 (0) 2 (0) 2
z 14 x

 D16  pz(0)   D18 ] p (0)


y  6 D2  p y   4[ D  p 
2 (0) 5 (0) 2
6 x

 D7  pz(0)   D12 ]  p (0)


y  ,
2 3

W3  2[ D5  px(0)   D7  p (0)
y   D10  p x  p  D p 
4 (0) 4 2 (0) 2 (0) 2
y 15 x

+D16  p (0)
y   D19 ] p z  6 D3  p z   4[ D  p 
2 (0) (0) 5 (0) 2
8 x

 D9  p (0)
y  +D13 ]  p z  .
2 (0) 3

where the parameters Di  i  1 ~ 19  are the imaginary part of the coefficients Di  i  1 ~ 19  .

75
Appendix C

The coefficients of ray velocity vector in viscoelastic ORT media:

Vx0  6 D01  px0   4  D04  p 0y   D05  pz0   D11   px0 


5 2 2 3

 
2  D06  p 0y   D08  pz0   D10  p 0y pz0   D14  p 0y   D15  pz0   D17  p x0 ,
4 4 2 2 2

 

E1  6D01  px0   4  D04  p 0y   D05  pz0   D11   px0   2D06  p 0y  px0


5 2 2 3 4

 
2  D08  pz0   D10  px0 pz0   p y2 D14  p y0   D15  pz0   D17  px0 ,
4 2 2 2

 

E2  30 D01  px0   12  D04  p 0y   D05  pz0   D11   px0   2 D06  p 0y 


2 2 2 2 4

 
2 D08  pz0   2 D10  p 0y pz0   2 D14  p 0y   2 D15  pz0   2 D17 ,
4 2 2 2

E3  4 2D04  px0   2D06  p 0y   D10  pz0   D14   px0 p 0y  ,


2 2 2

 

E4  4 2D05  px0   2D08  pz0   D10  p0y   D15   px0 pz0  ,


2 2 2

 

Vy0  6 D02  p 0y   4  D06  px0   D07  pz0   4 D12   p 0y   2 D04  p x0  p 0y


5 2 2 3 4

 
2  D09  pz0   D10  px0 pz0   D14  px0   D16  pz0   D18  p 0y ,
4 2 2 2

 

F1  6D02  p 0y   4  D06  px0   D07  pz0   D12   p 0y   2D04  px0  p 0y


2 2 2 3 4

 
2  D09  pz0   D10  px0 p 0y   D14  px0   D16  pz0   D18  p y0 ,
4 2 2 2

 

F2  8D04  px0  p 0y  8D06 px0  p 0y   4 D10 px0 p 0y  pz0   4 D14 px0 p 0y ,


3 3 2

F3  30 D02  p 0y   12  D06  px0   D07  pz0   D12   p 0y   2 D04  px0 


4 2 2 2 4

 
2 D09  pz0   2 D10  px0 pz0   2 D14  px0   2 D16  p z0   2 D18 ,
4 2 2 2

F4  8 D07  p 0y  pz0  8D09 p 0y  pz0   4 D10  px0  p 0y pz0  4 D16 p 0y pz0 ,


3 3 2

76
Vz0  6 D03  pz0   4  D08  px0   D09  p 0y   D13   pz0   2 D05  px0  pz0
5 2 2 3 4

 
2  D07  p 0y   D10  px0 px0   D15  px0   D16  p 0y   C33  pz0 ,
4 2 2 2

 

G1  6D03  pz0   4  D08  px0   D09  p 0y   D13   p z0   2D05  p x0  p z0


5 2 2 3 4

 
2  D07  p 0y   D10  px0 p 0y   D15  px0   D16  p 0y   D19  pz0 ,
4 2 2 2

 

G2  8 D05  px0  pz0  8D08 px0  pz0   4 D10 px0 pz0  p 0y   4 D15 px0 pz0 ,
3 3 2

G3  8D07  p 0y  pz0  8D09 p 0y  pz0   4 D10  px0  p 0y pz0  4 D16 p 0y pz0 ,


3 3 2

G4  30 D03  pz0   2 D05  px0   2 D07  p 0y   12 D08  px0 pz0   12 D09  p 0y pz0 
4 4 4 2 2

2 D10  px0 p 0y   12 D13  pz0   2 D15  px0   2 D16  p 0y   2 D19 .


2 2 2 2

77

You might also like