Sei sulla pagina 1di 360

A PRACTICAL UNDERSTANDING

OF
PRE- AND POSTSTACK MIGRATIONS
VOLUME 1
(POSTSTACK)
by
John C. Bancroft, Ph.D.
University of Calgary

A PRACTICAL UNDERSTANDING
OF
PRE- AND POSTSTACK MIGRATIONS
Volume 1
(Prestack)
by
John C, Bancroft, Ph.D.

Eighth Edition
Copyright January 2011

No figures or examples may be reproduced by any


means without permission from the author.

ii

Table of Contents
Table of contents .................................................................................................. iii
Abbreviations and symbols ................................................................................... vi
Dedication ............................................................................................................ ix
Preface ................................................................................................................. x
Acknowledgments ................................................................................................ xi
Thinking ............................................................................................................... xii
Chapter 1 Introduction
1.1

Introduction ............................................................................................. 1.2

1.2

Recording and processing ...................................................................... 1.4

1.3

Zero-offset migration ............................................................................... 1.8

1.4

Recording with variable source-receiver offset ..................................... 1.20

1.5

Relationship between Seismic and Geology ......................................... 1.24

1.7

Modelling .............................................................................................. 1.26

1.7

Points to Note in Chapter 1 ................................................................... 1.27

1.8

Quiz ...................................................................................................... 1.28

Chapter 2 Modelling with zero offset


2.1

Introduction ............................................................................................. 2.2

2.2

A simple numerical model (mod50.sgy) ................................................. 2.6

2.3

Constant velocity raytracing .................................................................... 2.8

2.4

Ray-tracing with compass ..................................................................... 2.10

2.5

Modelling with hyperbolic diffractions ................................................... 2.25

2.6

Fourier transform modelling .................................................................. 2.21

2.7

Complex models: the real Earth ............................................................ 2.33

2.8

The blocky Marmousi Model ................................................................. 2.40

2.9

Ray tracing in a blocky model ............................................................... 2.42

2.10

Smoothing the velocities (slownesses) ................................................. 2.50

2.11

Raytracing amplitudes .......................................................................... 2.54

2.12

Defining a diffraction shape for modelling ............................................. 2.56

2.13

Eikonal method of grid modelling .......................................................... 2.62

2.14

Exploding reflector model ..................................................................... 2.66

2.15

Huygens method of wave front modelling ............................................. 2.76

iii

2.16

Finite difference wave propagation ....................................................... 2.82

2.17

Modelling 3-D data ................................................................................ 2.84

2.18

Physical modelling ................................................................................ 2.88

2.19

Comparison of a physical model with a numerical model ..................... 2.90

2.20

Points to note in Chapter 3 ................................................................... 2.94

Chapter 3 Bits and pieces


3.1

Circles ellipses and hyperbolas ............................................................. .3.2

3.2

Velocities (instantaneous, interval, average, RMS, stacking,


migration, and DMO)............................................................................... 3.8

3.3

The Fourier transform ........................................................................... 3.31

3.4

Aliasing in one and two dimension ........................................................ 3.46

3.5

Trace interpolation to reduce aliasing ................................................... 3.68

3.6

The wave equation(s) ........................................................................... 3.70

3.7

Derivatives and the wave equation ....................................................... 3.72

3.8

Time and depth migration ..................................................................... 3.76

3.9

Points to note in Chapter 3 ................................................................... 3.84

Chapter 4 Poststack migration (zero offset)


4.1

Introduction ............................................................................................. 4.2

4.2

Migration using a compass ..................................................................... 4.4

4.3

Hagedoorn migration ............................................................................ 4.10

4.4

Wavefronts, raypaths, and diffractions for linear velocities ................... 4.12

4.5

Kirchhoff migration ................................................................................ 4.16

4.6

F-K direct Fourier transform migration .................................................. 4.54

4.7

Downward continuation migration ......................................................... 4.74

4.8

Various algorithms for downward continuation ..................................... 4.82

4.9

Methods based on the Phase-shift ....................................................... 4.94

4.10

Reverse Time migration ...................................................................... 4.112

4.11

Migration of 3-D data .......................................................................... 4.114

4.12

Testing migration ................................................................................ 4.126

4.13

Points to note in Chapter 4 ................................................................. 4.132

iv

Chapter 5 Examples of migration


5.1
Dip models with constant velocity ........................................................... 5.2
5.2

Practice examples for interpretation ....................................................... 5.8

5.3

Assorted migrations of real data ........................................................... 5.12

5.4

Assorted migrations of ray-trace model ................................................ 5.24

5.5

Migration examples with velocities that varying from 80% to 150% ...... 5.34

5.6

Comparison of 2-D and 3-D migrations................................................. 5.52

5.7

Migration tests with constant velocity .................................................... 5.56

5.8

Migration test with a velocity profile ...................................................... 5.62

5.9

Summary text ........................................................................................ 5.64

5.10

Points to note in Chapter 5 ................................................................... 5.66

Chapter 6 Special topics on migration


6.1

Special topics .......................................................................................... 6.2

6.2

Moving window migration........................................................................ 6.4

6.3

Side-swipe .............................................................................................. 6.6

6.4

Oblique reflectors .................................................................................... 6.8

6.5

Overlapping reflections ......................................................................... 6.10

6.6

Recording times and migration aperture ............................................... 6.12

6.7

Lateral resolution and Fresnel zones .................................................... 6.16

6.8

Migration of spatially aliased data ......................................................... 6.22

6.9

Aliasing in time to depth conversions .................................................... 6.32

6.10

Frequency changes due to migration .................................................... 6.34

6.11

Cascade migration ................................................................................ 6.36

6.12

Surface elevation, datums, and migration ............................................. 6.38

6.13

Example problem on recording time and aperture ................................ 6.56

6.14

Points to note in Chapter 6 ................................................................... 6.58

Appendix 1 Solving the wave equation .......................................................... Appen.-1.1


Appendix 2 The Fourier transform ................................................................. Appen.-2.1
Appendix 3 Diffractions, statics, and downward continuation ........................ Appen.-3.1
Appendix 3 Kinematic derivations for DMO-PSI and EOM ............................ Appen.-4.1
References: Author sort ......................................................................................... Ref. 1
Numerical sort...................................................................... Ref. 13

Abbreviations and symbols


AVO

amplitude verses offset

MO

moveout as in moveout correction (dipping and horizontal)

NMO

normal moveout (historically used for horizontal layering)


my term for moveout of horizontal layers

DD-MO

dip-dependent moveout,
my term for moveout of dipping layers

DMO

dip moveout, prestack partial migration


(historically used for moveout of dipping layers)

PSM

prestack migration

PSPM

prestack partial migration

COS

Constant (limited) offset section

CDP

Common depth point

CRP

Common reflection point

CMP

Common midpoint (preferred over CDP)

CSP

Common scatter point

P-P

Conventional seismic data considering only P-waves

P-S

Seismic data with incident P-wave and reflected S-wave

x, y

horizontal surface distances, y required for 3D surveys

vertical depth

t, T

recording time, usually two-way time

T0

zero-offset time (collocated source and receiver)

half offset distance between source and receiver

he

equivalent offset computed for prestack migration

Vins

instantaneous velocity, velocity of rock (P-wave, isotropic...)

Vnmo

normal moveout velocity used for horizontal layers

Vstk

moveout velocity for dipping layers best stack

Vave

relates total depth to total time

Vrms

enables curved rays on a horizontally layered depth section


to be considered linear rays on a time section.
Vrms = VNMO : short offsets in isotropic horizontal medium.

Vint

interval velocity, Vave, or Vrms over a convenient interval

Vmig

should be Vrms (not stacking) or interval velocities

Vdmo

should use RMS (not stacking) for NMO moveout

vi

(x, t)

dimensions usually for a 2-D seismic section

(x, z)

dimensions usually for a 2-D depth migrated structure

(x, h, t)

prestack volume of traces from a 2-D line

(x = h, t)

usually a source (shot) record

(x = xCMP, h, t)

usually a CMP gather

(x, h = h0, t)

constant offset section when h constant

(x, y, t)

dimensions for a 3-D seismic data set

(x, y, z)

dimensions for a 3-D depth migrated data set

F-K

Fourier transform of an (x, t) section into a frequency,


wavenumber (k) domain (F, K)

one-way time

traveltime from either a source or receiver to a scatter point or


reflector; two-way time divided by two

two-way time

total time to travel from source to reflector and back to the


receiver

section

seismic record displayed with horizontal distance and vertical


time axis (sometimes the vertical axis is in depth)

structure

representation of the geology

reflector

geological interface

reflection

seismic response to a reflector

wave equation

many forms of this equation are found and usually relate the
acoustical pressure (displacement) with position and time

Eikonal equation

relates time and position of wave fronts: slowness vectors

Eikonal raytracing Raytracing method derived from the Eikonal equation and
requires a smooth velocity model
Snell raytracing

Raytracing method that uses Snells law in a blocky model

imaging

usually refers to velocity analysis, migration and inversion

imaging condition location on a processed time section that corresponds to the


migrated output, usually at t = 0;
coincident location of the source wave front with back
propagated receiver part of the wave front (shot migration)
migration

use of seismic data to create a subsurface image that attempts


to match the reflectivity

inversion

estimate geological parameters from seismic data; modify


geological structure to optimize migration

focusing

collapsing diffracted energy to clearest image

positioning

moving seismic energy to the correct geological location

vii

kinematics

aspects of timing, (not amplitudes, phase, etc.)

amplitudes

duh (usually difficult to estimate accurately)

stationary

properties that dont vary with time or space


a typical abut poor assumption of seismic data

homogeneous

properties are independent of location

isotropic

properties are independent of direction

anisotropy

the properties vary with direction

acoustic

assumes only the P-wave is present, possibly a fluid

elastic

P- and S-waves are present, real world

slowness

reciprocal of the velocity


Smoothing the velocity implies smoothing the slowness

source (shot) location


receiver location
[ ]

reference number in reference list at the back of the text

viii

Dedication
To:
Janice Fay, John, James, Jeffory, Janenne, Jared, and Joel.
Additions:
Julia and Behn
Karen: Connery, Sterling, and Evalina
Christine: Elizabeth
Justin: Jordan, Jocelyn, Janelle, Jillian, and Jacob
Candace: Sophia and Caden
Laura: Linaya and Colby
Remember the time when I had my original drawings spread across the kitchen table...
and Joel (4 at the time) showed me how fast he could spin the mixer blades...
and the blades were fresh from a chocolate cake mix...
and...

Update: Times have changed, and now all the figures and diagrams are electronic.
Any smudges or miss-positioning within figures may be attributed to any of the above
hackers.

ix

Preface
Many years ago I was asked to gather material and prepare a series of lectures on
migration that would be suitable for people skilled in the area of seismic processing and
interpretation but lacking the mathematical skills usually associated with the topic.
These notes endeavor to present migration intuitively by relying on many diagrams.
Consequently, very little mathematical rigor is included, although some is retained in
Appendix 1, which contains examples of the 15 and 45 degree solutions to the wave
equation.
Many of the figures may be either a time section (x, t) with x the horizontal distance and
t vertical time, or a depth structure (x, z) with z the depth. Often the velocity is assumed
to be unity with equivalent time and depth scales that enable time and depth structures
to be plotted together, i.e., (x, t or z). It is also common practice to plot sections with the
vertical time scale calibrated to match the horizontal distance, enabling seismic dips to
be displayed with angles in degrees (rather than in slopes of time/distance). Care has
been taken to label the axis where appropriate, however, at times the labeling is left for
the participant to complete as an exercise. Sometimes the apparent ordering of the
figure parts (a), (b), (c) and/or (d) will not be in the usual order, but arranged in a
clockwise order to maintain a concept flow.
The notes begin with a brief overview in chapter 1 followed by poststack modelling
concepts in chapter 2. This chapter is important for three reasons; first, the modelling
principles are more intuitive than the migration principles, second, modelling becomes
part of depth migrations, and third, migration is the inverse operation of modelling.
Some background and related material such as geometry and aliasing are included in
chapter 3. Chapters 4 and 5 deal with poststack migration, and chapter 6 deals with
special topics that usually relate to poststack migrations.
Volume II deals with the prestack case where the source and receiver offsets vary.
Prestack modelling, dip moveout, and various prestack migration methods will be
presented.
The pages were prepared for course presentations by keep topics together on facing
pages, usually with text on the left and figures on the right. A large font was used to
enable simultaneous projection of the pages by an instructor, requiring little note taking
by the participant. A number of geometrical construction exercises have also been
included to help reinforce the basic concepts.
Two lists of the references are provided that are sorted by an ID number [..] and the first
authors last name. The author sort also uses the date for additional sorting and
requires a month and year for each reference. In some cases, when the month is not
defined or is not known, January is used, and in a few cases when the year is not
known, the year 1940 is used. Any assistance in clarifying any of the references or
adding ones that I have missed would be greatly appreciated. Not all references are
cited.

Acknowledgments
I would like to express my thanks to the many people who have helped in the
preparation of this manual and to the many people who have guided me in my
geophysical career. Special thanks go to Don Simpson for his encouragement in the
preparation of the course, to Don Chamberlain who made it possible, and to John
Hodgkinson who was my tutor and mentor as I entered the field of geophysics. The
efforts of Mark Ng, Jim Montalbetti, and Dave Oldroyd who assisted in the preparation
and processing of examples were greatly appreciated along with help from Lisa Buckner
and other who assisted in editing.
Two teachers from my high-school days were Mr. Scott and Mr. Anderson, who were
very instrumental in reinforcing the point that all mathematical equations are related to
physical principles and that I should always find an intuitive appreciation for all
mathematical expressions.
Acknowledgment is also due to my family that has had to endure a great deal of stress
due to the "bear" in the home while the material was spread throughout many rooms
during preparation. Special thank are also expressed to my wife, Fay, for all her
encouragement and dedicated support, and to my father who was always willing to
spend the time to teach the fundamental principles of engineering and of life itself.
Appreciation is expressed to various companies that have assisted by providing
examples of tests and seismic data. They are Geo-X, Husky, Mobil, Shell, Veritas
Seismic, Ulterra Geoscience, C and C Systems, TGS-NOPEC Geophysical Company,
SANTOS, and CREWES.

xi

The real glory of science


is that we can
find a way of thinking
such that the
law
is evident.
Feynman 1963 (Lectures on Physics Vol. 1, p26-3)

xii

Chapter One
Introduction to Migration

Objectives:

Recognize how data acquisition produces a distorted view of the subsurface

Realize that the distortion is removed by the migration process

Become aware of time or depth migration

Know that scatterpoints are useful in defining the acquisition response

Know that post-stack migration deals with semicircles

Know that prestack migration deals with ellipses

Identify the three main migration methods


1. Kirchhoff
2. F-K
3. Downward continuation

A Practical Understanding of Pre- and Poststack Migrations

1.1 Introduction
Possible scenario for finding productive hydrocarbons

Initial idea

Geological field work

Planning seismic programs

Seismic acquisition

Seismic processing

Interpretation

Drilling program

Well completion

Production

Seismic processing above is expanded to

Geometry

Amplitude recovery

Deconvolution

Statics

Noise attenuation

Velocities

Stacking

Migration

Inversion, AVO

Page 1.2

Chapter 1 Introduction to Migration

Migration above is expanded to

Time

Depth

Post-stack

Prestack

DMO (DIP moveout)

2-D

3-D

Constant velocity

Structured geology

Datum

Noise

Aliasing

Velocities

Migration algorithms

Migration algorithms above is expanded to

Kirchhoff time

Kirchhoff depth

F-K

Downward continuation

Finite difference

15 degree, 45 degree, ...

Phase shift

PSPI (Phaseshift plus interpolation)

X time

X depth

DMO

Pre-stack

Poststack

Page 1.3

A Practical Understanding of Pre- and Poststack Migrations

1.2 Zero offset reflection energy


1.2.1 Zero-offset recording with linear reflectors
The main objective of seismic processing is to produce an image, or section, that
represents the geological structure below the surface.
Figure 1.1 illustrates
1. the source is identified by "

"

2. the receiver is identified by "

"

3. the geological structure in (a)


4. the seismic response in (b)
5. the migration in (c), found by repositioning the data from (b)

Note the following:

The source and receiver are located at the same position.

The time response is plotted directly below source and receiver

One way travel times are shown (assume velocity = 1.0).

Flat event looks OK.

Dipping event has moved to a new position and appears to have a new dip.

An objective of migration is to reposition the dipping to appear more like the


geological structure.

For the migration to be correct, the vertical axis would be in depth. However, it
is common to display the migrated section with a vertical time axis.

Page 1.4

Chapter 1 Introduction to Migration

a)

time

b)

c)
Figure 1.1 Illustration of migration concepts with linear reflectors, a) geological
structure, b) a seismic section, and c) a migration of the seismic section.

Page 1.5

A Practical Understanding of Pre- and Poststack Migrations

1.2.2 Zero-offset recording with point reflectors


It is convenient to evaluate the effect of reflections from scatterpoints. This leads
to mathematical solutions for migration that collect the energy in the hyperbolas.
Figure 1.2 illustrates

source receiver raypaths to the scatterpoints in (a)

resulting "diffractions" or hyperbolas in (b)

effect of migration in (c) where the diffraction energy has been summed into
the scatterpoint.

From Figure 1.2b note the following:

Diffractions at the same depth have the same shape.

Deeper reflections have a broader shape.

The migration process collapses the energy in the diffraction back to the
position of the reflector.

The diffraction Figure 1.2b has a hyperbolic shape in a constant velocity medium.
In a smoothly varying velocity medium, the diffraction may be approximated by a
hyperbola.
In rugged velocity media, the diffraction shape may require ray tracing or wave
front analysis for evaluation.
In extremely rugged velocity media, the diffraction shape may require
simplification back to the hyperbola for evaluation.
The energy of the diffraction becomes weaker with increased offset. This effect is
not represented on kinematic illustrations.

Page 1.6

Chapter 1 Introduction to Migration

a)

b)

c)
Figure 1.2 Illustration of migration concepts using (a) geological structure with
scatterpoints, (b) a seismic section showing hyperboloids, and (c) a migration of
the seismic section.

Page 1.7

A Practical Understanding of Pre- and Poststack Migrations

1.3 Zero offset migration


1.3.1 Zero offset migration of a single point on a seismic section
A point or spike of energy on a single seismic section before migration would
normally be undesirable as it would represent noise. However, it serves a very
useful purpose as it aids in understanding of the migration process. (A collection
of spikes may represent a reflection.)

Question: Where are all the possible reflectors located if they all produced the
same reflection time?

Figure 1.3 shows the migration of points on a single trace. Since we don't know
where the final migrated position will be, we simply disperse the amplitude of the
energy in semicircles.

Constructive interference of these semicircles will create the migrated image.

Destructive interference will cancel unwanted smiles.

This effect is observed when smiles are generated after migrating a noisy
section.

Question: Do migration algorithms always construct a full semicircle?

Page 1.8

Chapter 1 Introduction to Migration

x
t

a)

x
t

b)
Figure 1.3 Migration of single points on the input trace is shown in (a) and the
semi-circular result of migration is shown in (b).

The migrated energy is only circular if the velocity of the medium is constant.

Page 1.9

A Practical Understanding of Pre- and Poststack Migrations

1.3.2 Zero offset migration of linear reflectors


The stacking process attempts to create a section that is equivalent to the zerooffset section. Migration after the stacking process is referred to as post-stack
migration, or as the assumed source-receiver offset is zero, it is also referred to
as zero-offset migration.
As shown in Figure 1.4, one objective of migration is to reposition the dipping
event of (a) to the migrated position in (b).
For convenience, we equate times with distance.

The velocity is constant.

The velocity has a value of unity.

One-way times are used.

For zero-offset, one-way times may be considered half the two-way time.

We will continue to use these assumptions throughout this chapter and in many
other areas of the notes.

Migration with a compass is illustrated in Figure 1.4b. Note:

Reflection point is plotted at its reflection time below the surface location.

Zero-offset reflecting rays are at right angles to the reflector.

Any reflection with the same time could come from any position on a
semicircle.

Construct semicircles at T1 and T2.

The tangent to these circles will represent the migrated position.

Page 1.10

Chapter 1 Introduction to Migration

x
t

T1

T2

a)

t or z

b)
Figure 1.4 The seismic event to be migrated is shown in (a). Compass migration
of a dipping event is shown in (b).

Page 1.11

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

1.3.3 Migration
M
n of real data
d
The following disp
plays in Figure 1.5 illlustrate:
1. a proces
ssed seism
mic section
n
2. a migrated time se
ection
3. a migrated depth section
s
Note th
he apparen
nt diffractio
ons on the seismic section in
n (a) and the confusion
caused by overlapping diffrractions. These con
nfused are
eas are gre
eatly impro
oved
on the time migrration (b). Depth miigration (c
c) illustrate
es to pote
ential abilitty of
seismic
c processing to prod
duce an im
mage of th
he geological structture below
w the
surface.

a)
e 1.5 Exam
mple of (a) a seismic section,
s
(b
b) a time migration an
nd (c) a depth
Figure
migration of the sam
me data

P
Page
1.12

Chapter 1 In
ntroduction to Migration
M

b)
X

c)

P
Page
1.13

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

Example
es of a seis
smic proje
ect in Indon
nesia

Figu
ure 1.6 Location map
ps for Mak
kassar Stra
ait, Indones
sia, data provided by
y
TGS--NOPEC Geophysica
al Company
y.

P
Page
1.14

Chapter 1 Introduction to Migration

a)

b)
Figure 1.7 Makassar Strait data showing a) a structural image in depth of a depth
migration from one seismic line, and b) the velocity that was estimated and used
in processing this data.

Page 1.15

A Practical Understanding of Pre- and Poststack Migrations

a) Time migration

(b) Depth Migration converted to vertical time scale for comparison to (a).
Figure 1.8 Close up Makassar Strait data comparing a) and c) time migrations
with b) and d) depth migrations. The vertical scale for the depth migrations have
been converted to time for comparison purposes.

Page 1.16

Chapter 1 Introduction to Migration

c) Time migration

d) Depth migration converted to vertical time scale for comparison.

Page 1.17

A Practical Understanding of Pre- and Poststack Migrations

1.3.4 The three main post-stack migration methods for 2D data


Figure 1.9 Illustrates the three major methods of post-stack migration. They are
Kirchhoff, F-K (or Fourier transform method), and downward continuation.
Kirchhoff migration
For every output sample, sum the energy along diffraction, or hyperbolic paths,
on the input section. The summed value becomes the amplitude value at the
output location. Additional scaling and filtering may be required. Kirchhoff
migration is an integral solution to the wave-equation.

F-K migration
This method makes use of the 2-D Fourier transform to convert the input section
into the 2-D Fourier domain where it is migrated with a simple algorithm. The
inverse transform provides the migrated structure. The F denotes the Fourier
transform of time to frequency, and K denotes the Fourier transform of space or
distance to wave number. Sometimes the wave number K may also be referred to
as frequency. F-K migration uses the Fourier transform to solve the waveequation.

Downward continuation
This method is quite different to the above two migrations as it works on a
volume of information rather than two planes of data. The input 2-D time section
(x, t) is the top surface of a cube at z = 0, and various processes continually
compute the next downward layer in depth until the volume is complete. Each
downward step represents the time section that would be produced if the sources
and receivers had also been lowered to the new depth. This volume of
information will then yield the migrated structure (x, z) at the end of the cube
where t = 0. The wave field is propagated from one layer to the next using various
solutions to the wave-equation, such as finite difference, phase shift, or X.
The wave-equation solutions can be modified to produce a migrated time
section (x, t), at the bottom of the volume where z = zmax.

Page 1.18

Chapter 1 Introduction to Migration

a)

b)

c)
Figure 1.9 Examples of the three major migration methods: (a) Kirchhoff
migration, (b) FK migration, and (c) downward continuation migration.

Page 1.19

A Practical Understanding of Pre- and Poststack Migrations

1.4 Recording with variable source/receiver offsets


1.4.1 Offset modelling
In practical seismic acquisition there are many surface recordings made from one
source or explosion. Consequently, there are variable distances (offsets)
between the source and receiver as illustrated in Figure 1.10. The sorting and
merging together of the many recorded traces constitute a major part of seismic
processing.

Figure 1.10 illustrates raypaths for a source gather, a CMP gather, and two source
receiver pairs from a dipping reflector.

To is the normal reflection from a co-located source and receiver.

Farther offsets have a longer travel times TMO.

This increasing refection time with offset is referred to as moveout (MO).

The trace is located midway between the source and receiver, at the common
mid point (CMP), that is often referred to at the CDP.

Traces with the same CMP but with different offsets require MO time
correction to realign the reflection.

CMP traces from dipping events contain reflections from different parts of the
reflector.

Special MO is required, dip-dependent moveout(DD-MO), to align the reflection


times from dipping events.

Special prestack processing is required to correct the dipping reflection point


smear (DMO or prestack migration).

Page 1.20

Chapter 1 Introduction to Migration

a)

MO

b)
Figure 1.10 Illustrations of linear reflectors and variable source receiver offsets
show (a) raypaths for a source gather, a CMP gather, and reflections from a
dipping linear surface. (b) shows the resulting traces from (a).

Page 1.21

A Practical Understanding of Pre- and Poststack Migrations

1.4.2 Prestack migration


Prestack migration (PSM) takes into account the location of the source and
receiver for each trace when determining the reflector location.
The construction for PSM can be difficult and is usually left for the computer.
The migration of a spike on a single trace will serve as an introduction as shown
in Figure 1.11. Note:

Velocity is assumed to be unity.

One-way time response T is shown at the CMP location.

Time response could come from any reflection point that has the same total
time for the raypath.

All possible reflection points lie on an ellipse with the source and receiver
locations being the foci of the ellipse.
1. The travel times are the same.
2. The incident and reflection angles are equal.

Prestack migration is accomplished by spreading the energy of each time


sample along its appropriate ellipse when the velocity is constant, or an
aplanatic surface when the velocity is not constant.

Page 1.22

Chapter 1 Introduction to Migration

CM

x
t

T
a)

CM

x
t

T
b)
Figure 1.11 Possible reflection locations a) for a single point on the input trace,
and b) the result of prestack migration.

Page 1.23

A Practical Understanding of Pre- and Poststack Migrations

1.5 Relationship between Seismic and Geology


Seismic data is recorded in time t then processed and stacked into a section in
(x, t) that should be equivalent to a zero offset section.
The stacked section has a better signal-to-noise-ratio (SNR) than a zero offset
section and unwanted coherent energy such as multiples are attenuated.
Diffractions are still present (though in an attenuated form), and dipping energy is
mis-possitioned.
Figure 1.12 shows a geological cross-section and a zero offset time response.
They are plotted orthogonal to each other to emphasize the difference between a
time and depth section.
When the velocity is constant, time and depth migrations are equivalent and
diffractions in (x, t) are simultaneously plotted with raypaths in (x, z).

Page 1.24

Chapter 1 Introduction to Migration

Figure 1.12 Illustration of a geological cross-section, a zero offset time section,


and a depth migration. The time axis is plotted orthogonal to depth to emphasize
the incompatibility.

Page 1.25

A Practical Understanding of Pre- and Poststack Migrations

1.6 Modelling
Modelling will be considered the forward process of starting with a known
geological structure and creating seismic time data. This tends to be a natural
process and the principles are straightforward, easy to understand, and natural to
program.
The migration process is the reverse process that uses the seismic time data to
build an image of the subsurface geological structure. This is especially the case
with downward-continuation migration where the wave field is moved back to the
reflector location. Migrations may also use forward modelling as part of their
algorithm. Consequently, the next chapters will deal with modelling in detail.
Precise methods of migration require an accurate knowledge the geological
model (i.e., velocities and structure). Modelling the kinematics of ray paths or
wave fronts is often part of these migration processes, especially depth
migration.

The following Geophysical inversion is the process that uses the seismic time
data to estimate the rock properties of the Earth. This is a much more complex
process.
models will be discussed in greater detail to introduce the various methods of
migration:

modelling with zero-offset

ray tracing with constant velocity

ray tracing with constant interval velocities

raytracing with variable velocity media

diffraction sum modelling

Fourier Transform modelling

exploding reflector model

wavefront modelling

source or shot record modelling (prestack)

constant or limited offset stacked sections (prestack)

Page 1.26

Chapter 1 Introduction to Migration

1.7 Comments
A time migration is relatively straightforward and is usually the responsibility of
the seismic processor. Time migration would typically use stacking velocities
that have been converted to RMS or interval velocities. The objective of a
time migration is to focus the data, especially in areas where it is difficult to
construct an accurate velocity model.
A depth migration (x, z) may require a team approach to aid in building the
velocity model that includes a geologist. The objective is to produce accurate
positioning of the data, and where possible, better focusing. However, accurate
positioning requires an accurate anisotropic velocity model.
Seismic imaging typically includes the processes of velocity analysis and
migration.
Economics, noise, and the geological complexity of the area under investigation
control the level to which the data is processed.
Two dimensional data (2D) data that may have been acquired along a one
dimensional (1D) road. These concepts may be extended to the real world of
three dimensions (3D) where sources and receivers are spread across 2D
surfaces of land, water, water bottoms, or in wells. 2D solutions to the wave
equation may not have the correct amplitudes that would be provided with a 3D
solution.
Constant velocities and one-way times allow time and depth sections to be
plotted on top of each other. We can also scale time sections to approximate
depth sections for simple geometrical constructions. However, we must always
remember that the time section is orthogonal to a depth section and care must be
taken when assuming they are similar.

Real-world note: The real-world migrations are more complex than indicated in
this initial chapter. We have assumed the velocities to be constant and unity,
the geological surface to be flat and horizontal, and that we have no noise. This
is not the case for real data, but these assumptions help clarify the concepts
and are valuable aids in testing the various migration algorithms.

Page 1.27

A Practical Understanding of Pre- and Poststack Migrations

Points to Note in Chapter 1

Seismic acquisition produces a distorted view of the subsurface.


The migration process reduces this distortion.

Migrations can be in time or depth.

Scatterpoints help define the migration response.

Post-stack migration produces semicircles in a constant velocity medium.

Prestack migration produces an ellipse (constant velocity medium) where the


source and receiver are located at the foci of the ellipse.

The three main migration methods are


1. Kirchhoff
2. F-K
3. downward continuation

Summing energy along a hyperbola is identical to spreading energy along a


semicircle (constant velocity).
2D and 3D data

For further information:


See chapter 2 for more information on modelling and references.
See chapter 4 for more detailed information on migration and references.

Page 1.28

Chapter 1 Introduction to Migration

1.8 Quiz 1
1)

Define
CRP?
CMP?

2)

What is the assumed source-receiver offset in post-stack migration?

3)

Sketch the zero-offset migration response of the trace below.


x
z
or
t

4)

Where could the reflection come from on the trace below?


s

x
z
or
t

5)

Define

Two-way time...

One-way time...

Page 1.29

A Practical Understanding of Pre- and Poststack Migrations

Page intentionally left blank

Page 1.30

25

Chapter Two
Modelling with Zero-offset

Objectives

Construct a post stack or zero-offset section using a compass

Know that seismic dips are defined by , and geological dips by

Understand modelling by summing diffractions, F-K methods, ray-tracing


methods, and the exploding-reflector model

Diffraction can be formed by many processes

Use the "exploding reflector" model of 2D data in (x, z) with the added
dimension of time to define the 3D volume with dimensions (x, z, t)

A Practical Understanding of Pre- and Poststack Migrations

2.1 Introduction
2.1.1 Basics
Seismic modelling is the forward process to create seismic data.
follows simple rules of physics to create simulated seismic data.

Modelling

The reverse process of seismic migration and or inversion is much more complex
than modelling. These processes take the seismic data and attempt to recreate
images that represent the geology of the Earth.
When the geological model is homogeneous, isotropic, and lossless, it is
relatively straight forward to convert to the seismic data and back to the model.
Modelling is an integral part of the migration process. We assume a geological
model and process the seismic data to form an improved geological model. We
then iterate the process until we have a satisfactory model of the subsurface, a
typical chicken and egg problem.
Some of the models may start as a constant velocity and iterate through a
stacking velocity, RMS velocity, and then to an interval velocity in depth. They
may also incorporate well log information and attempt to go to directly to a
subsurface image using an advanced inversion process.
If care is not taken constructive and destructive interference may tend to verify an
incorrect model
Seismic modelling using a known or defined geological structure may provide
insight to the seismic data, and to develop strategies for acquiring and
processing real data.
There are many modelling method used to create synthetic seismic data. One of
the simplest method is to assume the reflection and or scattered energy lies on a
hyperbolic path.
Moveout correction is part of the migration process. Combining the moveout
equation with the poststack Kirchhoff time migration operator is a prestack
migration.

Page 2.2

Chapter 2 Mo
odelling with ze
ero-offset

t4

t3

t1

t2

t0
2

Time

Figurre 2.1 Vario


ous modellling figure
es.

P
Page
2.3

A Practical Understanding of Pre- and Poststack Migrations

2.1.2 List of modelling processes

Convolving a wavelet with a reflectivity array to form a seismic trace.

Convolving a wavelet with an assumed reflection model, such as a dipping


reflection, or an hyperbola.

Moveout correction where we find a hyperbola to fit energy in a CMP gather.

2D Fourier transform

Defining the shape of a diffraction for Kirchhoff migration.

Ray tracing through a geological model to define a reflectivity.

Raytracing with Snells raypaths in a blocky model.

Raytracing in a smooth velocity model using an Eikonal ray tracer.

Using the first arrival time.

Using the maximum energy arrival time.

Using multiple arrival traveltimes.

Wave propagation using the wave-equation

Exploding a reflector and following the wavefield to the surface to produce a


zero offset section.

Explode a source point and follow the wavefield to the surface to define a
source (shot) record.

Page 2.4

Chapter 2 Modelling with zero-offset

Figure 2.2 Modelling methods

Page 2.5

A Practical Understanding of Pre- and Poststack Migrations

2.2 A simple numerical model (mod50.sgy)


Seismic data can be created from physical and numerical models.
Physical models of geological features may be built at a small scale that is
convenient for laboratory testing where seismic data can be acquired in a
controlled manner. The modelling uses very high frequencies that tend to match
the wavelengths of real seismic data in real geological structures. The advantage
(or sometimes disadvantage) of these models is that multiples and mode
conversions are present.
Numerical models are defined within a computer and can contain features that
are not present in real geological structures. These models can be very simple or
complex.
A numerical model that is known as mod50, and shown in Figure 2.3a, will be
used a number of times in these notes. It has a constant velocity of 10,000 ft/sec
and contains linear and dipping reflections, diffractions, and localized points of
reflection energy. The model has 150 traces with trace interval of 100 ft, and 750
time sample at a 2 ms sample rate for a total time of 1.5 sec. The reflection
events were defines at specific times, and amplitudes were linearly weighted
between time samples. After all the events were defined, the entire section was
bandpass filtered to a maximum frequency of 50 Hz.
Mod50 was padded with extra traces and sample to enable the 2D Fast Fourier
transform (FFT) to compute the FK plot that is shown in Figure 2.3b. Note the
band limiting to 50 Hz, and energy at all wave numbers. Evanescent energy that
is steeper than 45 degrees is also evident and comes from the points and the
truncated events. The aliased energy from the two steeper dipping events has
wrapped around as illustrated by the black outlines.

Page 2.6

Chapter 2 Modelling with zero-offset

a) The mod50.sgy model

Maximum Frequency
(Nyquist)
250 Hz

Aliased energy

b) FK amplitude spectrum of mod50


Figure 2.3 A constant velocity numerical model with its FK transform.

Page 2.7

A Practical Understanding of Pre- and Poststack Migrations

2.3 Constant velocity raytracing


Ray tracing is a simple procedure in which the path from the source to the
reflector and on to the receiver is defined. This procedure is simple when the
velocity is constant and the source and receiver are at the same location, as
illustrated in Figure 2.4.
Assume:

one-way times and unit velocity

source and receiver are collocated

the emitted and receiving rays travel the same path

collocated source and receivers will only produce right-reflections

a scatter point will reflect energy to any location of the source-receiver pair

reflections are plotted below the source-receiver location

Page 2.8

Chapter 2 Modelling with zero-offset

a)

b)

c)

d)
Figure 2.4 Examples of raypaths in (a) a horizontal reflector, (b) dipping reflector,
(c) a scatter point, and (d) a combination of the above.

Page 2.9

A Practical Understanding of Pre- and Poststack Migrations

2.4 Ray-tracing with compass


Construction:
Construct the zero-offset reflection from the reflectors for each surface location
in Figure 2.5.

Assume one-way travel times and unit velocity.

For each reflector:


1. Place the compass point at a s/r location.
2. Guess the reflection point on the reflector (right angle to ray).
3. Position the other point of compass at reflection point to get the travel
time.
4. Plot the travel time below the s/r position.

Highlight the reflection and then repeat the process for all the reflectors.

Ignore edge effects that could come from the ends of the reflectors.

Construct the zero-offset reflection from the two scatter points.


Measure the distance from a surface location to the scatter point.
Plot this distance (that now represents time) below the surface location.
Repeat for all surface locations and for both scatter points.

Observations:

Horizontal event remains in the same position.

Dipping events becomes shallower and reflection points move down dip.

The dipping reflector and reflection converge at the surface.

The reflection from the scatterpoints produces a hyperbolic curve referred to


as the diffraction shape.

The asymptotes to the hyperbola intersect at the surface.

The maximum dip is 45 degrees.

Page 2.10

Chapter 2 Modelling with zero-offset

CMPs

t
or
z

Figure 2.5 Construction with constant velocity and zero-offset.

Page 2.11

A Practical Understanding of Pre- and Poststack Migrations

2.4.1 Geometry of reflection


Using the ray-tracing information, we can construct the diagram in Figure 2.6 to
relate the dip angles between the reflector and the reflection.
Define:

the angle of the dipping reflection

the angle of the reflector

T two way time from B to R

From the construction note two triangles, ABC and ABR

VT
2

The length BR = BC is given by

s=

In triangle ABC

tan =

s
x,

and in triangle ABR

sin =

s
x

i.e.,

tan = sin
For more information see [788].

Page 2.12

(2.1)

Chapter 2 Modelling with zero-offset

s
s = tv/2
R
C

t or z

Figure 2.6 Geometry for computing the relationship between the dip angles of the
reflector and reflection.

Page 2.13

A Practical Understanding of Pre- and Poststack Migrations

2.4.2 Construction example of anticlines and synclines


Draw the raypaths and time response for the following structures.
x
z
or
t

a)

x
z
or
t

b)
Figure 2.7 Construction of anticline's response with (a) shallow, and (b) deep.

Page 2.14

Chapter 2 Modelling with zero-offset

x
z
or
t

a)

x
z
or
t

b)
Figure 2.8 Construction of synclines that give buried focus images, (a) shallow
and (b) deeper.
Conclusions

Anticlines have broader reflections.

Synclines may appear as anticlines.

Page 2.15

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

2.5 Modelling
M
g with hy
yperbolic diffra
actions
A reflec
ction may be consid
dered as a collection
n of scatter points as
s illustrate
ed in
Figure 2.9a. Th
he points in the ro
ows increa
ase in de
ensity to form a be
etter
approxiimation to a line. Each
E
pointt is replaced with a hyperbola
a as in (b) that
approxiimates a diffraction. Blurring of
o the imag
ge (c) remo
oves the cu
urves.

a)

b)

c)
Fig
gure 2.9 Modelling with
w
diffracttions show
wing in (a) input poin
nts, (b) the
prog
gressive efffect of difffraction he
ealing, and
d c) the effe
ect of a blu
urring filterr.
What fa
actors are involved
i
in
n removing
g the lines in (b)?

P
Page
2.16

Chapter 2 Modelling with zero-offset

The structure below was used as input for modelling with diffractions. Modelling
was accomplished by replacing each input point with an appropriate hyperbolic
shape, with the amplitude scaled to be proportional to the amplitude of the input
sample. The figures on the following pages illustrate this process by increasing
the number of modelled traces in each figure. Only one trace (encircled below) is
used in Figure 2.11a. Two additional traces are included in (b) that are midway to
the sides. The following figures continue to increase the number of modelled
traces until every input sample has been mapped into a seismic section as in (f).
Notice that constructive and destructive interference creates the final image.

Figure 2.10 Input structure for diffraction models that are displayed on the
following pages.
This modelling example also scaled each sample on the hyperbolic diffraction by
the cosine of the angle of incidence, i.e. the angle from the vertical to the raypath,
or by T0/T. Additional filtering and time varying scaling will be required to match
the amplitude and phase of flat events before and after modelling.
Advanced modelling may assume incremental reflecting elements that contain
the dip of the reflector for angle dependent reflectivities.

Page 2.17

A Practical Understanding of Pre- and Poststack Migrations

a)

b)
x

c)

d)

Figure 2.11 The number of diffractions is increased to illustrate modelling with


(a) one trace, (b) three traces, (c) 8 traces, (d) 16 traces, e) 32 traces, f) 64 traces,
and g) all 100 traces.

Page 2.18

Chapter 2 Modelling with zero-offset

e)

f)
x

g)

Page 2.19

A Practical Understanding of Pre- and Poststack Migrations

Diffraction modelling is

very simple to set up and program

has minimal edge-effect problems

is accurate for depth-varying velocities

was demonstrated with hyperbolic curves to represent diffraction

illustrates constructive and destructive reinforcement.


A stacked section may be considered to be made entirely from diffractions even
though they are not visible.
x
z

a)
t

b)
Figure 2.12 Computed diffractions for buried focus, (a) the input scatter points,
and (b) the resulting pattern formed from the diffractions.
Each diffraction in Figure 2.12 is visible. Blurring this image could remove the
hyperbolic diffractions. Complete removal of the diffractions to create the
specula reflection requires a temporal bandwidth that is limited by the trace
interval, velocity, and dip. This relationship will be discussed later when dealing
with aliasing.

Page 2.20

Chapter 2 Modelling with zero-offset

2.6 Fourier transform modelling


2.6.1 Introduction
The figures in this section illustrate that images may be stored in a computer.
These images are stored as well-organized numbers and may be manipulated to
produce various effects. An example of photo deblurring is shown in Figure 2.13,
image enhancement shown in Figure 2.14 and Figure 2.15, and a series of
pictures illustrates coordinate transforms in Figure 2.16.
These images are similar to seismic sections that may be manipulated for similar
effects.
The 2D Fourier transform may also be used on images to convert the information
into a different form for other applications as illustrated in Figure 2.17. Its use in
seismic modelling is illustrated in Figure 2.18 and Figure 2.19.
These processes are usually part of any image processing package.

Page 2.21

A Practical Understanding of Pre- and Poststack Migrations

2.6.2 Illustration of deblurring

a)
Figure 2.13 Image processing showing in a)a blurred image, and (b) the
deblurred image.

Page 2.22

Chapter 2 Modelling with zero-offset

b)
What seismic processing procedure would accomplish this same effect?
What causes the border around Figure 2.13b?

Page 2.23

A Practical Understanding of Pre- and Poststack Migrations

2.6.3 Illustration of image enhancement

a)

b)
Figure 2.14 Image enhancement from (a) the original, to (b) the improved image.

Page 2.24

Chapter 2 Modelling with zero-offset

Figure 2.15 Image enhancement of the moon.

The processes used to accomplish these effects are readily available in image
processing software. What seismic processing procedure would accomplish this
same effect?

Page 2.25

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

2.6.4 Illustration
n of coorrdinate tra
ansforms
s

r
a)

y
c)

b)

d)

Figure 2.16
2
Spec
cial image deblurring
g is shown
n with (a) a radially blurred im
mage
(x, y), (b)
( a transfform to (r, ), (c) de
eblurring in direction, and (d
d) the resto
ored
(x, y) im
mage from [92].

P
Page
2.26

Chapter 2 Modelling with zero-offset

2.6.5 2D Fourier transform examples


The Fourier transform is a mathematical process that converts the data in one
domain (i.e., time) to a different domain (i.e., frequency). The different domains
contain the same data but present the information differently.

Different information becomes more obvious in different domains.

Low-cut or high-cut filters are easier to define in the frequency domain.

Radio stations send out a time signal (traces).

We tune to different stations by selecting different carrier frequencies.

The Fourier transforms a time trace into a frequency array.

The Fourier transform of distance becomes "wave number."

The 2D Fourier transform, transforms an image into a new space that has special
properties such as:

temporal and spatial frequency content

dip content of the image

the alignment of events with the same dip at the same location,

the ability to modify all events at one dip to have a new dip.

The images on the following page show a photo on the left and its Fourier
transform on the right. The mathematical origin of the transform is at the center
of each image.

Page 2.27

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

Figurre 2.17 Various phottos and the


eir Fourier transform
ms.

The 2D Fourier transforms


s in these figures ap
ppear to contain
c
litttle informa
ation
and ma
ay not seem
m useful at
a this time
e. Howeve
er, the 2D transform does simplify
many processing operations as illustrrated in the
e following
g section.

P
Page
2.28

Chapter 2 Modelling with zero-offset

http://www.cs.toronto.edu/~jepson/csc320/notes/linearFilters2.pdf

Figure 2.17cont.

Page 2.29

A Practical Understanding of Pre- and Poststack Migrations

2.6.6 Example of seismic modelling with the 2D Fourier transform


x

kx

a)

b)
x

kx

d)

c)

Figure 2.18 Modelling of a buried focus by the Fourier Transform with (a) the
input (x, z) section, (b) the Fourier transform (kx, kz) of (a) with the origin in the
upper left corner, (c) the modelled transform (kx, f), and (d) the modelled image (x,
t).

Dip angle in (x, t) or (kx, f) space are defined using tan ( ) = vkx / f .

Page 2.30

Chapter 2 Modelling with zero-offset

kx
z
kz

a)

b)
x

kx
t
f

450

d)

c)

Figure 2.19 Hockey stick structure modelled by the 2D Fourier transform with (a)
the input structure, (b) the 2D Fourier transform, (c) the modelled Fourier
transform, and (d) the modelled image.
The large spike in Figure 2.19a has only positive amplitudes that have significant
low frequency (wavenumber) energy as indicated by the gray shading at the top
of the KK transform in (b). After modelling this energy is confined to dips that are
less than 45 degrees as indicated in FK plot (c), producing the diffraction in (d)
where the dips are less than 45 degrees.

Page 2.31

A Practical Understanding of Pre- and Poststack Migrations

2.6.7 Some comments on Fourier transform modelling

Gives ideal solution with constant velocity

May give an adequate model when the velocities are smoothly varied and a
depth to time is included after the last transform

May have wrap-around problems

Indicates that no energy should exist above the 45 degree dip in a modelled or
seismic section

Is very fast

The reverse or inverse operation of modelling is migration

Page 2.32

Chapter 2 Modelling with zero-offset

2.7 Complex models: the real Earth


2.7.1 Introduction
The previous models assumed linear raypaths where the acoustic medium has a
single mode, homogeneous, isotropic, and lossless. We now consider geological
models that are more complex. This section discusses these complexities and
attempts to identify the simplifying assumptions that are made in many of our
seismic processing algorithms.

2.7.2 The Issues


Velocity, density and acoustic impedances
Block and smooth velocities
Reflection
Transmission
Frequencies
Modes and conversion
Incidence angles
Inversion
Raytracing modelling
Wave propagation modelling
Absorption
Anisotropy
Fluid content
Permeability
Porosity
Viscosity

Page 2.33

A Practical Understanding of Pre- and Poststack Migrations

2.7.3 Modelling Simplifications


Velocities

Constant

RMS, average, stacking, interval, instantaneous

Blocky required for Snell raytracing

Blocky assume straight raypaths between interfaces

Smoothing required for Eikonal raytracing

Smoothing destroys reflectivity

Velocities must be smoothed by smoothing the slownesses

Density

Assumed to be constant

Required for salt imaging

Blocky or smooth models

Multiple layers and or segments with constant velocity and density

Smoothing required for raytracing

Smoothing required for finite difference modelling

Reflections

Assume a blocky model with velocity contrasts

Internal reflection usually ignored (multiples)

Transmissions

Assume all energy transmitted

No internal reflections

Frequencies

Assume high frequencies in wavelet

Wavelength less than the size of the blocky velocity segments

Smooth velocity model to lower the higher frequencies (wave-numbers)

Page 2.34

Chapter 2 Modelling with zero-offset

Modes

Assume only P-wave exists, i.e. acoustic medium

No conversion of energy at interfaces

Assume P- and S-wave energy exists

Assume mode conversion exist at the boundary of interest

Mode conversion exists at the ocean bottom: P to S then S to P.

Angle of incidence

Assume it is 90 degrees

OK to 30 or 40 degrees

AVO or AVA requires Zoeppritz Equations

AVO simplifications

Absorption (Q)

None

Ignored in spiking type deconvolutions

Gabor deconvolution

Anisotropy

Ignored

Absorbed into stacking velocities

Higher order moveout equations

Should be included in all depth migrations

Fluid content

Can it be seen from surface seismic

Time-lapse seismic or 4D recording

Requires sophisticated algorithms for correct transmission

Inversion methods are required

Page 2.35

A Practical Understanding of Pre- and Poststack Migrations

Raytracing methods

Used to compute traveltimes

Sometimes used to estimate the amplitude along traveltimes

First arrival times

Maximum energy arrival times

Multiple arrival times

May include anisotropy

Cannot be constructed or computed between two points on different sides


of an interface

Paraxial ray shooting

Fan shooting

Iteration in raypaths

Iteration in ray parameter

Snell raytracing: block model

Eikonal raytracing: smooth model

Wave propagation modelling

Assumes a wavelet

No absorption

Usually uses a simple wave equation (not adequate)

Should use an equation for variable impedances

Boundaries must not be aliased

Smoothing of a blocky model is required to reduce aliasing

Grid dispersion

2D versus 3D modelling amplitudes

May include anisotropy

Finite difference approximations

May use Fourier transforms

Page 2.36

Chapter 2 Modelling with zero-offset

Porosity

A ratio of pore volume to the total volume

Different porosities (connected etc)

Connectivity issues

Can it be estimate using surface seismic?

Permeability

The ease with which a fluid can pass through the pore space in a formation

Can it be estimate using surface seismic?

Viscosity

Resistance of a fluid to flow

Varies with temperature

Shear component

Gas, oil, heavy oil, water, honey,

Page 2.37

A Practical Understanding of Pre- and Poststack Migrations

2.7.4 Discussion of some assumptions


Blocky velocities
Correct geology
Required for Snell raytracing
Can produce multiples
Smooth velocities
May represent geology for band limited seismic data
Produces no reflections, multiples?
Smoothing is required for Eikonal ray tracing
Required for Finite difference wave propagation
RMS velocities
In horizontally layered isotropic media, RMS velocities allow simple calculations
of traveltimes along raypaths. In a depth section, the raypaths are curved, but on
a time section the raypaths are assumed to be linear.
Reflections
Reflections come from interfaces with sharp impedance contrasts where
impedance is the product of the velocity and density. We can get reflections from
an interface where only the density varies, and there will be no refraction of the
raypath. We may have the usual case where we assume the density is constant
and the velocity changes to give both reflection and refraction energy.
Amplitudes
Reflections depend on the thickness of the transition interval between two media
with contrasting impedances. A smooth transition zone over a larger distance
may produce no reflection. The reflectivity of the transition zone is dependent on
the transition distance and the frequency of the energy. Higher frequencies may
produce no reflection, while lower frequencies may be reflected.
The amplitudes of energy that is completely transmitted through a smooth
transition will be dependent on the impedance of the medium. (Remember that
the amplitudes of earthquake energy increases when entering areas with low
velocity, i.e. build on the rock.)
Modes
There are many modes of transmitting energy in an elastic medium. We usually
consider only the P and S body waves. The particle motion for P-waves is
longitudinal or in the direction of the raypath. The particle motion for S-waves is
transverse or at a right-angle the direction of the raypath.
In an acoustic medium, there are only P-waves and the energy of the S-wave is
assumed to be zero. An elastic medium contains both P-wave and S-wave
energy. Energy of one mode may be converted to another mode at an interface.
Typical seismic projects only seek and record the P-wave reflections. All the Swave energy is ignored. Ignoring the S-wave energy will affect the amplitudes of
the P-wave data.
Page 2.38

Chapter 2 Modelling with zero-offset

The S-wave energy contains additional information about the rock properties. It
may be estimated using the amplitudes of a reflection at various source-receiver
offsets, i.e. amplitude versus offset or AVO.
There is a trend in seismic acquisition to record both the P-wave and S-wave
energy to obtain more reliable information to aid in the exploitation of
hydrocarbon reservoirs.
Angle of incidence
The equations that are usually used to defined the reflection and transmission
coefficient assume a ray path that is normal to the reflector. However, these
amplitudes are dependent on the incident angles. The actual coefficients are
quire complex and are given by the Zoeppritz equations.
Multiples
Reflections, where possible, are be produced at every interface, but may be
ignored in some raytracing algorithms or eliminated by velocity (or slowness)
smoothing. Ignoring the multiple energy will affect the amplitudes.
Absorption
The wavelet is not stationary with time. When seismic energy propagates
through any medium, it loses energy through friction. The loss is proportional to
the particle velocity, and when the amplitudes are the same, the higher
frequencies will have a higher particle velocity. Consequently, the higher
frequencies are attenuated (absorbed) more than the lower frequencies. This is
evident in a seismic trace where the higher frequencies are in the shallow
reflections, while the deeper reflections contain only lower frequencies. A
measurement of this energy loss is defined by the parameter Q.
Anisotropy
In an isotropic medium, the velocity is the same in all directions. In an
anisotropic medium the velocity varies with direction.
An example is a
sedimentary basin where the layering tend to be horizontal. The bedding may be
composed of alternating layers of slower and faster layers. The velocity of a
vertical raypath will be related to the average of the layered slownesses. In
contrast, the velocity of a horizontal raypath will equal the velocity of the faster
layers.
Using the correct wave equation
Many modelling and migration algorithms are based on the scalar wave-equation
that assumes a constant velocity (impedance). This is a major error and concern
in the industry.

Now, lets talk about true amplitude migration...

Page 2.39

A Practical Understanding of Pre- and Poststack Migrations

2.8 The blocky Marmousi Model


In 1988 the Institute Francais du Petrole (IFP) created a complex 2D geological
model and generated synthetic seismic data. This model was distributed to
numerous institutions and companies for a blind test. The velocity model was
withheld from the participants until May 1990, when it was released at the EAEG
workshop on "Practical aspects of seismic data inversion".
Thirty-one invitations resulted in eight participants who were willing to compare
their blind tests. The proceeding of the conference [151] contains their results.
The model is complex and was based on a profile across the North Quenguela
trough in the Cuanza basin in Angola. Both density and velocity models with
high horizontal and vertical velocity gradients were used. The length of the
model was 9200 meters with a depth of 3000 meters. The grid size was 4 by 4
meters.
The seismic data simulated a marine acquisition from 3000 meters to 8975 meters
across the model, and consisted of 240 sources, each with 96 traces.
The results of the blind test show considerable differences between the final
velocity models, illustrating the problems in the depth migration process. One
interesting result, however, was the average of the models was superior to the
individual models. (See Versteeg's Doctoral thesis [152]).
Figure 2.20 on the opposite page shows the velocity model.
simple modelling programs is assumed to be constant.

The density in

The data has become a standard 2D test set that is used to compare or illustrate
the success of various pre or post-stack, time or depth migrations.
The 1990 EAEG workshop proceedings [151] and Doctoral thesis by Versteeg
[152] are an excellent source of material on the Marmousi experiment. (Note that
the title is in French, but the text is in English.)

Page 2.40

Chapter 2 Modelling with zero-offset

a) Vertical and horizontal axis equal

b) Exaggerated vertical scale


Figure 2.20 The Marmousi model a) with equal horizontal and depth scales and a
velocity legend, and b) an enlargement showing the velocity details. The
velocities are defined in the colorbar on the right side of (a).

More complex models of the Marmousi structure are now available.

Page 2.41

A Practical Understanding of Pre- and Poststack Migrations

2.9 Ray tracing in a blocky model


2.9.1 Introduction
Ray paths are no longer straight and require additional attention when estimating
traveltimes. The algorithm may depend on how the geological model is defined
such as:

Blocky model with constant interval velocities, or

Smooth velocities with no sharp velocity transitions.

In a blocky model, we use Snells law to moodily the trajectory angle of a ray path
through the different interfaces. The raypaths are assumed to be constant within
a block.
The following figure is a simplification of array arriving at an interface. The ray is
assumed to be part of a P-wavefront, and only produces a reflected P-wave and a
transmitted P-wave. Mode conversion to s-waves is ignored, or the medium is
acoustic (fluid).

V1
2

V2
T

Figure 2.21 Ray diagram illustrating only a single mode for reflection and
transmission.
In this figure,
1 = incident angle

2 = transmitted angle
V1 = velocity in the upper layer (P or S)
V2 = velocity in the lower layer (P or S).

Page 2.42

Chapter 2 Modelling with zero-offset

2.9.2 Snell raytracing with constant interval velocities.


We make use of Snell's Law of refraction that relates the angles of incidence and
transmission to the velocities in the respective layers.
and Snells law is given as

sin 1 V1
= .
sin 2 V2

(2.2)

We now include the density and define the impedance as I i = iVi to get the
normal incidence reflection coefficient R, i.e.,

R=

I 2 I1 2V2 1V1
=
I 2 + I1 2V2 + 1V

(2.3)

and the normal incidence transmission coefficients T,

T=

2 I1
2 1V1
=
.
I 2 + I1 2V2 + 1V

(2.4)

This simple model assumes the raypath is normal to the reflector, even thought
the raypaths are drawn at angles.
The density can be modified to make the acoustic impedance constant to prevent
reflections, but any change in the velocity will produce refractions.
The reflection and transmission coefficients do vary with angle, and become
quite complicated when we use both the S-wave and P-wave velocities in each
block. The actual coefficients are computed from the Zoeppritz equations as
indicated in the following section.
Simplifications to the Zoeppritz equations leads to prestack analysis where the
amplitudes of a P-P reflection, versus the source-receiver offset, leads to an
estimate of the S-wave velocity and then to other rock parameters.
Snell's law can be derived by a beam arriving at an interface in [550 page 63], or
from the minimum traveltime along a raypath [Stein, An Introduction to
Seismology, Earth Quakes, and Earth structure, 2005, Cambridge, pp 71 71.]

Page 2.43

A Practical Understanding of Pre- and Poststack Migrations

2.9.3 Amplitude versus angle: Zoeppritz Equations


A blocky model that contains the density, P-wave velocity, and the S-wave
velocity, requires a complicated algorithm to compute the reflected and
transmission coefficients as a function of incident angles. The diagram below
illustrates the complexity and is not intended as a recipe. The solutions are
known as Zoeppritz equations and are rarely presented in explicit form.

Figure 2.22 Illustration of the process required to compute solutions for


Zoeppritz equations.
http://www.google.ca/search?q=Zoeppritz+equations&rls=com.microsoft:en-us:IE-SearchBox&ie=UTF8&oe=UTF-8&sourceid=ie7&rlz=1I7SUNA_en&redir_esc=&ei=O8ILTeynLJO6sAOegfW9Cg

Page 2.44

Chapter 2 Modelling with zero-offset

To illustrate the complexity of the equations, assume a purely acoustic medium


with no S-waves. The reflection coefficient may be simplified to
1

V 2
2
Z 2 Z1 1 2 1 tan 2 (1 )
V

R=
1
V 2
2
Z 2 + Z1 1 2 1 tan 2 (1 )
V

(2.5)

as given by http://www.sal2000.com/ds/ds3/Acoustics/Wave%20Reflection.htm .
The following example is of an interactive CREWES Explorer to compute the
reflection and transmission coefficients for angles from 0 to 90 degrees. The blue
line is the transmission coefficient, the red line the reflection coefficient.

Upper layer

1 = 2000 kg/m3
Vp1 = 3000m / s
Vs1 = 1500m / s

Lower layer

2 = 2200 kg/m3
Vp2 = 4000m / s
Vs 2 = 2000m / s

Figure 2.23 An example of reflection and transmission coefficients as a function


of angle from an interactive CREWES explorer
This example illustrates the reflection and transmission coefficients are close to
the normal incident values for small offsets deviate wildly when the angel of
incidence approaches the critical angle.
The explorer may be accessed at:
http://www.crewes.org/ResearchLinks/ExplorerPrograms/ZE/ZEcrewes.html

[See also: The Rock Physics Handbook, by Mavco, Cambridge, 1999,pp.61 -64,
and Aki and Richards, Quantitative Seismology, 2002, pp 139- 145]
CREWES: Consortium for Research in Elastic Wave Exploration Seismology

Page 2.45

A Practical Understanding of Pre- and Poststack Migrations

2.9.4 Point to point raypaths


Except in a constant velocity medium,
it is not possible define a ray path explicitly between two points.

Some form of iteration is required.

Possible methods
1. A ray is started normal to a reflector and then traced up to the surface. The
error in the desired surface location can be used to modify the departure
location.
2. A ray can leave the surface and travel to an extended linear reflector where
the angle of incidence is defined at the intersection point. The departure
angle can be modified until the incident angle is normal to the reflector.
3. A family or fan of rays can be drawn from one point to the area of the target
location. The closest ray is either good enough or an iterative process is
used to modify the departure angle of another ray to minimize the error.
4. A family of rays from each surface point are traced to all parts of the
geological model. Traveltimes along the raypaths may then be converted
to traveltimes on a grid. This method saves considerable computation
time, but requires large amounts of memory to store the traveltime maps.
5. A ray parameter can be defined and then iterated numerically to find the
optimal path, as illustrated below in Figure 2.24.

x
2
3

Figure 2.24 Example of a ray traced between two points.

Page 2.46

Chapter 2 Modelling with zero-offset

2.9.4.1 Iterating the ray parameter


Rather than iterate the construction of raypaths, it is possible to iterate
numerically with the ray parameter p which is much faster and possibly more
accurate.
Consider raytracing in horizontally layered media that is composed of constant
velocity layers as illustrated in Figure 2.24. From Snells law, p is the same in
each layer, i.e.,

p=

sin i
sin 1 sin 2
=
= ... =
,
V1
V2
Vi

(2.6)

where 1 is the takeoff angle at the surface, i the incident angle of the raypath in
the ith layer, and Vi is the velocity in the ith layer.
The distance from the source to receiver is defined as X. The thickness and
horizontal component of the raypath in each layer are given by zi, and xi, where

xi =

pVi zi

(1 p V )
2

(2.7)

We need to solve for p such that


n

X = xi .
i =1

(2.8)

Unfortunately, p is implicit in equation (2.7) and must be solved iteratively. Less


than five iterations are usually required when using a Gauss-Seidel method.
Once p is known, the raypath is defined and traveltimes can be computed.
[See paper by Krebes, 2004, Seismic Forward Modeling, CSEG Recorder, April
2004, pp 28 39.]
The above derivation for p was based on horizontally layered media and a direct
raypath. It is possible to include multiples in the raypath as indicated by Krebes.
I believe that it is possible to use this method for all blocky media but have no
reference at this time. Problems of multi-pathing do exist and make the
programming of a solution more difficult.

Do we really need to be extremely accurate when raytracing?

Page 2.47

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

2.9.5 A simple raytracing


r
g model in a blocky mediu
um
A mode
el structurre is show
wn below in Figure 2.25 and the resultting ray-tra
aced
section in Figure 2.26. The
e modelling
g method applied Sn
nells law to
t normal rays
from re
eflectors, and
a
to sca
atter points on the central
c
vertical line. Diffuse rays
through
h the pinch
houts are ig
gnored.

Figure 2.25
2
Simp
plified ray--tracing sh
howing in the input structure with cons
stant
interval velocities
s.
Note:

Norm
mal rays le
eave reflecttors at righ
ht angles

They
y do not ta
ake into acc
count the energy
e
from internal reflections

This model wa
as not smo
oothed for the
t ray tracing

Poin
nt reflectors are in ve
ertical line

Poin
nt reflections are nott in a stra
aight line, and peaks
s move to the right with
incre
eased deptth

Diffrractions are not continuous and


d have dis
scontinuitie
es (why?)

Deep
p events th
hat are fla
at on the in
nput structture have pull-up on
n the mode
elled
section.

P
Page
2.48

Chapter 2 Mo
odelling with ze
ero-offset

a)

b)

Figure
e 2.26 The modelled time sectio
on a) and b)
b a spatia
ally compre
essed view
w of
(a), inc
cluded to emphasize
e
e the distorrtion in the
e diffractions.

Some claim
c
that rays
r
dont exist but only
o
waveffronts existt. A ray tra
aveling across
a linearr but rough
h interface may be sc
cattered alll any direc
ctions. Propagation of a
wavefro
ont across
s the same interface will rem
main coherrent (main
nly due to
o the
summin
ng effect of Huygens
ss principle). Rays should
s
be normal to the wavefrront;
what is the proble
em???
cing algorrithms ass
sume that velocity (actually th
he slowness S = 1/V
V) is
Ray-trac
smooth
h relative to
o the size
e of a wav
velet on a seismic
s
tra
ace. This smoothnes
s
ss of
the slow
wness is an
a assump
ption requiired to obttain the Eikonal equation from
m the
ray equation or th
he wave eq
quation.
Smooth
hing the sllowness te
ends to prreserves traveltimes
t
s [267]. The
T
size off the
smooth
her should be greaterr than the wavelength
w
h of the da
ata [716].
(Discon
ntinuities of
o the velo
ocity are re
equired to define the
e reflection coefficie
ents.
Reflectiion energy
y may be frrequency dependent
d
t, dependin
ng on the on the rela
ative
size of the
t transition zone between
b
tw
wo impedan
nces.)

P
Page
2.49

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

2.10 Smoothi
S
ng the velocitie
v
es (slownesses))
Smooth
hing the velocity
v
f
field
is re
equired when
w
raytrracing witth an Eik
konal
algorith
hm, or whe
en using a finite
f
difference wav
ve propagation algoriithm.
The Eik
konal raytrracing algo
orithm requires smo
ooth velociities, espe
ecially in areas
a
with complex stru
ucture. Sm
moothing tends to prrevents ray
ys from be
eing trappe
ed in
higher velocity
v
ch
hannels and to appro
oximate the
e propagattion of the wavefrontt.
Wavefro
onts are defined with a lower bandwidth
h and tend
d to smootth the effec
ct of
rapidly varying ge
eology. In contrast rays
r
follow
w the highe
er frequenc
cy structurre.
Conside
er a roug
gh interfac
ce that is struck by
b a numb
ber of pa
arallel rays
s as
illustratted below. The rays emerge in many different direc
ctions.
Wavefro
onts emerg
ging from the interfa
ace tend to
o smooth the
t next wa
avefront (h
heal)
as illus
strated by Huygens method. Energy propagates as waveffronts and
d not
rays.

a)

b)
Figurre 2.27 Comparison of a) ray propagation
n with b) wavefront
w
p
propagatio
on.

P
Page
2.50

Chapter 2 Modelling with zero-offset

The velocities should not be smoothed.


Smoothing the velocities introduces timing errors.
reciprocal of the velocities that should be smoothed.

It is the slowness or

The following model in Figure 2.28a was used to trace rays to the surface.
Traveltimes to the surface are plotted as black squares in the traveltime plot of
(b). The velocities were then smoothed and the resulting traveltimes plotted as
the red line. The slownesses were smoothed and plotted as the blue line.

a)

b)
Figure 2.28 Raypath traveltimes estimated from a Wedge model in a) and in b) the
traveltimes.
Smoothing the slownesses preserves the traveltime of the raypaths.

The slownesses should be smoothed.

Page 2.51

A Practical Understanding of Pre- and Poststack Migrations

2.10.1 Eikonal ray tracing with smooth velocities.


A summary of an accurate ray tracing algorithm from [716] is shown below. Also
see [781], [784] and [785].
Starting with the scalar wave equation where f(x, y, z, t)

2 2 2 1 2
+
+
=
or
x 2 y 2 z 2 c 2 t 2

2 =

1 2
1
or 2 = 2 2 ,
2
2
c t
c

(2.9)

we get the Eikonal equation

( )

1
t t
= 2 or + = 2
c
c
x z
2

(2.10)

This equation is expanded and the real parts collected to give the ray equation

( n ) = or

S ( r ) = S ( r )

(2.11)

where S is the slowness vector and s is a discrete increment along the raypath.
This equation may then be sorted into x and y coordinates giving;
Sx + S x x + S z xz S x = 0

Sz + S z z + S x xz S z = 0

(2.12)

Introducing two terms u and v as x = u and z = v we get four first order equations
x=u
Su + S u + S uz S = 0
x
z
x

z=v

u = 1 [ S x S x u S z uz ]

(2.13)

that finally give the finite difference equations


xn +1 = xn +

un +1 = un +

zn +1 = zn +

vn +1 = vn +
S

un
S x un ( un S x + vn S z )
vn
S z vn ( un S x + vn S z )

(2.14)

The ray now advances from (xn, zn) to a new location at (xn+1, zn+1), with un+1 and
vn+1 are sine and cosine of the incident angle at n+1.

Page 2.52

Chapter 2 Modelling with zero-offset

Note that the u and v terms are related to the angle of the raypath by u = sin
and v = cos .
In equation (2.8) we are extrapolating or predicting the future values of (xn+1, un+1,
zn+1, vn+1), and that the derivative is defined at the previously established location
(xn, un, zn, vn). Ideally the derivative should be known at a point midway between
the established and predicted location say at n+1/2. There are numerous
mathematical techniques to improve the accuracy of the predicted values such as
the Runge-Kutter algorithm contained in may texts such as Numerical Recipes
[187].
A simple approach is to use initial estimates of the n+1 values to estimate new
derivatives. Average the derivatives at n and n+1 to get an estimate for the
derivative values at the n+1/2 location, then recompute new values at n+1.
This is referred to as the improved Euler method.
An example of this algorithm in a linear V(z) medium is shown in Figure 2.29 that
traces five circular raypaths. The (+) signs are exact solutions, and the solid lines
are the solution using the Euler method using large ray increments (100m).

Figure 2.29 Five raypaths computed using the Euler method.


This result is quite significant when compared to a ray-tracing algorithm that is
based on Snells law which will not allow a ray to turn when the velocity is
constant. In Figure 2.29 the velocity is constant at the horizontal portion of the
raypath and it is the vertical component of slowness Sz that allows the ray to
curve up when using a solution based on equation (2.8).

Page 2.53

A Practical Understanding of Pre- and Poststack Migrations

2.11 Raytracing amplitudes


The amplitude along a simple hyperbola is usually define as T0/T where T0 is the
vertical two-way time and T the two-way time to a colocated source and receiver.
When ray tracing is involved, equation (2.10) may be expanded and the imaginary
terms collected to give the Transport equation. This equation will give the
amplitudes along the raypath, where, at the surface, it defines the amplitude at
the diffraction [716], and [Krebes].
Rather than use the Transport equation, some algorithms use Paraxial rays that
start at the same location and are very close together. In 2D, the amplitude at a
given time is defined from the distance between two rays d, i.e., Amp2 D 1/ d .
In Figure 2.30 we see linear, curved, and intersecting paraxial rays. In 2D, it is
possible that the rays intersect and the amplitude that result from this method
will tend to infinity, a real problem if this occurs at the surface.
In 3D, the amplitude is computed from three rays and the area of the triangle A,
defined at a given time along the raypaths, provides the amplitude
Amp3 D 1/ A . In Figure 2.31 we see three rays in 3D space with a triangle
drawn at a given time to represent the area. An ellipse is also drawn to illustrate
the area of the bundle or tube. In 3D space, rays rarely intersect.
The 3D amplitude is proportional to the area of the tube, and if the area is
represented by a square with side s, then the amplitude will be directly
proportional to the length of one side of the square, i.e., Amp3 D 1 / s .
In

homogeneous

medium,

and

are

proportional

to

R,

giving

Amp2 D 1/ R , and Amp3 D 1 / R .


Figure 2.30c illustrates two different raypaths that can arrive at the same point;
this is referred to as multi-pathing.
A diffraction may be more complex than a simple hyperbola or a single valued
function. Due to multi-pathing, it may fold back and forth providing three or more
traveltimes with different amplitudes at the same location on the surface. The
shapes produced by this effect are referred to as caustics or triplications, and are
illustrate in Figure 2.32. The amplitudes and phase vary significantly along the
caustics.
Many of the algorithms in commercial use may simply use the spherical
spreading factor of 1/R or 1/T where R is the distance along the raypath, or T the
traveltime along the raypath.

Page 2.54

Chapter 2 Modelling with zero-offset

a)

b)

c)
Figure 2.30 2D paraxial rays that are a) linear, b) curved, and c) curved but
crossing each other.

Figure 2.31 3D paraxial rays.

Figure 2.32 A diffraction containing multi-pathing or caustics..

Page 2.55

A Practical Understanding of Pre- and Poststack Migrations

2.12 Defining a diffraction shape for modelling


2.12.1 Introduction
Any point in the subsurface can be considered a scatterpoint. Energy arriving at
the scatterpoint is scattered in all directions. In exploration geophysics, we refer
to this energy, when recorded at the surface, as a diffraction.
These diffractions were combined in a previous section to produce a zero offset
seismic sections and were assumed to have a hyperbolic shape.
We also use the term diffraction to represent the reflection energy from a
truncated reflector.
This hyperbolic shape is the same for moveout correction of horizontal reflectors
or for Kirchhoff migrations where we refer to the diffraction as an impulse
response or a Greens function.
Consider Figure 2.33 where a scatterpoint is located in a constant velocity
medium at A. Rays emanate from the scatterpoint towards all points on the
surface. At B, the raypath is vertical with a distance Z, and to a point at C with a
distance L. The distance from B to C is xBC. The point D is defined to be below C,
at a depth equal to L. By allowing the point C to move across the surface, the
point at D will define the curve as constructed in the exercise of Figure 2.5. From
the triangle ABC we have
2
L2 = Z 2 + xBC
,

(2.15)

which defines the shape of the curve, and is the equation of an hyperbola.
If we assume a constant velocity V, we can scale and overlay a time section (x, t)
with horizontal distance x and the vertical coordinate time t.
This figure now represents traveltimes of the raypaths, and if we assume our
construction to be a zero offset section, we must consider the times along a
raypath from the surface at C to the scatterpoint A, and then back to the surface
point at C. This total time is referred to as the two-way time. For locations B and
C the two-way times are T0 and T. The horizontal distance x becomes a two-way
time measurement with t = 2 x / V , giving the hyperbolic diffraction equation

2x 2
T =T + 2 .
V
2

2
0

(2.16)

This hyperbolic equation is still the basis for most traveltime computations in
seismic processing, such as normal moveout (NMO) correction, or the double
square-root equation used in prestack Kirchhoff time migration.

Page 2.56

Chapter 2 Modelling with zero-offset

xBC
C

T 0 or Z

T or L
T or L

t or z

Figure 2.33 Traveltimes of a diffraction in a constant velocity medium.

In a layered media, the raypaths are curved and require complex calculations to
provide a traveltime. However, we can use an rms velocities that significantly
simplifies the problem by assuming the velocity to be constant and the
traveltimes computed using equation (2.16).
As seismic processing has evolved, our ability to build geological models of our
subsurface has increased, along with the computing power to compute more
accurately the raypaths and wavefronts in our complex media. Consequently
depth migration makes use of these enhancements.

The following sections illustrate the complexities in defining the diffraction


shapes by used by depth migrations.

Page 2.57

A Practical Understanding of Pre- and Poststack Migrations

2.12.2 Ray tracing a diffraction


We now consider the shape of the diffraction to be more complex and defined by
the medium through which it passes.
We use the Marmousi model to illustrate the complex raypaths that exist in a
blocky model that is smoothed. Eikonal raytracing is used from a scatterpoint at
(5500, 2450), with a family of 100 raypaths that are spread over departing angles
with equal increments from 0 to 180 degrees. The rays are traced in Figure 2.34
to the maximum extent at the bottom, sides or top of the model. Four different
size of smoothers are demonstrated with rays in red. The black curves connect
equal traveltimes to simulate wavefronts. The raypaths on the left of each figure
have a smoother appearance as the model tends to less structured, while in the
rays on the right are more scattered.
The curves in Figure 2.35 are diffractions that are constructed from the arrival
times across the surface. The arrival times are connected in order of the
departure angle. Note the complexity the curves as they move back and forth in
their arrival location at the surface. Significant smoothing (of the slownesses) is
required to produce a reasonable looking diffraction with caustics as in (c).
Amplitudes of the diffraction can be defined by the horizontal distance between
adjacent rays at the surface; more specifically, between rays defined by their
order of departure. Rays varying wildly at the surface will have very small
amplitudes while rays that are bundled together will have greater amplitudes.
(This is more evident in the next section).
These figures verify that ray tracing is a high frequency assumption. The rays
will faithfully follow a blocky model and not represent the actual wavefield as
illustrated in the next section. Consequently, smoothing of the slownesses is
required. A paradox is that we need a blocky model to define reflecting surfaces,
especially when including internal multiples.

An alternate method of smoothing uses an unsmoothed blocky model, then


smoothes the traveltimes at each surface location.

Page 2.58

Chapter 2 Modelling with zero-offset

a)

b)

c)

d)

Figure 2.34 Ray tracing and simulated wavefront construction of a scatter point
at (5500, 2450) with smoothing lengths of a) 50, b) 150, c) 250, and d) 500 m.

a)

b)

c)

d)

Figure 2.35 Effects of smoothness on traveltime computation with smoothing


lengths of a) 50, b) 150, c) 250, and d) 500 m. (Figures produced by Hassan
Khanaini).

Page 2.59

A Practical Understanding of Pre- and Poststack Migrations

2.12.3 Comparing raytracing with wave propagation modelling


An alternate method for defining a diffraction is to model the wavefield from a
scatterpoint and record the response at the surface as illustrated in Figure 2.36.
This required computing many wavefields at equal time increments of a minimum
phase wavelet leaving the scatterpoint. The location of the scatterpoint is the
same as the previous raytracing examples, i.e. (5500, 2450).
The recorded wavefield at the surface, is shown in Figure 2.36a, while (b) shows
the same recording with automatic gain control (AGC).
These figures also include the traveltimes and locations of the 100 raypaths from
the smoothed slowness model in Figure 2.34c and Figure 2.35c as blue points.
The times for computing the ray traced traveltimes was a few seconds, while the
wavefield traveltimes required six minutes.
A greater number of ray are required to fill in some of the areas along the
diffraction.
Note:

The complexity of the wavefield as it arrives at the surface. It is not a


single event as assumed by the first arrival times.

In Figure 2.36b that the arrival times from the smoothed raytracing
coincides with the first arrival times of the AGCd wavefield.

The spread between the raytraced points coincide with the amplitudes on
the wavefield.

The arrival times from raytracing do not match the location of the maximum
energy of the wavefield in Figure 2.36a.

A single valued diffraction could contain first arrival times or maximum


energy (amplitude) times.

Depth migrations will use either raytracing or wave propagation in their


algorithms.
Raytracing is much more economical and when used with multiple arrivals
produces reasonable results and is considered the work horse of the industry.
Wave propagation is much more expensive, but is capable of producing better
results.
Both these methods require an accurate velocity model.

Page 2.60

Chapter 2 Modelling with zero-offset

a)

b)
Figure 2.36 Wave propagation modelling compared to raypath modelling for
defining a diffraction. (Figures produced by Hassan Khanaini).

Page 2.61

A Practical Understanding of Pre- and Poststack Migrations

2.13 Eikonal grid method for computing traveltimes


2.13.1 Traveltimes computed on a grid in (x, z).
Traveltimes may be computed on a grid as illustrated in Figure 2.37. The velocity
in each gridded box may vary, allowing the computation of traveltimes in media
with complex velocity structures. The resulting diffraction shape will not be
hyperbolic.

c
b

Traveltime
contours

a)

b)

Figure 2.37 Traveltime modelling showing: a) a partial grid to located traveltimes


and the contours of traveltimes, and b) the resulting diffraction from the
traveltimes at the surface of the structure, i.e. z=0.

Page 2.62

Chapter 2 Modelling with zero-offset

2.13.2 Traveltimes computations


Traveltimes computations are illustrated in Figure 2.38:

Initially compute the closest four (or sixteen) grid points directly from the
source * in part (a). The known grid times are represented by black dots.

Locate the minimum traveltime on each side of the known traveltime shell as
highlighted by the yellow band in the second shell of part (b). At this location,
the wavefront is assumed to be parallel to the side or a ray normal to surface
illustrated in (d). See [136] for
The travel time to one point in the third shell (green) is computed using
perpendicular distance and the two velocities in the neighboring boxes.

The remaining unknown point (represented in (b) as circles) are then


computed from the three known points in shown in (e) and described in the
next section.

After each side is computed, the four corner point are then computed using
the three known points as illustrated in part (c).
A minimum
traveltime

Next point

a)

b)

c)
t4

t3

W. F.

t1

d)

t2

e)

Figure 2.38 Traveltime computations on a grid of points: a) the initial points, b)


expanding the shells, c) the corner points d) minimum traveltime corresponds to
tangential wavefront and normal ray, and e) solving for a fourth point on the grid.

Page 2.63

A Practical Understanding of Pre- and Poststack Migrations

2.13.3 Computing the fourth point


The key to the process is computing the traveltime t4 at a fourth point on a box
when the traveltimes t1, t2, and t3 are known at three points.

One method uses the geometry of a plane wavefront (Figure 2.39a) to


compute the unknown traveltime. It is assumed in this figure that t1 < t3
and t1 < t2 and that the t3 wavefront intersects the line connecting the two
point of t1 and t2. (The location of t3 could be anywhere on the horizontal
line through t1 and t2.) The wavefront through t4 and is parallel to the t3
wavefront. Given the geometry and times on the line through t1 and t2, the
desired time t4 can be computed. from

t 4 = t 2 + t3 t1 .

(2.9)

A second method (Figure 2.39b) assumes a circular wavefront and that the
time at the center of the circle is t0 that is not necessarily zero. The velocity
surrounding the box is assumed to be the same as that in the box. Finding
the location of the center of the circle (x0, z0) requires solving a fourth order
(quartic) equation (a difficult, but possible task). Once the center is found,
the traveltime to the fourth t4 point is trivial, Vidale [136]. (This problem is
similar to the navigation problem shown in Figure 3.6.)
The value of t0 may be expressed in an implicit form that is independent of
(x0, z0) and estimated using Newton-Raphson iteration techniques. Given
t0, (x0, z0) can be estimated and the t4 evaluated.
A new solution based on the Apollonius method greatly simplifies this
problem.

t4

t3
t1

t4

t2

t4

t3
t1

t2

t3 -t1

t3

t4 wavefront

t0

t3 wavefront

a)

b)

Figure 2.39 Computing the traveltime of the fourth point assuming: a) a plane
wave, and b) a circular wavefront.

Page 2.64

Chapter 2 Modelling with zero-offset

A third solution, also by Vidale [136] uses the finite difference method to
solve the Eikonal equation
2

t t
+
x z

1
,
v2

(2.10)

2h2
2
t4 = t1 +

t
(
)
,
3
2
v2

(2.11)

giving the solution

where v is the velocity of the box, (or the average at the four corners), and
h is the distance on each side of the box.
The Eikonal equation can be derived from the wave equation Scales [716 p79-81],
or from geometry Slotnick [542], or Nolet [383]. Applications to wavefront
modelling are found in [109], [268], expanding wavefront [143]-[423], [480], [515]
and [716].
The derivation of the Eikonal equation from the wave equation requires the
assumption of smoothly varying velocities, and is more accurate for higher
frequencies. The Eikonal equation also provides a foundation for ray theory
(section 2.4.2).
The Eikonal equation (2.10) is a square law relationship between the vector
components of slowness (reciprocal of velocity) as illustrated by the geometry of
a, b, and c in Figure 2.37a. The components a and b represent the x and z
components of velocity, and c the actual velocity.

The above method computes the traveltimes using square shell. This method
works well when the wavefront is a simple surface, but may not produce the first
arrival times in complex areas such as those encountered with refraction energy.
The progress of computing the unknown time should follow causality [136] and
for stability, should be larger than the times from which it is calculated.
More advanced methods (Qin [143] and Popovici [666]) update the location of the
minimum traveltime from all possible updated times, causing the traveltimes to
be updated on a surface that approximates the wavefront. See also [786]. Sethian
and Popovici March 1999 Geophysics

This method of modelling is an integral part of many prestack depth migrations.

Page 2.65

A Practical Understanding of Pre- and Poststack Migrations

2.14 Exploding Reflector Model


2.14.1 Introduction
The exploding reflector model, introduced by Loewenthal 1976 [8] is an extremely
important concept. It allowed the first migration of zero-offset (stacked) section.
It also allowed a variable velocity distribution.
The zero-offset case assumes the raypath from the source to the reflector is
identical to the raypath from the reflector to the receiver. Only one raypath is
required; the path from the reflector to the receiver, i.e. the path followed by an
exploding interface.
Assume:

the reflectors emit normal rays


the rays are traced at equal increments of time
the surface reconstructed by the raypaths is the wavefront
the energy and traveltimes at the surface are mapped to the seismic section
the model honors all structures

It is not necessary to keep track of all the rays, but it is necessary to know the
wavefront. The image of the wave font at discrete times, gives the illusion that
the reflector has exploded.
The amplitude on a plane-wave that leaves a linear reflector is made proportional
to the reflection coefficients. This same amplitude will travel to the surface
Usually one-way wave propagation is used to eliminate multiple reflections:
ignore losses at boundaries
ignore transmission loss
True amplitudes are not provided.
Do we want to preserve the internal or multiple reflections?
The exploding reflector model assume the initial condition is the geological
model, the seismic data as a boundary condition, and the wave equation as the
connecting link.
The power of the exploding reflector model lies in treating a stacked seismic
section as a measurement of a wavefield. Although only an approximation, it is
such a useful model that methods based on it still dominate migration. Gardner
[528] page 141.

Page 2.66

Chapter 2 Modelling with zero-offset

T=0

z
x

T=2

T=5

z
Wavefront
in depth
domain

T=8

x
2
Seismic in
time

5
8
t

Figure 2.40 Wavefronts of "exploding" reflectors are shown at T=0, T=1, and T=2.
Sketch the resulting seismic section in the space provided.

Page 2.67

A Practical Understanding of Pre- and Poststack Migrations

2.14.1.1 Photograph model


Figure 2.41 represents photographs of an exploding source by showing (a)
multiple exposure, and (b) a time sequence of photographs at constant time
intervals. In Figure 2.42, the photos are arranged in 3D order with time
orthogonal to the x/z plane. Inserting a large number of photos results in the
volume of information shown in Figure 2.43.

a)
x

x
z

x
z

T0
T1

T2
x

x
z

T3

T4

b)
Figure 2.41 Photos of an exploding point showing (a) a multiple exposure, and
(b) a series of photos taken at constant time intervals.

Page 2.68

Chapter 2 Modelling with zero-offset

Figure 2.42 Photos arranged in time sequence.

Figure 2.43 Continuum of photos in (x, z, t) volume.

Page 2.69

A Practical Understanding of Pre- and Poststack Migrations

2.14.2 Exploding scatterpoints


time

time

a)

b)
time

time

c)

d)
time

time

e)
f)
Figure 2.44 Examples of exploding reflectors at discrete times of: a) t=0 showing
the scatterpoint on a dipping event, b) t=1, c) t=3, d) t=5, e) t=7, and f) t=9, also
shown are a composite of the discrete times g) 1 - 9, and h) the continuous time
surface seismic response.

Page 2.70

Chapter 2 Modelling with zero-offset

time

x
g)
time

h)

Page 2.71

A Practical Understanding of Pre- and Poststack Migrations

2.14.3 Model examples of exploding reflectors


Figure 2.45 shows an example of an exploding reflector model with (a) a 3D view
of the volume and (b) the top of the output surface displayed as a 2D section.
Note:

Front surface (x, z) of the cube is the input structure

Top surface (x, t) now represent the output time section

Output traces on the section are horizontal, (top of photos)

The diffraction shape from the point

The horizontal event with its full end effects.

Amplitudes on the exploding reflectors may be set proportional to the reflection


coefficients. This method of modelling can be simple as in the following
examples with a constant velocity, or may be complex with variable velocities.
Typical mathematical algorithms used in exploding reflector modelling use a oneway wave equation that obeys Snell's Law of refraction. In addition, it does not
change the amplitudes of wavefront as they cross a velocity interface.
Other forms of the wave-equation solution allow reflections off velocity interfaces
and the surface to produce multiples. For example, these methods would be
used to produce shot models with surface reflections and multiples.
All methods must be concerned with boundary conditions at the surface, edges,
and bottom of the model.
The images from inside the volume may be displayed at various time intervals, as
shown in Figure 2.46. Ten internal "photos" from various times shown a more
complex syncline model. Note:

Energy from each reflector moves away from the original position.

Energy may be seen advancing to surface (z=0).

Energy advancing to the bottom of the images.

Figure 2.26k is a collection from the top of many "photos", and contains the time
record of the wave field as it reaches the surface (x, t, z=0).

Page 2.72

Chapter 2 Modelling with zero-offset

a)
x
t

b)
Figure 2.45 The model that results from a point and a flat reflector using the
exploding reflector model.

Page 2.73

A Practical Understanding of Pre- and Poststack Migrations

a) t=2

b) t=25

c) t=50

d) t=100

e) t=150

f) t=200

Figure 2.46 Exploding reflector model at various times as indicated from (a) to (j),
with axis x and z, and with the output time section (k) with axis x and t.

Page 2.74

Chapter 2 Modelling with zero-offset

g) t=300

h) t=350

i) t=400

j) t=450
x
t = 100

k) surface (x, t, z=0)

Page 2.75

A Practical Understanding of Pre- and Poststack Migrations

2.15 Huygens' method of wavefront modelling.


2.15.1 Introduction
How is the energy moved from one time image to another time? This is normally
done by using a solution to the wave equation and was the method used in the
previous examples. The diagrams could have been formed by hand using a
wavefront construction involving Huygens' method of modelling (illustrated in the
previous Figure 2.46).

An insight to this method will be gained from the following construction:

Using a compass, repeat the circles indicated for all points along the
"exploding" reflectors in (a) and (b) of Figure 2.47.

Note the formation of the new wavefronts.

The larger radius represents the wavefront at three time the initial increment

Note the formation of the buried focus cusp in b.

Extend the wavefront of the first image (a) to two and then three time increments.

Adding new circles centered on the newly constructed wavefronts.

Verify the similarity of the wavefront in (a) after three time propagations with
that obtained in (b).

In a constant velocity medium, the wavefronts could be propagated directly to


the surface for fast modelling.

Note:

The radius of the circles is proportional to the time increment and the interval
velocity.

The smaller time increments of the first image could have radii that vary with
velocity to follow a structure more accurately than the larger increment.

The buried focus cusp observed in this model and on the previous depth
sections is not the buried focus observed on the time section. The time section
buried focus is formed as the wavefronts cross the z = 0 surface.

Page 2.76

Chapter 2 Modelling with zero-offset

2.15.2 Construction of Huygens' model


t=3
t=2
t=1
r=1
x
z

a)
t=3

r=3

x
z

b)
Figure 2.47 Construction for Huygens' method of modelling for a small radius (a)
and a large radius (b).

Page 2.77

A Practical Understanding of Pre- and Poststack Migrations

2.15.3 Comparison construction using approximate raypaths.


The objective of this exercise is to evaluate wavefront propagation using crude
raypath locations with accurate traveltimes along these raypaths.
In Figure 2.48:
1. Draw with a ruler, or sketch free hand, normal rays that leave the reflecting
surface. The raypaths should be long enough to contain the three time
units.
These rays should leave the reflector at right angles, but in this exercise,
simply guess the angle and draw the rays rapidly, i.e. take only about 20
seconds to complete all the rays. The intent is to have an approximate
location for the raypath.
2. Set a compass to the length of the three time units, then scribe along each
raypath a small arc. Ensure that the arcs from the neighboring rays touch
or intersect.
3. How well do the small arcs recreate the wavefront?

Observations:
The wavefront constructed from approximate raypaths is surprisingly accurate.
The traveltime to the wavefront may be estimated from an approximate raypath.
We therefore conclude that reasonably accurate traveltimes can be computed
between any two points by using an approximate raypath.

Page 2.78

Chapter 2 Modelling with zero-offset

t=3

r=3

x
z

Figure 2.48 Construction for estimating the wavefront using approximate


raypaths.
For more information, see section 3.2.7 on Fermats principle

Page 2.79

A Practical Understanding of Pre- and Poststack Migrations

2.15.4 Huygens' method illustrating refraction


Huygens' method may also be used to illustrate the bending of rays or the
bending of wavefronts across a velocity boundary as illustrated below in Figure
2.49. The wavefronts are spaced at equal time increments. Note the change in
radius with change in velocity. A raypath normal to the wavefront is also shown.

V2
V1

Figure 2.49 Huygens' construction illustrating wavefront bending at a velocity


boundary. The wavefront is shown at equal time increments and propagates to a
lower velocity in the upper medium.
The triangles at the interface defined by the two raypath can be used to verify
Snell's law.
The angle if incidence, measured from the normal is the same as the angle of the
wavefront.
What is the velocity ratio in the above diagram?

Page 2.80

Chapter 2 Modelling with zero-offset

2.15.5 Huygens' modelling example


The image below is a computer-generated model using Huygens' method. All the
time increments are displayed as one time image to give a multiple exposure
effect. A wavefront may be seen crossing two velocity layers. Transmitted,
reflected and refracted energy may be observed.

Figure 2.50 Huygens' method of modelling of a wavefront crossing two velocity


boundaries.
Huygens method is usually used for illustration purposes. The normal method of
propagating a wave field from one time to an additional increment in time is the
wave-equation.

Page 2.81

A Practical Understanding of Pre- and Poststack Migrations

2.16 Finite difference wave propagation


This method is probably the most common way of computing the wave fields at
incrementing times. The examples in Figure 2.46 were computed using this
method.
The wavefields from previous two time increments are used to compute the wave
field at the next time increment. In the following figure, there are four panels that
represent the wave field at four times t = 1, 2, 3, and 4. The wavefield is defined at
the intersections of each grid line in these panels. The red point on the t = 1 wave
field and the five red points on the t = 2 wavefields are used to compute the new
sample on the t = 3 wavefield as represented by the black dot.
x

Time

Figure 2.51 Computing the wavefield at the next time increment.


The operator is moved to each grid point on the t = 3 panel to compute the
complete wavefield at t = 3. The wavefields at t = 2 and 3 are then used to
compute the wavefield at t = 4, and so on until the complete volume for all t is
computed. The three points in each orthogonal direction are approximations to
the second derivatives in the constant velocity scalar wave equation:

2P 2P
1 2P
+
= 2
.
v t2
x2 z 2

Page 2.82

Chapter 2 Modelling with zero-offset

The amplitudes at the red points are weighted and then summed to produce the
new value of the wavefield. In the next chapter we will show how to derive the
finite difference operator to the acoustic wave equation, but only for a constant
velocity. Adding more parameters to the model will increase the complexity of
the operator and increase the compute times.
One may think that using more neighboring points would improve the accuracy of
the finite difference computation. That is not a simple process, but there are a
few cases that we will consider.
1. The three horizontal point in the t = 2 panel represent the second derivative
of the wavefield in the x direction. The accuracy could be improved by
using five point in both the x and z directions as in Figure 2.52a and (b).
We could continue to increase the number of points to represent the
second derivatives, but there comes a point where it is better to use the
Fourier transform in the x and z direction to compute the second
derivatives.
2. I have mentioned the finite difference method of solving the wave
equation. There are other techniques such as the finite element method
that can be combined with the finite difference to produce more accurate
results. In this case, more points on the t = 2 panel can be used such as
nine points in a 3 by 3 grid as in Figure 2.52c, or possibly twenty five points
in a 5 by 5 grid. For more information on this method see Du 2007.
x

Time 2

a)

Time 2

Time 2

b)

c)

Figure 2.52 Variations in the operator for propagating the wave field. The
velocity (parameters) are defined at the central grid point, and can vary from point
to point, a) 3 point second derivative, b) five point second derivative, and c) a
finite element nine point solution.

Page 2.83

A Practical Understanding of Pre- and Poststack Migrations

2.17 Modelling 3D data


Three dimensional data is typically divided into

marine 3D

land 3D

Marine 3D
Marine 3D data is typically acquired by a boat towing a 2D streamer (many times)
across the area of interest. Each pass of the line may be considered a 2D
section. It is processed as a large collection of 2D data sets as illustrated in (a) of
Figure 2.53. Prestack processing is therefore quite similar to the processing of
regular 2D data. DMO may also be applied to the 2D lines. Migration would be
performed on the 3D volume of data.
Modern marine acquisition may drag a number of lines, and may have a number
of sources. These data sets are more complex and may require processing
similar to 3D land processing.
Marine 3Ds have problems with water-bottom multiples but are usually easier to
process than land 3Ds as they have negligible static problems and are acquired
at an elevation that varies only with the tide.

Land 3D
Land 3Ds are usually acquired with a number of receiver lines and a number of
source lines that are at right angles to the receiver lines as illustrated in (b) of
Figure 2.53. Many receivers may be laid out to cover a large area, but only a
patch around the source will be activated.
The receivers in a patch will have a large number of angles or azimuths to a
source that is at the center of the patch.

Page 2.84

Chapter 2 Modelling with zero-offset

a)

Figure 2.53 Surface descriptions of 3D acquisition for (a) marine and (b) land.

Page 2.85

A Practical Understanding of Pre- and Poststack Migrations

2.17.1 3D modelling
3D modelling may be accomplished by many methods, and a number of
commercial packages are available.
Some models may be very simple such as horizontal and dipping planes, while
others may be quite complex.
One method of modelling reasonably complex data is similar to the diffraction
method of section 2.2 but now uses a 3D form of the hyperbola called the
hyperboloid.
As in the 2D case, the amplitude information of each input point is dispersed on
the surface of the hyperboloid. The shape of the hyperboloid is determined by
the RMS velocity and time of the output sample. When modelling to a 2D line
from a 3D structure, only the hyperbola that intersects the output plain is
required, as illustrated in Figure 2.54.

Page 2.86

Chapter 2 Modelling with zero-offset

Figure 2.54 Modelling a 3D structure to a 2D line with hyperboloids. Only the


hyperbola is projected onto the output 2D line.

Page 2.87

A Practical Understanding of Pre- and Poststack Migrations

2.18 Physical modelling


Scaled models are used to represent real geological formations. These scaled
models fit into a reasonably size apparatus for simulated seismic experiments.
The scaling is on the order of 10,000 to 1. The source wavelet is approximately
500 kHz.
Physical models can be immersed in water to simulate marine acquisition and or
simplify surface coupling. The models can also be acquired directly on a
physical model.
The physical size of the modelling source or receiver is 1 cm or approximately
100m when scaled to real world dimensions. This limits how close the source
and receivers can be placed, have different effects on small changes in the
structure, and changes the relative size of the radiation pattern.
The physical models can be anisotropic, and will include all modes of wave
propagation.
A physical modelling system is shown below in Figure 2.55. It has two main
gantries for x positioning and these have secondary gantries for y positioning.
Vertical motion is controlled by the vertical gantries. Each gantry is driven by
linear digital motors. The main gantries can hold either the source or receiver, an
array of source and receivers, or both a source and receiver.
The physical models can be wheeled under the gantries as illustrated but the
marine model in Figure 2.56. A description of the model is shown in Figure 2.57,
and the resulting seismic section in Figure 2.58.

Figure 2.55 Physical modelling apparatus.

Page 2.88

Chapter 2 Modelling with zero-offset

Figure 2.56 A physical model in a tank of water for a simulated marine


acquisition. The source and receiver are on the same gantry to simulate a fixed
source receiver offset.

Figure 2.57 A schematic diagram of the physical model.

Figure 2.58 The simulated seismic response of the above physical model.

Page 2.89

A Practical Understanding of Pre- and Poststack Migrations

2.19 Comparison of a physical model with a numerical model


A wedge model has been included to illustrate the similarities and differences
between numerically modelled data and physically modelled data.
The model shown with field scale distances is shown in Figure 2.59. The physical
model zero offset section is shown in Figure 2.60, and the numerically model zero
offset section is shown in Figure 2.61.
The physically modelled section was acquired with a cross-line offset of 40 m,
and the scaling was 2,500;1.
The numerical section was modelled with an exploding reflector finite difference
algorithm that included multiples, but not converted wave energy.

Figure 2.59 Wedge model in field dimensions.


These examples are taken from:
Cooper, J..K., Lawton, D. C., and and Margrave, G. F., 2007, Multiples and multimode
events in a 2D marine zero-offset survey: a physical modelling example, CREWES
Research Report Volume 19 (2007)

Page 2.90

Chapter 2 Modelling with zero-offset

Figure 2.60 Physical modelled zero offset section.

Figure 2.61 Numerically modelled zero offset section assuming an acoustic


medium.

Page 2.91

A Practical Understanding of Pre- and Poststack Migrations

2.20 Points to note in Chapter 2

Poststack modelling assumes zero-offset input section.

= seismic dip (using the RMS velocity in the area of interest).

= geological dip ( really the only true dip ).

tan = sin .

A dipping linear event and its model converge at the surface.

Diffractions are hyperbolic in constant-velocity media.

Diffractions approximate hyperbolas in variable velocity media.

Diffraction asymptotes are at 45 degrees and intersect at the surface.

Stacked sections are composed of diffractions.

The 2D Fourier transform may be used to model and is perfect for constant
velocities.

A structure with any velocity distribution may be accurately modelled by the


"exploding reflector" method that creates a 3D volume (x, z, t). The input is (x,
z, t=0) and the output is (x, t, z=0).

How good is the exploding reflector model?

Good kinematic responses.

It assumes a one-way wave-equation solution where there are no reflections,


i.e. no boundary transmission losses.

For further information:


Ray tracing modelling, Gray [64]
2D Fourier transforms, Bolondi [39], Kosloff [65]

Page 2.92

Chapter 2 Modelling with zero-offset

2.21 Quiz 2.
1. Sketch the seismic section that results from the given structure.

x
t or z

2. The apparent dip on a seismic section will be steeper/shallower than the


structure dip.
3. Why did question 2 use the term "apparent dip"?
4. Where do the asymptotes of a diffraction converge?
5. Where will a dipping reflector and its reflection converge?
6. Given tan = sin ,

=
=

8. What is diffraction modelling?


9. Can you "image" process?
10. Given the following volume V(x, z, t), indicate: (a) V(x, z, t=0), (b) V(x, z=0, t)
t

x
z

Page 2.93

A Practical Understanding of Pre- and Poststack Migrations

This page is intentionally left blank.

Page 2.94

Chapter Three
Bits and Pieces

Objectives

Define and review the geometric shapes of the semicircle, ellipse, and
hyperbola from the perspective of modelling and migration.

Define and evaluate instantaneous, interval, average, stacking, RMS, DMO,


and migration velocities.

Gain insight to the Fourier transform and aliasing.

Become acquainted with the wave equation and be exposed to some methods
of applying it to seismic data.

Identify the differences between time and depth migrations.

A Practical Understanding of Pre- and Poststack Migrations

3.1 Circles, ellipses, and hyperbolas


Some simple observations of a circle

center of a circle represents the location of a source and receiver

reflections from any part of the circle have the same traveltime

circle is a special form of an ellipse in which the foci are co-located

the circle and hyperbola are directly related

a triangle whose base is a diameter and whose apex lies on the circumference
will have a right angle opposite the diameter.

x2 + z2 = r2

Figure 3.1 The circle and its geometry.

Page 3.2

(3.1)

Chapter 3 Bits and Pieces

3.1.1 Some observations of an ellipse

The foci are special points for an ellipse, and in seismic processing, they
represent the locations of the source and receiver.

A ray through one focus (source) will reflect off the ellipse and pass through
the other focus (receiver).

Length of the raypath from one focus to any location on the ellipse and on to
the other focus will always be constant.

Similarly, a raypath of fixed length that starts and ends at the foci will describe
an ellipse.

An ellipse will tend toward the shape of a circle when the distance between the
foci is much less than the raypath.

x2 z 2
+ 2 =1
2
a b

Figure 3.2 The ellipse and its geometry.

Page 3.3

(3.2)

A Practical Understanding of Pre- and Poststack Migrations

3.1.2 Some observations of an hyperbola

A diffraction pattern has the shape of a hyperbola as defined below.

The depth of the apex is Zo, and the half offset is h.

Slope of the hyperbola (diffraction) tends to 45o asymptotes when h >> Zo,

Asymptotes intersect at the surface and for all diffractions on the zero offset
section.
A hyperbola also has two foci and may be defined as having the
same difference in length from each of the foci.

z 2 x 2 =zo2

(3.3)

z 2 = zo2 + x 2

(3.4)

For two way times,

4x 2
T =T + 2
v
2

2
o

Figure 3.3 The hyperbola and associated definitions.

Page 3.4

(3.5)

Chapter 3 Bits and Pieces

3.1.3 Relationships between the circle and hyperbola

A diffraction hyperbola with its apex on a semicircle will also pass through the
bottom of the circle.

The apex of a diffraction hyperbola and any other point on the diffraction
hyperbola will lie on a semicircle with a center at time-zero above the point.

a)

b)
Figure 3.4 Relationship of the diffraction hyperbola and the semicircle (a)
showing the angles and (b) the intersections.

Page 3.5

A Practical Understanding of Pre- and Poststack Migrations

3.1.4 Relationships between the ellipse and hyperbola

The ellipse defines a constant total distance from the two foci, i.e.
d1 + d2 = d3 + d4.

The hyperbola defines the difference in distance between the two foci, i.e.
s2 - s1 = s4 - s3.

Figure 3.5 The relationship between the ellipse and hyperbola.

Figure 3.6 Two orthogonal antenna systems used for navigation. The difference
in time from each antenna system will locate four positions on corresponding
hyperbola.

Page 3.6

Chapter 3 Bits and Pieces

The two following figures illustrate non-geophysical applications of the ellipsoid.

Figure 3.7 The Mormon Tabernacle in Salt Lake City has an ellipsoid shape with
the pulpit at one focus. People sitting near the other focus can hear a whisper at
200 ft.

Figure 3.8 Lithotripsy: A method for breaking gallstones uses an ellipsoid cavity
with a fluid that matches the velocity of tissue. A spark at one focus will reflect
energy to the focus within the patient and break the gallstones [97].

Page 3.7

A Practical Understanding of Pre- and Poststack Migrations

3.2 Velocities (instantaneous, average, RMS, stacking,


interval, migration, and DMO)
It will be assumed that a geological formation has one acoustic velocity that is
defined by the instantaneous velocity and that a well log is a good source for
interval velocities.
At this time we assume P-waves in an isotropic medium.
Other types of velocity are defined for convenience and are described as follows:

Average velocities are used to convert directly from time to the corresponding
points in depth (or visa versa).

RMS velocities are defined to simplify the computation of raypath traveltimes


or moveout correction. .

Stacking velocities are used to create the best-looking stack. For isotropic
horizontal layers with small offsets, they should be equal to the RMS
velocities, VNMO = VRMS.. Dipping areas of the section Vstk require the stacking
velocities Vstk to be higher than the RMS velocities [115]. When the effects of
dip have been removed, moveout velocities are often considered equivalent to
RMS velocities (for all offsets), allowing the estimation of interval velocities.

Interval velocities are assumed constant over a small range. They may be
derived from well-log velocities using either an average or RMS computation,
depending on the application. Interval velocities may also be estimated from
RMS velocities (by way of stacking velocities) using Dix equation [442] [443].

Migration velocities produce the best-looking migrated section and should be


either RMS or interval velocities and not the stacking velocities. Other factors
that effect 2-D migration velocities are sideswipe and oblique reflectors [4].

When applying NMO prior to the DMO process, the NMO should use RMS
velocities and not the stacking velocities.

Transverse anisotropy may occur in horizontally layered media, where, at a given


location, the velocity may vary from 10% to 20% with dip.
Conventional processing ignores the effects of transverse anisotropy. Its effect
on traveltimes may be partially absorbed into stacking and migration velocities.
Azimuthal anisotropy may be caused by depositional environment. Its effect in
3-D processing may require gathers to be formed using an azimuthal parameter.
Shallow azimuthal anisotropy may also effect weathering analysis.

Page 3.8

Chapter 3 Bits and Pieces

Typical seismic processing defines a stacking velocity to give a best fit for NMO.
Occasionally, these velocities are assumed to be RMS with interval velocities
computed using Dixs equation [442] (see Figure 3.12). This method is subject to
serious error in structured data where the stacking velocities are increased by the
cosine of the dip to produce the best stack.
The inclusion of DMO and migration in the velocity analysis loop removes the
effects of dip on stacking velocities. These stacking velocities will tend toward
RMS velocities from which the interval velocities may be estimated.

Velocities are picked at discrete points. It is assumed RMS velocities are


linearly interpolated between picks, and that interval velocities are constant
between picks. Linear interpolation is considered accurate enough for standard
processing even though it may not be strictly accurate relative to the constant
interval velocity assumption.
RMS velocities, Vrms are defined from interval velocities for the purpose of
simplifying the computation of raypath traveltimes or moveout correction. They
enable the curved raypaths on an isotropic horizontally layered medium to be
approximated with linear rays on a time section [56]. These linear rays form
triangles, allowing the hyperbolic assumption of traveltimes. This definition is
strictly for small offsets as the real traveltimes are more complex and may require
a polynomial approximation for larger offsets. A best hyperbolic fit to the
traveltimes in an isotropic horizontally layered medium with any offset will use a
velocity defined as the normal moveout velocity VNMO , often equated to Vrms.
In the presence of anisotropy, it is the difference between well-log estimates of
Vrms, and VNMO that leads to estimates of the anisotropy.
The diffraction shape may be quite complex, even in horizontally layered
structures. Hyperbolic moveout that achieves a best fit using a stacking velocity
may have a zero-offset time that varies from the actual zero-offset time. See
Yilmaz [83] pages 159 to 168 for additional information.
Questions:

What type of velocity does the speedometer in a car read?

What type of velocity may be calculated from the odometer reading and
elapsed time?

Page 3.9

A Practical Understanding of Pre- and Poststack Migrations

3.2.1 Instantaneous velocity (Vins)


The true structure is defined by the instantaneous velocity. The instantaneous
velocity should be similar to the well-log velocity and represents the actual
velocity of the subsurface at a specific location. It is defined as a small increment
of distance divided by the time required for the acoustic wave to travel across
that distance.

Vins =

Small dist. d z
=
Time inc.
dt

Figure 3.9 Instantaneous velocity.

Remember that we are assuming isotropic media.

Page 3.10

(3.6)

Chapter 3 Bits and Pieces

3.2.2 Average velocity (Vave)


The average velocity relates a total distance with a total traveltime, and is not
concerned with what happens within the measurements. An average speed of
travel does not require knowledge of refueling stops or the maximum speed
traveled, but only with the total distance and total time. In geophysics, the
average velocity is used to relate a specific depth with a specific time in
time/depth conversions.
n= N
t =T

specific dist. 1
=
Vave (T ) =
Vins ( t ) dt
time required T t =0

V (n) t
n =0

int

n= N

t
n =0

(3.7)
n

Vave

Figure 3.10 Average velocity of the defined instantaneous velocity.

Average velocities are only valid for vertical velocity variations, and it should not
be used for structures with significant lateral velocity variations. When lateral
velocities are present, image-ray conversions should be used. See work by
Hubral [3].

Page 3.11

A Practical Understanding of Pre- and Poststack Migrations

3.2.3 RMS velocity ( Vrms)


The RMS velocity is an apparent velocity that takes into account Snell's law of
refraction and enables a simple calculation for normal moveout and diffractions.
RMS is derived from "root mean squared" and is defined in [56], [83] and [442] as
n= N

2
rms

(T )

1
T

t =T

V ( t ) dt
2
ins

t =0

V (n) t
n =0

2
int

n=N

n =0

(3.8)

tn

Vint2 ( 0 ) t0 + Vint2 ( 1) t1 + Vint2 ( 2 ) t2 + ... + Vint2 ( n ) tn


Vrms ( n ) =

t
+
t
+
t
+
...
+
t
0
1
2
n

(3.9)

Vave

Vrms
Figure 3.11 The velocity diagram showing the RMS velocity.
RMS velocities defined above are accurate for small offsets when computing
NMO correction. Larger offsets may require more terms in the MO equation (as
given in section 3.2.1 in [83]).
The term RMS velocities will also be used (in an approximate sense) to refer to:
best fit velocities for horizontal reflectors, Vrms Vstk, (isotropic media)
velocities extracted from stacking velocities in which the dip-dependent factor
has been removed, i.e. Vrms Vstk cos(dip), or
NMO velocities used prior to the application of dip moveout (DMO).

Page 3.12

Chapter 3 Bits and Pieces

3.2.4 Stacking velocities (Vstk)


Stacking velocities Vstk are chosen to obtain the "best looking" stack from the
NMO equation (3.10), where T and T0 are two-way times and h is half the sourcereceiver offset. The double quotes emphasize that velocity picking is subjective
and that stacking velocities are not necessarily equal to theoretical RMS
velocities.
2
nmo

4h 2
=T + 2
.
Vstk (To )
2
o

(3.10)

The use of the word "robust" is very applicable for stacking velocities. These
velocities cover many approximations that are made in seismic processing and
have enabled the processing of complex data structures well beyond the original
assumptions of NMO and stacking. Anisotropy is one approximation that is
partially absorbed into that stacking velocity.
Another robust feature of stacking velocities is its ability to improve the
appearance of dipping data by increasing the stacking velocities with

Vstk =

Vrms
.
cos ( )

(3.11)

Equation (3.11) enables NMO to stack energy of a dipping event at a CMP


location; however, this process does smear data along the dipping event. This
equation is derived in Volume II, Section 7.0.9 and is attributed to Levin [115].

Since the stacking velocities incorporate dipping information, care must be taken
when using them for other processes such as interval velocity analysis,
migration, or inversion. If used for velocity analysis, the results may be
unreliable and absurd to the point of negative interval velocities (Figure 3.12).

Remember that the stacking velocities are defined for a best-fit hyperbola, and
that T0 of the hyperbola may not be exactly the same as the actual zero-offset
time. See Yilmaz [83] pages 159 - 166 for more detailed information.

Page 3.13

A Practical Understanding of Pre- and Poststack Migrations

3.2.5 Interval velocities (Vint)


Interval velocities combine the instantaneous velocities over a defined interval.
The actual definition of interval velocities depends on the application. If the
application involves time to depth estimations, then an average velocity over the
interval T1,2 from T1 to T2 is used, i.e.

Vint (T1,2 ) =

specific dist.
1
=
time required T1,2

T2

T1

Vins ( t ) dt .

(3.12)

When the use of interval velocities are associated with NMO or migration
applications, the RMS definition is used, i.e.,

Vint2 (T )

1
T1,2

T2

T1

Vins2 ( t ) dt .

(3.13)

Interval velocities may be estimated from seismic processing using Dixs


equation and the RMS velocities that are derived from stacking velocities. Dixs
equation for interval n may be derived from equation (3.9) and is
2
2
Vrms
n ) tn Vrms
(
( n - 1) tn-1
Vint ( n ) =

tn tn1

(3.14)

The interval velocities derived in this manner tend to be continuous along events
and aid in the velocity analysis process.

When the interval velocity is estimated from stacking velocities of dipping data,
inaccurate results occur. The stacking velocity for dipping data is increased by
equation (3.11), and repeated as (3.15), i.e.,

Vstk =

Vrms
.
cos(dip)

(3.15)

Using the stacking velocity of equation (3.15) as RMS velocities in equation (3.14)
will produce interval velocities that may vary in an extreme manner, i.e., they may
be very high, or very low, and even negative as illustrated in Figure 3.12.

Page 3.14

Chapter 3 Bits and Pieces

a)

b)

c)

d)

Figure 3.12 Illustration of errors in interval velocity when estimated from stacking
velocities in area of dipping geology. Part (a) shows the instantaneous and RMS
velocities of the geological structure in (b). The resulting stacking velocity is
shown in (c), with the incorrect interval velocities in (d) estimated from (c).

Always convert stacking velocities to RMS velocities by removing


dipping effects before estimating interval velocities.

We typically use Vrms for Vstk when the structure is horizontal. This
assumes no anisotropy present. When anisotropy is present, Vrms
obtained from well information is not equivalent to Vstk.

Processing that includes DMO has significantly improved the science of velocity
analysis and by passes the "dip" problem. Velocity analysis that includes DMO
derives RMS-type velocities and has helped automate the process.

Page 3.15

A Practical Understanding of Pre- and Poststack Migrations

3.2.6 Assumption of linear raypaths


In a constant velocity medium, the travel times for NMO or a diffraction are
hyperbolic. When the velocity increase with depth, as illustrated in Figure 3.13a,
the raypath curves, adding complexity to the problem.

The raypaths in Figure 3.13a are curved because velocity increases with
depth.

The traveltimes of the NMO or diffraction curve in Figure 3.1a is described by a


N

t = an x 2 n = a0 + a1 x 2 + a2 x 4 + a3 x 6 + ... .
2

polynomial

n =0

The NMO or diffraction curve is usually approximated by the first two terms of
the polynomial giving a hyperbola t = t0 + a1 x , i.e. equation (3.10).
2

The hyperbolic assumption is possible when using the RMS velocities defined
in equations 3.8 and 3.9. The hyperbolic and RMS velocity approximation
assumes horizontally layered media and is accurate for small offsets.

The hyperbolic approximation is shown in (b), and from Pythagorean theorem,


the raypaths on the time section are linear.

The time section is often superimposed on top of the geological structure with
the vertical time T0 scaled to coincide with a reflection at depth z as illustrated in
(c).

Estimating the true depth z from T0 must use the average velocity Vave.

The use of the RMS velocities with T0 will only estimate a pseudo depth z0.

The assumed linear raypaths will have errors in the angle of incidence and
reflection.

The hyperbolic approximation of traveltimes is one of the most significant


developments in exploration seismology.
The travel time from any location on the surface to a scatterpoint, or
from any scatterpoint to any point on the surface,
can be approximated by using the RMS velocity and the geometry.
At this point we still assume horizontally layered media.
Dipping events are approximated using stacking velocities.
Even in structured media, the traveltimes of specula reflection energy tends to be
hyperbolic.
NMO correction and Kirchhoff time migration are based on this assumption.

Page 3.16

Chapter 3 Bits and Pieces

t
Diffraction

Hyperbola

T0

T0, z0

T
T0

T
x
z

a)

b)

T
T0
z0

True depth z
T
c)

Figure 3.13 Raypaths and traveltimes in a) are for the actual the depth structure
and b) the assumed linear raypaths on the time section. The actual diffraction in
(a) is approximated by the hyperbola in (b). Part c) illustrates overlapping time
and depth sections that confuse the pseudo depth with the actual depth.
Assume that the raypaths in Figure 3.13b are zero-offset to the scatterpoint.
Define the equation for the two-way traveltime T in terms of T0, x, and Vrms.

Assume the raypaths in Figure 3.1b are to the same midpoint. Define the two-way
traveltime T in terms of T0, h, and Vrms.

Page 3.17

A Practical Understanding of Pre- and Poststack Migrations

3.2.7 The power of the RMS assumption


In a review, I emphasize that in a multi-layered media, where we would normally
have to do raytracing to get the travel times; we dont. We simply use the RMS
velocity Vrms defined at the time T0 of the scatterpoint, then assume that the rays
are linear on the time section, and compute the traveltimes with a hyperbolic
equation of the form:

a1 x 2
t =T + 2
Vrms
2

2
0

(3.16)

Figure 3.14 illustrates the general concept in (a), a CMP gather in (b), post-stack
Kirchhoff migration in (c), and the double square-root equation in (d).
For a CMP gather, we compute the normal moveout correction for horizontal
reflectors by using the hyperbolic equation where h is the half offset:

4h 2
t =T + 2
Vrms
2

2
0

(3.17)

Dipping reflectors require the RMS velocity to be replaced by stacking velocities.


For a post-stack Kirchhoff migration, we compute the two-way traveltime with the
hyperbolic equation where x is the displacement from the scatterpoint to the
source-receiver location:

4x2
t =T + 2
Vrms
2

2
0

(3.18)

Prestack Kirchhoff time migration computes the individual one-way traveltimes of


the source and receiver ray paths to give the double square-root equation:

T02 hs2
T02 hr2
t = t s + tr =
+ 2 +
+ 2
4 Vrms
4 Vrms

(3.19)

where hs and hr are the offsets from the location of the migrated trace to the
location of the source and receiver.

Page 3.18

Chapter 3 Bits and Pieces

T0 Scatterpoint

a)
h

CMP

CMP
gather

T0

Scatterpoint

b)
x

Post-stack
migration

T0 Scatterpoint

c)
x

hr
hs

Pre-stack
migration

T0

Scatterpoint

d)
Figure 3.14 RMS velocity assumption for a) any rays to or from a scatterpoint, b)
a CMP gather, c) zero-offset migration, and d) prestack migration.

Page 3.19

A Practical Understanding of Pre- and Poststack Migrations

3.2.8 Fermats principle


What is the shortest path from P to S?

Figure 3.15 Direct and refracted raypaths.


Fermat's principle:
The seismic raypath between two points is that for which the first-order variation
of traveltime with respect to all neighboring paths is zero (Sheriff [543],[550]).
In simple terms a ray will usually follow the path of shortest time.
(A maximum path is optimum for zero offset rays to a buried focus.)

Even though the raypath PQRS is longer in distance than PS, the traveltime for
PQRS will be shorter as the raypath spends more time in the higher velocity
portion than in the lower velocity portion.

Page 3.20

Chapter 3 Bits and Pieces

Consider Figure 3.16 below.

What path would a lifeguard take to reach the person crying for help?

What path should a lifeguard take to reach the person crying for help?

What are the two main parameters that determine the path?

Figure 3.16 Lifeguard path.

We encounter the application of Fermats principle many time in nature, and our
intuition usually is correct.
Consider:

Hiking across fields that may be hard and easy to walk on and ones that
may have been plowed, and difficult to walk on.

Driving across a city:


o Do you take the shortest and most direct path, or a free-way that may
have a longer distance to travel.
o Do you want to optimize with respect to time, distance, fuel economy,
or some other parameter such as alternate shopping locations?
o Do you include probabilities in you plans such as the times of possible
traffic congestion.

Page 3.21

A Practical Understanding of Pre- and Poststack Migrations

An additional concept of Fermats principle deals with the minimum (or


maximum) traveltime. Consider Figure 3.17.
The thicker ray path in (a) represents the actual raypath with alternate paths
shown on either side. The traveltimes (T) for the various paths are shown by the
curve in (b). From (b) we can observe that the minimum traveltime along paths
between A and B is the actual raypath, verifying Fermats principle.

At the minimum, the slope of the curve is horizontal.

A small error in the refraction point location has negligible effect on the
traveltime. (See construction in Figure 2.28).

Actual
raypath

a)

b)

Figure 3.17 Minimum traveltime illustration with a) showing possible raypaths


and b) the corresponding of traveltime verses the refraction point location.

Fermats principle also tells us that small errors in the raypath location will have
negligible effect on the estimation of the traveltime.

Traveltime computations for depth migration may be simplified by using an


approximate raypath location to save on computational overhead.
Some
contractors have even used straight raypaths when estimating traveltimes for
depth migrations.

Page 3.22

Chapter 3 Bits and Pieces

3.2.9 Exercise for velocity estimation.


The structure is defined as shown below by its instantaneous velocity.

Figure 3.18 Structure for velocity calculations.


Compute the traveltimes TB, TC and TD at the depths B, C, and D (one-way time).
Instantaneous times TAB, TBC, and TCD are,

TAB =

dist 1000
=
= 0.5 sec
vel
2000

TBC = ... =

1000
= 0.25 sec
4000

TCD = ... =

1000
= 0.167 sec
6000

TB = 0.5 sec ,

TC = 0.75 sec

TD = 0.917 sec

Page 3.23

A Practical Understanding of Pre- and Poststack Migrations

Compute the average velocities at B, C, D.


Vave ( B ) =

dist
1000
=
= 2000 ft/sec
time
0.5

Vave (C ) =

...

2000
= 2667 ft/sec
0.75

Vave ( D ) =

...

3000
= 3272 ft/sec
0.917

Vave

Figure 3.19 Plot of Average velocity.

Page 3.24

Chapter 3 Bits and Pieces

Compute the RMS velocities at B, C, and, D using equation (3.8).


1

20002 0.5 2
=2000 ft/s
Vrms (B )=

0.5

20002 0.5 + 40002 0.25 2


=2828 ft/s
Vrms (C )=

0.5
+
0.25

2000 2 0.5 + 4000 2 0.25 + 6000 2 0.167 2


=3614 f/s
Vrms (D )=

0.5
+
0.25
+
0.167

Vrms

Figure 3.20 Plot of RMS velocity on simple structure.

Page 3.25

A Practical Understanding of Pre- and Poststack Migrations

3.2.10 Construction for time response of rays


The following constructions plot the traveltimes of rays that first ignore, and then
apply, Snell's law. The first construction uses shortest distance raypaths and
ignores Snell's law: the second includes refraction for a relative comparison.
Traveltimes using the RMS velocity and offset are computed to be plotted for
comparison with the construction.

3.2.10.1 Construction: Time response of straight rays (incorrect)


In Figure 3.21:

Equate a velocity of 2000 with unity.

In the first layer, use a time measure equal to the distance.

In the second layer, use a time measure that is one half the distance traveled
(two times the velocity or half the slowness).

In the third layer, use a time measure equal to one-third the distance traveled
(three times the velocity or one-third the slowness).

The horizontal dashed line in the figure represents one half and one third the
distance across the respective instantaneous velocities and should be used to
construct the traveltimes.

Construct the traveltimes by:

Measuring the incremental traveltimes of each ray in the different velocity


bands.

Plot the accumulated times for each ray on the time graph (b) below the
surface location of the ray.

The traveltimes on (b) define the straight raypath diffraction shape that is to be
compared to the true shape.

Page 3.26

Chapter 3 Bits and Pieces

a)

b)
Figure 3.21 Construction of incorrect diffraction shape found by assuming a
direct raypath; (a) shows the ray paths and (b) area for plotting travel times.

Page 3.27

A Practical Understanding of Pre- and Poststack Migrations

3.2.10.2 Construction: Time response of refracted rays (correct)


Use the same time to distance principles from the previous construction

Use a time measure equal to distance in the top layer.

Plot the incremental and total times in Figure 3.22b.

In the second layer use a time measure of one half the distance.

In the third layer use a time measure of one-third the distance.

Plot the five raypath traveltimes.

Add the traveltimes from Figure 3.21 to Figure 3.22.

Add to the plot, the computed values using RMS velocities that are found on
the following page.

Note how well the RMS velocities allow the NMO computation to match the actual
traveltimes. (They should if constructed accurately).

One interesting feature that may be observed from this construction is that
minimum distance raypaths also appear hyperbolic. An adjusted stacking
velocity would allow reasonable NMO or migration of the data. This concept is
used in some modelling based processes to improve run times.

Page 3.28

Chapter 3 Bits and Pieces

a)

b)
Figure 3.22 Construction of diffraction shape with refracted raypath.

Page 3.29

A Practical Understanding of Pre- and Poststack Migrations

3.2.10.3 Computation of traveltimes using RMS velocities


Using the zero offset traveltime T(D) from page 3.23 (which should be the same
time on both constructions), the RMS velocity at D from page 3.25, and the
various offsets in Figure 3.22, compute the traveltimes using the one-way NMO
equation,
2
h

T(D) = 0.9167,

= T

2
0

1) h 2
(
+ 2
Vrms (T0 )

(3.20)

Vrms(D) = 3614

T ( h = 1000 )

10002
2
= 0.9167 +

36142

1
2

T ( h = 2000 )

20002
2
= 0.9167 +

36142

1
2

T ( h = 3000 )

30002
2
= 0.9167 +

36142

1
2

T ( h = 4000 )

40002
2
= 0.9167 +

36142

1
2

= 0.958

= 1.071

= 1.237

= 1.437

Add these times to the plot of Figure 3.22 and compare their locations with the
times found from the construction.

Page 3.30

Chapter 3 Bits and Pieces

3.3 The Fourier transform


3.3.1 Introduction
The Fourier Transform (FT) deals with continuous time and frequencies.
often use this term when we are really using the discrete Fourier transform.

We

The Fourier Series (FS) deals with periodic and continuous time, but discrete
frequencies.
The Discrete Fourier Series (DFS) deals with discrete time and frequencies such
as that stored in a computer. This is what we really use.
The length of one domain defines the sample interval in the other. As the length
approaches infinity, the other domain has a very small sampling interval and may
be considered to be continuous.
When the length of a domain is shorter, then wrap-around or aliasing may
occur and the DFS is a poor match to the FT.
The Fourier transform is a mathematical process that converts a signal such as
time recording, into another form such as a frequency spectrum. Some
advantages of the transformed data are:

the complex process of convolution in the time domain but becomes a trivial
process of multiplication in the frequency domain,

derivatives and integrals also become multiplication in the frequency domain,

in the 2-D space events with the same dip transform to one location.

Some interesting points of the transform are:

the Fourier transform samples represent the magnitude and phase of a range
of sinusoidal signals,

the summation of these sinusoids reproduce the input signal,

a temporal, or time series of samples, (e.g., a trace) is transformed into a


sequence of frequencies which are designated by F or ,

a spatial, or distance series of samples is transformed into wave numbers that


are designated by K. The wave number may also be referred to as frequency.

Page 3.31

A Practical Understanding of Pre- and Poststack Migrations

3.3.2 Examples of summing sinusoids to create signals.


The following figures illustrates the summation of sinusoids to produce various
periodic waveforms, note:

The relative amplitude and phase of the sinusoids,

The periodic sinusoids produce periodic waveforms,

Adding more sinusoids with higher frequencies will improve the approximated
waveform.
Waveform

a)

b)

Figure 3.23 Summing a) two and b) three sinusoids to approximate a triangular


periodic waveform.

a)

b)

c)

d)

Figure 3.24 Summing five sinusoids to produce periodic waveforms; (a) a


triangular, (b) a square, (c) a rectified sine wave, and (d) a ramp.

Page 3.32

Chapter 3 Bits and Pieces

The magnitude and phase of the sinusoids can be varied to have special effects.
In the figures below, time zero is indicated at the center of each image, with the
left image containing only cosines that sum to approximate a sinx/x wavelet,
while the right side is composed of sines to produce a wavelet that has been
phase shifted by 90o.

Frequency

Frequency

Sum

T=0

a)

b)

Figure 3.25 Summing a) cosines to create a zero phase wavelet t = 0.0, and b)
sines to create a 90 degree wavelet. Zero time is at the center. (FOUTIERexamples.m )

Frequency

Phase
shift

Frequency

The phase may also be modified to change the location of the wavelet. Figure
3.26 shows two images with a positive and negative phase shifts that vary linearly
with frequency. The phase-shift is measured at t=0 on each plot.

a)

b)

Figure 3.26 Two linear phase shift applied to the cosines to displace the wavelet.

Page 3.33

A Practical Understanding of Pre- and Poststack Migrations

More frequencies produce a narrower wavelet as illustrated in Figure 3.27. (Time


in now plotted vertically and frequency horizontal.)
Note in all these figures that the lowest (non zero) frequency is always one
complete cycle. This fundamental frequency is also the interval between each
frequency. Therefore, the frequency interval f is related to the duration T of the
waveforms or the total time of the trace giving the relationship,

f =

1
.
T

(3.21)

At this point we could consider the time traces to be continuous, but in reality
there are closely separated points with a sample interval t or a sample rate of
Fsamp. The maximum frequency Fmax that can be recorded must be less than the
Nyquist frequency, Fnyq that is half the sampling frequency Fsamp i.e.,

Fmax Fnyq

Fsamp
2

1
.
2t

(3.22)

We can write the summation of these sinusoids f(t) over the time interval T as
2 t
4 t
6 t
f ( t ) = a0 + a1 cos
+ a2 cos
+ a3 cos
+ ...
T
T
T
.
2 t
4 t
6 t
+ b1 sin
+ b2 sin
+ b3 sin
+ ...
T
T
T

or

(3.23)

n= N

n2 t
n2 t
f ( t ) = a0 + an cos
+ bn sin

T
T
n =1

where the Fourier Series coefficients an and bn are computed from (assuming
continuous time)

1
an =
T

n 2 t
f ( t ) cos
dt
T

1
n 2 t .
f
t
sin
()
dt
T 0
T

bn =

(3.24)

The coefficients an and bn may be expressed as magnitude and phase where


b
cn = an + bn and . n = tan 1 n giving,
an
n= N

n2 t

f ( t ) = c0 + cn cos
+ n .
T

n =1

Page 3.34

(3.25)

Chapter 3 Bits and Pieces

Frequency

Time

Frequency

a)
Frequency

Time

Frequency

b)

c)

d)

Time

Frequency

e)
Figure 3.27 The size of the frequency band contributes to the shape of the
wavelet. More frequencies produce a narrower wavelet that will allow higher
resolution between similar wavelets. The number of sinusoids are 4, 9, 14, 18,
and 50.

Page 3.35

A Practical Understanding of Pre- and Poststack Migrations

The previous used sinusoids with the same amplitude. In general the amplitudes
vary as illustrated in the following figures. The amplitudes (cn) are plotted above
the corresponding sinusoids.
Figure 3.28a shows the sinusoids with an amplitude weighting of sinx/x (really
sinf/f) that produces a box-car weighting of a sinusoid. Figure 3.28b weights the
sinusoids with a triangular shape that produces a time trace that has a (sinx/x)2
weighting of a sinusoid.

Time

Frequency

a)

Time

Frequency

b)
Figure 3.28 The amplitudes of the sinusoids were multiplied in a) by a sinx/x
function that was centered over the central trace, and b) with a limited triangular
shape window.

Page 3.36

Chapter 3 Bits and Pieces

The amplitudes of the sinusoids, shown in the red box in Figure 3.29, is called the
spectrum. The magnitude and phase of the spectrum contain the complete
information that is needed to create the summed trace indicated in green.

Time

We do not need to show the sinusoids as their information is entirely contained


within the spectrum. Fourier processing enables us to use the discrete Fourier
transform (DFT) to convert directly from one domain (say time) to the other
domain (frequency) . The process can be reversed by using the inverse DFT
(IDFT) to convert the spectrum back to the time trace.

Trace
Frequency

a)
Spectrum

I-DFT
DFT

Trace

b)
Figure 3.29 Illustration of a) the spectrum and trace in sinusoidal format and in b)
the conventional form without the sinusoidal traces.
The an and bn coefficients may also be expressed in a complex form allowing the
sines and cosines to be expressed as an exponential. The exponential form is
very convenient when working with differential equations as the derivative of an
exponential is simply itself.

Page 3.37

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

3.3.3 Examples
E
s of 1-D Fourier tra
ansforms
s pairs

Figurre 3.30 Exa


amples of Fourier tra
ansform pa
airs wavefo
orms from
m Lathi [560
0].

P
Page
3.38

Chapter 3 Bits and Pieces

3.3.4 Fast Fourier transform


The FFT or Fast Fourier Transform is a fast algorithm for computing the Discrete
Fourier Series. It requires the number of input samples to be a power of 2, such
as 256, 512, 1024, etc. In geophysics, we often have a trace that is say, two
seconds long, and sampled at two milliseconds, giving 1000 samples. It is
therefore necessary to add 24 points to this trace to make 1024 samples before it
can be Fourier Transformed with the FFT.
An input trace with 1024 real samples will produce 512 complex samples in the
transformed space. The complex samples may be expressed in a real and
imaginary format, or as a magnitude and phase.

Sometime the input trace is complex, both real and imaginary, and in this case
1024 complex input samples produces 1024 complex samples in transformed
space. As each sample represents a "variable," the number of variables before
the transform must equal the number of variables after the transform.

Page 3.39

A Practical Understanding of Pre- and Poststack Migrations

3.3.5 Fourier transform of seismic data


The discrete Fourier transform requires the input signal (trace) to be periodic. We
therefore assume the input trace itself to be periodic, with a period that equals
the length of the input trace that includes the padding to make it FFT compatible.
Figure 3.31 repeats a figure introducing the Fourier transform and a periodic
seismic trace to emphasize the periodic.
Waveform

t = 0 ???

Tmax

Seismic trace

Figure 3.31 Seismic trace is assumed to be periodic when using the DFT.
The fundamental frequency of the seismic trace f0 above will have a period that is
equal to the inverse of the maximum time Tmax, i.e. f0 = Tmax. This frequency is
also the frequency interval f on the discrete Fourier transformed trace, i.e. f = f0.

Page 3.40

Chapter 3 Bits and Pieces

Figure 3.32 below shows in (a) a trace and (b) the corresponding amplitude
spectrum.
Amp
t

a)
Amp

Frequency

b)
Figure 3.32 A seismic trace showing in (a) the time domain and in (b) the
amplitude spectrum in the frequency domain.

Page 3.41

A Practical Understanding of Pre- and Poststack Migrations

3.3.4 2-D Fourier transform


A real time domain trace is transform using a 1-D Fourier transform into a two
sided Fourier transform. Each side (positive and negative frequencies) are
conjugately related. Consequently we usually only consider one side of the
spectrum.
A similar relationship occurs in the 2-D Fourier transform (FT). Data in the 2-D FT
should normally be considered with positive and negative wave-numbers and
frequencies as illustrated in Figure 3.33a. Because the input 2-D data (x, t) is real,
there is diagonal conjugate symmetry in the transform space (kx, f).
Because of this symmetry, only two quadrant are displayed. The origin of the (kx,
f) space can be located in a number of locations, as illustrated in (b), (c), and (d).
Displays with the origin in the center of the upper or lower space are typical in the
industry. The boundaries of the (kx, f) space are the Nyquist wave number and
frequencies, and energy should not cross that boundary.
Great care is taken to ensure that a 1-D trace is not aliased. However, when a
number of traces are combined into 2-D space, 2-D aliasing can and quite often
does occur.
Consequently, I like to locate the origin located in the upper right of (d). In this
display, the spatial Nyquist wavenumber is in the vertical center and 2-D aliasing
is easier to comprehend. Note that the vertical scale is considered to be positive
frequencies. (Data in the fourth quadrant is often considered to have positive kx
and f values.)
2-D FFTs may swap the energy in the quadrants because of the different
methods of defining the Fourier transform; i.e. the sign on the exponential.
Consider testing a dip in (x, t) space and identifying which quadrant it is
transformed to.

Page 3.42

Chapter 3 Bits and Pieces

fnyquist

2nd quadrant

Ist quadrant

Congugate
symetry

knyquist

-knyquist
Origin

rd

th

3 quadrant

4 quadrant
-fnyquist

a)

-knyquist

fnyquist
2nd quadrant

knyquist

Origin

Ist quadrant

3rd quadrant
-knyquist

4th quadrant
fnyquist

knyquist
Origin

b)

c)
-knyquist

knyquist

Origin

rd

th

3 quadrant

4 quadrant

fnyquist

d)
Figure 3.33 Various views of the (kx, f) space and the orogin.

Page 3.43

A Practical Understanding of Pre- and Poststack Migrations

3.3.5 Exercise comparing data in (x, t) and (Kx, ) transform domains.

a)

b)

c)
Figure 3.34 2-D Fourier transform pairs in (a) to (f).

Page 3.44

Chapter 3 Bits and Pieces

d)

e)

f)

Page 3.45

A Practical Understanding of Pre- and Poststack Migrations

3.4 Aliasing in one and two dimension.


3.4.1 Sampling in one dimension
Numbers stored in a computer may represent many different things and must be
interpreted by the user for what is intended. Consider the following figure:

a)

b)

c)

d)
Figure 3.35 Interpretation and interpolation of samples in (a) depends on the
application, such as (b) a staircase, (c) a picket fence, or (d) a continuous multivalued signal.

Page 3.46

Chapter 3 Bits and Pieces

The samples of Figure 3.35a may be interpreted to represent a staircase, the tops
of a picket fence, or a multi-valued function. One logical approach would be to
join the samples with straight lines. Another approach may use a more complex
method to interpolate the point between the samples. How the points are
interpolated may have a significant effect on:

the sampling rate,

how data is stored,

or the way programs are written to interpolate the data.

Complicated shaped signals must have a sufficient number of samples to define


the desired shape. If not enough points are used, the implied interpolation will
not represent the desired signal. This error is called aliasing.
The data must be sample with a high enough sampling frequency Fs to prevent
aliasing. This is referred to as the Nyquist criteria. One-half the sampling
frequency is referred to as the Nyquist frequency Fnyq, and to prevent aliasing, the
maximum frequency of the signal must be less than the Nyquist frequency.

Fnyq =

Fs
1
=
2 2 t

(3.26)

The Nyquist criteria requires:


The sampling frequency to be greater than twice the
maximum frequency.
or
"There must be at least two samples for the period of
the highest frequency in the signal."

The Nyquist criteria sets a theoretical limit for the lowest possible sampling
frequency. It is possible but difficult or expensive to recover the original signal
when sampled at this rate. Usually a much higher sampling rate is used to enable
the interpolation to be easier and faster.
There becomes a compromise between a higher sampling rate that requires more
memory but has a faster run time, verses a lower sample rate that requires less
memory but a more elaborate interpolation scheme that may require a longer run
time.

Page 3.47

A Practical Understanding of Pre- and Poststack Migrations

3.4.2 Aliasing construction


Aliasing is illustrated by the construction in Figure 3.36. Sinusoids are sampled
at various rates.
1. Connect the samples with straight lines.
2. Compare the interpolated curve with the original curve and note the
appearance of any new apparent frequencies.
3. Estimate how many samples are required per cycle for linear
interpolation.

Comments and conclusions

Linear interpolation may be adequate for seismic traces if the sample rate is
more than six times the maximum frequency.

The Nyquist theorem states that the sample rate must be at least two samples
per period, but from the construction, linear interpolation will fail at these
rates.

When sampling below the Nyquist rate, a sinusoid will appear at a lower
frequency.

The original continuous time signal with a maximum frequency that is close to
the Nyquist frequency can be exactly recovered if a Sinx/x is used.

Software could be written to take advantage of a reduced sampling rate. It


would be a compromise between speed and size of storage.
Sinx/x
interpolators may take ten times as long as linear interpolators.

Linear interpolation may be OK is the sampling rate is greater than six times the
maximum frequency.

Page 3.48

Chapter 3 Bits and Pieces

Figure 3.36 Exercise to illustrate aliasing; connect the dots with straight lines to
evaluate linear interpolation.

Page 3.49

A Practical Understanding of Pre- and Poststack Migrations

3.4.3 Sampling and signal recovery (interpolation)


Assume we have continuous time signal that is band limited in Figure 3.37a and
the amplitude spectrum in (b).
Now assume a periodic sampling in time (c) with its periodic sampling in
frequency (d).
Sampling is accomplished by multiplying the two continuous time signals in (a)
and (c) to get (e). The sampled spectrum in (f) is obtained by convolving the
corresponding spectrums in (b) and (d).
We can recover the original continuous signal in (a) exactly by applying a box-car
filter in the frequency domain to isolate the original spectrum, as illustrated by
the dashed box in (f).
Multiplying by the box car in the frequency domain is equivalent to convolving
the sampled signal in (e) with a sinc or sinx/x waveform as illustrated in (g).

The main points:


1. A un-aliased sampled signal can reconstruct the exact continuous original
signal by convolving with a sinc function.
2. The sinc waveform extends to infinity in each direction and requires a very
long sampled signal.
3. Convolving with a very long operator will require a large computational
effort.
4. Data stored in a computer has a finite length and will require the sinc
interpolator to be truncated.
5. Sampling at a higher rate will enable the use of a shorter convolutional
waveform.

Page 3.50

Chapter 3 Bits and Pieces

Convolve

Multiply

Figure 3.37 Sequence of figures illustrating the sampling and reconstruction of a


continuous signal from Lathi [560].

Page 3.51

A Practical Understanding of Pre- and Poststack Migrations

3.4.4 Increasing the sample rate


A larger sampling frequency fm = 1/t, will separate the periodic
spectra as illustrated in Figure 3.38a below which contains twice the
sampling frequency of Figure 3.37d. The increased separation makes
it easier to isolate the central spectrum with an interpolation operator
that is not a boxcar in the frequency domain.
The interpolating filter is continuous in the time domain and therefore
is not periodic in the frequency domain. It therefore attenuates all the
periodic spectra beyond the central spectra.

a)
Interpolation
filter

b)

Figure 3.38 Illustrating a) double sampling frequency of that shown in Figure


3.37d producing in b) an increased separation between the periodic spectra for
easier isolation.

We desire the interpolator to have a spectrum that is reasonably flat in the area of
the signal spectrum, but tends to zero by the time it reaches the other periodic
spectra.

Page 3.52

Chapter 3 Bits and Pieces

3.4.5 Sinc function in the discrete domain


Our discussion about a sinc function is only correct for the Fourier transform of
continuous functions. In the discrete world, where sampling has produced a
maximum frequency, the infinite extent of the sinc function will wrap around in
either domain, .i.e., the sinc function is replaced with its periodic form

F (t ) =

A1 =

sin ( xN )
xN

sin ( xN )
.
N sin ( x )

A2 =

sin ( xN )
N sin ( x )

(3.27)

N = 21

Figure 3.39 Comparison of the continuous sinc function and its corresponding
discrete function. This figure can represent both time and frequency domains.
(2005\SincSinNt.m)

The sinc function has infinite extent, while the discrete form is the sinc function
convolved with the periodic impulse train. The discrete function contains the
overlapping tails and is only slightly different from the continuous sinc function
over the first half of the spectrum as illustrated in Figure 3.39. When N is smaller,
the difference increases. Note that N should be an odd value.

Page 3.53

A Practical Understanding of Pre- and Poststack Migrations

3.4.6 Interpolation operators


The sinc interpolator is very efficient in the frequency (wavenumber) domain by
isolating the desired spectrum with a box-car shape. However, it is inefficient to
apply in the time domain. We will now consider the interpolating functions
defined in the time domain that have a finite extent as illustrated in Figure 3.40.
A boxcar interpolator in the time domain is chosen to be the width of the sample
interval then convolved with the data resulting in a blocky interpolation (a). The
effect in the transformed domain is to multiply the periodic spectra with a sinc
function. The attenuation of the periodic spectra is quite poor as there is
significant ring to the sinc function.
A triangular interpolator has a base with of two sample intervals and can be
formed by convolving a boxcar with itself (or convolving the data twice with a
boxcar). The effect on the time data is a linear interpolation (b) between each
point. In the transformed domain, the spectrum is the product of the two sinc
functions or sinc2. There is a better attenuation of the higher frequencies, but the
pass band is smaller.
We can also convolve the triangle interpolator with the boxcar to produce a
quadratic interpolator (c) that produces a smooth curve. The spectrum is sinc3.
There is excellent attenuation of the higher frequencies but more attenuation in
the pass band.
These three interpolators and their transforms are illustrated in Figure 3.41. Note
the location of the line at a relative frequency of 0.1 that represents 20% of the
Nyquist frequency. The relative attenuation of the higher frequencies in this band
can be observed from the three sincn spectra. When n > 2, smoothing occurs and
the interpolated points will not pass through the original sample points.
A spline interpolator fits a polynomial curve to the samples.
polynomial

yi ( x ) = ai + bi x + ci x 2 + d i x 3 ,

A trivial cubic
(3.**)

could use four point yi ( xi 1 , xi , xi +1 , xi + 2 ) to define a curve that is only used in the
window between the two central points yi ( xi , xi +1 ) . However, the curvature would
be discontinuous from one window to the next. The cubic spline uses the two
central points to fit yi ( x ) and two additional constrains that use the second
derivative ( yi ( x ) = ci + d i x ) to ensure continuity of curvature at the two central
points yi ( xi , xi +1 ) . A set of equations from each window is inverted to get the
corresponding coefficients for each yi ( x ) . Each equation is then used to fill the
corresponding window. This method ensures a fit at each point. A feature of the
cubic spline is that the matrix is tri-diagonal for easier inversion. All spline
interpolators can also be used with arbitrary intervals between each sample.

Page 3.54

Chapter 3 Bits and Pieces

a)
b)
c)
Figure 3.40 Interpolation illustrations using a) boxcar, b) linear, and c) quadratic
functions

a)
b)
Figure 3.41 Three interpolators in a) with a boxcar, linear, and quadratic shape,
and b) their corresponding Fourier transforms. (N = 11)
Continued convolution of boxcar with itself produces interpolators that tend to
the shape of a Gaussian function (Central limit theorem). A three point boxcar
interpolator is self convolved up to nine times with the results illustrated in
Figure 3.42. Convolving with a box car a number of times (say three) is also an
efficient method of smoothing that is often used to smooth elevations to produce
a floating datum: the size of the boxcar would equal a spread-length.

Figure 3.42 Continued convolution of a three point boxcar with itself produces an
interpolator that tends to the Gaussian shape. (N = 3)

Page 3.55

A Practical Understanding of Pre- and Poststack Migrations

3.4.7 Wrap around effects of aliasing


Use of the Discrete Fourier Series assumes both time and frequency domains are
periodic. When the Nyquist criteria is violated, the periodic nature of the
spectrum causes overlap with some of the frequencies as illustrated in Figure
3.43 and Figure 3.44.
A seismic trace sampled with 2-millisecond sample rate, or 500 Hz sampling
frequency will have a Nyquist frequency Fnyq of 250 Hz. The usual seismic signal
has a frequency content less than 100 Hz, well below the Nyquist rate and has
about five samples per cycle (of the maximum frequency). When the maximum
frequency is 80 Hz, there are 6 samples per cycle.
If a signal is sampled at a rate such that the Nyquist frequency is less than the
maximum frequency, the frequencies above the Nyquist rate will appear to be
reflected back at the Nyquist frequency as illustrated in Figure 3.43.

a)

b)
Figure 3.43 Aliasing in the frequency domain with (a) a normal sample rate, and
(b) when the sampling frequency is halved.

Page 3.56

Chapter 3 Bits and Pieces

The aliased frequencies may also be illustrated by the periodic spectrum


illustrated in Figure 3.44 with a comb filter illustrated in (a) and in (b) a two-sided
spectrum repeated over the interval Fs1. When Fs1 is halved to Fs2, the spectra
overlap as shown in (c). The arrows show the periodic locations of the new Fsamp.
The periodic nature of the spectrum may also be considered as a single spectrum
wrapped in a circle with the circumference defined by the sampling frequency, as
illustrated in (d). Shrinking the diameter will cause the spectrum to overlap.

a)
Amp

-Fnyq

Fnyq
b)

Freq

Lower sampling
frequency

Amp

-Fs2

-Fnyq

Freq

Fs1

Fnyq

2Fs2

Fs2

c)
Fs1
Fs2

d)
Figure 3.44 Aliasing viewed from repeated spectra, with a) the comb filter of
sampling theory, b) normal sampling, c) half the sampling frequency, and d)
illustrating the circular wrap-around concept where the circumference defines the
sampling frequency.

Page 3.57

A Practical Understanding of Pre- and Poststack Migrations

3.4.8 Aliasing in two dimensions


Two dipping events are shown in Figure 3.45. The positive dip 1 contains higher
frequencies than the negative dip 2. Data in each time trace is well sampled, and
not aliased. However, horizontal traces will be aliased as illustrated by the
horizontal traces at the bottom of the figure. It shows amplitudes taken from the
position indicated in the main figure. Only one sample defines the aliased dip.

Figure 3.45 Two dipping events with different bandwidths. The horizontal trace
across the bottom shows the amplitudes taken from the main figure.
Aliasing energy can be identified in the 2-D Fourier transform of the above figure.
A schematic representation of the FK domain is shown in Figure 3.46. Note:

The aliased event 1 crosses over the spatial Nyquist frequency Kn.

The aliased frequencies are indicated in gray.

The maximum frequency before spatial aliasing is Fc.

The maximum frequency after aliasing is F1.

Page 3.58

Chapter 3 Bits and Pieces

Fc or Kc

F1
Fmax

Figure 3.46 Examples of FK data illustrating spatial aliasing.


Aliasing could be prevented by limiting the frequency content of the dipping
event to Fc. The frequency Fc may be defined from

tan ( ) =

Kn
Kc

(3.28)

where is the dip before or after migration, Kn is the spatial Nyquist frequency,
and Kc the cut off frequency expressed in depth parameters, i.e.

Kn =

1
2 CMPx

, and

Kc =

2
Fc
V

(3.29)

giving a cut-off frequency fc by the aliasing equation:

fc =

V
4 CMPx tan ( )

(3.30)

This important equation is also used to define the station interval (2 * CMPx) in
field designs. Some alternate forms of this equation replace with , then tan()
with sin(). Care must be taken to know whether fc is before or after migration.

Additional information on aliasing may be found in Bancroft [554], [555], [625],


and Lumley [310], [445], and [462].

Page 3.59

A Practical Understanding of Pre- and Poststack Migrations

3.4.9 Time domain aliasing in two dimensions


Aliasing and wrap-around also occurs in the times and space domain as
illustrated in Figure 3.47. This time domain image shows an input structure and
the resulting section obtained from Fourier Transform modelling. Note that
energy of the dipping event departs from one side of the model and re-enters on
the other side.
This effect is more pronounced when the figure repeated many times and with
higher amplitudes as illustrated in Figure 3.48. Events are seen leaving one
section and appear to enter the neighboring section, which in reality is the same
section.
x

a)

b)

Figure 3.47 Time domain wrap-around illustrated by (a) showing an input


structure and (b) a model formed by using the Fourier transform.

Page 3.60

Chapter 3 Bits and Pieces

Figure 3.48 The wrap-around in Figure 3.24 is repeated many times illustrating
wraparound when using the Fourier transform.

Page 3.61

A Practical Understanding of Pre- and Poststack Migrations

The following figure contains an example of a poor FK migration algorithm that


contains many artifacts of wraparound when using the mod50.sgy data..
The main migrated event that is encircled by the widest blue line has many other
pseudo events at the same dip. Three are identified by the thinner blue lines.
Also note the numerous smiles and frowns with various radii that result from the
lower spike on the input section.

Figure 3.49 FK migration (very poor quality) that illustrates spatial wrap-around.
The FK migrated section is repeated tile fashion in Figure 3.50 to aid in
identifying the source of the artifacts. The energy in each tile is migrated relative
to the neighboring zero depths; both upper and lower.
Exercise:
1. Locate the relative source of the artifacts in the encircled events.
2. Locate the center of curvature for the many smiles and frowns that
originate from the lower spike on the input section.
3. Note that the reflection and the reflector intersect at the apparent
surface.
4. Blue reflectors indicate the apparent surface is above the actual surface,
while green indicates an apparent surface below the actual surface.

Page 3.62

Chapter 3 Bits and Pieces

Figure 3.50 Tiled version of the FK migration on the previous page to illustrate
wrap-around.

3.4.10 Examples of aliasing and its effects on migration


Figure 3.51 shows three images of a) an input section, b) the 2-D Fourier
transform, and c) the 'frequency' domain migration. Note

Page 3.63

A Practical Understanding of Pre- and Poststack Migrations

The origin is at the upper left corner.

The Nyquist wave-number occurs at the center of the horizontal transform


space.

The energy of the dipping event crosses the Nyquist wave number in (b).

After migration, only the left portion of the dipping energy in (c) is correctly
migrated. The remaining energy is distorted by the migration and becomes
noise to the remaining part of the migration. (A full discussion of this effect is
given later in section 4.4)
x

kx

a)

b)
kx
f

c)
Figure 3.51 Illustration of aliasing or wraparound in the frequency domain
showing (a) an input time section, (b) the 2-D Fourier transform and (c) the
migration in the frequency domain. Spatial aliasing occurs at the center vertical
line.

Page 3.64

Chapterr 3 Bits and Pieces

One me
ethod of controlling
c
g aliasing is illustra
ated in Figure 3.34 in which the
aliased frequenciies are re
emoved wiith a high
h cut filterr that is applied
a
to
o the
tempora
al frequenc
cies.

kx
t

a)

b)
kx
x
f

c)
Figure 3.52 Simiilar to Figu
ure 3.51 except that the
t input section has
s been filte
ered
wiith a low pass freque
ency filter to
t prevent aliasing.

P
Page
3.65

A Practica
al Understandin
ng of Pre- and Poststack
P
Migrations

Figure 3.53 illusttrates an extreme case


c
of aliasing witth tempora
al frequen
ncies
extende
ed to the Nyquist
N
freq
quency (as
s well as th
he Nyquistt wave-num
mber). Now
w the
effects in the freq
quency dom
main migra
ation are more
m
appa
arent. The migrated time
domain in (d) dis
splays the migrated noise clos
se to the original
o
po
ositions off the
diffractiion and dip
pping even
nt.

kx
f

a))

b)
x

kx
t

c)

d))

Figure 3.53
3
Illustration of th
he conseq
quence of aliasing.
a
T input section
The
s
(a)) has
a very high
h
frequ
uency conttent, in (b)) the aliase
ed energy visible in the transfform
and (d) the migrated
domain is severe
e, with (c) showing migration
m
d time sec
ction
showing
g aliasing noise.

P
Page
3.66

Chapter 3 Bits and Pieces

x
t

a)
kx

b)
Figure 3.54 Repeat of Figure 3.53 illustrating a) the correctly migrated and b)
aliased dipping energy.

Page 3.67

A Practical Understanding of Pre- and Poststack Migrations

3.5 Trace interpolation to reduce aliasing


The design of an acquisition project requires the trace spacing to be close
enough to prevent aliasing. However, some 3-D projects are deliberately
acquired with a larger trace spacing for to reduce costs, with the anticipation that
aliasing effects may be reduced with trace interpolation.
Kirchhoff and FK methods may be written to deliberately migrate aliased data,
thereby preserving the frequency of the aliased data, however, aliasing noise will
be present [445]. Choices are made between acceptable noise and desired
frequency content of dipping events.
It is also possible to insert or interpolate traces to reduce aliasing. This may be
accomplished by a number of methods

Inserting a dead traces between existing traces

Linear interpolation between neighboring traces

Dip interpolation by cross correlating neighboring traces,

2-D Fourier transform methods

X methods

Use of DMO

The worst form of trace interpolation, adding dead traces can preserve the
frequency of dipping events with a controlled addition of aliased noise. This is
illustrated in Figure 3.55 which shows the x-t and FK domains a) before and b)
after one dead trace was inserted between the live time domain traces, while c)
show an ideal interpolation. Note:

Dipping event in the input data (a) is aliased, where the maximum usable
frequency is F1, and the aliased portion shown in red.

Adding dead traces in the time domain causes the FK domain to double, with
the original FK transform repeated twice, as illustrated in (b).

The left dipping event in (b) is still aliased.

The right dipping event is not aliased and has a maximum frequency of F2.

FK filtering (or migration) should remove a large amount of aliased noise


above 45 degrees.

Page 3.68

Chapter 3 Bits and Pieces

45
Aliased
energy

a)

45
Aliased
noise

b)

45

c)
Figure 3.55 Illustration dead trace interpolation, a) is the original xt, and FK
domain of an aliased dip, b) the result after adding dead traces between each live
trace in the time domain, and c) an ideal interpolation.

Interpolation should be performed before migration when the reflection energy is


less than 45 degrees. The migration process itself should help to remove the
noise. Some migration algorithms, such as finite difference, are unable to
remove dipping noise and an FK filter may be required before migration.

Page 3.69

A Practical Understanding of Pre- and Poststack Migrations

3.6 The wave equation(s)


There are many forms of the wave equation and many assumptions that are made
to derive these many different forms. The most common assumptions are the
medium is acoustic and that only P-waves exist (no shear waves) and that energy
is propagated in one direction (no multiples). The following is a brief introduction
to various types of the scalar "wave equations".
Using Kosloff's [295] notation, the acoustic two dimensional wave equation, with
velocity v(x, z), pressure P(x, z, t), bulk modulus K, and density (x, z) is,

1 P
1 P
1 2P

=
x x z z
K t2 .

(3.31)

The usual form of the wave equation expressed with constant density is,

2P 2P
1 2P
+
= 2
v t2
x2 z 2

(3.32)

Another form is represented in the Fourier transform domain with second order
partial derivatives of P with respect to x, z, and t, giving,

k +k = 2 ,
x
z v
2

(3.33)

where the k and terms are the Fourier transforms of distance and time.
The paraxial wave equation, or single square root equation may be written as,

k z = 2 k x2
v

(3.34)

or when inverse transforming kz to the derivative we get


1

2
P
v2 2
2
= i 2 k x P = i 1 2 k x P .
v
z
v

(3.35)

Approximate solutions to the square root in the above equation lead to the
development of the parabolic wave equation, also known as the 15 degree
solution, where the variable P is now replaced by a time shifted Q, i.e.,

Q
iv 2
kx Q ,
=
z 2

Page 3.70

(3.36)

Chapter 3 Bits and Pieces

or,

Q v 2Q
=
.
z 2i x 2

(3.37)

Another approximation to equation 3.28 yields the 45 degree approximation,

- i k x2
Q
Q
=
z vk x2 .
2
v 2

(3.38)

The above methods solved for the depth derivative dQ/dz. This derivative may be
used with the wavefield at depth z to compute a new wavefield at depth z+z.
Using the following definition of the derivative, i.e.,

dP (z ) P (z + z ) P (z )

,
dz
z

(3.39)

the terms may be rearranged to solve for the wave field at P(z+z), i.e.,

P (z + z )P (z ) + z

P ( z )
.
z

(3.40)

Many other solutions exist that are based on alternate approximation to the
square root found in the above equations. Some are in Appendix 1.

The Eikonal equation is similar in some manners to the wave equation 3.25 and is
included for a comparison.

t
+
x

1
V2

(3.41)

The Eikonal equation does not involve the pressure wave P but is an equation of
slowness used for computing local travel times for depth migration.

Page 3.71

A Practical Understanding of Pre- and Poststack Migrations

3.7 Derivatives and an approximate solution to the full wave


equation
The wave equation is a partial differential equation that is usually solved with
advanced mathematical principles. The basics are quite simple to understand
and simulate in a computer. This following is a brief introduction to illustrate the
finite difference method. (A 15 degree solution is presented in section 5.6.1, with
additional solutions in Appendix 1).
Differential equations relate "things" such as position, velocity, and acceleration,
or other "things" such as temperature, and the rate at which temperature
changes. Recall

Velocity is the rate of change of distance with respect to time.

Acceleration is the rate of change of velocity with respect to time.

The rate of change of "something" such as x relative to time t is defined by the


derivative with the mathematical symbol,

dx
,
dt

(3.42)

where dx approximates a small part of x, and dt approximates a small part of time


t. The small distance divided by a small time gives the velocity. Similarly, a small
change in velocity divided by a small change in time gives the acceleration. i.e.,

distance

x,

velocity

v =

dx
,
dt

a =

dv
dt

acceleration

d dx
dt dt

d2x
= 2 .
dt

The scalar wave equation relates the rate of changing pressure in the x and z
directions with the rate of changing in pressure over time, i.e.,

2P 2P
1 2P
+
=
.
x2 z 2
v2 t 2

Page 3.72

(3.43)

Chapter 3 Bits and Pieces

The relationship between two "things" may be plotted on a graph as illustrated in


Figure 3.56 The geometry for a) the first and b) the second derivatives. When
relating distance with time, the slope or first derivative would give velocity. The
velocity at a point midway between t7 and t8 is approximately,

vel (7 ) =

dx
dt

distance

x ( 8 ) x (7 )
.
t ( 8 ) t (7 )

(3.44)

distance

x8

x8

x7
x6

x7
x6

t6

t7

t8

time

t6

t7

t8

time

Figure 3.56 The geometry for a) the first and b) the second derivatives.
The second derivative or acceleration at t7 would be found from,

x ( 8 ) x (7 ) x (7 ) x ( 6 )

d x vel (7 ) vel ( 6 )
t
t
accel (7 ) = 2
=
t
t
dt
2

x ( 6 ) 2x (7 ) + x ( 8 ) T x (7 )

=
t 2
t 2

(3.45)

where T is the second derivative operator (1, -2, 1) on three equally spaced
values.
Now back to the wave equation, each of the second derivatives may be
approximated by

d 2 p Px 1 2Px + Px +1 T Px

= 2
dx 2
x 2
x

with z, and t constant,

d 2 p Pz 1 2Pz + Pz +1 T Pz

= 2
dz 2
z 2
z

with x, and t constant,

d 2 p Pt 1 2Pt + Pt +1 T Pt

= 2
dt 2
t 2
t

with x, and z constant.

Page 3.73

A Practical Understanding of Pre- and Poststack Migrations

These second derivatives are taken at a point (x, z, t) within the volume and along
the respective orthogonal axis illustrated in Figure 3.57. In these equations the
value of P at (x, z, t) is simplified to Px, Pz, and Pt, relative to the respective axis.

Pz-1
Pt+1

Px+1

P
Px-1

Pt-1
Pz+1

Figure 3.57 Orthogonal direction of x, z, and t for solving the wave equation.
The wave equation requires the finite difference terms to balance, i.e.,

Px 1 2P + Px +1 Pz 1 2P + Pz +1 Pt 1 2P + Pt +1
+
=
.
x 2
z 2
V 2 t 2

(3.46)

If we know six of the seven points we can find the seventh point by rearranging
the above equation to solve for that point. When modelling, we have two vertical
planes of P at times at t =0 and t =1 and desire to find the value at the next time
plane at t =2. We therefore solve equation 3.39 for Pt+1, first getting

P 2P + Px +1 Pz 1 2P + Pz +1
+
Pt 1 2P + Pt +1 = V 2t 2 x 1
,
x 2
z 2

(3.47)

P 2P + Px +1 Pz 1 2P + Pz +1
Pt +1 = 2P Pt 1 + V 2t 2 x 1
+
.
x 2
z 2

(3.48)

then,

The process is repeated for all points on the computed plane P(t=1) and then the
process repeated for all points on the next plane P(t=2) and so forth. This method
was used in the modelling examples in chapter two. Given the first two "photos,"
all the points in the next photo were computed using the above equation, and
then the next photo, and so on.

Page 3.74

Chapter 3 Bits and Pieces

Equation (3.39) may be modified to solve for P at (t-1) to perform a reverse time
migration as in Fricke [98]. It may also be modified to solve for P at (z+1) to
compute a new depth layer in downward continuation migration.

A solution for Pz+1 could be used for full wave equation migration, but is not
normally used as it propagates energy in all directions, including multiples from
velocity interfaces. This may seem an advantage, but at the present time it is
very difficult to define the structure accurately enough to collapse the multiples.
The one-way wave equations do not have this problem.
An additional problem encountered by this method is the requirement of the two
starting layers.

The scalar wave equation defined in this section is only valid for constant
velocity (impedance) models. Its use in variable velocity models may, or will,
produce amplitude and phase errors. It may appear to be working correctly as
the reflection amplitudes appear to be correct but the phase may be wrong. The
transmission coefficients will show errors in the amplitude when crossing an
interface. Any algorithm using this equation should be thoroughly tested.
Variable impedance algorithms are available such as equation 3.31
Note the following regarding differentials:

Partial differential equations may also be solved with the aid of the Fourier
transform where derivatives may be computed exactly by a frequency
multiplication and phase shift.

The first derivative is difficult to accurately represent in the time domain,


(why ?).

The second derivative is easier to represent in the time domain than the first
derivative, (why ?).

Page 3.75

A Practical Understanding of Pre- and Poststack Migrations

3.8 Time and depth migration


3.8.1 What is time migration?
It assumes:

assumes a horizontally layered medium

the output is defined as a time section

relative positioning errors in positioning errors may occur in areas with lateral
velocity variations,

velocities are estimated for a local area to focus the data.

A time migration velocity model is based on the RMS velocities that can be
extracted from the stacking velocities.
Kirchhoff time migration:

Usually defines the diffraction shape as a hyperbola from RMS velocities.

Assumes the diffraction is symmetrical.

Collapses energy to the apex of the diffraction.

A higher order polynomial for the diffraction shape may be used.

Downward continuation time migration:

Slowly collapses a diffraction at the same time location on the time section
with each downward step.

The velocity for each downward step is an interval velocity that is computed
from the RMS velocity.

A band of data from each time section becomes part of the migrated section.

Time migration is fast.


Local or relative positioning is assumed.
It is usually the responsibility of the processor.

Page 3.76

Chapter 3 Bits and Pieces

3.8.2 What is depth migration?


Depth migration assumes:

a known complex velocity structure,

the output is in depth,

correct absolute positioning of the data.

Practical depth migration estimates a velocity model by starting at the surface


and estimating the velocity to, and the location of, the first horizon. When that
horizon is defined, the velocity to the next horizon is estimated, and the location
is then defined and then This is known as the layer stripping method. Errors
in any shallow layer will effect those below.
Kirchhoff depth migration:

defines the diffraction shape from ray tracing or wavefront modelling.

Downward continuation depth migration:

Slowly collapses diffraction on the time section with each downward step.

Moves the diffraction towards the top of the time section as the downward
step is increased.

After a downward continuation step to depth z, the top time samples P(x, t=0)
are copied to form the migrated section at each depth layer P(x, z).

FK migration:

Tends to a depth migration because of the time to depth stretching.

It is only valid for smoothly varying velocities.

Depth migration assumes a known depth structure.

Velocity anisotropy should always be considered.

Many iterations may be required to define the depth structure.

Usually one depth layer is defined for each iteration.

Tomographic solutions enable the velocities of multiple layers


to be estimated simultaneously.

Depth migration is the responsibility of an interpreter.

Page 3.77

A Practical Understanding of Pre- and Poststack Migrations

3.8.3 Advantages and disadvantages of time migration


Time migration, even though it is an approximation, it is still the main work horse
in the industry and is used much more than its depth counterpart.
The following reasons highlight the durability of time migration.

The objective is best focusing and correct relative positioning [743].

The stacked time section and the migrated time section appear quite similar
with respect to the position of the data.

Diffractions collapse to the same position.

Velocities only need to be within 3 to 10 percent.

Small change in velocity only affects the focusing of the data, not the time
position.

Velocities may be focused independent of time or depth.

Complex structures in a shallower part of the section do not effect the


migration of the data below.

It is usually computationally faster.

Requires fewer runs to produce desired results.

Is the responsibility of the processor (not interpreter).

A robust process.

Disadvantages of time migration are:

Output is in time, not depth.

Structured data may not be positioned correctly.

Does not position dipping data as accurately as depth migration.

Might not focus structurally complex data.

When comparing sections acquired with various propagation modes (P-P, S-S,
or P-S), reflections from the same event will occur at different times.

The advantages and disadvantages of time migration become the disadvantages


and advantages of depth migration. The following lists are presented to
emphasize some of the points of depth migration.

Page 3.78

Chapter 3 Bits and Pieces

3.8.4 Advantages and disadvantages of depth migration


The advantages of depth migration are:

The objective is to have correct absolute positioning [743],

It is displayed in depth,

It may be structurally correct (if the velocity model is correct),

When comparing sections acquired with various propagation modes (P-P, S-S,
or P-S), reflections from the same event should occur at the same depth.

Disadvantages of depth migration are:

Requires many iterations to converge on a suitable depth model.

Highly dependent on the structure above the point of migration.

May require long run-times.

May converge to an incorrect geological model.

May require special hardware to handle interactive processing.

Requires velocities to be within 0.5 to 1.0 percent.

Depth positioning errors are directly proportional to velocity errors.

The geological model becomes the responsibility of the interpreter.

The geological model must include anisotropy.

The objective of seismic migration is to produce a section that represents the


geological structure of the Earth. This implies that the migration should only
output a depth section. Fortunately, the development of migration introduced
time migration, one that is robust, and in most areas approximates the benefits of
depth migration.
There are areas in which the data is too complex for time migration, or in which
the depth migration is superior. Even in these cases the time migration has been
very useful in helping to develop the depth models used in the depth migrations.
In highly structured data, it may not be possible to define a depth model, and only
time migration solutions that attempt to focus the data may be practical.

Page 3.79

A Practical Understanding of Pre- and Poststack Migrations

3.8.5 The effect of lateral velocity changes on diffractions


Figure 3.58a illustrates a geological structure with a lateral velocity change
defined by a dipping interface. The rays traced through this dipping layer define
a tilted diffraction shape that is illustrated in part (b). Note

the diffraction is tilted,

the apex has moved horizontally to the right,

the apex is below the point where the image ray meets the surface, and

the tilted diffraction is now non-hyperbolic.

Image ray

a)

Apparent
location

b)
Figure 3.58 Lateral velocity change in (a) causes a tilt on the diffraction (b).
Collapsing this diffraction using vertical hyperbolas will result in a misspositioning and smearing of the assumed diffracting point. The positioning error
may be corrected by tracing the image ray back to the scatter point.

Page 3.80

Chapter 3 Bits and Pieces

3.8.6 A pictorial summary


Time Migration

Depth Migration

Hyperbola

Computed diffraction
a)

Time section

Depth section
b)

Z=0

Diffractions at same time

Diffractions converge to surface


c)

Figure 3.59 Pictorial comparisons of time and depth migrations.

Page 3.81

A Practical Understanding of Pre- and Poststack Migrations

3.8.7 Image ray conversion from time to depth migration


Time migration is usually the starting point for depth migration. Image ray
processing introduced by Hubral [3] converts the time migrated data to an
approximate depth migrated structure.
The shortest time raypath from a point reflector to the surface is an image ray.
This image ray will also represent the apex (minimum time) on a diffraction. Time
migration collapses energy to this point at the apex of diffractions. The image ray
may curve in areas with lateral velocity variation, but will arrive normal to the
surface, or in a vertical direction. Tracing this raypath from the surface on a
depth model enables a lateral correction and a time to depth conversion from
which, samples in the time migrated section may be moved to approximate a
depth migration.
The approximate depth structures, and the interval velocities from time migration,
help to define the initial depth model.
Papers by Parkes [80] and Loveridge [79] use ray tracing to relate the time
migration and depth migration, and their diagram is repeated below.

Figure 3.60 Time and depth migrations related by normal and image rays.

Page 3.82

Chapter 3 Bits and Pieces

Normal rays are the familiar zero offset rays that reflect at right angles from
the reflectors

Image rays are rays that leave the surface vertically and travel to the reflector,
as illustrated in the following diagram.

Figure 3.61 Description of 'normal' and 'image' rays.


Use of the traveltimes with vertical and lateral positions along an image ray will
allow the conversion of a time migrated section to an approximate depth migrated
section. That is implied in Figure 3.61 where time migration tends to position the
migrated energy at the apex of the diffraction or below time zero of the image ray.
Note however, that the energy of the tilted diffraction will be dispersed since it
will not align exactly with the time migration hyperbola.
Care must be taken to ensure the time section is a true time section and not a
hybrid migration produced by time to depth stretch in an FK migration.
Why is the image ray a minimum travel time from a scatterpoint to the surface?
1. Consider a wavefront emanating from a scatterpoint. The first arrival of the
wavefront at the surface will be tangent to the surface (otherwise another
point will have reached there first). Rays are normal to the wavefront and thus
normal to the surface.
2. Imagine a colocated source and receiver at the scatter point and the surface
as a reflector; the raypath (minimum time) must be normal to the reflector.

Black [323, 324 , and 331] confirms that image ray correction is not an exact
process, and is only accurate for moderate dips. Bevc et al [582] identify errors
with image ray corrections.
Image ray correction is of great value in establishing an initial depth model.

Page 3.83

A Practical Understanding of Pre- and Poststack Migrations

3.9 Points to note in Chapter 3

A semicircle represents all possible reflectors for a given time response when
the source and receiver are located at the same position (constant velocity).

An ellipse represents all possible reflectors for a given time response when
the source and receiver are located at the foci of the ellipse (constant
velocity).

The shape of a diffraction may be described by a hyperbola and is accurate


when using the RMS velocity in a vertically varying velocity.

Instantaneous velocity defines the geological structure.

Average velocity is used for direct time to depth conversion.

Stacking velocity is part of NMO and gives the best stack.

Stacking velocities should be RMS velocities in horizontal areas and faster in


dipping areas.

RMS velocities are related to the interval velocities.

Migration should use RMS or instantaneous velocities where appropriate,


however some migrations use smoothed forms of the velocity.

DMO processing requires the NMO to be calculated with RMS (and not
stacking) velocities.

The Fourier transform presents a different view of the same information.

A wave form may be reconstructed by summing sinusoids.

Signals may wrap-around when using the Fourier Transform.

2-D aliasing is controlled by dip, CMP interval, frequency and velocity.

There are many forms of the wave equation. The constant velocity scalar
wave-equation should not be used in a variable velocity medium.

Derivatives may be approximated by simple expressions.

Time migration is based on the assumption of hyperbolic diffraction shapes.

Depth migration involves a complex process to accurately define the


diffraction shapes based on a velocity model.

Image rays may be used to convert a time migrated section into an


approximate depth migration.

Page 3.84

Chapter 3 Bits and Pieces

For further information


Conics: Schaum's Mathematical Handbook [382]
Fourier transform: Digital Signal Processing [384]
Numerical Recipes [187]
The Fourier Transform and its Applications [385]
Wave equations: Claerbout's "Imaging The Earth's Interior" [294]
Finite difference solutions: Claerbout [294] Kosloff's [295]
Finite difference modelling: Fricke [98]
Time and depth migration: Parks [80]

Page 3.85

A Practical Understanding of Pre- and Poststack Migrations

Blank page.

Page 3.86

Chapter 4 Poststack Migration ( Zero Offset)

Chapter Four
Post-stack Migration (Zero Offset)

Objectives

Identify the three main methods of migration: Kirchhoff, FK, and downward
continuation.

Estimate migration results with a compass.

Understand the principles behind Kirchhoff migration, and recognize the


range of algorithms from very simple to complex.

Identify the features of FK migration and know how and why the parameters
are defined.

Recognize the principles of downward continuation migration, distinguish


between a time and depth migration, and identify different algorithms.

Know the difference between one pass and two-pass 3-D migration
algorithms.

Recognize that all migration algorithms should be tested.

Page 4.1

Chapter 4 Poststack Migration ( Zero Offset)

4.1 Introduction
There are many different methods of migration that exist in the geophysical
literature and some of these methods will be discussed in considerable detail.
The three most common methods follow (as initially described in chapter 1).

Kirchhoff migration:
For every migrated sample, energy is summed along the diffraction, or hyperbolic
paths, on the input section. The summed value becomes the amplitude value at
the output location. Additional scaling and filtering may be required.
FK migration:
This method makes use of the 2-D Fourier transform to convert the input section
into the 2-D Fourier domain where it is migrated with a simple algorithm. The
inverse transform provides the migrated structure. The F denotes the Fourier
transform of time, and K denoted the Fourier transform of space or distance.
Downward continuation:
This method is quite different to the above two migrations as it works on a
conceptual volume of information rather than two planes of data. The input 2-D
time section becomes the top surface of the cube, and various processes
continually compute the next downward layer until the volume is complete. The
time section at each downward step represents data that would be produced if
the sources and receivers had also been lowered to the respective depth. This
volume of information will then yield the migrated structure at the end of the cube
where t = 0 for a depth migration, or from the last times section at z = zmax for time
migration. Some downward continuation algorithms are referred to as finite
difference, phase shift, or, X.
Reverse-time migration is similar to downward continuation in that it computes
the entire volume (x, z, t) however, it computes constant time images (x, z,
t=tconst), the last one being the front surface of the cube at (x, z, t=0). An
advantage of the reverse-time algorithm is that is uses the full wave equation and
allows steeper dips to be migrated. This algorithm starts at (x, z, t=tmax) with all
data equal to zero, and draws energy from the surface as the images or photos
progress toward t = 0. The full wave equation may introduce unwanted multiple
energy

Page 4.2

Chapter 4 Poststack Migration ( Zero Offset)

Sum energy and move


x

x
t
or
z

a)

Kx

Kx

Kx

x
Kz

Kz

FFT

Mig

IFFT

b)

c)
Figure 4.1 Sketches illustrating the three major poststack migrations, a)
Kirchhoff, b) FK, and c) downward continuation.

Page 4.3

Chapter 4 Poststack Migration ( Zero Offset)

4.2 Migration using a compass


4.2.1 Compass migration of dipping event
This simple method of migration was commonly used by interpreters before the
computerized versions were available and is still very useful in estimating
migrated results from stacked sections. The section must be plotted with scales
to match the apparent vertical distance with the horizontal distance in the area of
interest. For Figure 4.2:

We want to find the reflector position near PQ that produces the reflection
image CD.

The reflection at C could come from any point on a semicircle centered at A.

The reflection at D could also come from any point on a semicircle centered at
B.

Construct these two circles and draw a tangent to them to define the location
of the reflector at PQ.

As PQ is a tangent, the angle APQ is a right angle.

Extend the lines CD and PQ past the surface. Where do they intersect?

The above definition of the point P is not very accurate.


improved with the following construction:

Its location can be

Project the line CD to the surface at S, then get M midway between A and S.

Construct a semicircle with center at M and radius MA.

The intersection of the circles centered at A and M define the point P; the
angle APS is a right angle because the triangle APS contains the diameter AS.

Special note:
The time section has a vertical time scale, giving dips the parameter of slowness
(inverse of velocity). This vertical scale may be modified with the local velocity to
approximate a vertical scale of depth and allow dips to be estimated as slopes or
as angles designated by . It is assumed that time sections are plotted with equal
distances on the vertical and horizontal scales to define values of and that they
may overlap depth section plotted with similar scales. Exact overlaps may be
made with constant velocities, but only approximations apply when the velocities
vary in x or z.

Page 4.4

Chapter 4 Poststack Migration ( Zero Offset)

B
D

x
z
or
t

P
C

a)

x
z
or
t

P
C

b)
Figure 4.2 Construction for migration, a) uses two points and a dip while b) uses
one point.
In the above construction, we have ignored the presence of diffractions that will
occur on seismic data from reflectors that may be truncated.

Page 4.5

Chapter 4 Poststack Migration ( Zero Offset)

4.2.2 Dips before and after migration


The previous construction leads to one of the most basic and important
equations in migration: the relationship between the angle before and after
migration. The equation is the same one developed earlier in section 2.2, and is
referred to as the migrators equation (Robinson [456] page 35). See also [788].
Let

be the dip on the stacked section,


be the dip on the migration.

Figure 4.3 Geometry for relationship between angles.


For the angle ASC = , and the angle ASP = ,

Page 4.6

Chapter 4 Poststack Migration ( Zero Offset)

Table 4.1 Comparison of dips before and after migration.

10

10

15

16

20

21

25

28

30

35

35

44

40

57

45

90

Note:

No seismic reflection can be greater than 45.

The reflection from a vertical reflector will be at 45.

The asymptotes to any diffraction will also be at 45.

Dips after migration go to 90.

Angles on time sections assume pseudo depth conversions with velocities


defined at the time of interest; true depth requires average velocities.

Figure 4.4 Limits to the reflection dips.

Page 4.7

Chapter 4 Poststack Migration ( Zero Offset)

4.2.3 Migration of one point on the input section


The objective of migrating one point is to locate all the possible reflectors that
would produce that reflection. In a constant velocity medium, those reflection
points lie on a semi-circle, with the center of the circle at the surface.
Plot all these possible reflections in Figure 4.5.
This example also describes the Greens function or impulse response of
migration. It is the shape and amplitudes of the migration when the only input is
one sample point. In a constant velocity medium, this function is a semi-circle.
Energy from the source point on the surface will propagate to the reflector, then
will reflect back and then focus the energy at the original point on the surface.
The energy propagating to the reflector is identical to a wavefront, and the return
path is simply the reverse. In practice, where the medium has a complex velocity
structure, the shape of the migration impulse responses are estimated by a using
wavefronts.
As in the hyperbolic modelling process, a seismic reflection can be considered to
be a collection of points, and each point will produce a corresponding semi-circle
after migration. The energy on the semi-circles will combine to re-enforce energy
on a true reflector and then cancel energy at all other locations. The cancellation
of energy will require the parameters defined by the aliasing equation.
The main point of this concept is to emphasize that migration produces smiles
from data that has noise or discontinuities. The smiles are an inherent part of
migration.

Page 4.8

Chapter 4 Poststack Migration ( Zero Offset)

x
z
or
t

Figure 4.5 Migration of one point.

Page 4.9

Chapter 4 Poststack Migration ( Zero Offset)

4.3 Hagedoorn migration (distribution operator)


The previous section illustrated relocating the reflection energy with a compass.
We did know to some extent where the energy was moving. However, if we don't
know where the reflectors are, simply spread the energy along the semicircles
and let "wave form" reconstruction compose the reflector position. This method
is referred to as Hagedoorn migration [1] or the Karcher method (see page 69
[528], but is only valid for constant velocities (i.e. the semicircular shape). It does
however, provide a valuable insight to the migration process, i.e.,

Where necessary the semicircles re-enforce to create a migrated event, and


cancel where there is no energy.

The main point to note is that the semicircular smiles seen on a migrated
section are a natural result of the migration process, and usually emanate
from noise.

Construction 4.1:

Draw a complete semicircle from the surface to the reflector.

Repeat the process by drawing semicircles for each surface location.

z
or
t

Figure 4.6 Construction to migrate a dipping event with semicircles.

Page 4.10

Chapter 4 Poststack Migration ( Zero Offset)

Construction 4.2

Migrate the diffraction in Figure 4.7 with a compass by drawing the semicircles
for each trace.

What should the result be?

Figure 4.7 Construction to migrate a diffraction pattern.

What criteria could be established from Figure 4.6 and Figure 4.7 to enable the
constructive and destructive interference to recreate the migrated image?

There must be sufficient surface points or aperture.

The bandwidth of the wavelet (not shown by these kinematic sketches) must
be compatible with the trace spacing, to prevent aliasing.

Page 4.11

Chapter 4 Poststack Migration ( Zero Offset)

4.4 Wavefront , raypaths, and diffractions for velocities that


vary linearly with depth
4.4.1 Wavefronts
In a constant velocity medium, wavefronts from a point source are circular,
raypaths are straight lines, and diffractions are hyperbolic.
In a variable velocity medium, these shapes change and the migration impulse
response corresponds to a wavefront (an isochron, or an aplanatic surface).
A special case exist when the velocities vary linearly with depth, i.e. v(z) = v0 + kz,
where v0 is the velocity at the surface and k is a constant.
In this medium:
1. wavefronts are still circular, and
2. raypaths are also circular, as indicated in Figure 4.8a and b
The center of the circle for the wavefront in (a) is below the surface and the center
increases in depth (b) as the time of the wavefront increases.
The center of the circle for the ray paths is fixed at an elevation above the surface
where the velocity goes to zero. Note that the raypaths turn upwards and
becomes known as a turning ray. These turning rays can image dips steeper
than 90 degrees.
A family of wave fronts is plotted in (c) with a vertical axis in depth, and represent
time increments of 100 ms in a media with velocity v(z) = 2000 + 2z. These same
wavefronts are plotted in (e) with a vertical axis converted to time using the
average velocity.
The wavefronts of (e) are shown as impulse responses for poststack time
migrated energy from scatterpoints. (In depth migration they would be circular.)
We can use the information in (a) to construct circular raypaths from a
scatterpoint to the surface and accurately define the shape of diffractions. They
can also be computed analytically as discussed by Kleyn (1984) pages 57 to 61
[300].
We will see that the FK method of migration also has a constant velocity
assumption. It uses time to depth stretching or stretching to a constant velocity
in many seismic applications to accommodate the varying velocities.

Page 4.12

Ch
hapter 4 Poststtack Migration ( Zero Offset)

Velocity

Z=-k
-Vo/k
k

Vo

z=0

Z=0

Ray Patth

W
Wave front

Depth z

a)

b)
1

10
000

-2000

-1000

1000

0.5

2000

-10
000

-2000

-1000

1000

2000

-20
000

-0.5

Depth
h

Time

-30
000

-1

c)

d)

e)
Figure 4.8
4 Waveffronts and raypaths in a) and b)
b are circu
ular in line
ear V(z) me
edia.
The wav
vefronts are eccentrric b) and c)
c with the
e centers in
ncreasing in depth. The
wavefro
onts in d) and
a
e) hav
ve the vertiical axis in
n time with
h e) illustra
ating migra
ation
wavefro
onts from scatterpoin
s
nts.

P
Page
4.13

Ch
hapter 4 Poststtack Migration ( Zero Offset)

4.4.2 Diffraction
D
ns
Wavefro
onts and raypaths are circular in linea
ar v(z) me
edia where the velo
ocity
increases linearly
y with dep
pth. Figurre 4.9a sho
ows a sub
bsurface point
p
P1 an
nd a
point on
n the surfa
ace p2. A bisecting line of P1
1-P2 define
es a point P3 on the
e v=0
horizon
ntal line. P3 is the center off the circle
e for the raypath frrom P1 to
o P2.
Circularr wavefron
nts (not sh
hown) can
n be drawn
n from the
e wavefron
nt vertical line
through
h each point to defin
ne the trav
veltimes. Figure 4.9
9b contains
s the rayp
paths
for two other surfface points
s that are defined
d
at z = 0.
Equatio
ons can be
e used to accuratelly define the
t
shapes of diffra
actions in this
media [300].

a)

b)
Figure 4.9
4
Raypa
ath and difffraction co
onstruction in a) for linear v(z)) media an
nd b)
for thre
ee raypaths
s. Diffracttion for the
e same me
edia in Fig
gure 4.8 are shown in
n (c)
for dips
s of 140 de
egrees, d) limited to 90 degree
es, and e) with
w
largerr separatio
on to
illustratting the ap
pproximate
e hyperbolic shape.

P
Page
4.14

Chapter 4 Poststack Migration ( Zero Offset)

-8000 -6000 -4000 -2000


-0.5

2000 4000 6000 8000

-1
-1.5
-2
-2.5
-3
-3.5
-4

c)
-2000 -1500 -1000 -500

500

1000 1500 2000

-2000 -1500 -1000 -500

500

-0.2

-0.2

-0.4

-0.4

-0.6

-0.6

-0.8

-0.8

-1

-1

d)

1000 1500 2000

e)
continued.

The traveltimes from five scatterpoints at 100 m depth intervals are shown in
Figure 4.9c, plotted to a dip of 140 degrees. The same media for Figure 4.8 was
used. Many more diffractions are shown in (d) but are limited for a maximum dip
of 90 degrees, while (e) has increase the separation to 500m to better illustrate
the shape of the diffractions.
We can observe that the shapes of the diffractions in Figure 4.9e are
approximately hyperbolic, even though the wavefronts in Figure 4.8c are noncircular.
Points to note:

Dips well beyond 90 degrees can be present on a diffraction (not possible


with linear velocities).

The above figures contain dipping energy to 90 degrees, well beyond that
required for most migrations. Limiting these dips further will improve the
hyperbolic approximation.

We have assumed increasing velocities.


decreasing velocities.

Most linear v(z) media exhibits hyperbolic diffraction shapes

Page 4.15

The opposite effect applies to

Chapter 4 Poststack Migration ( Zero Offset)

4.5 Kirchhoff Migration


4.5.1 Introduction
The Kirchhoff method [37] has descended from one of the oldest methods of
migration, the diffraction stack. It is considered by many to be the best. Its
implementation may vary from a very simple algorithm to one that is complex.
An important signal detection method of electrical engineering is the matched
filter, where signals are estimated by correlating with the known impulse
response [693] pp. 714-721 and [560] pp. 394-400. The early (1950s) migration
method of diffraction stacking [1] was based on this principle. The seismic model
was reflection coefficients convolved with diffractions, so the diffraction was
correlated (summed) over the seismic section to estimate the reflective structure.
Schneider in 1978 [37] showed the diffraction sum method could be an exact
solution to the wave equation if scaling and filtering were included. His method
was based on the Kirchhoff integral solution used in optics. The term Kirchhoff
migration has been used since then for all algorithms that used the summation
method.
It should be noted numerical inversion methods could also be used to estimate
the reflective structure. These methods produce results that are similar to the
Kirchhoff approach.
Kirchhoff migration can

Be a time migration

Be a complex depth migration

Handle 2-D or 3-D projects

Do prestack migrations

Handle conventional (P-P), converted wave (P-S), or shear (S-S)

Reduce noise by eliminating dipping energy that is too steep

Migrate all input data to any desired output window

Handle data that is unevenly sampled

Control aliasing

Be an efficient interpolator.

Page 4.16

Chapter 4 Poststack Migration ( Zero Offset)

The following Kirchhoff migration principles are common to all the above.

Define a diffraction curve at a migration output location.

Weight and sum the energy within that diffraction shape.

Insert the summed value at migrated position.

Apply a wave-shaping filter.

Repeat the procedure for each migrated output sample.

How the diffraction curve is defined and what weights to apply (and when)
become topics of considerable interest. For example:

Defining the diffraction curve as a hyperbola results in time migration.

Defining the diffraction curve from ray tracing or wave front modelling of a
depth structure results in depth migration.

Kirchhoff migration is more general than Hagedoorn migration:


Hagedoorn semicircles are only valid for constant velocities.
Energy could be spread along non-semicircular wavefronts.
The hyperbolic approximations of Kirchhoff migration extend to
complex areas that suit the RMS velocity model.
See the noncircular wavefronts and diffraction curves below.

-2000 -1500 -1000 -500


-2000

-1000

1000

500

1000

1500

2000

2000

-0.2

-0.2

-0.4

-0.4
-0.6

-0.8

-0.6

-1

-0.8

-1

Supplemental information may be found in Lumley [54], Wiggins [68], Larner


[104], Schneider [37] Jakubowicz [53], and in Docherty [121].

Page 4.17

Chapter 4 Poststack Migration ( Zero Offset)

4.5.2

Kirchhoff migration algorithm

The basic steps for a Kirchhoff migration algorithm are:

Start with a migration sample location.

Compute the travel-times T at each offset to define the diffraction curve.

Interpolate and possibly filter the input data at time T.

Multiply the input interpolated sample by a weighting factor.

Accumulate the weighted values.

Insert the summed value on the output structure.

Repeat the procedure for each migrated sample location.

After summation is completed, additional scaling and filtering may be


required.

The size of the diffraction, or aperture, is usually limited by the dip angles or
which act as dip limiting filters, and/or some maximum offset xmax. The data at
the dip limit may also be tapered to zero over a small range of dips as
illustrated on the left side of Figure 4.10, that shows a family of diffraction curves
for a constant velocity.
xmax

Taper

zone

Figure 4.10 Family of hyperbolic diffraction shapes for a constant velocity


Kirchhoff migration.

Page 4.18

Chapter 4 Poststack Migration ( Zero Offset)

The diffraction shape for a time migration is found from,


where x is the distance between the input and migrated trace, To the two-way time
at zero offset, and V the RMS velocity defined at T0. The amplitude Amp(x) is
usually weighted by,
and tapered to zero between some dip limit such as 40 to 60 degrees. Additional
amplitude scaling is applied after the summation and depends on T0.
The maximum half offset xmax, illustrated at BD in Figure 4.11, may be established
from the triangle ABC, and taking the tangent of max at the region of interest T0,
xmax

Figure 4.11. Angle relationships with hyperbola.

The processing speed may be increased by using the same diffraction shape for
samples with the same time T0 and velocity V(T0). In contrast, vector processing
may be more efficient if an entire input trace is migrated to the output trace.

Page 4.19

Chapter 4 Poststack Migration ( Zero Offset)

4.5.3 Kirchhoff migration aperture


An aperture in seismic processing is considered the spatial range of data that
contributes to calculation (Sheriff 2002). This usually implies a fixed offset range,
but, we can see from Figure 4.12 that a dip limitation on the Kirchhoff migration
provides an aperture that is time variant.
We must therefore be careful when referring to the migration aperture and should
specify either a range aperture in Figure 4.12a or an angle aperture in (b).

The migration aperture is illustrated with a Kirchhoff migration, but it really


applies to all migrations that have a dip limit.
Note from (a) the migration aperture decreases below T0-max-app.

Page 4.20

Chapter 4 Poststack Migration ( Zero Offset)

Range aperture

a)

Angle aperture
T0-max-app

b)
Figure 4.12 Migration apertures.

Page 4.21

Chapter 4 Poststack Migration ( Zero Offset)

4.5.4 Aliasing and Kirchhoff migration


Aliasing may occur in:

the input data,

the migration algorithm, or

the output data.

The Kirchhoff algorithm allows great flexibility and can:

Reduce aliased input data by dip filtering.

Reduce and even eliminate operator aliasing.

Pass aliased input data for maximum frequency content, at the expense of
aliased noise.

Prevent all aliasing on regular grid data.

Optimally prevent aliasing on irregular data.

Output data to any grid, but anti-alias filter to any input grid.

Each input trace, intersected by the summation diffraction, requires the data
around the time of the intersection, to be locally low pass filtered. This local
filtering will prevent the input data or the operator from aliasing.

Figure 4.13a is a high frequency diffraction created from a single value of T


computed at every trace. In this case the amplitude was weighted between the
sample on each trace, then each trace bandpass filtered using (1, 5, 200, 249) to
limit the maximum frequency to the Nyquist frequency. The sampling rate is 2
ms, trace interval 20 m, and the velocity 3,000 m/s. Recall from section 3.4 that
dip, maximum frequency fm, trace spacing CMPx, and velocity V, are related by the
aliasing equation,
giving a maximum frequency for a 45o dip to be 37.5 Hz. This point can be
identified in (b) of the sides of the FK plot . Aliased energy folds back into the
plot with energy extending to the maximum frequency. This operator is aliased.
Energy from a zero dip event will lie vertically in the center of the FK plot. When
used in a migration, this aliased operator will add high frequency horizontal
events (as illustrated in Figure 4.38).

Page 4.22

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)
Figure 4.13 A diffraction a) filtered to the Nyquist frequency and b) its FK
amplitude.

Page 4.23

Chapter 4 Poststack Migration ( Zero Offset)

4.5.4.1 Displacement based antialiasing filter


A diffraction can be considered to be compose of many short linear events. The
dip of these linear events increase with the displacement from the peak value at
To. An antialiasing filter can be designed for each of these dipping elements base
on the aliasing equation. The dip can be computed for the displacement x for
each input trace and the input data locally filtered around T.
The size of the anti-aliasing filter (AAF) to prevent aliasing is illustrated in Figure
4.14 which shows half the period of the maximum frequency dt, to be increasing
with dip.
Unfortunately, the AAF, at a given input location, is dependent on the diffraction
location, and will be different for all diffractions that pass through a given input
location. That means that the same input point will be locally filtered many times,
each with a different bandpass filter. One special bandpass filter of the input data
will not work. Consequently, there are many techniques that are used for dealing
with operator aliasing.

Antialias filtering can be very expensive and time consuming. A high quality
antialiasing filter may make the runtime increase by a factor of twenty. Many
migration algorithms simply ignore the aliasing problem and simply accept the
noise produced by aliasing. This is one major reason why other migration
algorithms may appear to be superior.
An aliased operator can also be a benefit and be deliberately used in areas where
the input data has known aliased dips. The aliased dipping energy can be
migrated with the unaliased data. This aliased energy can also be migrated as
noise.
There are model-based migrations available that will only migrate a portion of the
diffraction, (say a Fresnel zone size), based on the dips defined by the velocity
model. This will also reduce aliasing noise. Caution must be used with these
algorithms as they will tend to verify as correct an incorrect velocity model.
(An Fk migration also has the ability to perform as an aliased migration operator
and migrate aliased dipping events.)

Page 4.24

Chapter 4 Poststack Migration ( Zero Offset)

Figure 4.14 The half-period dt of the anti-aliasing filter shown on the diffraction.

The stepped nature of Figure 4.14 illustrates the frequency content of the antialiasing filter is established by the input trace spacing. Even if the output trace
spacing is much finer, the anti-aliasing filter should be set to the input trace
spacing (or to the output trace spacing if it is larger). [359]

Page 4.25

Chapter 4 Poststack Migration ( Zero Offset)

4.5.4.2 Antialiasing filters


Aliasing may be controlled by a number of methods:

Lowpass filtering the entire section to the maximum frequency of the


maximum dip (a poor choice).

FK filtering the aliased energy, at the sacrifice of removing some dipping


energy.

Low pass filtering the input data a number of times (say 5 to 20) with
maximum frequencies defined by the aliasing equation and the range of dips
expected on the summation diffraction. When summing data along the
diffraction, the slope of the diffraction, at that point, defines which of the input
filtered section will contribute the sample [163].

Local filtering the input trace at the location intersected by the diffraction. A
number of local filters may be used, such as:
box car filter,
triangular filters [445, and 462], or
sin(x)/x filter.

The best choice for removing aliased energy without affecting the reflection
energy is the local sinx/x filter as it provides a boxcar shape in the frequency
domain to sharply attenuate aliased energy. This solution requires significant
computer effort and will slow a migration by a factor of ten over a local boxcar or
triangular filter.
Boxcar and triangular filters approximate low pass filters but will attenuate some
of the reflection signal and pass some of the aliased energy. Fast versions of
these filters will be described in following sections.
The choice of an anti-aliasing method may become a compromise between data
storage, run time, and acceptable noise. Quite often, the anti-aliasing filter is not
included to reduce the computational time of the migration.

Page 4.26

Chapter 4 Poststack Migration ( Zero Offset)

4.5.4.3 Antialiasing filter options


Anti-aliasing methods are illustrated in the frequency domain of Figure 4.15. The
input section (a) has dipping events that cross the spatial Nyquist axis in the
center. Figure (b) illustrates aliasing removal by high cut filter, and is a poor
choice due to loss of high frequencies in the shallow dips. Figure (c) show an
optimum FK filter method that double the frequency content of shallow dip data,
and (d) local high cut filters that are designed using equation (4.7). The right side
of (d) shows the frequency domain filter effect of using local sinx/x, boxcar, and
triangle time domain filters.
knyq

45

a)
knyq

b)

k
45o

knyq

45

k
45o

Sinx/x

45

k
45o

c)

Triangle

45

k
45o

Boxcar

knyq
o

d)

Figure 4.15 Anti-aliasing filter possibilities viewed in FK domain, a) the aliased


input, b) single high cut filter, c) 45 degree cut filter, and d) sharp cutoff at the
Nyquist frequency, possibly using an ideal sinx/x .

Local filtering of the input traces is presented by Silva [251], and Lumley [445]
[462]. The filtered input section method is discussed in Gray [163]. Aliasing in
prestack Kirchhoff is discussed in Bancroft [554]. Some additional papers with
aliasing problems are [395] and [532].

Page 4.27

Chapter 4 Poststack Migration ( Zero Offset)

4.5.4.4 Triangle filter


The triangle filter is a reasonable AAF and the process is optimized by integrating
the trace twice, then selecting and summing only three samples to get the same
result as applying a scalar version of the filter.
Consider a trace x(t), a filter w(t) and the filtered trace y(t), such that

y (t ) = w (t ) x (t )

(4.8)

Y ( f ) =W ( f ) X ( f )

(4.9)

or in the frequency domain,

Differentiating the filter and integrating the trace and we get

dw ( t )
y (t ) =
(4.10)
x ( t ) dt
dt

The integral and differential operators in the frequency domain are 1 / j and j
respectively, and the filter equation becomes

W ( f ) j X ( f ) = Y ( f
Y(f )=
j

(4.11)

which, when simplifying, gives the original filtered function Y(f).


We now apply a boxcar filter w(t), with width 2h, to a trace at time t1 as illustrated
in Figure 4.16. This is accomplished by summing all the data within the range
from t1-h to t1+h. Now, assume the input trace is integrated and the boxcar
differentiated. The integrated trace at t1-h is the sum of the input data to t1-h, and
t1+h. The difference between these two summed values is the sum over the
boxcar. Differentiating the boxcar produces the two impulses as shown and
define the points to be differenced.
We obtain the sum over the boxcar by differencing the two integrated values at
the end of the boxcar at times t1-h, and t1+h.
t

w(t)
t1-h
d w(t)
dt

t1

t1+h
t1+h

t1-h

Figure 4.16 Differentiating a boxcar into two delta functions.

Page 4.28

Chapter 4 Poststack Migration ( Zero Offset)

We can apply any number of integrations n to the input trace and with the same
number of differentiations applied to the filter operator, and end up with the same
original filtered trace, i.e.,

1 n

n
Y ( f ) =
W
f

( j ) X ( f ) = Y ( f )
(
)

(4.12)

Consider a triangle filter w(t) as shown in Figure 4.17. This operator requires a
scalar multiplication for each point within the range of the operator. However,
two differentiations reduce the filter operator to three impulses. If the input data
''
is integrated twice, i.e. x ( t ) =

x ( t )dt

, then the triangle filter becomes equivalent

to four summations, [ '' x ( t1h ) '' x ( t ) '' x ( t ) + '' x ( t1+h ) ] with no multiplications.

w(t)

t
t1-h

t1

t1+h

d w(t)
dt

d2w(t)
d t2

Figure 4.17 Differentiating a triangle filter twice produces impulses at the


beginning, center and end.
Note that the locations of the impulses are dependent on the size of the filters
and that a long operator will require the same time as a short operator.
When used with Kirchhoff migration, the modified triangle filter can be applied
efficiently at all points along the diffraction, with the size of the filter defined by
the antialiasing criteria.
Higher order filters can also be constructed such as a quadratic (n = 3) that has a
bell shaped curve. These filters are produced by convolving a boxcar with itself a
number of times that will tend to a Gaussian shape in both the time and
frequency domains.

Page 4.29

Chapter 4 Poststack Migration ( Zero Offset)

4.5.5 Time interpolation and Kirchhoff migration


The Kirchhoff diffraction usually intersects an input trace between time samples.
Interpolation is required to eliminate broadband quantization noise. Linear
interpolation will usually produce acceptable results when the Nyquist frequency
is three times the maximum frequency content of the data. Higher frequency data
may require sinx/x interpolation.
Interpolation should also be used with any anti-aliasing filter.

Figure 4.18 Illustration of need for interpolation between time samples.

Approximate interpolation schemes contribute broadband noise to the migrated


trace. Phase and frequency filters, such as the Kirchhoff differentiators, that may
be applied after migration will amplify this high frequency interpolation noise.
These phase and frequency filters will also amplify the effects of truncation at the
edge of the migration aperture and the edges of AAF boxcar filters.

Page 4.30

Chapter 4 Poststack Migration ( Zero Offset)

4.5.6 Kirchhoff phase and amplitude correction


4.5.6.1 Phase distortion
Figure 4.19 illustrates a seismic section with one flat event in (b) containing a
wavelet described in (a). Migration should have no effect on the flat event.
However, when T0 is above the event, some energy will always be picked up by
the summation diffraction and inserted into the migrated trace of (c). This energy
will distort the migrated wavelet. No energy will be added when T0 is below the
event. The migrated wavelet requires amplitude and phase correction to match
the input (a).

a)

b)

c)

Figure 4.19 Distortion that results from Kirchhoff migration, a) contains the input
wavelet, b) the Kirchhoff summation into one trace, and c) the resulting wavelet.
An example of the actual seismic wavelet follows in Figure 4.20. The numerically
input data was a horizontal event that was filtered with a zero-phase 5-14-40-50 Hz
bandpass filtered in (a), then Kirchhoff migrated. Note the change in the shape of
the wavelet after migration in (b). This distortion is removed by using a rho
filter.

a)

b)

Figure 4.20 The input zero-phase wavelet a) and b) the wavelet after Kirchhoff
migration.

Page 4.31

Chapter 4 Poststack Migration ( Zero Offset)

4.5.6.2 Kirchhoff integral solutions to the wave equation


Early 2-D diffraction stack migrations corrected the phase distortion with a 45degree phase shift. When the Kirchhoff integral solution to the wave equation
was recognized, the amplitude and phase corrections became apparent. The 3-D
and 2-D Kirchhoff integral equations as given by Gazdag [102] are

p3 D = ( x1 , y1 , z1 , t = 0 ) =

p2 D = ( x1 , z1 , t = 0 ) =

cos
p x, y, z = 0, t =
2 t

r
dxdy
c

(4.13)

r
cos 2
p x, z = 0, t = dx
c
2 rc t

(4.14)

These equations indicate the input data need to be differentiated for 3D data and
a square-root derivative operator needs to be applied to 2D data. These
equations were simplified by Silva [251] giving equations (4.10) and (4.11) that
move the differential operators outside the integral and are performed on the
output data after the summation. These equations use the more familiar terms
used in these notes, i.e.,

p3 D

1
= ( x1 , y1 , z1 , t = 0 ) =
t V 2T0
1

p2 D

T0
T p ( x, y, z = 0, t = T ) dxdy (4.15)

2
1
T0
= ( x1 , z1 , t = 0 ) =
t V T 0 T

p ( x, z = 0, t = T ) dx

(4.16)

In these equations, T is the time on the input section, and T0 is the time at the
apex of the diffraction and is equivalent to the depth of the scatterpoint at Z0 on
the migrated section.
Equation (4.16) defines the 2D data with three dimensional coordinates, but only
two are actually used. The input time section p(x, z=0, t) is defined at the surface
z = 0. The output migrated depth section p(x, z, t=0) is defined with time t = 0.
These surfaces are illustrated as the surfaces in the volume (x, z, t) as illustrated
Figure 4.21, that was previously viewed in the modelling section.
Equation (4.16) also defines the shape of the diffraction with a simple description
of t = T, where T should be defined as a function of x. For a time migration, T
could have been written in hyperbolic form as T = T02 + 4 ( x x1 ) / v 2 . The depth
2

of the migrated sample is defined by the average velocity Vave or z 0 = Vave T0 .

Page 4.32

Chapter 4 Poststack Migration ( Zero Offset)

Seismic data
p(x, t) z = 0

Geologic data
p(x1, z2) t = 0

Figure 4.21 Zero offset and migrated sections as surfaces on 3D volume (x, z, t)
to aid in the description of equation (4.16).

In words, the input samples p(x, z=0, T), of the stacked section, are weighted and
summed into the migrated sample at p(x1, z1, t=0) of the migrated section. After
the summation, additional scaling and filtering may be applied.

The square-root derivative is referred to as the Rho or rj filter.


The migration process may introduce high frequency noise due to aliasing and
interpolation. The amplitudes of these high frequencies may be enhanced with
the application of the rho filter.
Consequently, it is my opinion that the rho filter should be applied to the input
data rather than the output data.
Another reason for applying the rho filter to the input data is the initial integral
definitions in equations (4.13) and (4.14) requires it that way. Migration does
move the energy to different frequencies, and applying the rho filter after
migration may introduce slight errors. These errors may be insignificant at the
lower dip energy and may be less than those introduce with approximations to
the rho filter

Page 4.33

Chapter 4 Poststack Migration ( Zero Offset)

4.5.6.3 Rho filter


Equations 4.8 through 4.11 indicate that Kirchhoff migration requires a differential
type operator or filter be applied to the input data. However, this filter is often
applied after the migration summing.
The differential filter for 3D data can be applied with a simple convolutional
operator that can be defined as ( 1, -1) or -1, 0 , 1). The square-root derivative or
rho filter, required for 2D migration, is more complex. But what is the squareroot of an operator?
Basically it means that applying a square-root operator twice, produces the same
result as applying the original operator once.. It is easily described in the
frequency domain for the operator P(f) and the square-root operator PS(f) where

PS ( f ) PS ( f ) = P ( f )

(4.17)

Multiplication in the frequency domain implies convolution in the time domain


where the operators p(t) and ps(t) imply

ps ( t ) ps ( t ) = p ( t )

(4.18)

Finding the time domain convolutional operators can be quite difficult. However,
the Rho filter is easily defined in the frequency domain as the derivative is
defined as j . This is a filter the weights the data proportional to the frequency
and applies a 90 degree phase shift, i.e. 90 o . The square-root of this operator
is 4 5 o , and is straight forward to implement in the frequency domain.
1

If applied in the time domain, it is desirable to have a short operator that is fast to
compute. An old rule of thumb indicated that a time domain operator need to be
less than 30 samples to be faster than an equivalent frequency domain process.
One filter design technique uses the inverse Fourier transform of the frequency
domain operator, then the time domain result is windowed with a cosine or raised
cosine window. Accurate operators for the first and second derivatives can be
achieved with 3, 5, or 7 pint operators, but the Rho filter requires correction for a
considerable low frequency distortion.
A practical asymmetrical correlation Rho filter is shown in Figure 4.22 with the
samples listed below. Note that this operator has been reversed to be applied as
a correlation process and does not require the time reversing in a convolution
process. The operator has 21 samples, but is asymmetrical with time zero at the
16th sample highlighted in red. There are 15 negative time samples and 5 positive
time samples.
The results of applying this Rho filter is illustrated in Figure 4.23, which shows in
(a) the original input data, then (b) after Kirchhoff migration, and (c) after
application of the Rho filter.

Page 4.34

Chapter 4 Poststack Migration ( Zero Offset)

-15 to +5 rjw operator


1

Amplitude

0.5

-0.5

-1
-15

-10

-5

Figure 4.22 Practical Rho correlation filter with samples from -15 to +5.
Sample point of correlation operator for Figure 4.22.
-0.0010 -0.0030 -0.0066 -0.0085 -0.0060 -0.0083 -0.0107
0.0103 -0.0194 -0.0221 -0.0705 0.0395 -0.2161 -0.3831
0.5451
0.4775 -0.1570 0.0130 0.0321 -0.0129

a)

b)

-0.0164

c)

Figure 4.23 A few wavelets from a) an input section, b) the 2D Kirchhoff migrated
output, and c) after application of the above Rho filter.
Modelling with diffractions also produces a change in the wavelet of a reflection
similar to, but opposite in time, to that of migration. A reversal of the above filter
will correct the diffraction modelled wavelet.
It may be interesting to note that modelling with diffractions, and Kirchhoff
migration (with no Rho filter) will cancel the phase effects, but not the frequency
scaling.
The rho filter produces a constant phase shift for all frequencies, which will be
the same if applied to the input or output data. The frequency dependent
amplitude scaling will be different when applied before or after the summing, but
may not be significant if deconvolution is applied after the migration.

Page 4.35

Chapter 4 Poststack Migration ( Zero Offset)

4.5.6.4 Efficient amplitude scaling


Dellinger et al. [717] showed the amplitude scaling could be divided into two parts
where
1. the amplitudes involving T may be applied before summation and
2. the amplitudes involving T0 after summation.
This procedure eliminates all scaling during the summation process and
significantly reducing runtimes.
A modified form of Silvas equation becomes

2 T0 p x, t = ...
p2 D ( x1 , z1 ) =
3

dx

3
t V
T

Page 4.36

(4.19)

Chapter 4 Poststack Migration ( Zero Offset)

4.5.6.5 Comments
The integral solution is basically a constant velocity assumption that indicates
what should be done in practical applications.

The RMS velocity allows us to assume the velocity is constant at the


location of the migrated sample, and that it can vary at each migrated
location. The diffraction shaped traveltimes T may be estimated from the
hyperbolic equation, or from more advanced equations that incorporate
anisotropic parameters.
Amplitudes are estimated using the above
equations using T and T0. This result is a Kirchhoff time migration.

Depth migrations assume that T in the above equations is estimated from


either ray tracing or traveltime mapping. The amplitudes may be estimated
from the spreading between bundled raypaths that emanate from the
source location and travel to the area of the scatterpoint, (3 for 3-D and 2
for 2-D).
Amplitudes may also be estimated analytically using the
transport equation, that is obtained from the wave equation.

The Kirchhoff integral solutions assume continuous data with no physical


boundaries. Truncated data and migration apertures produce migration
artifacts.

Note that the antialiasing filter (AAF) requirement is not represented in the
Kirchhoff integral equations.

References for the amplitude and phase correction filters (post- and prestack)
may be found in Bleistein et al. [81], Berryhill [6], Carter [69], Docherty, [121],
Schleicher [746],Tygel et al. [736 - 740], and Hood [77]. An excellent discussion is
also found in pages 257-260 of Yilmaz [83]. See also Gray [747] and the other
papers in the August 1999 The Leading Edge.

Page 4.37

Chapter 4 Poststack Migration ( Zero Offset)

4.5.7

Kirchhoff migration pseudo code

4.5.7.1 Diffraction summation


1
1

-N

(i)

n
x

N
j

1
1

Migrated data
Migdata

Input data
Indata

Figure 4.24 Illustration of basic Kirchhoff migration, summing over a diffraction.


Assume a trace interval dx, a time sample interval dt and fixed aperture of -N to
+N traces
The following algorithm sums over the diffraction.
Loop migrated trace

for i = 1 : I

Loop sample on the migrated trace

V = Vel(i,j)

Get time of diffraction apex

T0 = j*dt

Loop migration aperture trace

for n = -N : +N

Sum = 0

Define relative offset of trace

x = n*dx

Define time on the trace

T = sqrt( T02 + (4x2/V2 )

Check aperture

if 2x/TV < cos(betamax)

Define time sample number

m = T/dt

Define diffraction trace location

k = i + n

Check bounds of the input data

if m < M + k > 0 + k <= K

Cosine weighting, plus others

W = T0/T

Sum data

Sum

= Sum + W*Indata(m,k)

End

end

Save migrated sample

Mig(j, i) = Sum

End
End

for j = 1 : J

Get the velocity at that location

end
end

Page 4.38

Chapter 4 Poststack Migration ( Zero Offset)

4.5.7.2 Trace by trace summation


The previous algorithm summed over each diffraction and may be inefficient as
the search for input samples may extend over large areas of computer memory.
A more efficient approach is to change the order of the loops to sum all the data
from one input trace into the migrated trace, and repeat the process for all input
traces. The migration aperture becomes all input traces (1 to K).
(i)

1
1

x
To

Migrated data
Migdata

Input data
Indata
M

Figure 4.25 Illustration of basic Kirchhoff migration, migrating each input trace to
a migrated trace.
Loop migrated trace

for i = 1 : I

Migrated trace geometry

xm = i*dx

Loop input traces

for k = 1 : K

Input trace geometry

xin =k*dx

Migration geometry

x = xm-xin

Loop migrated samples

for j = 1 : J

Get the velocity

V = Vel(i,j)

Define migrated sample

T0 = j*dt

or z0

Define time on input trace

T =

Check migration aperture

if 2x/TV < cos(betamax)

Define time sample number

m = T/dt

Check bounds of the input data

if m < M

Cosine weighting, plus others

W = T0/T

Sum data to migrated trace

Mig(j,i) = Mig(j,i) +

sqrt(T02+(4*xap/V)2)

W*Indata(m,k)
End
End
End

end
end
end

This algorithm is quite general and in the prestack routines can take any
arrangement of input data, i.e., source gathers, CMP gathers, or constant offset
sections.

Page 4.39

Chapter 4 Poststack Migration ( Zero Offset)

4.5.7.3 First loop over the input trace


The order of the loops could also be modified to loop over the input data first.
Rather than summing the input data to one migrated trace, we will now sum the
one input trace into all the neighboring migrated traces. When the velocity is
constant, one input sample is spread along a semi-circle to the neighboring
migrated traces.
1

(i)

T0

j
Input data
Indata

Migrated data
Migdata

T
J

Figure 4.26 Illustration when the first loop of the algorithm is over the input data.
Loop input traces

for k = 1 : K

Input trace geometry

xin =k*dx

Loop migrated trace

for i = 1 : I

Migrated trace geometry

xm = i*dx,

Migration geometry

xap=abs(xm-xin)

Check migration aperture

if xap < aperture

Loop migrated samples

for j = 1 : J

vel =

Get the velocity

V = Vel(i,j)

Define migrated sample

T0 = j*dt

or z0

Define time on input trace

T =

Check migration aperture

if 2x/TV < cos(betamax)

Define time sample number

m = T/dt

Check bounds of the input data

if m < M

Cosine weighting, plus others

W = T0/T

Sum data to migrated trace

Mig(j,i) = Mig(j,i) +

sqrt(T02+(4*xap/V)2)

W*Indata(m,k)
End
End
End

end
end
end

The velocity must be defined at the migrated trace at time T0.

Page 4.40

Chapter 4 Poststack Migration ( Zero Offset)

4.5.7.4 Comments on the code


1. In each of the three Kirchhoff examples, we compute the time T using the
time T0. This is very significant as the velocity is defined at time T0 and not
time T.
2. In the first version of the code, we sum all the data into Sum then insert
this value into the migrated sample. We could carry with this sum a count
for the fold to balance areas with poor subsurface coverage.
3. The last two versions loop through the migrated locations and sum into it
the input values. The total number of samples is not known until he last
input values has been added, therefore another array may be required to
keep track of the fold in each migrated sample.
4. The code in all three examples does not include a specified dip limit, but
could be incorporated into the aperture. A practical dip limit should have a
taper, say from 60 to 65 degrees to limit artifacts.
5. The migrated time value T0 is defined at a specific location with no error,
however the input sample value T is compute from as a real number that
will lie somewhere between two input samples, say Tn and Tn+1. An
accurate migration may require interpolating an input value at T that lies
between Tn and Tn+1.
6. When an interpolator is not used, a broad bandwidth noise is added to the
migrated section that is known as quantization noise.
The high
frequencies of this noise may be amplified by the rho filter that may be
applied after the migration summing. (The rho filter should, in theory be
applied to the input data, but is usually applied after the migration as it is
more efficient.)
7. An antialiasing filter should be applied locally to the input data to obtain
the input sample value at T that is summed into the migrated value.

Page 4.41

Chapter 4 Poststack Migration ( Zero Offset)

4.5.8 Kirchhoff time migration


4.5.8.1 When time migration is used.

The velocities of a structure vary smoothly.

Depth migration is not really required.

As a starting point in defining a structured model for depth migration.

The structure is too complex for depth migration.

4.5.8.2 Defining the diffraction shape


The time of the diffraction shape is computed using the Pythagoras theorem

4x 2
T (x) = T + 2
Vrms ( T0 )
2

2
0

(4.20)

where the velocity is defined at T0 for all values of T, and x is the migration offset.
This equation implies that the raypaths on a time section are linear as illustrated
in Figure 4.27. Higher order equations may be used that include anisotropic
effects.
The RMS velocity is defined at the scatterpoint or migration point at time T0. Note
that this is RMS and not the stacking velocity. The stacking velocity should only
be used for stacking and not migration. In areas with dipping reflections the
stacking velocities must be converted to RMS velocities.
We do assume that the RMS velocity is locally constant for that scatterpoint, even
though it may be different for the surrounding scatterpoints. This implies that all
the raypaths, to and from a scatterpoint, travel through the (x, t) medium with a
fixed velocity defined at the scatterpoint. A neighboring scatterpoint may have a
different RMS velocity, but the raypaths will travel through virtually the same
medium. This concept enables geology with laterally varying velocities to
produce reasonably accurate time migrations, especially when an accurate
velocity model cannot be created.
In highly structure areas, it may not be possible to build an interval velocity
model that is accurate enough for a depth migration. In these cases, only an
approximate RMS velocity field may be defined to produce a migrated section.
A migrated time section may be used as a starting point for building a depth
migration velocity model.

Page 4.42

Chapter 4 Poststack Migration ( Zero Offset)

x
t

T0

vrms

a)
x
t

vrms

T0

Figure 4.27 Traveltimes computed using a) a bundle of raypaths defined with


Vrms, and b) the corresponding diffraction defined by the traveltimes at the
surface of (a).

Lateral varying velocities above a migration point may (will) tilt the diffraction.
This tilted can be estimated to refine a time migration. Note that the specular
energy will reside in a small area of the diffraction that is defined by the Fresnel
zone. This specular reflection energy tends to be hyperbolic, implying that a
combination of time and depth migration may produce an optimum result.
Limiting the migration aperture to an order of the Fresnel zone becomes a model
based migration that will reduce significant noise, but errors in the model may
tend to be inappropriately validated.

Page 4.43

Chapter 4 Poststack Migration ( Zero Offset)

4.5.8.3 Lateral positioning distortions in time migrations


The previous material has assumed the diffraction shape to be hyperbolic and
based on the RMS velocities.
Structurally complex geology may produce diffractions that are not centered at
the location of the scatterpoint as illustrated in Figure 4.28. However, time
migrations place the summed energy at the apex of the diffraction producing a
lateral positioning error. This is to be expected when using a time migration.
Depth migrations that compute the correct shape of a diffraction should correct
this positioning error. However, the depth migration must include anisotropy
when estimating traveltimes for a more accurate positioning.
It is also assumed that the diffraction is continuous and single valued in time. In
the following sections with depth migrations, we may have discontinuous
diffractions and or multiple values in time. These result from choosing the first
arrival times of energy, or the arrival times of the maximum energy. Choosing the
arrival times of the maximum energy requires either

use of the transport equation to monitor amplitudes along a ray path,

tracking bundles of raypaths whose separation distance relates to the


amplitude or,

propagating wavefield energy.

When a wavefront just reached the surface, it will be tangent to the surface and
that location will have the minimum time across the surface. Since the ray is
normal to the wavefront (isotropic material) we conclude that a ray that is vertical
to the surface will have minimum traveltime. It is shown in blue in Figure 4.28,
and arrives at the surface with a vertical direction. This is an image ray and
represents the apex or peak of the corresponding diffraction. Notice that this
surface point is laterally displaced from the scatterpoint.
Time migration places the summed energy at the apex of the diffraction that is
vertically below the arrival point of the image ray. Consequently, time migrations
will have laterally positioning errors.
Use of image ray modelling will allow the correction of a time migration to a depth
migration.

Page 4.44

Chapter 4 Poststack Migration ( Zero Offset)

x
z

vlow
vhigh

a)

x
z

vlow
vhigh

b)

x
t
or
z
Peak of
diffraction

vlow
vhigh

Figure 4.28 Traveltimes computed using a) a bundle of raypaths, b) gridded


traveltimes, and c) the resulting diffraction of traveltimes recorded at the surface.

Page 4.45

Chapter 4 Poststack Migration ( Zero Offset)

4.5.8.5 Dips in a medium with vertically varying velocities


Sedimentary basins tend to have velocities that may be approximated with
velocities that vary with depth V(z) and are ideally suited for time migration.
These velocities may be approximated to be linear with depth allowing analytic
solutions for raypaths and wavefronts that are circular.
The curved rays aid seismic imaging by allowing:
steeper dips to be imaged, and
migration algorithms to migrate dips beyond their rated dip limit.
Even though the velocities vary with dip, the shape of the diffractions (and
moveout) still tend to be hyperbolic enabling accurate traveltime estimates using
the RMS velocities and raypaths on time sections to be approximated by straight
rays as illustrated below in Figure 4.29. Dips estimated using linear raypaths will
appear to be much less than the actual dip the curved raypath at the reflector.
Dips on the diffraction will actually be much steeper than indicated by the straight
ray assumption and the resulting migration will move energy to dips beyond the
range of the algorithm.
x
Angle of
raypath at
reflector

x
t
or
z

d
T

b = (q + d )

q
T0

V2 V1
V2 + V1

d
T
Geologic
dip

Angle of
raypath at
surface

a= (q - d)

Figure 4.29 Dips along the diffraction are really much greater than expected from
straight raypaths.

Linear raypaths require infinite offsets to image a 90-degree dip.

The raypath curvature enables imaging of dips that exceed 90 degrees with
reasonable offsets.

Migration algorithms that limit imaging geological dips will actually be


migrating much steeper dips, as illustrated in Figure 4.30 to Figure 4.32Error!
Reference source not found..

Hyperbolic assumptions may extend to geological dips beyond 65 degrees.

AAFs should be designed with input data dips a, not the geological dips b.

Page 4.46

Ch
hapter 4 Poststtack Migration ( Zero Offset)

a)

b)

Figure 4.30 Exam


mples of diffraction curves
c
a) with
w
no ray bending and
a
b) with
h ray
bending
g using Gulf of Mexic
co velocitie
es as given
n in Larnerr [94].

Figurre 4.31. Co
omparison
n of linear raypath
r
dip
ps with curved raypa
ath dips in a
v
velocity
me
edium thatt increases
s linearly with
w
depth from
f
Dona
ati [559].
Geologiical structures wherre the velo
ocity decre
eases with depth hav
ve the rev
verse
effect. An examp
ple is carbo
onates ove
er shales or
o ground penetratin
ng radar (G
GPR)
where the radar velocities decrease
d
w
with
depth due
d to the moisture content.
c

P
Page
4.47

Ch
hapter 4 Poststtack Migration ( Zero Offset)

a)

b)

c)

d)

Figure 4.32 Turn


ning ray migration
m
o a spherre in a line
of
ear velocitty gradien
nt, a)
shows the circle reflector and
a
circula
ar raypath
hs of the depth
d
model, b) the time
section, c) the migration
m
u
using
hype
erbolic, an
nd d) an 8th order Kirchhofff MO
equation from Ma
artin [687].

An exam
mple of tu
urning ray migration
n using the phase-shift metho
od is given
n by
Hussein
n [741].

P
Page
4.48

Chapter 4 Poststack Migration ( Zero Offset)

4.5.9 A fast Kirchhoff pseudo depth migration


Time migration does not position the migrated data correctly when there are
lateral velocity variations. The method presented below compensates for the
lateral velocity variation and uses a time migration to produce a migrated section
that approximates a depth migration. Consider Figure 4.33a.

Depth z is defined at a time Tz0 at convenient place on the section.

The time Tz of this depth is computed for each trace.

The time-shift Tz0 - Tz is applies to the hyperbolic times to modify the shape of
the diffraction.

This method is efficient and quite accurate when the velocity changes are
smooth.

Results of this method tend toward image ray corrected time sections by
tilting the diffraction.

a)

Tz - Tz0

hyperbola
modified diffraction

b)
Figure 4.33 Lateral velocity correction by a) constant depth compensation and b)
the time correction on the hyperbola.

Page 4.49

Chapter 4 Poststack Migration ( Zero Offset)

4.5.10 Kirchhoff depth migration.


Kirchhoff depth migration computes the shape of the diffraction from a depth
model using either

ray tracing techniques, or

using wave front travel-times on a grid.

An early method of defining the diffraction shape was to compute the traveltimes
from scatterpoint to the surface as illustrated by the rays in Figure 4.34a. This
method required a family of rays (or gridded traveltimes) to be computed at every
migrated sample on every migrated trace. N = Ntr x Nsamp
Then came a big change, or a legitimate use if the term a paradigm shift.
It was recognized that travel-times should be pre-computed from each surface
location to all subsurface locations, and stored in memory. This is illustrated in
Figure 4.34b where the traveltimes are computed by ray tracing or wavefront
propagation and entered into a grid. The travel-times along a diffraction, could
then be found by looking up the corresponding traveltime at the scatterpoint on
each of the pre-computed travel time panels, as illustrated in Figure 4.34c. The
significantly reduced the number of traveltime computations to only the number
of surface locations. N = Ntr
Fermats principle indicated that a raypath is defined by the minimum (or
maximum) traveltime between two locations. Consequently traveltime maps are
normally computed using minimum traveltimes. However, it is possible to have
more than one raypath between two locations, each with a local minimum (or
maximum) traveltime in geologically complex areas.
Consequently, some
Kirchhoff migration algorithms do allow for multiple arrivals. (In contrast,
downward continuation algorithms unravel the complex raypaths (wavefronts)
and automatically include all possible raypaths).

For raypath computation see Zelt [386] and Moser [139].


For wave front modelling see Vidale [136] - [138], Reshef [109], Moser [139],
Notfors [141], and Fischer [217].

Page 4.50

Chapter 4 Poststack Migration ( Zero Offset)

x
z

vlow
vhigh

a)

x
z
Ray

Wavefronts

b)

Ray

Ray
Ray

c)
Figure 4.34 Traveltimes used in Kirchhoff depth migration a) shows old method
of rays from a migration point, b) traveltimes estimated using ray and wavefronts
from a surface location to all subsurface scatterpoints, and c) three additional
panels of traveltimes emanating from other surface locations.

Page 4.51

Chapter 4 Poststack Migration ( Zero Offset)

4.5.10.1 Multivalued diffractions


It is commonly assumed that the traveltimes on the diffraction are single valued
in time for each spatial location, i.e. a smooth curve with a hyperbolic type shape.
An example from Chapter 2 is repeated in Figure 4.35 this point in a constant
velocity medium.
That may not be the case in depth migrations where there may be multiple values
of traveltime for the same displacement x. Usually there are three values that
produce a caustic curve in a diffraction. The amplitudes and phase in these
areas are complex and are not evaluated accurately in a Kirchhoff migration.
However, in a wave propagation migration, such as downward continuation or
reverse time, these caustics are gradually removed and eventually eliminated
when the wavefield arrives at the reflection point. This is the main difference
between a Kirchhoff and wave propagation migration.

Figure 4.35 Traveltimes used in Kirchhoff depth migration a) shows old method
of rays from a migration point, b) traveltimes estimated using ray and wavefronts
from a surface location to all subsurface scatterpoints, and c) three additional
panels of traveltimes emanating from other surface locations.
Constant velocity model of an exploding scatterpoint in a) showing the expanding
wavefields in time lapse photos, and b) a continuum of wavefields that produce
a hyperbolic diffraction on the surface.
A complex wavefield reaching a subsurface layer is illustrated in Figure 4.36. A
similar wavefield from a scatter point to the surface would represent a complex
diffraction arriving at the surface.

Page 4.52

Chapter 4 Poststack Migration ( Zero Offset)

z = ZA

a)
x

z = Za
z = Za

b)
Figure 4.36 Wavefield propagation in a) an inhomogeneous medium (Marmousi),
and b) the wavefield at the surface showing a complex first arrival time and
caustics.

Page 4.53

Chapter 4 Poststack Migration ( Zero Offset)

4.5.11 Kirchhoff migration examples of mod50.


A Kirchhoff migration was applied to the numerical mod50.sgy. The software
contained a number of anti-aliasing algorithms that can be turned on and off at
any desired location (a feature unique to Kirchhoff migration.)
In the first example below, only a small window of data was migrated; i.e., traces
65 - 85 and times 500 to 900msec. This is referred to as a porthole. All the
input data was used for migrating data into the porthole. Note the energy of the
diffraction has collapsed to a well defined spike.
The time to migrate the data was imperceivable. A more sophisticated algorithm
would allow the porthole to be interactive with a slider to modify the velocities,
and or other parameters, to focus the data in the porthole.
The second example Figure 4.38 a was migrated over the first half of the section
with no AAF, while the second half used a sinx/x interpolating AAF. The first half
took 1 second using a standard PC, while the second half took 12 s. The FK plot
of (a) is shown in (b). Note the difference in the aliased energy for the dipping
events.
Why are there two aliased events on the right side of Figure 4.38b???
The horizontal noise on the left side above the horizontal reflector in (a) is caused
by operator aliasing.
The aliased steep dip on the left side has higher amplitude and frequency due to
the correct migration of aliased energy.

Porthole

Figure 4.37 Example of a Porthole Kirchhoff migration.

Page 4.54

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)
Figure 4.38 A Kirchhoff migration in a) with an antialiasing filter only applied to
second half of the section, and b) the FK amplitude spectrum.

Page 4.55

Chapter 4 Poststack Migration ( Zero Offset)

4.5.12 Kirchhoff migration exercise.


Use the family of constant velocity diffraction curves in Figure 4.39 to migrate a
linear reflection (b) and a single energy point (c). The objective is to trace energy
that aligns with the diffractions, then move that energy to a location defined by
the apex of the diffractions. Repeating this procedure for all possible diffractions
will construct the migrated image.

Place the diffraction curve page behind the page containing Figure 4.40a.

Align zero time of part (a) with zero time on the diffractions. Slide the page to
the left or right until a diffraction becomes tangent to the dipping reflection.
Mark on the top sheet the location of the apex of that diffraction.

Repeat for all possible diffractions.

Align the diffractions behind the scatterpoint of (b) to migrate a single point of
energy. Ensure that the zero times are aligned.

The energy picked up by the diffractions should be relocated at the apex of


the corresponding diffraction.

Do the results meet your expectations?


Tangential energy picked is

specula energy.

IF only the specula energy is migrated, why sum over the entire diffraction?

Figure 4.39 Family of diffraction curves for exercise.

-4

-3

-2

-1

1
-0.5
-1
-1.5
-2
-2.5
-3
-3.5
-4

Page 4.56

Chapter 4 Poststack Migration ( Zero Offset)

z
or
t

a)
x

z
or
t

b)
Figure 4.40 Diffraction curves to migrate data, a) a straight line, and b) a single
point.
Note that the area of tangency between the dipping reflection and the diffraction
increases with increasing time. The amplitude of the energy summed within this
area would also increase with depth, indicating a need for the obliquity term or
the T0/T amplitude scaling factor.

Page 4.57

Chapter 4 Poststack Migration ( Zero Offset)

Blank page to aid in tracing of previous exercise.

Page 4.58

Chapter 4 Poststack Migration ( Zero Offset)

4.5.13 Summary of Kirchhoff migration

Hagedoorn migration spreads energy on a semicircle.

Kirchhoff migration sums energy along a diffraction.

(Modelling spreads energy on a diffraction.

Modelling sums energy on a semi-circle).

A time migration assumes the diffraction is hyperbolic.

A depth migration computes raypath or wavefront traveltimes to estimate the


shape of the diffraction.

The lateral extent of the area summed along the diffraction path is limited by a
time-varying aperture; usually defined by a maximum dip limit.

The spatial size of the time-varying aperture is usually controlled by dip limits.

Anti aliasing filters are required to prevent the operator from aliasing. (Even
when the input data is not aliased, e.g. a horizontal event, spurious energy
may result if an AAF is not used: see Figure 4.38a.)

Special shaping filters are required by Kirchhoff migration. (rjw, rho filter, etc)

In a constant velocity migration, the kinematics of distributing energy along a


semi-circle and the summation along a hyperbola are equivalent.
In variable velocity media, the distribution is no longer a semi-circle, but is
distorted on the time section. However, the summation diffraction may still be
approximated by a hyperbola.
Kirchhoff time migration uses one velocity to approximate the diffraction shape.

Page 4.59

Chapter 4 Poststack Migration ( Zero Offset)

4.6 FK direct Fourier transform migration


4.6.1 Introduction
The direct Fourier Transform migration is a very fast method and was introduced
by Stolt in 1978 [21]. The term FK is derived from the Fourier transform of time to
frequency F, and distance to wave-number K.
The FK method is ideal when the velocities are constant and will migrate
accurately to 90 degrees.
The FK domain loses the identity of the x and z positions of the data, and can
only migrate a section with an apparent constant velocity. Variable velocities
may be accommodated with a pseudo time-to-depth conversion that is performed
on the input time section. (Strictly speaking, the input section remains a time
section with constant velocity.)
The typical procedure for the migration is

Time to pseudo-depth conversion (constant velocity time section).

2-D Fourier transform.

Migration in FK to KK space (kernel).

Inverse 2-D transform.

Pseudo-depth to time conversion.


x
t
Input time

kx

x
Time
to
depth

Constant vel.
time section
(Pseudo depth)

t
(or z)

2D-FFT

f
FK space

Vsmooth
Migration
x

Migrated time

kx

x
t Depth
to
time

Migrated
pseudo
depth

z
2D-IFFT

kz
KK space

Vsmooth

Figure 4.41 Block diagram of FK or KK migration flow diagram.

Page 4.60

Chapter 4 Poststack Migration ( Zero Offset)

4.6.2 Time to pseudo depth conversion


In order to take advantage of the speed of FK migration, the input time section
was converted to a pseudo depth section by vertical time stretching with a
smoothed velocity field. After stretching the input section should still be
considered a time section with a constant velocity.
The velocities must be smoothed to maintain the shape of the diffractions.
Early attempts to achieve smooth velocities were to use 90% of RMS velocities,
apply only one velocity function to the data, or to apply a large mix to the
velocities.
Yilmaz [ 83 ] and Claerbout [294] have defined a velocity smoother that is similar
to the "interval to RMS" conversion. The input to the equation is the RMS
velocities and output is the smoothed pseudo depth velocities, i.e.,
2
Vmig
(T ) =

1
T

2
Vrms
( t ) dt

(4.21)

Figure 4.42 illustrates in (a) the input structure with a lateral velocity change, (b)
the time response, and (c) the pseudo depth response. The pseudo depth
stretching tends to correct the tilt on the diffraction to produce a depth migration.
These images are only sketches, but the actual outputs may be seen in the next
chapter.
The success of pseudo depth conversion for FK to produce a near depth
migration lead some contractors to use the time to depth stretch prior to other
migration methods. These were also called and sold as depth migrations.
It must be emphasized that the migrated output after the inverse 2-D FFT should
not to be considered a depth migration, as the velocities used were not
necessarily the correct velocities. The depth to time conversion is required to
produce migrated section.
The time to pseudo depth conversion has an effect similar to the thin lens term of
wave equation solutions.
Image ray conversion should not be used on FK migrated sections as they tend
to have lateral velocity correction due to the pseudo depth conversion. A more
appropriate method would be vertical stretching using average velocities.

Page 4.61

Chapter 4 Poststack Migration ( Zero Offset)

Vertical time to pseudo depth conversion is illustrated in the following figure.

a)

b)

c)
Figure 4.42 Time-to-depth illustrated with a) a structure with a lateral velocity
change, b) the time model, and c) the vertically stretched pseudo depth.

Page 4.62

Chapter 4 Poststack Migration ( Zero Offset)

While the time to depth conversion has the benefits of approximating a depth
migration, caution must be used with the positioning of steeper dips. The pseudo
time to depth conversion will distort the shape of diffractions as illustrated below
in Figure 4.43. In this figure

Diffractions with a constant velocity do not overlap (a) and tend to one
asymptote.

Diffractions with increasing velocity overlap (b).

Time-to-depth conversion is to make (b) appear as (a), an impossible task.

Pseudo depth conversion stretches the lower part of the diffraction causing
non-hyperbolic distortion.

Note however, that Snell's law will still help to migrate steeper dips in areas
with increasing velocities.

a)

b)

c)
Figure 4.43 Time-to-depth of diffractions a) the ideal constant velocity case, b)
the increasing velocity case, and c) the time-to-depth stretched of (b).

Page 4.63

Chapter 4 Poststack Migration ( Zero Offset)

4.6.3 FK time migration with large vertical velocity changes


The basic FK migration algorithm may be used to achieve excellent time
migrations in areas with large vertical velocity variations with the following:

Use the same input FK section (no stretching).

Migrate the input section many times with different constant velocities, but
only output the (x, t1-t2) portions that are valid for the migration velocity as
illustrated in Figure 4.44.

Reconstructed the final migrated section from the constant velocity sections
according to the velocity structure.

The result is an accurate time migration as the diffractions are unaffected by the
depth stretch distortion.
The method is quite fast as there is no depth stretch, and the input 2-D FFT is
only required once.
This method is a true time migration and can use image rays for conversion to a
migrated depth section.

a)

b)

c)

Figure 4.44 Accurate FK migration in areas with vertical velocity variations.

A comparison between normal FK and the above method may be seen in Figures
5.60 and 5.61.
This method is not the same as prestack Stolt migration.
Another method by Mikulich and Hale [157] also uses multiple FK's for a depth
migration with steep dips.

Page 4.64

Chapter 4 Poststack Migration ( Zero Offset)

4.6.4 The kernel of FK migration


Dips on the input section are also identified as dips on the FK section. Figure
4.45 illustrates a simple procedure that allows in the FK space of (b) to be moved
to the new migrated dip on the KK section (c). These dips obey tan = sin .

a)

b)

d)

c)

Figure 4.45 The FK migration kernel showing a) a depth section, b) the 2-D
transform in KK space, c) the migrated KK space, and d) the migrated depth
section.

???
??? How are the data moved from dips to in figures (b) and (c) ???
???

Page 4.65

Chapter 4 Poststack Migration ( Zero Offset)

The data movement in KK space may be illustrated by Figure 4.46 .

Recall that dips before and after migration hinge at the surface.

Note the dip spacing x in (a) and (d), and the related Kx, in (b) and (c).

The different dip spacing for t in (a) and z in (c), and Kz and Km.

Migration moves an FK point vertically to a new KK position.

kx=1/x

a)

b)

kx=1/x

kx
kz

kz

d)

c)

Figure 4.46 Illustration of movement on the FK and KK plane with a) showing a


dip model, b) the FK transform, c) the KK migrated data, and d) the resulting
migration on the (x, z) plane.
The actual migration position is quantified using Figure 4.47.

tan =

delz
k
( x,z ) = x = sin
delx
k

(4.22)

therefore,

k =
2
z

2
v

k x2 = k 2 k x2

Page 4.66

(4.23)

Chapter 4 Poststack Migration ( Zero Offset)

Figure 4.47 Migration in the KK plane. The input data must be interpolated before
it is moved to a known migrated sample location.
Data at the 45-degree line will move to the top at the 90-degree position.
The migrated position kz is defined on a grid, with data at k is computed and
interpolated from the complex samples around k. Interpolation of this complex
value may be difficult. Real linear or sinx/x interpolation applies to real data, but
now the interpolator is complex, and requires a linear phase shift (equivalent to
half the trace length). Padding extra time samples may aid in the interpolation.
Additional help in complex interpolation may be found in page 275 of Claerbout
[294].
The amplitude of the migrated sample is scaled by the term Amp

Amp =

kz
k

??? What happens to data above 45 degrees???

Page 4.67

(4.24)

Chapter 4 Poststack Migration ( Zero Offset)

4.6.5 Noise removal and retention


On the FK plane, data above the 45-degree line is noise. The removal of all this
noise may make a migrated section appear too smooth or mixed. To some
interpreters it appears unnatural. The migrated section may be created to appear
more "natural" by leaving some noise on the migrated KK space. Noise close to
the kx Nyquist axis may also be used for noise removal and retention.

a)

b)

c)

d)

Figure 4.48 Plots showing the movement of data on the KK plane. Part a) shows
an input diffraction on (x, t) plane, b) the KK plane, c) the migrated plane with all
noise removed, and d) a partially migrated KK plane retaining noise.

Page 4.68

Chapter 4 Poststack Migration ( Zero Offset)

4.6.6 Choosing parameters


Some parameters that are used to control the "appearance" of the migrated
section are the dip limit , %k, and %F as illustrated in Figure 4.49a. These
parameters should be tapered at the edges of the migrated data. Various
amounts of noise in Figure 4.49 b-d may be added back to modify the appearance
of the migrated section.

a)

b)

c)

d)

Figure 4.49 Parameter definition for FK migration with a) showing percentage


values, b) the dip noise, c) the Kx Nyquist noise, and d) high frequency noise.

Page 4.69

Chapter 4 Poststack Migration ( Zero Offset)

4.6.7 Padding traces and trace samples


Most Fourier Transform programs require special numbers of the form n=2m
such as 4, 8, 16, ... 1024, etc. To meet these requirements, traces and samples
must be added to the input section before the 2-D transform.
In addition other factors that may affect the number of traces and samples are:
1. Additional traces may be required to prevent data from wrapping around from
side to side, or top to bottom.
2. Additional samples are often added to the time traces to aid the interpolation
of the complex samples. The additional time on the input trace gives a finer
sampling in the frequency domain aiding the complex interpolation.

Some FFT's use numbers that are not powers of two. These algorithms usually
require more run time and the savings with a fewer number of samples is
questionable. The saving of memory however, may be of value.

Page 4.70

Chapter 4 Poststack Migration ( Zero Offset)

4.6.8 Examples of FK migration


The following pages illustrate FK migration by using a diffraction and a single
dipping event. The dip angle and frequency content are varied to illustrate
different effects of the method.
The first set of displays in Figure 4.50 illustrate:

shallow dip,

no aliasing.

Figure 4.51 shows an increase in frequency content of the wavelet and the
dipping event may be seen to reach the aliasing limit in (b). The slight distortion
that results from not migrating the central band is evident in (c). Very slight noise
may be seen around the spike in (d).

The dip has been increased in Figure 4.52 such that half the dipping event is now
aliased in (b). Panel (c) illustrates the effect of migration.

Part of the unaliased dip is migrated correctly.

Central panel is still in the original position.

Aliased energy is migrated but with respect to the wrong dips.

Note the presence of the aliased noise in (d).

The final Figure of FK migration in Figure 4.53 is an example with the dip greater
than 45 degrees. This may be an example of ground noise. Note:

The dipping event in (b) may be seen to be steeper than the diffraction energy.

The extreme amount of aliasing that has also wrapped around in (b).

Migration is only able to remove the unaliased portion of the dip.

The remainder of the dipping event is aliased.

This illustrates why it is so difficult to dip filter surface type noises. Note the
remaining energy in (d).

Page 4.71

Chapter 4 Poststack Migration ( Zero Offset)

kx
w

a)

b)
x

kx
z

d)

c)

Figure 4.50 FK migrations, shallow dip, and low frequency a) input, b) FK


transform, c) migrated KK, and d) the migration.

Page 4.72

Chapter 4 Poststack Migration ( Zero Offset)

kx
t
w

a)

b)
x

kx
z

kz

d)

c)

Figure 4.51 FK migrations, shallow dip and high frequency a) input, b) FK


transform, c) migrated KK, and d) the migration.

Page 4.73

Chapter 4 Poststack Migration ( Zero Offset)

kx
t

a)

b)
x

kx
z

d)

c)

Figure 4.52 FK migrations, steep dip and high frequency a) input, b) FK


transform, c) migrated KK, and d) the migration.

Page 4.74

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)

d)

c)

Figure 4.53 FK migrations, dip over 45, and high frequency a) input, b) FK
transform, c) migrated KK, and d) the migration.

Page 4.75

Chapter 4 Poststack Migration ( Zero Offset)

4.6.9 Examples using mod50 data.


The mod50.sgy numerical model was migrated with a basic FK algorithm that
contained a number of interpolation options. The new position of the data at Kz is
defined at a sample location, but the data to be moved at K can is a computed
number and will be a floating point number that located the desired data between
sample points. Consequently, an optimum migration will need to interpolate the
complex data from the neighboring samples. This is a complicated process and
is illustrated with a few examples. The first three images are plotted with high
amplitudes to allow the artifacts to be more visible.

The first image Figure 4.54a used a integer for the address (truncated from the
floating point address). In addition, amplitude scaling was not applied. Note the
upper semi-circle contains approximately equal amplitudes.. Also note the lower
spike is migrated to many semi-circles that both smile and frown. These semicircles have shapes relative to the periodic time zeros.
The second image (b) uses a linear interpolation between the complex samples.
(Linear interpolation of the real and imaginary values.) Noise is significantly
reduced, but the frowns to the lower time zero are still present. Amplitude
scaling was applied.
The third image (c) used a special complex interpolator based on the spectral
window. Only a small amount of wrap-around energy remains.
An example of the migrated FK transform is shown in (d). This image in similar
for all the FK migrations. Note that the aliased energy is migrated to the wrong
location.
Figure 4.55 shows an FK migration was formed by padding more zeros to the
input data (Nz = 2048) Linear interpolation and amplitude scaling was also
applied as in Figure 4.39b. Note the significant reduction in the amplitude of the
frowns and that the wraparound energy is spread over a larger area.
This doubling of time in the time domain halves the sample interval in the
frequency domain, aiding in the complex interpolation that is required between
the complex FK samples. Note the improvement over the previous linear
interpolation that used Nz = 1024 samples.

Page 4.76

Chapter 4 Poststack Migration ( Zero Offset)

x
z

a)

b)
Figure 4.54 Examples of FK migrations with various interpolation parameters, a)
no interpolation and no amplitude scaling, b) linear interpolation and amplitude
scaling, c) two sample complex interpolation, and d) complex sinx/x interpolation
(N = 25) Complements Zaiming Jiang.

Page 4.77

Chapter 4 Poststack Migration ( Zero Offset)

c)

d)
Figure 4.54 continued.

Page 4.78

Chapter 4 Poststack Migration ( Zero Offset)

Figure 4.55 An FK migration similar to Figure 4.42b, but with twice the number of
time samples.

Page 4.79

Chapter 4 Poststack Migration ( Zero Offset)

4.7 Downward Continuation Migration


4.7.1 Introduction
Downward continuation migration is a very important method that enables simple
time migration, or depth migrations, with complex velocity distributions. Recall
the (x, z, t) volume in Figure 4.56 from modelling.

a)

b)

c)

Figure 4.56 The (x, z, t) volume in a) with b) showing the input structure (x, z),
and c) the resulting time section (x, t).

Page 4.80

Chapter 4 Poststack Migration ( Zero Offset)

Recall from previous chapter on modelling:

Modelling uses the depth structure (x, z, t=0 ) to compute the volume (x, z, t)

A seismic section was obtained at ( x, z=0, t).

"Downward continuation migration" is very similar to the modelling program.


1. It starts with the top surface seismic section ( x, z=0, t ).
2. Computes the entire volume ( x, z, t ).
3. Extracts the structure ( x, z, t=0 ).

Note that the output migration is in depth, the most accurate representation of the
subsurface.

This is a depth migration.

Some methods of downward continuation output portions of the depth slices


(x, z = z0, t) into a time section. This may form part of an output for a time
migration.

There are many methods of computing the different layers in the volume.

The volume may also be computed from the rear to the front, referred to as
"reverse time migration".

Page 4.81

Chapter 4 Poststack Migration ( Zero Offset)

4.7.2 Downward continuation depth migration


The following Figure 4.57 illustrates downward continuation depth migration. In
figure (a) note:

Time sections occur at different depth layers.

The t=0 edge of the time section becomes part of the depth section (arrow).

The depth increment z becomes the sampling rate on the output migrated
section.

The number of time layers to achieve the depth sampling.

Bottom of the time sections move toward t=0 in time as the depth increases.

Velocities used at each depth layer should be the interval velocities at that
depth.

The time section at each depth level would be identical to a section recorded
with the sources and receivers at that depth.

Figure 4.57b contains some sliced time sections that illustrate the data movement
of a diffraction at various depth levels. These diffraction may be represented as
slices on a cone. Note also on these figures

Diffraction moves toward the surface.

Energy of the diffraction is concentrated as it moves toward the surface.

Energy is focused at a point when it reaches the surface.

Bottom of the time section moves toward the surface at increasing depths.

The size of the depth increment z needs to be:

large enough to minimize the number of depth layers,

small enough to prevent aliasing on the output depth migrated trace.

Some processes use a z close to the Nyquist rate to propagate one depth layer
to an other, then interpolate to fill in the data between the depth layers.

Page 4.82

Chapter 4 Poststack Migration ( Zero Offset)

Migrated
structure

a)
t
x
dz
z

b)

c)

Zeros

Figure 4.57 Illustration of a) downward continuation depth migration, b) the


movement of diffraction energy towards t=0, and c) time sections at different
depth layers, with the arrow indicating the new bottom of the section. The gray
area indicates the area with no data and the red area the energy at t = 0 that is
copied to the output migration..

Page 4.83

Chapter 4 Poststack Migration ( Zero Offset)

4.7.3 Downward continuation time migration


The output for downward continuation time migration is not the t=0 edge, but a
band of time samples, wide, from the time section as illustrated in Figure 4.58.
Also note:

As the depth increases, the bottom of the time section remains at the bottom.

The top of the migrated part of the time section is decrease by time
increments with each downward layer.

The output from each downward layer is defined by the time band that is at
the th time level.

Some sliced time sections are sketched in (b) to illustrate:

The diffraction remains at the same time level.

The energy in the diffraction is focused to a point at a time that corresponds to


the top of the migrated part of the time section.

The size of the step depends upon

The accuracy required for the velocity model.

The dip limitations required by the algorithm.

An approximate downward depth increment.

A 15 degree algorithm may require a step of 50 milliseconds to achieve 15


degrees of migration, while a 65 degree algorithm may require a step of 2
milliseconds to achieve 65 degrees of dip of migration (see [106]).

The velocity for each downward layer is defined:

At that particular time level of the step.

May vary for each trace.

It is constant for the traces at this step.

Page 4.84

Chapter 4 Poststack Migration ( Zero Offset)

Migrated time section

a)
t
x
dz

T0

Saved
i

t dd t

b)

c)
Figure 4.58 Illustration of a) downward continuation time migration, b) the energy
of a diffraction tends to the apex of the diffraction and c) time sections at different
depth levels with the arrow indicating the new top for migration. The gray area
represents data in which the migration is complete.

Page 4.85

Chapter 4 Poststack Migration ( Zero Offset)

4.7.4 Comparison of downward continuation in depth and time

Vertical movement of data


a) Data moves up toward t=0 with increased depth.
b) Horizontal data remains at the same time level (dips will move).

Amount of output with each depth layer


a) One sample per trace in depth migration.
b) Band of time samples per trace in time migration.

Downward increment is defined by


a) A depth increment small enough to prevent aliasing.
b) A step depends on velocity model, algorithm, and accuracy.

A separate storage area is required for the migrated depth output

Data above the th time level remains unchanged by the migration process,
and may be used to store the migrated data.

One main difference in the downward layers of time and depth migration is the
small vertical shift that is applied to the depth migration. This vertical shift (some
times referred to as the thin lens term) is responsible for applying small relative
corrections to data in areas with lateral varying velocities. The accumulation of
these small relative corrections will adjust the tilt on diffractions to correctly
position the apex at the reflection point.
Some algorithms do not apply the whole vertical time shift, but only the small
relative corrections due to the lateral velocity change. These algorithms may be
referred to as a hybrid time migration with more accurate spatial positioning.
Image ray corrections should not be used.
Construction
The differences between time and depth migrations are illustrated in Figure 4.59
where the energy of a diffraction collapses towards T=0 in a depth migration (a),
but remains at the same time for time migration in (b). Construct new diffractions
mid-way to the focus points by dividing the distance between each point on the
diffractions to the focus points as illustrated. Compare the shape, position, and
extent of the new diffraction.

Page 4.86

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)
Figure 4.59 Construction of diffraction movement for a downward continuation
migration in (a) depth and (b) in time. Move the diffraction halfway to the image
point as illustrated in each figure.

Page 4.87

Chapter 4 Poststack Migration ( Zero Offset)

4.8 Various algorithms for downward continuation


4.8.1 Example of a 15 finite difference solution
The 15 finite difference migration is one of the early types of migration and
A more detailed
provides an example of the finite difference algorithm.
description of other solutions to the wave equation is included in Appendix 1.
The wave equation

2P 2P 1 2P
+
=
,
x2 z2 v2 t2

(4.25)

or in the transform domain,

k +k = 2 ,
x z v
2

(4.26)

is not used with the downward continuation method as it naturally creates


reflections and attenuation losses. Although this might seem desirable, it would
only be effective for prestack data, and only then if the "statics" and velocity
model were perfect, and the processing ideal [295]. The wave equation is "broken
apart" into upward moving and downward moving components that still obey
Snell's law, but are not reflected or attenuated at velocity interfaces. This new
equation is the "one-way wave equation," and is found from

kz =

2
v2

2
x

v 2 k x2
1
+ ... ,
2
v
2

(4.27)

or,

P
= i P
z
v

iv k x2
P.
2

(4.28)

The i / v term is the time shifting term that distinguishes between migrations of
depth P(x, z) and time Q(x, t). Removing this term and continuing with time
migration now defined as Q we get,

Q v ( x,z ) 2Q
=
.
z
2i x 2

(4.29)

or in standard (x, z ,t) coordinates,

2
v 2
Q=
Q,
zt
2 x2

Page 4.88

(4.30)

Chapter 4 Poststack Migration ( Zero Offset)

A finite difference solution is derived in the Appendix 1 to be,

pn+1, j = pn, j+1 + (1 + wT/ x2 ) ( pn, j + pn+1, j+1 )

(4.31)

where n is in the depth direction, and j the time direction, T is (1, -2, 1) - the
second derivative operator, and w = v*z*t/4,. (For two-way times and using a
downward times step tau (), w = v**t/16 ). Applying the T term with i in the x
direction we get the full solution,

pn +1, j, i = pn, j, i + pn +1, j +1, i pn , j +1,i +


w
,
(p
2pn, j, i + pn, j, i+1 ) + (pn +1, j +1, i-1 2pn +1, j +1, i + pn +1, j +1, i+1 )}
2 { n, j, i-1
x

(4.32)

as illustrated in Figure 4.60.


pn,j+1,i
t
x

pn,j,i-1

pn,j,i

pn,j,i+1
z

pn+1,J+1,i

pn+1,j+1,i+1

pn+1,j+1,i-1
pn+1,j,i

Figure 4.60 Finite difference solution of a 15 degree finite difference solution with
(a) showing partial solution in the (t, z) plane, and (b) the complete (x, z, t)
solution.

Page 4.89

Chapter 4 Poststack Migration ( Zero Offset)

4.8.2 Procedure for implementing the finite difference solution


The solution on the previous page uses the data on the nth depth layer to
compute data on the next downward layer (n+1) as illustrated in Figure 4.61.
Compute the n+1th layer from nth layer:

The process starts at the bottom of the time section (with jmax) as shown in (a).

New samples at jmax are computed for each trace at the (n+1) time layer.

The next row at jmax-1 is computed using the previous layer, and the new
samples at jmax, as shown in (b).

The procedure is repeated for each trace and time sample in the n+1 layer (c),
until j=0 is reached.

i.e. The new time section is created from the bottom (in time) up toward the
time that corresponds to the depth level.

After the new tike section at n+1 is completed, the entire process in repeated for
the next downward level or step level at n+2, until each downward layer is
completed.

For stability, / x < 0.5 , providing a limit on the size of the maximum
downward step .

Some implementations of the algorithms may have a discontinuity at the


step and require filtering.
(Deconvolution after migration may cause
problems.)

Finite difference migrations are usually referred to as the 15, 45 or 60


solutions. These implied dip limits must be used with caution as the actual
dip limit is also controlled by the algorithm implementation, velocity, trace
spacing, sampling interval, and the tau step as illustrated in [106].

The 45 or 60 algorithms may be optimized to increase their accuracy to be


equivalent to 65 and 80 respectively.

Cascaded migrations may be used to obtain dips greater than implied by the
algorithm name. A 15 algorithm may be able to migrate dip over 80 with a
large number of cascaded migrations.

Explicit solutions may be solved directly, i.e. one solution sample from a
number of input samples. Implicit methods cannot be solved directly, and
require matrix type approaches such as the tri-diagonal solution.

Page 4.90

Chapter 4 Poststack Migration ( Zero Offset)

Start at
maximum

dx

jmax

jmax
jmax - 1
n

dt

Depth
layers

n+1

n+1

a)

b)

Traces
i-1

i+1
j-1
j

n
x
z
i
j-1
j
n+1

c)

Migrated rows
in time section
move toward
zero time

n+1

n+2

Downward
Continuation
(depth)

d)
Figure 4.61 Illustration of the 15 solution with (a) the kernel on the first row at (x,
z, tmax ) volume, (b) the second row, (c) in the middle of the two layers, and (d) in
the (x, z, t) volume.

Page 4.91

Chapter 4 Poststack Migration ( Zero Offset)

4.8.3 The 45 degree solution


The 45-degree solution requires a few more tools that are described in Appendix
1. A few of those results are shown on these two pages. Using a 45-degree
solution of the form;

3 p 2 3 p
4 3 p
+

=0
x 2z c x 2t c 2 zt 2

(4.33)

we get an intermediate finite difference solution


2
j
j
j
j
2 ( pn +1 pn ) 2 2 2 (1 z ) j
4 4 (1 z ) ( pn +1 pn )
+
p 1
=0
x 2
z
c x 2 t (1 + z ) n + 2 c 2 t 2 (1 + z )2
z

(4.34)

Now, using new variables a, b, , and T, (see Appendix 1), we get;

(1 z ) p j p j = 0
T
T (1 z )
pnj+1 + (1 ) pnj
b
pnj+1 pnj ) + a
(
n)
2 ( n +1
1+ t
1 + t (1 + z )
(1 + z )
2

(4.35)

then,

bT (1 + Z )

(p

j
n +1

pnj )

+ aT (1 2 Z 2 ) ( pnj+1 + (1

) pnj )

(1 Z ) (1 + T ) ( pnj+1 pnj )
2

(4.36)

( k + T ) ( p ) = all the rest

or

n +1, j

(4.37)

We still have the 'T' term in the equation implying three unknowns at pi -1, pi, and
pi +1 at each ith location. Substituting gives the three unknowns in the form,

pn +1,i 1, j + k 2 pn +1,i , j + pn +1,i +1, j = all the rest ,

(4.38)

The solution at this point is still not trivial as the three solution points are implicit,
or can't be solved directly as in the 15-degree method. (i.e. we need the two end
points (pi -1 and pi +1) to compute the center point pi. Fortunately this solution can
be represented in a tri diagonal matrix form that can be solved by standard
procedures that may found in many sources including Brysk [100], Carnahan
[296], Claerbout [294] and Press [187] (Numerical Recipes).

Page 4.92

Chapter 4 Poststack Migration ( Zero Offset)

An example of an eight x values is illustrated below.

k
1

0
0

0
0

0
where

k = 2+

1
k
1
0

0
1
k
1

0
0
1
k

0
0
0
1

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

1
0
0
0

k
1
0
0

1
k
1
0

0
1
k
1

0
1

k
0
0
0
0

px1
v1
p
v
x2
2
px 3
v3


px 4 = v4
px 5
v5


px 6
v6
p
v
x7
7
p
x8
v8

1
b + a

The p vector contains the new samples to be computed at the jth time level. The
values of the vector v are found from all the coefficients and their known values
from the surface above, and from previous times.
An example of the operator at the jth time layer is illustrated below. All samples
on the upper layer and those at time samples greater than j are know. The three
unknowns are identified by the three larger spheres.

i-1

i+1
j+2
j+1
j
n layer

time
j
n+1 layer
x

p vector

Figure 4.62 Illustration of the finite-difference 45 degree operator on tow depth


layers.

Page 4.93

Chapter 4 Poststack Migration ( Zero Offset)

4.8.4 Effect of step on algorithm performance


The step used in finite difference time migrations has a significant effect on the
performance of the algorithm. The following figure illustrates the effect of
increasing the size of the step from 2 ms to 20 ms and 80 ms for a 45 degree
algorithm (enhanced to 65 degrees). Note the degradation in the dip limit of the
respective operators and the addition of aliasing noise. What step do you use?
A red semi-circle defines the migration objective.

Figure 4.63 Various impulse responses for a 65 degree algorithm with tau steps
of a) 2 ms, b) 20 ms, and c) 80 ms from Cary [106].

Page 4.94

Chapter 4 Poststack Migration ( Zero Offset)

4.8.5 Examples of finite difference migration


Figure 4.64 to Figure 4.66 use the 15 solution described above , to illustrate the
dip limitation and resulting noise.
In Figure 4.64, note:

An input time section (a) and resulting 15 time migration (b).

Note the migration noise that result from dips on the diffractions that are too
steep for the algorithm.

Figure 4.65 shows time sections at various depths.

The dark bands identify the steps and the top of the last migration pass.

Data above the dark bands remains unchanged.

Note the gradual focusing of the energy in the bottom diffraction as the
migration surface approaches the diffraction point.

Figure 4.66 illustrates the effect of the imaging condition.

The entire time section is migrated at each downward level, rather than
stopping at the output location.

The 8 images show the location of the imaging condition (output time) by an
arrow on the side.

Data above the arrow is over migrated.

Data below the arrow is under migrated.

Data at the arrow could be left in place (providing there was no migration
above the arrow), or moved to another output section.

It is important to recognize that there is only one time level in each section
that is correctly migrated, i.e. the imaging condition.

Page 4.95

Ch
hapter 4 Poststtack Migration ( Zero Offset)

x
t

a)
x
t

b)
ure 4.64 Th
he input se
ection at a), and b) th
he final 150 migration
n. Note the
e
Figu
excess
sive noise due to mig
gration limitations. Both
B
figure
es have dim
mension (x
x, t).

P
Page
4.96

Ch
hapter 4 Poststtack Migration ( Zero Offset)

x
t

a)

b)
x

x
t

c)

d)
x

e)

f)

ows the tim


me section (x, t) at diffferent dow
wnward ste
eps.
Figure 4.65 This figure sho
The da
ark bands on
o (a) to (ff) identify the
t step width
w
and the
t top of the migrattion
proce
ess. Data above
a
the dark bands remains unchange
ed.

P
Page
4.97

Ch
hapter 4 Poststtack Migration ( Zero Offset)

x
t
t

a)

b)
x

x
t

c)

d)

4
These
e 8 images
s show the
e time secttion at incrreasing do
ownward le
evels
Figure 4.66
in whic
ch the mig
gration pro
ocess did not stop at the app
propriate time
t
level,, but
continu
ued to the surface.
s
Th
he migratio
on is only valid at the time leve
el indicated by
the red box. Ove
er migration is eviden
nt. The number of passes from
m (a) to (h)) are
0, 10, 25
5, 50, 75, 100, 150, an
nd 200.

P
Page
4.98

Ch
hapter 4 Poststtack Migration ( Zero Offset)

x
t

e)

f)
x

x
t

g)

h)

P
Page
4.99

Chapter 4 Poststack Migration ( Zero Offset)

4.9 Methods based on the Phase-shift


4.9.1

Basic Phase-shift method

The downward continuation step can include greater simplicity and accuracy
when using the 2-D Fourier Transform as introduced by Gazdag, [30], [32] and
[102]. This method (as described) results in a depth migration.

The identity of an x location is spread over the transform Kx, preventing the
velocity form varying laterally with the x location.

The velocity may vary with each depth layer so the method works well with
vertically varying velocities.

It handles very steep dips correctly, and is an excellent depth migration,


especially for marine or synthetic data.

The Phase-shift method is very simple and straight forward. Starting with the one
way scalar wave equation,

kz =

2
v2

k x2 ,

(4.39)

we get the simple first-order differential equation in z, where and kx are


independent of z, i.e.,

d P ( z)
2
= j 2 k x2 P ( z ) = aP ( z ) ,
dz
v

(4.40)

The solution is an exponential function,

P ( z ) = e az ,

(4.41)

Adding a z term to z (for a downward continuation step) we get

P ( z + z ) = e

a ( z+z )

= eaz eaz

= P ( z ) eaz ,

(4.42)

The downward extrapolation then becomes,

P ( k x , z + z, ) = P ( k x , z, ) e

Page 4.100

-j z

2
v2

k x2

(4.43)

Chapter 4 Poststack Migration ( Zero Offset)

Converting from one depth layer to the next depth layer only requires the
multiplication of a Phase-shift to each value in the (kx, ) plane. The Phase-shift
is defined by the exponential term in equation (4.30).
The output of the new Z layer could require the inverse 2-D transform to output
the energy at (x, t=0), however since only a single line of data is required, the
process is much simpler.

Sum all the frequencies for a given kx to get a 1-D array (kx, ) where is
equivalent to t=0, i.e. (kx, t=0).

Inverse Fourier transform the 1-D array to get the output (x, t=0) at the new
depth z+z.

z
kx

z+z
kx

a)

Sum

z+z
kx
x

b)
Figure 4.67 Phase-shift method illustrated with a) the phase shift, b) the
frequency summing and the inverse 1-D transform.

Page 4.101

Chapter 4 Poststack Migration ( Zero Offset)

4.9.2 Phase shift plus interpolation (PSPI)


This method allows lateral velocity changes [303]

Computes a number of sub-layers at different velocities for the next depth


level.

Each new sub-layer is inverse Fourier transformed into (x, ) domain sublayers.

The output layer is interpolated from the different velocity sub-layers using the
appropriate lateral velocity.

The new layer is transformed back to the (Kx, ) domain for the next
downward step.

Another important algorithm based on the phase shift method is X depth


migration [165], (not to be confused with a finite difference migration). This
method is described in the following section.

kx

kx

z+z

v1
v2

Interpolate with velocity


New z + z

Figure 4.68 Illustration of PSPI downward extrapolation in (kx, w) space.

The phase shift method is capable of migrating dips steeper than 90 [294]. This
is also referred to as turning ray migration. See Claerbout [294] pages 272-273.
For examples, see Hussein [741].
A PSPI method is described by Gazdag in [303].

Page 4.102

Chapter 4 Poststack Migration ( Zero Offset)

4.9.3 X depth migration


There are a number of different migration algorithms that use the name X. Some
use a finite difference approach (see Appendix 1), while others are based on the
phase shift method.
In the following illustration, an operator was defined by the phase shift method
and inverse transformed back to the x domain. A band of data, at a fixed
frequency, is weighted and projected onto one sample at the Z + z level. These
operators act in a convolution fashion to allow the extrapolation of a new layer
with varying lateral velocities. (Note that the operator assumes that the velocity
is locally constant over the spatial extent of the operator, but can vary with each
spatial location of the downward continued sample.)

z
x

z+z
x

Figure 4.69 Illustration of X downward extrapolation in (x, w) space.


Other properties and comments

Limitations are dependent on the width of the operator in the x direction.

The number of operators is defined by the range of velocities and the rate at
which the velocities are incremented and/or interpolated.

The design of the X operator must address four issues, (Al-Saleh 2006)
1. stability,
2. accuracy,
3. efficiency,
4. and the maximum dip limit of the migration.
References on this method are Holberg [88], Hale [159] and Nautiyal [216]

Page 4.103

Chapter 4 Poststack Migration ( Zero Offset)

4.9.4 Xt depth migration


The downward continuation may also be performed in the (x, t) space where the
operator is inverse Fourier transformed from X space into (x, t) space. Now a
patch of data is weighted and mapped to one point at the new depth. Note that
this operator is a mini-Kirchhoff diffraction shape that can vary with the x location
as discussed in the next section. The operator however will be the same at each
time location for a given x, allowing the velocity to vary in x and z.
The spatial size of this operator can be defined by the expected dip limit on the
data.

z
x

z+z
x

Figure 4.70 Illustration of Xt downward extrapolation in (x, t) space.


t

Page 4.104

Chapter 4 Poststack Migration ( Zero Offset)

4.9.5 Phase-shift operator


The Phase-shift operator has an interesting shape as illustrated in the following
figures. Figure 4.71 shows in (a) the phase shift (b) the amplitude operator, and
(c) a quiver plot where the small vectors identify both the amplitude and phase.
The wavenumber and frequencies are in radians, the velocity 1.0 and the depth
increment 1.0.

a)

b)

c)
Figure 4.71 Plots of the Phase-shift operator, a) the amplitude, b) the phase, and
c) a quiver plot showing both amplitude and phase.

Page 4.105

Chapter 4 Poststack Migration ( Zero Offset)

4.9.6 Phase shift example


The MOD50.SGY model data was migrated using the phase-shift method with the
result shown in Figure 4.72. The 2-D FFT required 256 traces and 2.048 secs
(1024 samples) so the input data was padded with extra traces and time samples.
Any output depth increment that prevented aliasing in the vertical direction could
be used. The depth migrated section was plotted with the same software to plot
the time sections, so a depth increment of 10 ft, equivalent to the time sample
was used.
The data was downward continued to a depth of 10,000ft, corresponding to 2.0
seconds on the output time scale.
The migrated result is very similar to the FK method, but has no complex
interpolation problems. It does contain wrap around of energy that could be
reduced by adding extra zeros to the bottom of all the input traces (i.e. N = 2048),
or by a gradual tapering of the energy at the bottom of the section. (Tapering
requires a complete inverse 2-D FFT at each downward step that increases the
runtime.)
Figure 4.73 contain examples of the downward continued time sections. The
depths of these time sections are 1.0 ft, 2,000 ft, 3,000 ft, and 5,000 ft. These time
sections could result from many downward steps, but because the velocity is
constant, each time section was produces by one depth step.
The first time section was chosen at 1.0 ft to illustrate the dip-limit filtering action.
Dip greater than 45 degrees in (x, t) space have been reduced. This is noticeable
at the ends of the dipping events, and at the input spikes.
The second time image corresponds to a depth of 1,000 ft that has an upward
time shift of 200 ms.
The third time image corresponds to a depth of 3,000 ft with an upward time shift
of 600 ms. At this depth, the upper diffraction has focused its energy to a point
at time zero.
The fourth time image corresponds to a depth of 5,000 ft that has an upward time
shift of 1,000 ms. At this depth the horizontal reflector has moved to time zero.

Page 4.106

Chapter 4 Poststack Migration ( Zero Offset)

1
1000

2000
3000

3000

5000

5000

Depth
(ft)

10,000

Figure 4.72 Phase shift migration of mod50.

a) I.0 ft

b) 2000 ft

c) 3000 ft

d) 5000 ft

Figure 4.73 Time sections at various depth of a) 1.0 ft, b) 2,000ft, c) 3,000 ft, and
d) 5,000 ft. Note that (c) and (d) are plotted with high amplitudes to emphasize
the wrapped energy.

Page 4.107

Chapter 4 Poststack Migration ( Zero Offset)

4.9.7 Comments on the phase shift method


The phase shift method:

It is a depth migration, (however it may be modified to be a time migration).

Assumes constant lateral velocities within the depth increment.

Is accurate for vertically varying velocities.

Data may wrap around in time and depth ( Kx and ) and thus require special
software, or extra padding of traces and samples.

The phase shift method may be simplified to the FK method (see pages 33-34
of [294]) when the velocity is constant.

The size of the depth increment should be established by the local (interval)
velocity within the depth increment. Any depth increment can be used when
the velocity is constant. (Not so for the finite difference method)

May incorporate a Stolt FK migration between depth layers and output a band
of data to reduce run times.

Is the basic theory behind most of the wave propagation algorithms.

Figure 4.74 The top edge (t = 0) of each time section is mapped to the
corresponding depth level on the migrated section

Page 4.108

Chapter 4 Poststack Migration ( Zero Offset)

4.9.8 Preparing the phase shift method to other domains


The perspective views of the phase-shift operator provide significant insight into
the energy at the Nyquist boundaries that are at the edges of these figures. The
amplitudes should be zero at these boundaries.
Sharp
frequency
transition

Sharp Kx
transition

Sharp
frequency
transition

Aliasing

kx-max = 0.5/x

Figure 4.75 The amplitude of the phase-shift operator with problems identified at
the Nyquist boundaries.

a)

b)

Figure 4.76 Amplitude plots illustrating a) band limiting the frequencies, and b)
reducing the size of the trace interval.
Simple corrections are identified in the above figure that provides band limiting
with a raised cosine bandwidth filter in the frequency direction, and a smaller
trace interval of 3 m in the spatial direction. We point out that the size of the trace
interval is not determined by the phase-shift operator but by the data being
migrated, causing potential problems.

Page 4.109

Chapter 4 Poststack Migration ( Zero Offset)

4.9.9 The phase-shift operator in (kx, ), (x, ), and (x, t) space


The simplicity of the perspective views of the phase-shift operator in (kx, ) space
has lead naturally to similar perspective views in (x, ) and (x, t) space as shown
in Figure 4.77.
Data in (x, w) space was created by 1-D inverse Fourier transforming the (kx, )
data at each frequency . The data still remains complex, with real and imaginary
components shown in (c), and (d), with the magnitude in (e) Viewing the data in
(x, ) space is very useful is aiding the design of X operators. These operators
are truncated in the spatial direction to a reasonable size for efficient migrations.
The shaping and design of these operators is of continual interest.
The inverse 1-D transform continues, but the data must now be conditioned to
convert the complex frequency data into real time data. This was accomplished
by creating a conjugate symmetry image of the (x, ) space, and then inverse
Fourier transforming each spatial array x. Views in this space show a mini
Kirchhoff operator that can also be used to downward continue data. The 2-D
size of this operator may also provide an indication of the spatial extent and its
effect on dip limiting on the X operators.

Page 4.110

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)

c)

d)

e)

f)

Figure 4.77 The phase-shift operator showing a) the (kx, ) amplitude, b) the (kx,
) phase, c) the (x, ) real, d) the (x, ) imaginary, e) the (x, ) magnitude and f)
the (x, t) operator.

Page 4.111

Chapter 4 Poststack Migration ( Zero Offset)

4.9.10 Visualization the Kirchhoff operator defined by the phase-shift


operator
The perspective view of the phase-shift operator in (x, t) space provide an
excellent image of an ideal Kirchhoff migration operator. Figure 4.78 shows
various perspective views of the (x, t) operator with an extrapolation depth of
250m. Note the characteristic amplitude and shape of the operator. The
frequencies of the operator were band limited to some extent and display a
wavelet shape. The side view hints at a typical phase rotation of the 45 degree or
Rho filter for 2-D data. This (x, t) operator could be used by using a 2-D
convolution for each migrated point, but would be very inefficient, even if the
convolution area was limited to a band around the central amplitude. In contrast,
a traditional Kirchhoff migration only performs a weighted line integral along the
central peak of the (x, t) operator, and then applies the Rho filter to the migrated
section.

Figure 4.78 Various perspective view of the phase-shift operator in (x, t) space
for a depth increment of 250 m.

Page 4.112

Chapter 4 Poststack Migration ( Zero Offset)

The previous view in Figure 4.77f of the phase-shift operator in (x, t) space was
for a small depth increment and was difficult to visualize. Consequently the
parameters have been modified for a maximum frequency of 1000 Hz, and a
spatial increment of 1 m to produce the high resolution image shown below. The
depth increment is 10 m.

Figure 4.79 Two perspective views of a high resolution phase-shift operator in (x,
t) space for a small depth increment of 10 m.
---------------------------------------------------------------------------------------------------------------

a)

b)

Figure 4.80 Identification of one frequency on the phase shift operator that will
be used to build an wX operator.

Page 4.113

Chapter 4 Poststack Migration ( Zero Offset)

4.9.11 Problems with the wX operator


The sharp corners of the phase-shift operator in (kx, ) space cause problems in
designing an wX operator. When the phase-shift operator is truncated or
windowed in x, instabilities occur. The inverse transform of the window operator
is convolved in the kx dimension, producing an overshoot of the amplitude in the
area where it should be 1.0 or unity. This overshoot is commonly referred to as
Gibbs phenomenon.
The over shoot is not a real problem for a single application of the operator, but
this operator will be applied many times, i.e. at each point for every downward
extrapolation. The wX operator in convolved in x space, but multiplied in kx
space. Any amplitudes greater than 1.0 will become very large numbers, causing
distortion. The following images illustrate this problem using an extreme case
with a box-car window. The figures progress from the phase-shift operator in kx,
then x, then the window in x and kx, close ups of the complex windowed operator
in x, the effect on operator in kx, and finally the kx operator raised to powers of 10
and 100 where extreme distortion is already visible

a) Magnitude Kx

c) Magnitude in x

f) Window in x

b) Phase

d) Real in x

e) Imaginary in x

g) Window amplitude in Kx

Page 4.114

Chapter 4 Poststack Migration ( Zero Offset)

h) Magnitude

i) Real

k) Magnitude

j) Imaginary

l) Real

m) Amplitude power 10

n) Power of 100

Figure 4.81 Phase-shift operator progressing from a) kx amplitude, b) kx phase, c)


x magnitude, d) x real, e) x imaginary, f) x box-car window, g) kx a sinc of the boxcar, h) the x magnitude of the windowed operator, i) x real operator, j) x imaginary
operator, k) kx magnitude, l) the kx phase, m) the kx amplitude raised to a power of
10 and n) the amplitude raised to a power of 100. (PhaseShiftHanning.m)
The above example used a boxcar window to illustrate the problem. Better
shaped windows will produce less overshoot that allow the maximum power to be
raised to a much higher value, say 500 or more.

Page 4.115

Chapter 4 Poststack Migration ( Zero Offset)

4.9.12 Kirchhoff downward continuation migration


There are many other methods for downward continuation migration such as
Huygen's wave front or the Kirchhoff method [37] illustrated below in Figure 4.82,
which illustrates extrapolations in a depth and time. Note:

The shape of the Kirchhoff operator is the same for all times at one x location,
and at a given step, but may vary laterally as the velocity changes with x.

The time shifting part of the operator for a depth migration in (a).

The lowering of the "surface" for a time migration in (b).

Spreading operator

a)

New t = 0

Spreading operator

Summation operator

b)
Figure 4.82 Kirchhoff method of downward continuation illustrating in (a) depth
extrapolation, and in (b) time extrapolation.

Page 4.116

Chapter 4 Poststack Migration ( Zero Offset)

4.9.13 Boundary conditions for downward continuation


In the previous work, finite difference solutions did not mention the problems that
are encountered at the edges of the various time sections. If care is not taken
energy will reflect off the edges and distort the data. Special effort must be
extended to ensure that these boundary conditions are handled in an appropriate
manner. Some methods are
1. Modify the equations at the boundaries.
2. Append large numbers of additional traces.
3. Append a few additional traces and include a taper.

Adding a large number of traces, such as 100, to each end of 2-D project may not
be too severe, but in 3-D projects the increased run time could be a serious
problem. For example a 3-D project that is 150 traces by 150 traces has a total of
150*150 = 22,500 traces in the volume. Appending 100 traces to each edge of the
volume raises the volume trace count to 350*350 = 122,500, nearly five and a half
times the original number.

A simple method for adding only a few trace (from 10 to 20) is included as it is
simple, and is quite successful.

First append 10 to 20 null traces at each end of the section.

After each downward step, multiply the appended traces by a very gentle
tapering function, to attenuate the energy that has moved into these null
spaces.

The function should be 1.0 at the inside edge and something like 0.9 at the
far edge.

The function may be linear, cosine, or exponential.

Repeated application of this attenuation function in each downward layer


provides a good taper without severe boundary reflections.

Note at the edge, 0.950 = 0.005.

The above description is based on a paper by Cerjan [60]. Other references on


boundary conditions are Clayton [17] and [237], Randall [154], and Sochacki [66].

Page 4.117

Chapter 4 Poststack Migration ( Zero Offset)

4.10 Reverse Time migration


Reverse time migration [98, 757] is similar to the downward method as it
computes the volume (x, z, t) to arrive at the t=0 surface. However, rather than
start at the top layer as in downward continuation, it starts at the rear of the
volume (x, z, t=tmax) and works toward the front surface as illustrated below.

Figure 4.83 Reverse time migration.

Page 4.118

Chapter 4 Poststack Migration ( Zero Offset)

The algorithms may be simple, and may require an extra high density of traces to
maintain accuracy. Trace spacing similar to the distance of the time samples is
usually required, i.e. x=tV/2.

Reverse time migration is:

A steep dip depth migration.

May use the full wave equation, or a one-way wave equation.

May work with multiples with normal implementations.

By allowing the density to vary inversely with the velocity, maintaining a


constant impedance to eliminates internal reflections.

References on reverse time migration are [453] by McMechan, [374] by Baysal, or


[98] by Fricke. See also [103], [123], [368], [488], and [489].

Zhang, Yu, and Sun, James, 2009, Practical issues in reverse time migration: true
amplitude gathers, noise removal and harmonic source encoding, First Break, V.
27, p53 59 (January)

Page 4.119

Chapter 4 Poststack Migration ( Zero Offset)

4.11 Migration of 3-D Data


4.11.1 Introduction
Three dimensional seismic provides a much clearer view of the subsurface than
is possible from 2-D seismic. One of the main benefits of 3-D migration is the
elimination of sideswipe and obliquity problems.
Assume a 3-D zero-offset volume. Recall a point reflector in 2-D will produce a
hyperbola. In 3-D, a point reflector produces a hyperbolic type of surface called
hyperboloid as illustrated below in Figure 4.84.
Kirchhoff migration would simply weight and sum all the samples that lie on the
hyperboloid to produce the output sample. This procedure is expensive. The
usual 3-D migration slices the volume as illustrated in the following pages.

Figure 4.84 Hyperboloid or 3-D diffraction.

Page 4.120

Chapter 4 Poststack Migration ( Zero Offset)

The migration of a point in 3-D will produce a hemisphere or bowl shape as


illustrated in Figure 4.85.

Figure 4.85 Complete 3-D migration of an input point becomes a hemisphere.


A time slice through the hemisphere will result in a circle of data. An examination
of this circle allows an evaluation on the performance of 3-D migration.
The migration of a single live trace in the center of a 3-D volume will produce a
series of concentric hemispheres. A time slice will now produce a series of
concentric circles.

Problems with the migration algorithm, such as dip limits, show up on these
concentric circles as an uneven distribution of energy. Consequently much effort
is extended to ensuring that these concentric circles are all round. Concentricity
should only be expected below the dip limit of the algorithm.

Page 4.121

Chapter 4 Poststack Migration ( Zero Offset)

4.11.2 Two pass 3-D method


Any vertical slice of the hyperboloid will be a hyperbola.
If the entire volume is sliced vertically in one direction (as a loaf of bread) as
shown in Figure 4.86 and each slice migrated, the data in each slice will collapse
into a point. The resulting points will form a hyperbola in the orthogonal
direction as illustrated in Figure 4.87.

Figure 4.86 Sliced bread image of hyperboloid.

Figure 4.87 Two-dimensional migration of hyperbolas in Figure 4.86.

Page 4.122

Chapter 4 Poststack Migration ( Zero Offset)

When the 3-D volume is sliced in the orthogonal direction, the points of Figure
4.88 will form a new hyperbola illustrated in Figure 4.50.

Figure 4.88 After migration in one direction data align in an orthogonal


hyperbola.
When the 3-D volume is sliced and migrated in the orthogonal direction, the new
hyperbola collapse to one single point as illustrated in Figure 4.89.

Figure 4.89 3-D migration after passes in both directions.


This two-pass method was illustrated with a single hyperboloid, but it will also
migrate all the possible hyperboloids in the volume.

Page 4.123

Chapter 4 Poststack Migration ( Zero Offset)

4.11.3 Two pass velocity error


A problem with the two pass method results when the velocity varies vertically.
During the first pass the deeper diffractions at the edge of the hyperboloid will be
migrated with velocities of deeper events, and not with the velocities at the top of
the hyperboloid as illustrated in Figure 4.90.

Figure 4.90 Two pass velocity errors.

These errors are not considered serious for shallow dips. Hood [77] reports that
a depth of 2.0 seconds, and dip of 45 degrees, the error is equivalent to a velocity
error of 1%. Others claim that there is still sufficient error to completely avoid the
two pass method.
This error may be eliminated by a one pass migration in which the downward
continuation process proceeds in both directions for each downward increment.
It should be noted that the dips at which these errors are significant may coincide
with the limits of the algorithms. If a steep dip 3-D migration is required, a one
pass method with steep dips should be used.

Page 4.124

Chapter 4 Poststack Migration ( Zero Offset)

4.11.4 One-pass 3-D migration


The one-pass method of 3-D migration is usually more accurate than a two-pass
method because of extra care in the design. It is still only accurate to the dip
limit, but it does eliminate the two-pass velocity errors.
True one-pass algorithms are available, but the most common one-pass method
is a modified version of the two-pass method as outlined below.

Slice the data as in the two-pass method.

Migrate each slice, but only for one downward increment.

Slice the data in the orthogonal direction.

Migrate this other direction for the same downward increment.

Repeat the above process by slicing the data in the original direction.

Migrate one downward increment.

Re-slice etc. ...

Repeat the procedure for all downward increments.

This method requires large CPU data memory to prevent excessive data
movement to and from disk.
The Stolt FK method of migration uses the pseudo depth conversion to convert
the input time section to a constant velocity section. This method is therefore,
not susceptible to the two pass velocity error.
A 2-D Kirchhoff migration used in the two-pass method would be subject to this
two pass error, but a 3-D direct one-pass method would not.

Page 4.125

Chapter 4 Poststack Migration ( Zero Offset)

4.11.5 Concentric distortion in 3-D migration


Two pass finite difference migrations have difficulty in maintaining an even
amplitude and circular shape on the time slice of a migrated point. This occurs at
steeper dips and usually at an angle half way between the migration directions as
illustrated in Figure 4.91 below, with an actual example in Figure 4.92.

a)

b)

Figure 4.91 Time slices illustrating the difference between a) an ideal 3-D
migration and b) a finite difference two-pass method at large dips.
This distortion may not be as serious as it may occur at the dip limit of the
algorithm.
Some 3-D seismic projects have been oriented specifically to align the source
and receiver directions with the geological features (faults) to minimize the two
pass migration errors.
Concentricity of slices does not guarantee an ideal operator. Consider the slices
across a cabbage and a carrot; both are circular, but only the cabbage has a
hemispherical shape.

Further study on these topics may be found in Yilmaz [131], Li [214], Cary [278],
and Dickinson [365].

Page 4.126

Ch
hapter 4 Poststtack Migration ( Zero Offset)

Figure 4.9
92 Example of a time
e slice thro
ough a mig
grated 3D volume
v

Pa
age 4.127

Chapter 4 Poststack Migration ( Zero Offset)

4.11.6 Fast 3-D Kirchhoff time migration


Figure 4.93a illustrates 3-D post-stack Kirchhoff time migration in which the input
data is summed over a hyperbolic surface.

Assume a small 3-D that is 100 by 100 traces. An output trace in the center would
sum over all 10,000 traces.

A brute force method for one migration sample would sum 10,000 input samples.
Each of these input sample requires computation of the offset and sample
position, the time shift, an antialiasing filter applied at that time, interpolation
between samples, and scaling.

A faster method is possible.

The offset and sample position will be the same for each trace, as illustrated
by the circle in (b).

All the samples within the circle could be summed and then scaled, and time
shifted as one sample.

This could be repeated for all time levels within the hyperboloid, i.e., all traces.

All traces in the migration aperture (say 10,000) are gathered according to
their migration offset into bins of a migration gather (of say 100 traces) in (c).

Antialiasing filters, scaling, and time shifting may now be done on the
migration gather for a huge saving in computer time.

The time shifting is the same as NMO and conventional routine may be used.

In areas where there is structure and visible diffractions, the migration gathers
may be used (as conventional CMP gathers) for migration velocity analysis.

(Could these savings be passed onto prestack migration?)

Note, binning of the input traces will combine a number of traces with an offset
range that is the same size as the bins. Combining these offsets is of benefit to
the process as it acts as a natural antialiasing filter.

Page 4.128

Chapter 4 Poststack Migration ( Zero Offset)

a)

b)

c)

Figure 4.93 Speedups for 3-D post-stack migration, a) the migration ellipsoid
showing the same offset, and b) the concentric rings of traces at the same offset,
and c) the migration stack of the concentric traces.

Page 4.129

Chapter 4 Poststack Migration ( Zero Offset)

4.11.7 3-D velocities


Velocity estimation for 2-D seismic may often require extensive testing because
of sideswipe, and oblique reflections require different velocities to focus the
energy. Typical migration velocities for 2-D data are 90 percent of the RMS
velocities estimated from the stacking velocities.
The problems of side swipe and obliquity do not occur in 3-D migrations.
A good place to start 3-D migrations is with 100 percent RMS velocities, or the
equivalent interval velocities.
The inclusion of DMO in the velocity estimation process has greatly improved the
estimation of velocities for migration.
Velocity anisotropy is often present in the vertical direction and is attributed to
geological layering. This form of anisotropy is usually absorbed into the velocity.
Azimuthal anisotropy is considered a potential problem, especially in the near
surface.

Page 4.130

Chapter 4 Poststack Migration ( Zero Offset)

Blank page.

Page 4.131

Chapter 4 Poststack Migration ( Zero Offset)

4.12 Testing Migration


Simple migration methods may be adequate for many applications, but when the
structure becomes complex, advanced migration algorithms are required. The
migration algorithms should be tested to ensure they meet the desired
performance. The following tests have proved valuable.

Test #1.

Constant velocity model with point, diffraction, and dips.

Look for

loss of energy

aliased data

incorrect positioning of dips

focusing of diffraction back to a point

points should migrate to a semicircle.

a)

b)

Figure 4.94 Constant velocity structure for test #1, a) the input time section, and
b) the migrated section.

Page 4.132

Chapter 4 Poststack Migration ( Zero Offset)

Test #2.

Migration of a single trace.

A single trace is inserted into a null section and migrated with either a constant
velocity, or a velocity from a project.
The migration should:

Produce semicircles for a constant velocity out to the appropriate dip of the
migration algorithm.

Produce smiles that will deviate according to the interval velocity.

Noise may be evaluated by applying automatic gain (AGC) to help identify


1. noise around the edges of the section
2. wrap around noise
3. step discontinuities
4. aliasing noise
5. edge effects.

The single trace is spatially aliased, and does emphasize aliasing noise. This
aliasing noise can be reduced by using three input traces. The same input trace
is used with amplitude weightings of 0.25, 0.5, and 0.25.

a)

b)

Figure 4.95 A single trace a) before and b) after migration.

Page 4.133

Chapter 4 Poststack Migration ( Zero Offset)

Test #3.

Test of vertical velocity accuracy for algorithms.

This test uses the RMS velocities from a project to create a series of diffractions
at various times. These diffractions are then migrated with available algorithms,
and varying parameters. The test is fast, and gives a measure of the expected
resolution that may be achieved after migration.

Use the RMS velocity to create a series of vertical diffractions at various


intervals (e.g. 100 milliseconds).

The migration should focus all diffractions back to points (in the ideal case).

Percentage variations of the velocities may be tried along with various


migration algorithms.

a)

b)

c)

Figure 4.96 Migration test of velocity functions at various times, a) the velocity
function, b) the diffractions, and c) one of the migrations.

Page 4.134

Ch
hapter 4 Poststtack Migration ( Zero Offset)

Test #4
4

Testt for laterrally varyiing veloc


cities.

st is usua
ally applie
ed to a nu
umber of test mode
els with known velo
ocity
This tes
structurre. The equivalent
e
zero-offse
et stack is
s produced
d from an accurate and
known method. The objec
ctive is to
o accurate
ely represe
ent the orriginal velo
ocity
e migration
n. Feature of this tes
st may be
structurre with the

positioning of point refle


ectors with
h a time or depth mig
gration

focu
using of difffraction to
o a point

dip limits

conttinuity of re
eflectors.

a)

b)

c)

d)

Fig
gure 4.97 Migration
M
t
test
for late
eral velocitty variation
ns, a) the geological
g
struc
cture, b) se
eismic mo
odel, c) time migration, and d) id
deal depth
h migration
n.

Pa
age 4.135

Chapter 4 Poststack Migration ( Zero Offset)

Test #5

3-D migration test.

The purpose of this test is to evaluate the concentricity of a 3-D migration. This
is accomplished by inserting a single trace into the middle of a null 3-D volume as
illustrated in Figure 4.98. The performance of the migration may be evaluated by
examining time slices.
The positive and negative peaks on a trace form a stacked bowl effect (a), which
in turn creates the rings on a time slice. The innermost ring corresponds to the
bottom of a bowl at zero dip, and the outer rings come from deeper bowls with
steeper dips. A vertical section through the center (a) will identify the ring with
the dip to evaluate the migration.
A hemisphere may be defined by a semicircle in one direction, and a series of
varying size semicircles attached to and orthogonal to the original semicircle.
The hemisphere is complete as long as the semicircles in both directions are
complete. When the semicircles start to distort, the hemisphere will also distort,
with the first notable distortion occurring on the 45 angle.
It is repeated that concentricity does not ensure an accurate migration.
As in 2-D migration, a singe trace in a 3-D volume is aliased and may produce
excessive aliasing noise. The aliasing noise may be reduced by spreading the
energy of the single trace to its neighbors. An example would be 1/6 of the
energy to the four traces in the x and y directions, and a value of 2/6 to the central
trace. The amplitude is then similar to the original single trace.
Figure 4.98b is a time slice from a test migration that incorporates the aliasing
reduction defined above.

Page 4.136

Ch
hapter 4 Poststtack Migration ( Zero Offset)

a)
x

b)
Figure
e 4.98 A 3--D migratio
on of a sing
gle trace, a)
a showing
g a verticall slice thou
ugh
the cen
nter of the hemispher
h
res, and b)) the resultting time slice.

Pa
age 4.137

Chapter 4 Poststack Migration ( Zero Offset)

4.13 Points to note in Chapter 4

Migrations may be classified as "time" or "depth", with the possibility that


some methods may approximate depth migrations.

Migration results may be estimated with a compass when the section is


plotted with a vertical time scale to match the horizontal distance.

The maximum dip of a seismic event on a section before migration is 45


degrees.

The migration process creates semicircles or smiles.


These usually
reconstruct to form the geologic structure, but may appear from noise.

The three main methods of migration are Kirchhoff, FK, and downward
continuation.

Kirchhoff migration methods may vary from simple to very complex.

The Kirchhoff method sum specula energy that is tangent to the summation
curve.

Downward continuation migration may be accomplished by many different


methods.

Most one pass 3-D migration uses the basic 2-D algorithm.

All migration algorithms should be tested.

Al-Saleh, S., 2006, PhD dissertation, Department of Geology and Geophysics,


University of Calgary.

Additional information:
An excellent review of migration may be found in a paper by Gazdag that was
published in Proceedings of the IEEE [102]. A First Break paper by Whitmore
[165] also provides a good overview of the basic algorithms.
A very interesting paper by Larner that was written in the 1976, but not published
till 1990 [104]. It contains an interesting discussion that shows Kirchhoff and
finite difference migrations to be solutions of the wave equation.
Another good overview paper is by Larner is "In Quest of the Flank" [94].
An excellent book with detailed math is Claerbout's "Imaging the Earth's Interior"
[294] and has been referenced a great deal in this chapter.
Another excellent work that has many excellent figures and is necessary for all
professionals working in the seismic industry are the two volumes by Yilmaz
"Seismic Data Processing" [83].

Page 4.138

Potrebbero piacerti anche