An Introduction To Seismic Diffraction. Advances in Geophysics

Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

CHAPTER ONE

An introduction to seismic
diffraction
Benjamin Schwarz*
Department of Earth Sciences, University of Oxford, Oxford, United Kingdom
GFZ German Research Centre for Geosciences, Potsdam, Germany
*Corresponding author: e-mail address: [email protected]

Contents
1. Introduction 1
2. Seismic diffraction 4
2.1 Optical roots 5
2.2 Diffraction types and properties 9
3. Diffraction separation and imaging 13
3.1 Coherence and wavefronts 14
3.2 Adaptive reflection subtraction 17
3.3 Diffraction focusing 19
3.4 Faults, fractures, and unconformities 25
3.5 Diffraction imaging in the field 29
4. Implications and new directions 36
4.1 Spatial wavefield sampling 37
4.2 Symmetries and correlations 41
4.3 Deciphering the coda 48
5. Conclusions and outlook 52
Appendix. Kirchhoff’s diffraction integral 54
Acknowledgments 59
References 60

1. Introduction
The first accounts on seismic diffraction date at least back to the 1950s,
where the concept was recognized to form a very useful and important
ingredient in initial attempts at migration (Hagedoorn, 1954). In addition,
it was first identified as a primary carrier of information when faults and
fractures are investigated (Krey, 1952). In particular in the 1970s and early
1980s, seismic diffraction phenomena were studied quantitatively by

Advances in Geophysics, Volume 60 # 2019 Elsevier Inc. 1


ISSN 0065-2687 All rights reserved.
https://doi.org/10.1016/bs.agph.2019.05.001
2 Benjamin Schwarz

computationally solving the forward problem of modelling seismic waves in


complex media (Berryhill, 1977; Buchen & Haddon, 1980; Hilterman,
1970, 1975; Trorey, 1977). Kunz (1960) was arguably the first to emphasize
the potential value of diffraction for dedicated seismic fault interpretation.
When the field of optics is concerned, however, it becomes apparent that
a rigorous theoretical framework for the treatment of diffraction phenomena
is even of significantly older origin and can be traced back to the pioneering
works of Grimaldi, Huygens, Babinet, Fresnel, and Kirchhoff (a concise but
compelling account can be found in Born and Wolf, 2013). Aptly named,
the process of Kirchhoff migration—be it in the depth or time domain—
represents a computationally feasible adaptation of integral representations
first formulated by Kirchhoff for the treatment of electromagnetic wave
phenomena to seismic imaging (Schneider, 1978; Wiggins, 1984).
Kirchhoff’s construction can be viewed as a mathematical manifestation
of the Huygens–Fresnel principle, which simply states that any imaginable
wavefield and its interaction with changes in medium properties can be
described by the interference of infinitely many point sources excited with
every progression of the wavefront. The drastic implications become imme-
diately obvious when Young’s celebrated slit experiments are considered,
in which striking and initially unexpected interference patterns result from
wave energy entering geometrical shadows of obstacles, which is not
predicted by the older framework of geometrical optics. Interestingly,
roughly at the same time as when the importance of seismic diffraction
was first investigated, a powerful extension of geometrical optics was
suggested, that can likewise accurately describe observable diffraction phe-
nomena (Keller, 1962). Largely based on this far-field approximation, a first
theoretical framework for the treatment of diffraction in seismic exploration
was formulated (Klem-Musatov & Aizenberg, 1984, 1985) culminating in
a standard text book, which until the present day must be considered a
reference on the topic (Klem-Musatov, 1994).
Triggered by these important theoretical developments, the first impact-
ful applications of seismic imaging with diffractions appeared in the litera-
ture, which impressively suggested that the diffracted wavefield bears the
potential to facilitate high-resolution imaging of important small-scale geo-
logical features (e.g., Kanasewich & Phadke, 1988; Landa, Shtivelman, &
Gelchinsky, 1987). Despite these initial successes, however, the following
decade largely saw a stagnation of developments, owing in large part to spec-
tacular developments occurring in the fields of migration and first successful
attempts at full-waveform inversion, in which diffraction is principally
An introduction to seismic diffraction 3

honored, but arguably does not take a prominent role (e.g., Etgen, Gray, &
Zhang, 2009; Virieux & Operto, 2009). In addition, the introduction of
insightful quantitative attributes in seismic interpretation, which, for exam-
ple, enabled the detection of faults through image processing in the image
domain, seemed to make the sometimes tedious investigations into the weak
diffracted wavefield superfluous (e.g., Bahorich & Farmer, 1995; Marfurt,
Kirlin, Farmer, & Bahorich, 1998). It was not until the late 1990s that it
was discovered that in particular in the context of seismic monitoring, rel-
evant small-scale structural changes can reliably be detected with dedicated
diffraction processing (Landa & Keydar, 1998), which again sparked inter-
esting new developments in the field with continued stimulation until the
present day.
Without a doubt one of the major stumbling blocks for the success of
diffraction imaging is the extreme weakness of the wavefield when com-
pared with reflections and diving waves. Likewise severely complicating
quantitative investigations, diffractions are known to heavily interfere with
themselves and the rest of the recorded wavefield, which makes them barely
or not at all recognizable on individual traces. In addition, the steep inclina-
tion of diffracted contributions together with their overall weakness often
tends to lead to the misconception that uncorrelated noise rather than coher-
ent signal is observed. Confronting these major challenges, the successful and
noninvasive separation of diffractions evolved into a major cornerstone of
the field. The important works of Khaidukov, Landa, and Moser (2004),
Fomel, Landa, and Taner (2007) and Moser and Howard (2008) arguably
were the first to rigorously establish integrated strategies for the successful
separation and focusing of diffracted energy to arrive at highly resolved
images of the subsurface. Extending the formerly dominant mindset of
migration-type imaging, Sava, Biondi, and Etgen (2005) and Fomel et al.
(2007) additionally suggested to use the separated diffracted wavefield for
velocity model building. From there onwards, a variety of new separation
and imaging workflows were developed, which helped diffraction-based
imaging and inversion gain momentum in recent years (e.g., Bansal &
Imhof, 2005; Bauer, Schwarz, & Gajewski, 2017; Berkovitch, Belfer,
Hassin, & Landa, 2009; Dafni & Symes, 2017; Dell & Gajewski, 2011;
Klokov & Fomel, 2012; Santos, Mansur, & McMechan, 2012).
Originating from the works of Berkovitch et al. (2009), Dell and Gajewski
(2011) along with Bakhtiari Rad, Schwarz, Gajewski, and Vanelle (2018), it
was recently observed that local coherence measurements can lead to an auto-
mated and highly noninvasive separation of weak diffracted contributions
4 Benjamin Schwarz

from the rest of the wavefield (Schwarz, 2019; Schwarz & Gajewski, 2017a).
In this framework, in contrast to previous investigations, the reflected
rather than the diffracted wavefield is specifically targeted through coherence
measurements and subsequently subtracted to reveal less dominant formerly
hidden contributions and make them accessible for further processing.
Through its ability to suppress particularly strong interfering reflections, this
approach shares strong similarities with the process of plane-wave destruction
(Fomel, 2002; Fomel et al., 2007). As the concept of coherence, in addition to
targeted wavefield separation, also allows the enhancement and amplification
of weak signals and because—just like diffraction itself—it is deeply rooted and
intuitively treated in the field of optics, it underpins essentially all consider-
ations that will be made in this chapter.
Here the general aim is to complement existing theoretical accounts
(Klem-Musatov, 1994; Klem-Musatov, Hoeber, Moser, & Pelissier,
2016a, 2016b), with an intuitive easy-to-access introduction to the topic
of seismic diffraction. In contrast to more complete accounts, which have
an emphasis on the forward problem, i.e., the modeling of wavefields, here
I take a more pragmatic, data-driven approach, in which the practical
utilization of diffracted signatures is the dominant incentive. A more rigor-
ous and complete treatment of diffraction phenomena, in terms of
derivations from first principles can be found in the aforementioned works.
Following a brief introduction of diffraction types and definitions and the
deep connections to optics, it will be demonstrated by means of coherence
arguments and conventional Kirchhoff migration, how seismic diffraction
paints a detail-rich complementary picture, backed up by controlled
synthetic investigations and complex academic and industry-scale field data
examples. Following this treatment of coherent diffraction separation and
imaging, additional unique properties of the diffracted wavefield and their
potential, as well as current limitations imposed by insufficient acquisition
design are illustrated through insightful examples.

2. Seismic diffraction
Seismic subsurface imaging has been strongly influenced by optical
imaging. This becomes particularly apparent when wavefield-based
processing and, in particular, the concept of migration are concerned. While
many developments in optics utilized the assumption of monochromatic
waves, a rigorous framework for the treatment of pulsed, partially coherent
light has also been formulated (Born & Wolf, 2013), which closer resembles
An introduction to seismic diffraction 5

the case of typically localized disturbances that seismic waves represent.


The process of diffraction was first encountered in the study of light phe-
nomena and found its strongest experimental manifestation in Young’s slit
experiments. In addition, the wavefield concepts of coherence and interfer-
ence will prove to be very useful in confronting practical challenges that arise
in seismic diffraction imaging. To provide a solid foundation and a deeper
intuition for diffraction phenomena in general, the following passages briefly
review some important principles of optics that will turn out to be useful in
seismic applications. For a more detailed historical overview and a deeper
treatment of classical optics, I refer the reader to the classic text book of
Born and Wolf (2013) or the recently published work by Klem-Musatov
et al. (2016a).

2.1 Optical roots


According to historical accounts Leonardo da Vinci was the first to have
observed that some portion of the light enters the shadow of an obstacle,
which is not explained in the frame of geometrical optics (Meyer, 1934).
However, it was only centuries later that Francesco Maria Grimaldi
(1665) sought a first systematic treatment of the process and coined the
term diffraction to refer to it (e.g., Cecchini & Pelosi, 1990). Christiaan
Huygens, one of the first proponents of the wave theory of light then intro-
duced a construction, in which arbitrary wavefronts could be thought of as
envelopes of many elementary circular wavefronts originating from imaginary
point sources that are locally excited by the incoming energy. A fundamental
limitation of the Huygens construction was the fact that only wavefronts—
likewise a concept complementary to the ray picture—was concerned.
Although ray-theoretical, asymptotic (infinite-frequency) solutions to the
wave equation have proved to be very successful in their range of applicabil-
ity, diffraction can only be accounted for, if finite-frequency wavefields
are considered. After Young’s famous slit experiments, which led to the
development of a first theory of interference, it was Fresnel who in his memoir
combined Young’s treatment with the wavefront envelope construction to
arrive at a first complete (classical) wave theory of light that is principally
capable of explaining all observable phenomena (e.g., Born & Wolf,
2013). The resulting Huygens–Fresnel principle simply states that arbitrary
wavefields and their interaction with structural changes can be described
by viewing them as a superposition of elementary wavefields emitted by
point sources.
6 Benjamin Schwarz

It is a unique and very powerful property of the wave equation that


the sum of separate solutions likewise forms a solution. This superposition
principle can be physically observable, in the form of noticeably different
wavefield components interfering with each other, but can also be of concep-
tual nature, in that coherent wavefields of arbitrary shape can be thought of
as being composed of simple ones. It is exactly this question of observability
that turns out to be at the heart of seismic diffraction imaging. Building on
the unifying work of Fresnel, at the end of the 19th century, Kirchhoff
provided a solid and practical mathematical foundation that is still influenc-
ing a variety of imaging applications across different disciplines (e.g., Daniel,
Moreau, & Nicol, 2003; Lyrintzis, 1994; Zhuge, Yarovoy, Savelyev, &
Ligthart, 2010). Kirchhoff’s integral formulation and its adaptation to
seismic problems can be considered one of the major milestones, as even
nowadays a majority of seismic depth imaging workflows are still based
on it (e.g., Etgen et al., 2009). It can be viewed as a practical approximation
to the Huygens–Fresnel principle and can be formulated as a generalized
diffraction stack
Z Z
∂Dðx, tÞ
I ðξÞ ¼ w δ½t  tdiff  dt dx, (1)
∂t

where I denotes the reconstructed image amplitude. In Eq. (1), ξ ¼ (x0, z0)
is the image point in depth, whereas ξ ¼ (x0, t0) represents its counterpart
in time migration (e.g., Hubral, 1977; Schleicher, Tygel, & Hubral, 1993a;
Schneider, 1978). In practice, surface observations of a scalar wavefield D
(representing, for example, pressure) are integrated over a limited spatial
aperture, ideally corresponding to the projected first Fresnel zone. To
arrive at the correct source pulse after this summation, somewhat not
intuitively, the temporal derivative of the wavefield rather than the
wavefield itself needs to be considered (Newman, 1990). In Eq. (1) δ rep-
resents Dirac’s delta function, w denotes an amplitude weight, and x is a
vector parameterizing the lateral positions of sources and receivers located
in the limited surface aperture considered in the migration. The Kirchhoff
integral thus links the wavefield in depth (or image time, indicated by
the subscript 0) with measurements carried out at the acquisition surface.
The traveltime tdiff ¼ ts + tg corresponds to diffraction at a point-scatterer
and ts and tg are the constituent one-way traveltimes linking the image
point ξ with the surface positions of the sources and receivers, respectively
(compare Fig. 1A).
An introduction to seismic diffraction 7

A B

Fig. 1 A comparison of the underlying imaging principle of Kirchhoff migration


(A) with the direct observation of a physically real scattered wavefield (B). While for
reflection, Eq. (1) amounts to Huygens’s envelope construction, the Kirchhoff operator
is accurate for the diffraction case. Rs and Rg denote local wavefront curvatures at source
and receiver, which can be defined geometrically, when time migration is concerned.

By incorporating a weight function w and taking the temporal derivative


of the wavefield in Eq. (1), amplitude-preserving migration with the
Kirchhoff integral is enabled (Bleistein, 1987; Schleicher et al., 1993a).
However, as was demonstrated in the context of diffraction imaging, if only
structural information is desired and absolute amplitudes are not of strict
importance in the migrated result, an unweighted diffraction stack (w ¼ 1
and ∂D=∂t ¼ D) leads to drastic simplifications (Khaidukov et al., 2004;
Moser & Howard, 2008). Owing to the fact that traveltimes are invoked,
which in practice are typically computed by means of ray tracing or
solving the eikonal equation, Eq. (1) is intrinsically of approximative nature
(e.g., Etgen et al., 2009). However, it was found by other authors and
will become apparent in the following that Kirchhoff migration is well-
equipped for the task of diffraction imaging. Fig. 1 compares the under-
lying imaging principle, which is of conceptual nature for reflections,
with the case of observable diffraction for which the Kirchhoff operator
provides the accurate kinematic description. In A.1 a detailed discussion
and derivation of Kirchhoff’s diffraction integral (1) is provided.
Likewise originating from optics, Babinet’s principle states that a small
localized scattering structure and a small gap, i.e., a small disruption in an
otherwise smooth material interface, are equivalent in that they both cause
a kinematically equivalent diffraction pattern. Its importance becomes
immediately obvious when typical geological structures are considered.
Small gaps, faults or intrusions constitute important structural features of
the Earth’s crust and are likewise honored by Kirchhoff’s theory. Figs. 2
and 3 illustrate the validity and importance of the Huygens–Fresnel
principle for synthetic wavefields diffracted at a varying number of point
scatterers, simulated in the Fourier domain with accurate time integration
8 Benjamin Schwarz

A Lateral distance (km) B Lateral distance (km)


2 4 6 2 4 6

2 2
Depth (km)

Depth (km)
4 4

Wavefield (10 scatterers) Wavefield (25 scatterers)


C Lateral distance (km) D Lateral distance (km)
2 4 6 2 4 6

2 2
Depth (km)

Depth (km)

4 4

Wavefield (50 scatterers) Wavefield (300 scatterers)


Fig. 2 Wavefields diffracted at (A) 10, (B) 25, (C) 50, and (D) 300 point scatterers located
at a constant depth of 3 km. While individual diffraction patterns can be distinguished
for a limited amount of scatterers, the validity of the Huygens–Fresnel principle can be
observed for a sufficiently high scatterer count.

via the rapid expansion method (REM, Tessmer, 2011). Depending on the
seismic wavelength, the process of reflection and transmission can be fully
reconstructed, if a sufficient number of point scatterers (exciting an equal
number of elementary waves) is considered (Fig. 2). Emphasizing the special
role the process of diffraction plays, it can be observed in Fig. 3 that the tran-
sition from concept to physical reality likewise translates into the data (time)
domain. Circular wavefronts in object space or the geology appear as hyper-
bolic features in data space or the image that can be clearly distinguished for a
limited amount of scatterers. In full analogy, Babinet’s equivalence principle
An introduction to seismic diffraction 9

A Lateral distance (km) B Lateral distance (km)


0 2 4 6 0 2 4 6

2 2
Time (s)

Time (s)
3 3
Data (10 scatterers) Data (25 scatterers)
C Lateral distance (km) D Lateral distance (km)
0 2 4 6 0 2 4 6

2 2
Time (s)

Time (s)

3 3
Data (50 scatterers) Data (300 scatterers)
Fig. 3 Wavefields diffracted at (A) 10, (B) 25, (C) 50, and (D) 300 point scatterers in the
data (time) domain, where circular wavefronts translate to hyperbolic signatures,
corresponding to projections to the registration surface (compare Fig. 2). In the transi-
tional regime (for the case of 50 scatterers) diffraction tails do not fully interfere destruc-
tively resulting in a nonnegligible contribution to the coda of the main event.

is likewise illustrated in Fig. 4 by means of the accurate simulation of diffrac-


tion occurring at a single scatterer (Fig. 4A) and a single slit in an otherwise
continuous interface (Fig. 4B).

2.2 Diffraction types and properties


In the framework of conventional geometrical ray theory, the process of
diffraction is not explained and can only be postulated (Keller, 1962). In
fact, diffraction can be summarized and defined as the collection of all
10 Benjamin Schwarz

A Lateral distance (km) Lateral distance (km)


2 4 0 2 4 6

2
2
Depth (km)

Time (s)
4

3
Wavefield (1 scatterer) Data (1 scatterer)
B Lateral distance (km) Lateral distance (km)
2 4 0 2 4 6

2
2
Depth (km)

Time (s)

3
Wavefield (1 slit) Data (1 slit)
Fig. 4 Illustration of Babinet’s principle of correspondence between diffraction at (A) a
localized point scatterer, and (B) a gap in an otherwise continuous interface,
corresponding to a seismic slit experiment. Both processes, albeit causing different ampli-
tude behavior, result in equivalently shaped wavefields in space (left) and time (right).

observations that appear forbidden in the classical ray picture (Born & Wolf,
2013). However, once their occurrence is postulated for small-scale
structural features, in the far field diffracted waves, just like reflections or
direct arrivals, can often be reliably approximated by rays (Keller, 1962;
Klem-Musatov, 1994). Following the systematic classification of Keller
(1962), different types of diffraction can be distinguished. Some of the most
important diffraction phenomena are illustrated in Fig. 5). While the classic
text book on diffraction imaging by Klem-Musatov (1994) closely follows
the definitions provided by Keller (1962), there sometimes appears to be
An introduction to seismic diffraction 11

A B C

D E

Fig. 5 A selection of different diffraction types important in seismic investigations.


Shown are (A) the highly symmetric case of point diffraction (scattering), as well as dif-
ferent varieties of diffraction at edges ((B), (C), and (E)). The special case of surface dif-
fraction (D), occurring when incident wavefields graze a locally (partially) smooth
interface, is of particular importance in earthquake seismology (Aki & Richards, 2002).

conflict in how the process of diffraction is defined in the geophysical liter-


ature. In the seismological community, for example, diffraction at the
Earth’s core-mantle boundary can routinely be observed (compare Fig.
5D). According to Keller (1962), this observation can be identified with
the process of surface diffraction (Aki & Richards, 2002). However, it is noted
by Keller (1962) that the same waves are also referred to as creeping waves in a
different context. So it can be concluded that terminology needs to be
treated with care to prevent confusion.
Similar to the example of surface diffraction, depending on the context,
it is widely debated whether edge diffraction, i.e., diffraction occurring at sharp
edges (Fig. 5B or C), or point diffraction occurring, for example, at structure-less
point-like objects (Fig. 5A) can be considered more fundamental. Without
intending to add to this debate, I believe it is somewhat helpful to consider
earthquakes, which are the most energetic natural sources of seismicity and
have arguably been studied much more intensely than seismic diffraction
processes. In the far-field, which corresponds to the ray-theoretical realm
of validity, the concept of a point source is useful, whereas in the near
field, ray densities increase significantly leading to the need for corrections
to be applied. The same holds true for diffraction. The notion of a point
12 Benjamin Schwarz

diffraction, in a sense, can be related to the ray-theoretical view underlying


Huygens elementary wavefront construction, which ignores variation in
amplitude and phase, while the process of edge diffraction is physically
more complete and honors the full wavefield properties, which lets it appear
more closely related to Fresnel’s combination of interference with the
wavefront picture.
For the sake of simplicity, and to illustrate the remarkable robustness and
usefulness of Kirchhoff migration, all the examples presented in this chapter
will utilize the point diffraction or scattering picture of uniform radiation,
which is well describable through wavefronts. It is important to note that
in realistic media the concept of omni-directional scattering often represents
an idealization. Diffraction at edges and at localized disruptions of geological
horizons represent more common causes and the amplitude behavior is often
more complicated (e.g., Greenhalgh & Manukyan, 2013). However, from
a wavefront perspective, i.e., taking a kinematic viewpoint, it can be argued
that more intricate diffraction patterns—following the superposition
principle—can be separated into two parts, one honoring Snell’s law, the
other representing a highly illuminating, uniformly radiating component. It
is this illuminating component that encodes very localized (high resolved)
information about the scattering geometry and provides superior illumination
that can be utilized in seismic imaging.
It can generally be argued that, in contrast to conventionally utilized
reflected energy, diffraction can only be fully understood, if wave fields rather
than single isolated recordings are considered. Its occurrence is intimately
linked to abrupt, very localized changes in medium properties on the order
of the predominant seismic wavelength, laterally and vertically. Consequently,
in contrast to reflection which occurs when abrupt changes are encountered
locally in only one spatial direction, diffraction is caused by discontinuous
features that have a radius of curvature that is small compared to the seismic
wavelength. Two of the most characteristic properties of diffracted fields
originating from point-like features or edges are their overall weakness
(due to the rapid geometrical spreading) and the fact that, amplitude and phase
considerations aside, their overall shape in space and time is essentially inde-
pendent of the encountered discontinuity. Although, again for simplicity,
examples will be restricted to two dimensions, it is important to appreciate
that diffraction is intrinsically three-dimensional and that more types can be
distinguished spatially (compare Fig. 5E). However, for the same reasons as
before, the processes of point or edge diffraction remain particularly useful
in that the connected wavefronts radiate uniformly and can be viewed as a
An introduction to seismic diffraction 13

physical (observable) manifestation of elementary waves originating from sec-


ondary sources whose locations are related to Earth structure.
Returning to Figs. 2 and 3, it becomes clear that in the Huygens–Fresnel
picture of interference, a terminating reflector, up to the point of termina-
tion, can be viewed as a collection of a large number of closely spaced
point diffractors, whose superposition results in the reflected wavefield.
At the termination point, however, the destructively interfering counterpart
is missing, resulting in an observable diffracted wavefield that albeit its faint-
ness can be utilized for further processing. In this way diffractions indeed
share some crucial characteristics of Green’s functions (Dirk Gajewski,
personal communication) and it comes as no surprise that migration tech-
niques, dealing with squeezing out the propagation effects and leaving only
structural information in the seismic image, have been shown to be naturally
equipped constructs for diffraction imaging in time and depth (e.g., Khaidukov
et al., 2004; Moser & Howard, 2008; Silvestrov, Baina, & Landa, 2016). As
will be shown in the next section coherence represents a collective wavefield
property that helps to couple the useful wavefront picture to waveform
characteristics.

3. Diffraction separation and imaging


Research in seismic diffraction has so far mostly been concerned with
two overarching goals: (1) the migration-type imaging of small-scale discon-
tinuities and related to that, (2) the successful separation of the often orders
of magnitude weaker diffracted contributions from the rest of the wavefield.
Starting with the pioneering works on wavefield imaging (Hagedoorn,
1954), crustal fault characterization (Krey, 1952; Kunz, 1960), and the
development of a more rigorous theoretical framework (Berryhill, 1977;
Klem-Musatov, 1994; Klem-Musatov & Aizenberg, 1984, 1985; Trorey,
1977, 1970), the problem of interference of diffracted with the stronger
reflected wavefield components was soon recognized. As a result, first strat-
egies for their extraction and/or practical utilization started to emerge in the
1970s and 1980s (Hubral, 1975; Kanasewich & Phadke, 1988; Landa et al.,
1987; Schilt, Kaufman, & Long, 1981). Generally, two different schools of
thought exist—one first aims at separating the diffractions directly in the
time domain using some sort of coherence argument (e.g., Bansal &
Imhof, 2005; Berkovitch et al., 2009; Dell & Gajewski, 2011; Fomel,
2002; Fomel et al., 2007; Schwarz & Gajewski, 2017a) followed by subse-
quent focusing; the other suppresses the reflected energy directly during
14 Benjamin Schwarz

migration (e.g., Dafni & Symes, 2017; Klokov & Fomel, 2012; Moser &
Howard, 2008; Yin & Nakata, 2017). While both these mindsets have their
advantages, it can be argued that a distinct time-domain separation step gen-
erally leaves the user with more flexibility. In addition, insufficient knowledge
of the velocity structure does not automatically compromise the extraction.
On the contrary, time-domain prestack diffraction extraction principally
enables dedicated diffraction-driven velocity inversion workflows that utilize
the unique illumination properties of these weak wavefields (Bauer et al.,
2017; Fomel et al., 2007; Santos et al., 2012; Sava et al., 2005).
One of the main benefits of dedicated diffraction processing lies in the
potential for high-resolution imaging of small-scale discontinuities. This
chapter on seismic diffraction is aimed at widening this view to other
potential impactful applications. However, as the former conventional branch
has matured considerably over the past decades and, more importantly,
because it without a doubt adds value in the pursuit of the ultimate goal
of an accurate characterization of the subsurface, this section will briefly
review a straightforward way of achieving goals (1) and (2) in an integrated
fashion. Building on the intuitive notion of wavefronts, the presented
framework is reasonably simple to implement and should therefore be
accessible and reproducible for people in both, industry and academia.
After explaining the inner workings of this method, the section will be con-
cluded with a brief discussion of a simple class of time-domain focusing
operators and the application to a variety of challenging synthetic and field
data examples.

3.1 Coherence and wavefronts


Before the recent rise of ambient noise interferometry in earthquake seis-
mology and seismic exploration, a signal was regarded useful for imaging
if it is coherent. Even surface-related multiple reflections or the effect of
ground roll, both previously considered coherent noise, nowadays serve a
purpose and can be utilized for imaging and inversion. Originating from
optics, coherence is a fundamental property of lasers, which emit very
intense narrow beams of electromagnetic radiation. Owing to their finite
duration, seismic excitations are only partially or locally coherent, but similar
to lasers, through constructive interference, even weak emerging energy can
be amplified through beam forming or stacking (Mayne, 1962). In turn,
it was found by Taner and Koehler (1969) and Neidell and Taner (1971)
that the coherence of a seismic wavefield can be systematically investigated
by means of trial data summations within a predefined aperture. As a discrete
An introduction to seismic diffraction 15

and reduced version of the Kirchhoff integral, such a data summation over all
traces located at x in the neighborhood of a reference location x0 can be
written as
1X n
Ck ðx0 , t0 Þ  D½xi , tk ðxi Þ: (2)
n i¼1

where Ck denotes the reconstructed coherent wavefield, D is the input data


volume and t0 stands for the reference traveltime for which the reconstruction
is performed. The index k is introduced to honor conflicting coherent con-
tributions with different emergence slopes and curvatures that might intersect
at the reference data point (x0, t0). The trajectory tk(xi) describes the
wavefront-consistent traveltimes of the kth locally coherent event, as they
are observed at neighboring measurement locations xi. Corresponding to w
in Eq. (1) the division by the number of traces entering the data fold in the
aperture n represents a normalizing weight of the summation. As is illustrated
in Fig. 6A, when diffractions are concerned, the presence of conflicting
event dips is very likely as they usually are not an isolated phenomenon
but rather appear in numbers and tend to interfere with each other and more
energetic reflected contributions. Consequently, the data can be written as a
superposition of interfering coherent wavefields and uncorrelated noise N ,
X
Dðx0 ,t0 Þ ¼ Ck ðx0 , t0 Þ + N ðx0 , t0 Þ: (3)
k

Fig. 6 A simple synthetic example illustrating the typical challenge of interference that
has to be confronted for targeted diffraction imaging. Shown are (A) the input data,
(B) the reflection reconstruction via stacking, and (C) the adaptive subtraction of
(B) from (A).
16 Benjamin Schwarz

Within the framework of automated data enhancement and wavefield recon-


struction (e.g., Baykulov & Gajewski, 2009; Gelchinsky, Berkovitch, &
Keydar, 1999a, 1999b; J€ager, Mann, H€ ocht, & Hubral, 2001; Xie &
Gajewski, 2017) this was considered a fundamental problem of coherent
summation. However, there exist several attempts of honoring multiple event
dips, ranging from operator extrapolation (Baykulov & Gajewski, 2009;
H€ ocht, Ricarte, Bergler, & Landa, 2009) to Fresnel volume inspired superpo-
sition strategies (Walda & Gajewski, 2017; Xie & Gajewski, 2018). All of these
approaches are phrased as optimization problems, in which, conventionally,
the semblance norm
( )2
X Xn
D½xi , tk ðxi Þ
1 δt i¼1
S k ðx0 , t0 Þ ¼ X Xn , (4)
n
D2 ½xi ,tk ðxi Þ
δt i¼1

is maximized using either local or global solvers (or a combination of the


two). The additional summation in expression (4), performed within a time
window δt around the reference time t0, honors the band-limited nature of
the signal under investigation. The semblance norm has gained strong atten-
tion over the past decades and is still frequently favored over other coherence
measures. This can partly be explained by its relative ease of implementation,
its computational efficiency and its overall robustness. From a physical per-
spective it can be viewed as the normalized beam energy or the ratio of sta-
cked to overall intensity contained in the data. Therefore it intuitively
reflects a wavefield’s potential to self-interfere.
If sufficiently local apertures are considered, a closed-form relationship
between the traveltimes tk(xi) and the reference t0 can be found. For
the most general 3-D prestack case, paraxial two-way traveltimes for
g g g
source separations Δxsi ¼ xsi  xs0 and receiver separations Δxi ¼ xi  x0
can be computed via
g g g
tk2 ðΔxsi , Δxi Þ ¼ ðt0 + psk Δxsi + pk Δxi Þ2
g g g
+ t0 ½ðΔxsi ÞT Msk Δxsi + ðΔxi ÞT Mk Δxi  (5)
sg g
+ 2 t0 ðΔxsi ÞT Mk Δxi ,
g g
where psk and pk represent two-dimensional slope vectors and Msk and Mk
are 2  2 matrices encoding the observed curvature of the considered kth
An introduction to seismic diffraction 17

coherent event on the source and on the receiver side, respectively


(Schleicher, Tygel, & Hubral, 1993b). Eq. (5) corresponds to a second-order
Taylor series expansion of the squared traveltime. The components of the slope
g
vectors psk and pk represent horizontal slowness and have the dimension s/m.
For common subconfigurations, like, for example, the zero-offset section (i.e.,
the poststack domain), the common-source or the common-receiver gather,
the hyperbolic traveltime formula (5) reduces to formally equivalent expres-
sions that relate to auxiliary one-way wave propagation and, consequently,
can likewise be employed in passive seismic investigations (e.g., Diekmann,
Schwarz, Bauer, & Gajewski, 2018; Schwarz, Bauer, & Gajewski, 2016;
Schwarz & Gajewski, 2017b). So as a by-product, the use of Eq. (5) in con-
junction with Eq. (4) allows for the automated extraction of first- and second-
order attributes of the event’s shape, which can be directly related to the slope
and curvature of the emerging wavefront causing the observed trace-to-trace
traveltime differences.

3.2 Adaptive reflection subtraction


Fig. 7 shows the estimated maximized semblance as well as the absolute
wavefront emergence angle and a normalized version of the curvature attri-
bute for the 2D synthetic example introduced in Fig. 6A. All three attributes
were estimated automatically without any user involvement aside from the
predefinition of the semblance time window δt and the size of the data

Fig. 7 By-products of the suggested coherence analysis: (A) maximized semblance


norm, (B) absolute dip angle, and (C) normalized curvature of the emerging wavefront.
Reflections are generally characterized by moderate dips and curvatures, which helps to
additionally discriminate between both wavefields.
18 Benjamin Schwarz

aperture. As can be observed throughout, the reflected energy, characterized


by small curvature and emergence angle estimates, is consistently favored by
the coherence analysis. Accordingly, as will be more thoroughly demon-
strated later, we often have to a good approximation

1X n
Cref ðx0 , t0 Þ  D½xi , tref ðxi Þ, (6)
n i¼1

where Cref denotes the coherent reflected contribution at data point (x0, t0)
and tref is the corresponding traveltime trajectory. Rather than aiming at
honoring the interfering wavefields’ full complexity, which is computation-
ally demanding and challenging to implement successfully, it was recently
suggested by Schwarz (2019) to make use of the undesired directional filter
characteristics imposed by expression (2) and, consequently, the semblance
(Eq. 4) to extract diffractions from pre- and poststack seismic, as well as
ground-penetrating radar data. In contrast to previously proposed strategies
(e.g., Bakhtiari Rad et al., 2018; Berkovitch et al., 2009; Dell & Gajewski,
2011) which directly target the weak diffracted wavefield, a conventional
reflection stack is suggested, followed by its subsequent adaptive subtraction
from the full input wavefield,

Cdiff ðx0 , t0 Þ + N ðx0 ,t0 Þ ¼ Dðx0 ,t0 Þ  α0 Cref ðx0 , t0 + τ0 Þ: (7)

To accommodate data imperfections, local waveform variations and more


general deviations from the underlying assumption of local coherence,
the subtraction process in Eq. (7) is formulated in an adaptive fashion through
the introduction of perturbative time shifts τ0 and local amplitude scaling coef-
ficients α0. In order to preserve even the faintest diffracted signatures, the
adaptation, just like the coherent summation and the semblance estimation,
is performed within a data aperture. The residual misfit of this adaptation,
as well as the scaling coefficient and time shift fields are displayed in Fig. 8.
All three quantities show distinct correlations in regions where diffractions
and reflections heavily interfere. In cases where diffraction is the dominating
contribution, as might, for example, be the case at the high-impedance con-
trast observed at complex sediment–salt interfaces, the estimated slope and
curvature attributes can be used to design diffraction-preserving wavefront
filters that can objectively inform the coherent subtraction. For more details
on the adaptation step and implementational aspects I refer the reader to the
work by Schwarz (2019).
An introduction to seismic diffraction 19

Fig. 8 By-products of the adaptive subtraction of the reflection stack: (A) least-squares
residual misfit of adaptive reflection stack and input data, (B) optimized absolute ampli-
tude scaling, (C) absolute time shift estimates minimizing the misfit. Bright colors indi-
cate regions where strong reflection–diffraction interference is sensed by the analysis.

In summary, the results of the automated coherent summation and the


subsequent adaptive subtraction alongside the full wavefield data input are
presented in Fig. 6. To demonstrate that this methodology is reasonably
flexible to handle different data configurations, the results of the coherence
maximization and the diffraction separation, performed on a set of neighbor-
ing common-source gathers simulated for the complex Sigsbee 2a synthetic
model, are shown in Fig. 9. Again, the coherence analysis consistently favors
the reflected contributions, resulting in a successful amplitude-preserving
separation of the diffracted wavefield in the prestack domain.

3.3 Diffraction focusing


Kirchhoff depth migration uses ray tracing or eikonal solvers to compute
diffraction traveltimes for trial subsurface image locations ξ ¼ (x0, z0) and
known velocity structure. In contrast to that the discussed multidimensional
coherence analysis, which can be expressed by a very similar mathematical
construct, is based on an adaptive and flexible traveltime kernel (Eq. 5) that is
likewise suited to describe diffracted and reflected contributions. While the
former, even when formulated in the time domain, is in need of detailed
knowledge of a velocity field for computing diffraction traveltimes, the latter
in turn can facilitate the estimation of velocity structure from the estimated
wavefront attributes (Bauer et al., 2017; Duveneck, 2004).
In fact, when considering the diffraction case and a zero-offset reference
g
frame, i.e., midpoint and half-offset ðΔxsi ¼ Δxi Þ rather than source and
20 Benjamin Schwarz

Fig. 9 Prestack data (A), estimated maximized coherence (C), and diffraction separation
result (B) for a collection of common-source gathers of the Sigsbee 2a synthetic dataset.

receiver coordinates and confinement to the diffraction case and vanishing


g
wavefront inclination (i.e., psk ¼ pk ¼ 0), the number of degrees of freedom
reduce drastically and the traveltime operator (5) can be used for Kirchhoff-
type time migration (Zhang, Bergler, & Hubral, 2001). However, previous
An introduction to seismic diffraction 21

investigations suggest that double-square-root-type expressions for the


traveltime are more suited to accurately describe and incorporate the response
from highly curved subsurface features and diffracting structures. Conse-
quently, a range of extensions to the zero-offset subset of operator (5) have
been introduced in recent years (e.g., Fomel & Kazinnik, 2013; Landa,
Keydar, & Moser, 2010; Schwarz, Vanelle, Gajewski, & Kashtan, 2014).
While these differ in the way the wavefront attributes appear in the expres-
sions, they share the overall double-square-root shape and are equally accu-
rate for the limiting diffraction case (Schwarz & Gajewski, 2017c). In fact,
if parametrized consistently they were demonstrated to provide equivalent
accuracy and can be considered largely equivalent descriptions (Schwarz &
Gajewski, 2017b; Walda, Schwarz, & Gajewski, 2017). For an effective aux-
iliary medium they incorporate the conventional Kirchhoff time migration
formula
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2 rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2
t02 ðΔxmi  hi Þ t2 ðΔxmi + hi Þ
tdiff ðΔxmi , hi Þ ¼ + + 0+ , (8)
4 v 4 v

where Δxmi ¼ xmi  x0 is the midpoint displacement and hi denotes half the
source-receiver offset of the ith trace with respect to the time image location
ξ ¼ (x0, t0). The migration velocity v corresponds to the root-mean-square
velocity which can be directly calculated from the wavefront attributes esti-
mated during coherence analysis. Analogous to conventional Kirchhoff
depth migration, the diffraction traveltime defined in Eq. (8) can be divided
into two distinct contributions, one corresponding to the source the other to
the receiver leg of the full ray-path.
In academia, target depths typically exceed the maximum source-
receiver distances by a factor of two or even considerably more. As a result,
crucial lateral illumination is not recorded, thereby intrinsically limiting lat-
eral resolution. Also, detailed a priori information required for successful pres-
tack depth imaging is rarely available, which makes the related computational
demands difficult to justify. In general, due to the aforementioned limitations,
pragmatic yet stable and computationally efficient imaging approaches with
a good performance with respect to low signal-to-noise ratios are ideally
suited in these scenarios. Despite its unique properties, the diffracted wavefield
is still rarely utilized in academic practice. The main benefit of a distinct
time-domain separation strategy is that firstly, velocity uncertainties do not
immediately affect the extraction, and secondly, the subsequent imaging
workflow can either operate in time, which decreases accuracy but increases
computational efficiency and stability, or in depth, if the circumstances are
22 Benjamin Schwarz

favorable. To ensure comparability and to demonstrate that diffraction imag-


ing does not have to be extremely sophisticated to produce highly resolved
and useful images, the following examples utilize a time migration kernel
and, consequently, make use of expressions (1) and (8).

3.3.1 Diffraction vs reflection images


As a first simple example of data-driven diffraction imaging, Fig. 10 shows
a close-up of the Sigbsee 2a synthetic dataset which was made available to
the public by the Subsalt Multiples Attenuation and Reduction Joint
Venture (SMAART JV). Aside from an extended salt structure mimic-
king the Sigsbee escarpment below the Gulf of Mexico, the model
depicted in Fig. 10A reveals complicated localized structures in the form
of faults, and a horizon of small-scale velocity perturbations which are
deemed to cause diffraction patterns in the data. After wavefield separa-
tion is performed through the coherent strategy described above, the
(conventional) full-wavefield image, comprised of reflected and diffracted
contributions (Fig. 10B), is compared with a reflection-only migration
(Fig. 10C). The reflection image lacks resolution in essentially all regions
that deviate from the picture of a smoothly layered medium. Although
this reconstruction is clean in appearance and might add value in tracking
horizons, it lacks sufficient detail on where the geodynamically important
faults are located in the model.

A Lateral distance (km) B Trace number C Trace number


0 2 4 0 200 400 0 200 400

4 4

4 2.5
Depth (km)

Time (s)

2.0

1.5
6 6

6
Velocity model (km/s) Full-wavefield image Reflection image
Fig. 10 Comparison of (A) an excerpt of the complex Sigsbee 2a velocity model with
(B) the full-wavefield image and (C) the reflection-based image reconstruction. The lack
of diffracted energy results in an overall smoother appearance of the reflection image,
which, however, lacks crucial detail on model complexity.
An introduction to seismic diffraction 23

It has been emphasized by the authors of several previous works that


reflection and diffraction imaging are supposed to complement each other
and that diffraction images alone will not necessarily be of value to an inter-
preter (e.g., Khaidukov et al., 2004; Moser & Howard, 2008). Confirming
this notion of complementarity, the corresponding diffraction image, pres-
ented in Fig. 11A, reveals detailed features of the fault system, provides an
account of the localized perturbations in the lower half of the model, and
even manages to accurately reproduce a detailed pattern of distinguishable
diffracting edges (indicated, for example, by the apparent phase reversal near
the edge). Although these diffractions were unintentionally introduced
through the too coarse definition of the synthetic model, it can be concluded
that the diffraction separation was sufficiently successful to enable imaging
of features at a resolution approaching the model’s spatial discretization.
Of course, the diffracted wavefield has not only virtue on its own, but also
contributes to the perceived sharpness of the full-wavefield imaging result
shown in Fig. 10B. However, the identification and detailed structural char-
acterization of the model coarseness and the highly relevant extent of the
fault system is much easier to accomplish when the diffraction image is
considered.

3.3.2 Imaging attributes


When only structural imaging is concerned, a noticeably improved signal-
to-noise ratio can be achieved when the semblance norm (Eq. 4) rather than
the conventional wavefield sum is estimated during migration (e.g., Heincke,
Green, van der Kruk, & Willenberg, 2006; Landa & Keydar, 1998; M€ uller,
2000). This is generally possible for Kirchhoff-type migrations, as summation
over an aperture is directly related to the definition of semblance (Neidell &
Taner, 1971). It is important to note the critical difference in treatment
that reflected and diffracted contributions experience during Kirchhoff migra-
tion. Reflections are merely repositioned and the summation operator is only
tangent to the wavefield (meaning that only a subset of the fold within the
aperture contributes constructively). Diffractions on the other hand—accurate
velocity information provided—fully contribute to the resulting migrated
amplitude. This seems reasonable, as diffracted energy, due to its more uni-
form radiation, was distributed over a larger area and, consequently, must
be gathered and refocused using a larger aperture. Semblance, however, rep-
resents a normalized quantity and turns out to be largely independent of the
absolute amplitude of the signal, which leads to a natural suppression of
reflected energy during migration.
24 Benjamin Schwarz

A Trace number B Trace number C Trace number


0 200 400 0 200 400 0 200 400

4 4 4
Time (s)

Time (s)

Time (s)
6 6 6

Diffraction image Focusing semblance N-th root semblance


Fig. 11 Comparison of (A) the diffracted wavefield image with reconstructions based on
(B) the conventional semblance norm and (C) its nonlinear (10th root) version. Both
coherence images are cleaner in appearance, while residual artifacts are least
pronounced in (C).

Upon closer inspection, the migration of diffraction-only data strongly


resembles the procedure of passive-source back-projection which relies on
focusing and provides a robust and highly resolved means of detection and
localization of natural seismicity in well-instrumented areas in earthquake
seismology studies (Ishii, Shearer, Houston, & Vidale, 2005). Inspecting
the earthquake seismologists’ toolbox, the notion of the so-called Nth root
stack, which is often used for the detection of weak events in very challenging
low signal-to-noise scenarios (Rost & Thomas, 2002), complementary to the
varimax norm (Fomel et al., 2007; Wiggins, 1978), can serve to introduce
nonlinear selectivity and improve robustness:
( )
X X n pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2
N
D½xi , tk ðxi Þ
1 δt i¼1
SNk ðx0 ,t0 Þ ¼ X X n n pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffio2 :
(9)
n N
D½xi , tk ðxi Þ
δt i¼1

In Fig. 11 this Nth root semblance is compared with conventional diffracted


wavefield focusing and the linear semblance norm. As expected, both
coherence measures reveal clearer delineation of the conjugate faults and
the highly localized features in the bottom half of the model (compare
Fig. 10A), while indicating to be less prone to artifacts. In the 10th root
semblance image (Fig. 11C), an overall higher signal-to-noise ratio and
An introduction to seismic diffraction 25

improved reconstruction of the edge diffractions due to the model dis-


cretization in the shallower regions can be observed. To prevent undesired
destructive interference due to an amplitude reversal at the apex, which typ-
ically can be encountered for edge diffraction, both semblance versions
are evaluated twice, once with and once without correcting for this phase
change. In the final image, for each sample, the maximum value of these
two fields is displayed. Aside from normalized quantities like the linear or
Nth root semblance norms, tools from complex trace analysis, such as the
trace envelope and its first- and second-order derivatives (Fig. 12A–C)
can be employed to additionally sharpen the diffraction imaging results.
Due to the perceived robustness in the presented example, in the following,
diffraction images were achieved using the newly introduced Nth root sem-
blance measure (9).

3.4 Faults, fractures, and unconformities


From the previously conducted systematic comparison of full-wavefield and
diffraction images it can be concluded that—and this is not entirely
unexpected—seismic wave diffraction is responsible for the reconstruction
of a majority of the laterally resolved features of the model. In fact, as these
comparisons likewise suggest, the diffracted wavefield provides an overall
sense of sharpness, which cannot be achieved with the reflections alone.
Various other authors have commented on the high-resolution imaging
potential of diffractions and essentially all of them conclude in their respec-
tive analysis that a successful separation is crucial for these highly resolved

A Trace number B Trace number C Trace number


0 200 400 0 200 400 0 200 400

4 4 4
Time (s)

Time (s)

Time (s)

6 6 6

Focusing envelope First envelope derivative Second envelope derivative


Fig. 12 The application of complex trace analysis tools can help to additionally sharpen
diffraction images. Displayed are the trace envelope (A), alongside its first (B), and
second derivatives (C).
26 Benjamin Schwarz

features to be distinguishable from artifacts that arise from data imperfec-


tions, too coarse spatial wavefield sampling and inaccuracies in the retrieved
subsurface velocity model (Grasmueck, Weger, & Horstmeyer, 2005;
Khaidukov et al., 2004; Moser & Howard, 2008; Silvestrov et al., 2016).
A comparison of full-wavefield and reflection image, as for example pres-
ented in Fig. 10, suggests that, in a complementary sense, the separated
reflected wavefield is overall less artifact prone and appears smoother and
somewhat cleaner, which among other things might prove useful for the
reliable tracking of horizons in seismic interpretation.
Conversely, diffraction images highlight and emphasize features that are
very localized laterally, like abrupt truncations of the aforementioned hori-
zons, erosional surfaces, the carving imprints of river channels, and, most
prominently, faults and fracture systems cutting through sedimentary strata
or the crystalline basement of the Earth’s crust. An integrated (full-
wavefield) view may provide a somewhat complete picture when velocity
model estimates are accurate and imaging algorithms are sophisticated and
well implemented. However, owing to the general faintness of diffractions,
their voice tends to be overheard and often does not carry sufficient weight
to sustainably have an impact on the image quality and to be accurately hon-
ored in the conventional, reflection-biased processing chain. It is only through
their separation that targeted reflection and diffraction processing can be fully
tailored to the peculiarities and distinct properties and challenges of these
insightful wavefields. As a result, a complementary picture of the subsurface
emerges, that in some sense, may provide more insights than the mere sum
of its parts.

3.4.1 Diffraction in interpretation


From a material science or optics perspective, isolated diffraction patterns are
often directly related to material imperfections. In an Earth science context,
it is exactly these imperfections that are of interest to an interpreter of seismic
images, as they, in some sense, encode the dynamic history of the crust.
Conventional and established attributes such as image coherence or image
curvature (Bahorich & Farmer, 1995; Chopra & Marfurt, 2007; Marfurt
et al., 1998) have proved to bear the potential of being reliable indicators
of faults. However, the successful and precise delineation of discontinuous
features such as faults and erosional unconformities still remains an area of
active research, where recent improvements have been made with the help
of conventional and more specialized image processing and analysis
An introduction to seismic diffraction 27

techniques. Among others these include spectral analysis (Liu & Marfurt,
2007), structural tensor analysis (Wu, 2017; Wu & Hale, 2016), optimal
surface voting (Wu & Fomel, 2018), or deep learning (Araya-Polo
et al., 2017).
Despite the tremendous value these purely image-driven approaches
have delivered, it was recognized early on that they all intrinsically rely in
some way on the success of the preceding imaging step (Berkovitch et al.,
2009; Khaidukov et al., 2004; Moser & Howard, 2008). Diffraction images
are different in that they are naturally coupled to the aforementioned discon-
tinuities. As illustrated in Section 2, diffraction can be viewed as the birth of a
new wavefield that focuses—just like a passive source—at the discontinuous
structure that caused it. In consequence, diffraction images represent phys-
ically derived attribute maps. In addition, the use of robust image functions
such as the suggested Nth root semblance norm (Eq. 9) permit an increase in
the signal-to-noise ratio and reduce artifacts directly at the imaging stage,
which in turn might improve subsequent image and data-driven processing.
Consequently, there exist first attempts to use diffraction imaging as a robust
alternative for seismic interpretation (e.g., Decker, Janson, & Fomel, 2014;
Dell, Hoelker, & Gajewski, 2018; Schoepp, Labonte, & Landa, 2014;
Sturzu, Popovici, Pelissier, Wolak, & Moser, 2014; Tsingas, El Marhfoul,
Satti, & Dajani, 2011; Tyiasning, Merzlikin, Cooke, & Fomel, 2016).
In the following, by means of two challenging yet controlled synthetic
examples, I want to illustrate how robustly dedicated diffraction imaging
can reveal structurally important small-scale subsurface feature whose
details normally tend to be drowned out by the strong amplitude reflection
foreground. The first example is concerned with another close-up of the
Sigsbee 2a synthetic dataset which was also investigated when discussing
the mechanism of diffraction focusing. This time, another part of the model
near the top of the complicated salt body is investigated. Fig. 13 shows the
comparison of the actual seismic velocity grid alongside its reconstruction
based on the diffracted wavefield that was automatically separated before
imaging.
To illustrate the overall robustness of this approach, the image represents
a brute-force constant-velocity (1.6 km/s) focusing result employing the
newly introduced Nth root semblance. As before, aside from the clear delin-
eation of a fault at the top of the salt tip, largely every single stair step in the
discrete velocity grid can be resolved. In addition, the steep flanks of the rug-
ged top-of-salt in the bottom part of the model are clearly revealed through
diffraction imaging. Due to the fact that the Nth root semblance (Eq. 9),
28 Benjamin Schwarz

Lateral distance (km) Trace number


16 18 20 0 200 400

2
4
Depth (km)

Time (s)
2.0

1.5
3 5

Velocity model (km/s) Diffraction image


Fig. 13 Close-up of the top-of-salt region of the complex Sigbsee 2a model (left) and the
corresponding diffraction-based reconstruction (right). To illustrate robustness, a brute-
force constant focusing velocity was used.

just like its conventional linear counterpart, is a normalized quantity,


amplitude-strong features, connected to significant impedance contrasts
appear equally well-resolved and bright as, for example, the stair steps of
the velocity grid, which are the cause of extremely faint signals that can
barely be recognized on individual traces (Heincke et al., 2006; Meyer,
1934; Schwarz, 2019). Although absolute amplitudes can also be used for
imaging, destructive interference does not occur for this case, which often
results in noticeable migration artifacts. Semblance and its nonlinear
counterpart are different, in that they uniquely combine the notion of inten-
sity, which is most familiar to what we perceive with our eyes, with phase
information leading to destructive interference for nonphysical migration
trajectories and uncorrelated noise.
The Sigsbee 2a synthetic dataset was simulated based on a moderate
macro-gradient with small-scale stratigraphic velocity variations. Realistic
but controlled variable-density benchmark datasets likewise are publicly
available and can be used for the investigation of the potential of diffraction
imaging of small-scale density perturbations in the subsurface. One of
these challenging datasets is the 2004 BP velocity benchmark (Billette &
Brandsberg-Dahl, 2005). In contrast to the Sigsbee example that represents
a detailed reproduction of a real-world geologic scenario, the BP model can
be viewed as a hybrid in that it synthesizes a variety of different geologies,
important in the context of hydrocarbon exploration and likewise challeng-
ing to image seismically. The Sigsbee velocity structure can be considered
reasonably complicated and fine-structured, whereas the structural
An introduction to seismic diffraction 29

complexity of the BP model in large parts is not connected to the relatively


simple and well-behaved velocity field, but rather arises from a very detailed
and heterogeneous density distribution.
In Fig. 14—again for a top-of-salt regime—a close-up of the velocity
(Fig. 14A) and the respective density model (Fig. 14B) together with the
diffraction reconstruction (Fig. 14C) are displayed. Like before, to empha-
size the robustness of coherence-based diffraction imaging, a largely inaccu-
rate constant velocity (1.6 km/s) was chosen for wavefield focusing. Despite
the crude assumption of a constant focusing velocity, a stunningly detail-rich
reconstruction of the rugged top of the salt body and small-scale disconti-
nuities within the sedimentary overburden, primarily caused by density
fluctuations, results, which provides a complementary, highly resolved
structural image.

3.5 Diffraction imaging in the field


Generally, there is a consensus that seismic wave diffraction routinely occurs in
realistic media and that these weak signals are likewise routinely recorded in
the field. However, with some impressive exceptions (e.g., Sturzu et al., 2014)
it can be argued that diffraction imaging still does not form an integral part of
the seismic processor’s toolbox (Landa, 2012). One of the reasons is arguably
that the overall weakness of their defocused wavefields, as they appear in the
raw recorded data, makes diffractions barely recognizable and their impact on
image quality is only perceived indirectly. Additionally, the aforementioned
strong interference with the conventionally utilized reflected contributions
makes a successful separation—deemed necessary for targeted imaging—very
challenging in practice. Also, in particular in the framework of non-
commercial surveys with a purely scientific objective, data are generally
imperfect and acquisitions only record a limited portion of the back-scattered
energy, making focusing strategies prone to artifacts whose amplitudes often
approach the strength of the weak diffracted component itself.
Using a noninvasive extraction strategy, Schwarz and Gajewski (2017a)
recently demonstrated that a faint background wavefield can be recovered,
even if single-channel field data recorded in a highly complex geological
setting are considered. Owing to their special illumination capabilities,
the detection and characterization of diffractions, even under these unfavor-
able conditions, can have profound implications for a majority of the
reduced measurements that are nowadays conducted in academia. The
inherent kinematic symmetries and correlations will be more closely exam-
ined in the final section of this chapter.
30 Benjamin Schwarz

A Lateral distance (km)


0 2 4 6 8 10 12 14 16
Depth (km)

2 2.0

1.5

Velocity model (km/s)

B Lateral distance (km)


0 2 4 6 8 10 12 14 16
Depth (km)

2 2.5

2.0

Density model (g/cm3)


Trace number
C
0 500 1000 1500 2000 2500

2
Time (s)

Diffraction image
Fig. 14 Close-up of the top-of-salt velocity structure (A) and the corresponding density
distribution (B) for the challenging 2004 BP velocity benchmark. The diffraction image
(C), like for the Sigsbee 2a example, was gained through brute-force focusing with a
constant velocity.
An introduction to seismic diffraction 31

3.5.1 Industry-scale long-offset acquisition in offshore Israel


Here, I want to follow up on previous work (Schwarz, 2019; Schwarz &
Gajewski, 2017a) and demonstrate that coherent diffraction separation
and focusing, in the simple way it is described in this chapter, can lead to
highly resolved images that provide interpretational value that is highly
complementary to the full-wavefield migrated images which are normally
consulted. To investigate applicability for both extremes of high and low-
fold seismic field data, an industry-scale acquisition and the very reduced
single-channel acquisition already studied by Schwarz and Gajewski
(2017a) are subsequently analyzed. The former constitutes in a sophisticated
multichannel measurement campaign conducted by TGS off the coast of
Israel in the Eastern Mediterranean, whereas the latter was performed by
academia near Santorini in the Aegean Sea. By means of more conventional
processing, both datasets were already demonstrated to contain information
on very intriguing tectonic features either relating to complicated tectonics
that are connected to the formation of massive salt complexes (e.g.,
Netzeband et al., 2006), or arose from nearby persistent volcanic activity
(e.g., H€ ubscher, Ruhnau, & Nomikou, 2015).
A conventional full-wavefield prestack time migration of the industrial
dataset is shown at the top of Fig. 15. As can be deduced from this image,
underlying structures are very heterogeneous and the subsurface can broadly
be divided into at least three distinct regimes. In particular as observed in the
left half of the section, the signature of a massive salt body originating from the
Messinian salinity crisis (Gradmann, H€ ubscher, Ben-Avraham, Gajewski, &
Netzeband, 2005; Krijgsman, Hilgen, Raffi, Sierro, & Wilson, 1999) is over-
lain by a thick sheet of sediments. From these mostly horizontally stratified
sediments upwards to the sea bottom at 1.8 s two-way time a complicated
turbulent complex characterized by chaotic but comparably strong reflectivity
patterns is followed again by horizontally smooth stratification of sediments.
From left to right (corresponding to a progression toward the coast of Israel)
firstly, a thinning of the salt sheet and secondly, tectonic tilting of different
lateral units, separated by large but barely visible faults are revealed. In addi-
tion, it may be argued that the overall reflectivity of the subsurface decreases,
which may in part be attributed to the fact that structures become more
laterally complex and events are generally steeper as the coast is approached.
Displayed at the bottom of Fig. 15 is the corresponding diffraction image,
resulting from full prestack diffraction separation and focusing based on a
simple vertical velocity gradient and the newly introduced Nth root sem-
blance (Eq. 9). Strikingly, and this becomes even more apparent in the
32 Benjamin Schwarz

Fig. 15 The first considered field data example comprising an industry-scale multichannel
acquisition performed by TGS in the Eastern Mediterranean. Displayed are the conven-
tional time-migrated image (top) and the corresponding diffraction image (bottom).

close-up investigated in Fig. 16 (indicated by the frame in Fig. 15), the


diffracted wavefield provides highly resolved structural detail in parts of
the section, where the conventional full-wavefield image lacks resolution.
This is in full accordance with the previously mentioned overall comple-
mentarity of reflection and diffraction imaging. In particular, the chaotic
turbulent features and faults, both of which are important indicators of
dynamic processes, appear bright and detail-rich in the diffraction image,
whereas the undisturbed sedimentary (sub) units are largely transparent to
the diffracted wavefield. For more details on the interaction between the salt
body and the overall tectonic regime, I refer the reader to the work of
Gradmann et al. (2005) and Reiche, H€ ubscher, and Beitz (2014).
An introduction to seismic diffraction 33

Trace number
200 400 600 800 1000

2.5
Time (s)

3.0
Full-wavefield image
Trace number
200 400 600 800 1000

2.5
Time (s)

3.0
Diffraction image
Fig. 16 Close-up of the complicated salt-tectonic regime in which a predominant salt
body is overlain by faulted sediments and a turbulent, highly diffractive layer. Compared
are the conventional full-wavefield migration (top) and the diffraction-only reconstruc-
tion (bottom).

3.5.2 Academic single-channel acquisition near Santorini


To illustrate the versatility and robustness of coherent diffraction imaging, a
very reduced academic marine field dataset, consisting of only a single
near-offset channel, is briefly presented and described. The seismic data were
recorded above the Anydros Basin, close to the island of Santorini in the
Aegean Sea. The region was found to be largely influenced structurally
through prevailing volcanism (H€ ubscher et al., 2015; Nomikou, H€ ubscher,
Ruhnau, & Bejelou, 2016). One of the dominant features is a large submarine
volcano—the Kolumbo seamount—whose activity sustainably shaped the
tectonic landscape of the region and is even thought to have been involved
34 Benjamin Schwarz

in Tsunami generation (Nomikou, Druitt, et al., 2016). As examples, Figs. 17


and 18 show two different sedimentary environments adjacent to the
Kolumbo seamount. Owing to the availability of only a single channel, con-
ventional velocity analysis was rendered impossible, so all images in both fig-
ures were gained through a brute-force approach with a constant velocity of
1.6 km/s. Like before, the diffraction images show the normalized Nth root
semblance, which again proves to be a robust and largely artifact-free diffrac-
tion imaging attribute. In Fig. 17 the crystalline basement of the crust, rising
to the upper few kilometers can be identified as a largely diffuse body. While
the strong velocity contrast between the sediments and this basement can

Trace number
200 400 600 800 1000
0.6

0.8
Time (s)

1.0

1.2
Full-wavefield image
Trace number
200 400 600 800 1000
0.6

0.8
Time (s)

1.0

1.2
Diffraction image
Fig. 17 Single-channel migration (top) and the corresponding diffraction image (bottom)
of a tectonically shaped sedimentary regime near Santorini in the Aegean Sea. Both
images carry complementary information and faults and erosional unconformities appear
clearly reconstructed through the diffracted wavefield.
An introduction to seismic diffraction 35

Trace number Trace number


200 400 200 400

0.8 0.8
Time (s)

1.0 1.0

1.2 1.2

Full-wavefield image Diffraction image


Fig. 18 Close-up of a small sedimentary basin at the flank of the Kolumbo submarine
volcano which represents a dominant feature in the Anydros Basin largely shaped
by volcanic activity. Displayed are the conventional full-wavefield image (left) and its
diffraction counterpart (right).

clearly be identified in the full-wavefield image, the lower portions are dom-
inated by migration artifacts due to the lack of precise velocities and their
severe underestimation during the reconstruction.
The diffraction image on the other hand, owing to the nonlinearity of
the Nth root semblance norm, is generally less prone to artifacts, so internal
unconformities and faults, which are mostly hidden in the conventional
image, become distinguishable. In particular in the right part of the section
at around 1 s two-way time, a detail-rich system of faults can be delineated in
the diffraction image, which remains largely obscured in the conventional
migration. Once more supporting the notion of complementarity, Fig. 18
shows a comparison of conventional migration and diffraction imaging
for a small confined sedimentary basin at the flank of the submarine volcano.
The conventional full-wavefield image shows its strength in the largely
undisturbed sedimentary sequences revealing strong reflectivity, whereas
the diffraction image helps to delineate a sequence of erosional surfaces
and internal faults and fractures, which were previously hidden.
Both field data examples suggest that, provided the spatial sampling of
the emerging seismic wavefield is reasonably dense in at least a single data
subdomain, diffraction imaging becomes feasible and has the potential to
provide valuable additional information particularly relevant for the study
of dynamic processes related to erosion and tectonics. The lack of laterally
resolved velocity information can be considered a fundamental problem in
academic studies, where targets are typically deep and lateral illumination is
36 Benjamin Schwarz

very limited. In addition to permitting high-resolution imaging of faults,


fractures, and unconformities, the diffracted wavefield radiates uniformly
and encodes valuable lateral velocity information which can be extracted
with tailored inversion strategies (e.g., Bauer et al., 2017; Fomel et al.,
2007). Thus, diffractions may help to overcome the traditional shortcomings
of these low-fold recordings. In the following, the unique properties of dif-
fractions and their potential impact on seismic data acquisition and imaging
will be discussed in more detail.

4. Implications and new directions


It was previously demonstrated that with the help of noninvasive
separation techniques, such as the suggested adaptive subtraction strategy,
a rich diffracted wavefield can be utilized for targeted imaging. In many
ways, diffractions turn out to be simpler wavefields, in that with conven-
tional Kirchhoff migration, for example, using reasonably accurate velocity
information, the utilized operators are not only tangent but coincide with
the full response (or a major portion of it). So naturally, Kirchhoff migra-
tion utilizes the full redundancy and coherence of the diffracted wavefield,
which even for short-offset acquisitions and despite their overall weakness,
leads to improved signal-to-noise ratios and a high resolution in the result-
ing images. In addition to this localized highly resolved nature, diffracted
wavefields provide the same illumination that a point source excited at
the diffractor location would provide. Accordingly, diffracted wavefields
in the far-field approximation only encode propagation effects from the
scatterer location to the receivers, which makes them very suitable for
velocity inversion (e.g., Bauer et al., 2017; Fomel et al., 2007; Sava
et al., 2005).
These approaches of diffraction-based velocity model building often
rely on wavefield focusing—a process that was discussed extensively in
the context of imaging in the previous section. Here, I want to draw the
attention to additional striking properties of diffractions, which directly
follow from their uniform radiation. As diffracted contributions often reveal
significant inclination angles when emerging at the registration surface,
spatial aliasing constitutes an important and often under-appreciated practical
challenge. Accordingly, I will briefly start with discussing sampling criteria
that need to be met and how unintentional wavefield under-sampling can
compromise the reconstruction of subsurface structure, in particular when
diffractions are concerned.
An introduction to seismic diffraction 37

4.1 Spatial wavefield sampling


Diffraction is a wavefield phenomenon and to optimally utilize its potential
this fact has to be honored in how we process the data. For wavefields to be
accurately reconstructable, dense spatial recordings are needed. It can be argued
that the process of diffraction, aside from surface waves which are commonly of
lower frequencies, causes the steepest possible wavefront inclination observable
in seismic sections. Accordingly, and because of their overall weakness,
diffracted wavefields are the first to suffer from unintentional but sometimes
unavoidable wavefield under-sampling. As it was appreciated that diffraction
is intimately linked to resolution in the reconstructed image, it was concluded
that too coarse sensor spacing also leads to lower image resolution (Grasmueck
et al., 2005).
This problem of under-sampling becomes arguably most severe, when the
imaging of near-surface structures is concerned, where diffraction, due to
medium complexity, becomes particularly likely and constitutes a primary car-
rier of information. Approaching the limiting case of surface-wave radiation,
inclination angles generally become higher the shallower the scattering
structure is located. Spatial under-sampling, just like its more commonly dis-
cussed temporal counterpart, has a characteristic signature in the frequency-
wavenumber (f-k) domain. According to simple geometrical considerations,
a Nyquist condition for the spatial reconstructability of a signal containing
the smallest wavelength λmin can be formulated

λmin
Δx  ΔxN ¼ , (10)
4 sinαmax

where Δx is the trace spacing, ΔxN is the spatial analog of the Nyquist fre-
quency and αmax corresponds to the largest inclination angle of the emerging
wavefront measured from the horizontal (e.g., Grasmueck et al., 2005;
Yilmaz, 2001). As was convincingly argued by Grasmueck et al. (2005) in
the context of ground-penetrating radar imaging, Eq. (10) finds its direct
correspondence in a criterion formulated by Claerbout (1985) describing
the maximally achievable lateral resolution in a migrated image. Fig. 19
shows thinned out versions of a properly sampled diffracted wavefield and
the corresponding f-k spectra. Even if only every fifth trace is considered
(j ¼ 5) aliasing effects can already be observed.
In particular in earthquake seismology, where typical inter-station dis-
tances, even in well-instrumented areas, can reach hundreds of kilometers,
spatial aliasing can be considered a major reason why wavefield-based
migration-type imaging still rarely leads to convincing results. To illustrate
38 Benjamin Schwarz

Trace number Trace number Trace number Trace number


100 200 20 40 10 20 5 10

1
Time (s)

2
Wavenumber (1/km) Wavenumber (1/km) Wavenumber (1/km) Wavenumber (1/km)
–20 0 20 –5 0 5 –2 0 2 –1 0 1
0
Frequency (Hz)

20

j=1 j=5 j = 10 j = 20
Fig. 19 Diffraction data (top) and the corresponding frequency-wavenumber (f-k)
distributions (bottom) for sufficient spatial sampling (j ¼ 1) and for different degrees
of under-sampling (only every jth trace is considered). Inclination angles are the highest
at the diffraction’s flanks, which provide the superior illumination and, accordingly, can
be viewed as the most precious part of the event.

that similar problems are encountered in near-surface prospecting, in


Fig. 20 a series of vertically aligned diffracted events that could correspond
to the signature of a fault are displayed. In particular the response of the
shallow diffractors is heavily aliased for insufficient sampling. As a result,
when diffracted energy is not isolated but accompanied by other coherent
Trace number Trace number Trace number Trace number
100 200 20 40 10 20 5 10
0
Time (s)

2
j=1 j=5 j = 10 j = 20
Fig. 20 The response of vertically aligned diffracting structures (as they could be caused by
a subvertical fault) for the same degrees of data sparsity as in Fig. 19. As the shallowest
diffraction caused the steepest inclinations, effects of aliasing are particularly pronounced.
An introduction to seismic diffraction 39

contributions, miscorrelations and unphysical coherent signatures can be


identified potentially leading to a flawed interpretation.
In analogy to our eyes, migration is an image forming system, which
is prone to mis-correlations when spatial sampling criteria are not met
(Gray, 2013). This has to do with the fact that arguably all migration types
build on the assumption of local coherence. In particular in Kirchhoff migra-
tion (Eq. 1), where seismic energy is repositioned through constructive and
destructive interference by summing along diffraction traveltime trajectories,
artifacts due to spatial aliasing represent a common problem and extensive
research efforts went into suppressing them (e.g., Abma, Sun, & Bernitsas,
1999; Biondi, 2001). Just as illustrated in Fig. 20, migration aliasing constitutes
a major challenge in near-surface imaging, where the migration trajectories,
just like the physical diffracted wavefronts themselves, are most inclined. So,
wavefield sampling is a topic not only important in the context of diffraction,
but of wavefield imaging as a whole.
To demonstrate the challenges diffraction imaging faces especially in the
near-surface zone, Fig. 21 shows the result of Kirchhoff migration applied to
the set of vertically aligned diffractions introduced in Fig. 20. While even
for the supposedly well-sampled case (j ¼ 1), minor artifacts can be observed,
they become increasingly more pronounced if sparser data are considered. In
line with what our visual inspection of the datasets already suggested, the

Trace number Trace number Trace number Trace number


100 200 100 200 100 200 100 200
0

0.5
Time (s)

1.0

1.5

j=1 j=5 j = 10 j = 20
Fig. 21 Kirchhoff migration results achieved for dense sampling (j ¼ 1), and for reduced
acquisitions, where only every jth trace of the well-sampled example is considered.
While for j ¼ 5, the deeper diffractors, due to lower dips in the respective response
remain well-imaged, artifacts due to spatial aliasing compromise the image quality in
the shallower region. In the extreme scenario of j ¼ 20, it becomes almost impossible
to reliably identify the exact scatterer locations.
40 Benjamin Schwarz

severe aliasing resulting from only considering every 10th or 20th trace
causes strong artifact contamination which is especially pronounced in
the near-surface region.
The need of large-aperture recordings and the overall weakness of the
wavefield means that diffractions often are almost unrecognizable in the
migrated images. In addition to an improved signal-to-noise ratio and a nat-
ural enhancement of diffracted energy, the use of coherence measures can
likewise enhance image quality, when significantly under-sampled data
are considered. To illustrate this, Fig. 22 shows the migration result for
the extremely unfavorable case of j ¼ 20, i.e., when only every 20th trace
is included in the migration (compare Figs. 20 and 21). Even though the
migration, in particular in the upper half of the image, results in strong arti-
fact contamination and an overall poor reconstruction of the true diffractor
locations, the use of the conventional semblance norm helps to clearly delin-
eate the vertically aligned scatterers. As can especially be observed for the
uppermost diffractor, the use of the nonlinear (Nth root) semblance norm
results in the cleanest image and a reliable reconstruction of all scatterer loca-
tions. As both, linear and nonlinear semblance represent normalized
quantities, i.e., only take values between 0 and 1, they can also be used as
a weight to suppress artifacts in conventional migration (see the rightmost
panel in Fig. 22). Despite these improvements and the availability of pow-
erful wavefield reconstruction algorithms (e.g., Hennenfent & Herrmann,

Trace number Trace number Trace number Trace number


100 200 100 200 100 200 100 200
0

0.5
Time (s)

1.0

1.5

Migration Semblance Nth root semb. Weighted migration


Fig. 22 Illustration of the impact normalized coherence measures can have on the
focused image, even if only every 20th trace (j ¼ 20) is considered during migration.
Displayed are the conventional Kirchhoff migration, the focusing semblance, its
nonlinear (Nth root) counterpart and a coherence-weighted migration image.
An introduction to seismic diffraction 41

2008; Xie & Gajewski, 2017; Zwartjes & Sacchi, 2006), it can nevertheless
be concluded that sufficiently dense spatial sampling should be ensured
already in the field for the potential of diffractions to be fully accessible.

4.2 Symmetries and correlations


Seismic imaging using diffracted wavefields has gained some momentum
in recent years. However, novel dense acquisition geometries using, for
example, fiber-optic strain sensing (e.g., Daley et al., 2013) and improved
noninvasive separation and extraction techniques promise the practical
realization of even more sophisticated strategies to exploit the unique prop-
erties of diffractions, including but not limited to velocity inversion. As was
emphasized earlier, diffraction is ultimately linked to resolution and diffrac-
tion imaging indeed lets one arrive at a highly resolved reconstruction of
the Earth’s abrupt property changes and imperfections (see, for example,
the illustrative comparison of conventional migration and diffraction
focusing in Figs. 10 and 11). However, as I will seek to demonstrate, even
more profound implications are to be expected when the multidirectional
radiation of scattering is systematically investigated. Ultimately linked to
illumination, provided that the assumption of a point scatterer is reasonably
justified at least for a subset of the considered wavefield, the recorded pres-
tack diffraction response turns out to be highly symmetric and redundant
(e.g., Bauer, Schwarz, & Gajewski, 2016; Schwarz & Gajewski, 2017b).
To illustrate this symmetry and the redundancy of information in differ-
ent data subdomains, the problem of common-source (CS) migration is
considered. Due to reciprocity, uniform radiation after scattering also
implies that no matter from where the energy arrives before the scattering
process, the diffraction response, kinematically encoded in its moveout, will
always look the same. Consequently, it can be argued that in contrast to
reflection imaging the migration of different CS gathers is not of comple-
mentary value, but rather (provided the reconstruction is successful)
delivers a multitude of copies of the same image. To illustrate this, infor-
mation encoded in a word formed by individual diffractors leads to a seem-
ingly different response, when sources located at opposite ends of the
acquisition line are considered (Fig. 23). However, the application of
CS Kirchhoff migration to the individual records reveals that exactly the
same structural image with the same information content can be
reconstructed. For sufficiently repeatable seismic sources this fundamental
redundancy of diffraction in CS gathers, owing to reciprocity, likewise
applies to the common-receiver (CR) domain.
42 Benjamin Schwarz

Lateral distance (km) Lateral distance (km)


0 2 4 6 8 0 2 4 6 8

2
Time (s)

CS gather (left) CS gather (right)


0 2 4 6 8 0 2 4 6 8
1
Time (s)

Reconstruction (left) Reconstruction (right)


Fig. 23 Image reconstruction by Kirchhoff migration (bottom) of two diffraction-only
common-source (CS) gathers (top), where one source was excited at the left and the
other at the right end of the acquisition line (indicated by a vertical red line in the respec-
tive gather). From an imaging perspective, both records contain the same information
and lead to a comparable reconstruction.

As a consequence, if diffractions can be reliably detected, the uniform


radiation of point scatterers potentially allows to image and successfully
characterize subsurface structures with drastically reduced acquisition effort.
As the comparison of the reconstructions based on the left and the right CS
gathers suggests, a multicoverage acquisition does not add more information
but rather reproduces the diffraction results gained in individual subdomains.
Although two-way moveouts are concerned in the zero-offset domain,
this means that also single- or small-offset acquisitions can potentially lead
to diffraction images comparable with industry-scale multicoverage
datasets (compare the academic and industrial data examples discussed in this
chapter). To further support this claim of redundancy, the migration results
of three additional adjacent CS gathers—this time strongly contaminated
with uncorrelated noise—are displayed in Fig. 24. Again, equivalent recon-
structions could be achieved and noise could effectively be suppressed by
employing the Nth root semblance norm.
An introduction to seismic diffraction 43

Trace number
500 1000 1500 2000

2
Time (s)

3
Three adjacent CS gathers (noisy)
500 1000 1500 2000
1
Time (s)

Reconstruction
Fig. 24 Reconstruction attempts (bottom) for three additional adjacent common-
source (CS) gathers (top), clearly emphasizing the redundancy of diffraction information
contained in conventional multicoverage recordings. The use of the nonlinear sem-
blance norm leads to resolved and clear images despite the fact that the gathers are
severely contaminated with uncorrelated noise.

To fully appreciate this redundancy of information in the prestack


diffraction response, the simple case of a single point diffractor embedded
in a smooth background medium is considered. Two CS gathers—with
the first source located at the far left and the second at the center of the
receiver line—are displayed in Fig. 25. The trace-wise cross-correlation
of the two gathers not only reveals that the overall shape of the diffraction
is independent of the source location, but also that the constant time shift
is directly encoded in the event’s moveout. So, in principle, a single
diffraction record is sufficient to deduce the event’s shape and position
in time at other locations.
We now consider the full prestack diffraction response shown on
the left side of Fig. 26. If we follow the same strategy and cross-
correlate all CS gathers with the first one (indicated by the red frame),
we arrive at perfectly flattened gathers for all considered sources and
in the common-receiver (CR) domain the estimated time lags again
44 Benjamin Schwarz

Trace number Trace number Trace number


50 100 150 50 100 150 50 100 150
-0.5
0.5

Time shift (s)


Time (s)

0
1.0

0.5
CS gather (left) CS gather (centre) Cross-correlation
Fig. 25 Two common-source (CS) gathers and their trace-wise cross-correlation for a
single-diffraction experiment. Owing to the fact that point scatterers radiate uniformly,
the diffraction response has the same shape in different data configurations, with the
only difference of a constant time shift, which itself is encoded in the event’s moveout.
The respective shot positions are indicated by the red vertical lines.

Receiver position (km) Receiver position (km)


)

)
m

m
(k

(k

0 1 2 0 1 2
n

2 2
tio

tio
si

si
po

po

1 1
ce

ce
ur

ur
So

So

0 0
0.6 0.6
Time (s)

0.85 0.85

1.1 1.1

Prestack data Interferometric reconstruction


Fig. 26 The same single-diffraction experiment as in Fig. 25, but now the full prestack
response is considered. Compared are the actual prestack data (left) and the full-
prestack reconstruction (right) based on a single common-source (CS) gather (indicated
by the red frame).

reproduce the event’s moveout (shown on the left side in Fig. 27). In anal-
ogous fashion, the cross-correlation of all CR gathers with a reference
gather has the same behavior as was observed in the CS domain (compare
Fig. 27). This lets us arrive at a simple recipe for the reconstruction of the
full prestack response from a single gather (a similar strategy was formulated
for passive events by Diekmann, 2018). For conciseness, only the CS
An introduction to seismic diffraction 45

Receiver position (km) Receiver position (km)

)
m

m
(k

(k
0 1 2 0 1 2

n
2 2

tio

tio
si

si
po

po
1 1
e

e
rc

rc
u

u
So

So
0 0
–0.4 –0.4
Time shift (s)

–0.2 –0.2

0 0

0.2 0.2
Cross-correlation of CS gathers Cross-correlation of CR gathers
Fig. 27 Illustration of the kinematic redundancy of the prestack diffraction response.
Shown are the trace-wise cross-correlations of all individual common-source (CS)
gathers with the first (left) and the respective cross-correlation results of the
common-receiver (CR) gathers (right). The flat lines indicate that the only difference
between different gathers consists in a constant shift in time.

domain is considered in the following. The trace-wise cross-correlation


X ij of the ith with the jth CS gather can be written as
Z
X ij ðx, τij Þ ¼ DCS
i ðx,tÞ ? D CS
j ðx,tÞ ¼ DCS
i ðx,tÞ Dj ðx, t + τ ij Þ dt:
CS
(11)

Owing to the invariant shape of diffracted events in different configurations,


the cross-correlation only differs from the autocorrelation by the constant
time shift τij which can be extracted from the event’s moveout in any
common-source gather via
τij ¼ tkCS ðxj Þ  tkCS ðxi Þ, (12)
where xi and xj denote the ith and the jth receiver locations, respectively
(which coincide with the positions of the ith and the jth source). The index
k indicates that the traveltime differences do not necessarily have to be esti-
mated in the ith or the jth gather, but are encoded in any (kth) gather avail-
able. Following from Eqs. (11) and (12), a combination of cross-correlation
(denoted by ?) and convolution (denoted by *)

DCS
j ðxÞ ¼ ½Di ðxi Þ ? Di ðxj Þ * Di ðxÞ
CS CS CS
(13)
allows us to fully reconstruct the observations of the jth CS gather
solely based on the knowledge of the ith gather. In practice, this means that
kinematically, the full prestack diffraction response can be reconstructed
from only a single gather. The applicability of Eq. (13) is demonstrated in
46 Benjamin Schwarz

Fig. 26, where only the first CS gather indicated by the red frame was used to
accurately reconstruct the full prestack diffraction response (shown on
the right). While the chosen example suggests the overall practicability of
Eq. (13), complications arise if more than one diffraction is recorded in
the CS gather (Fig. 28A). In order for the reconstruction to be successful,
spurious arrivals connected to inter-event correlations need to be prevented.
Fortunately—in contrast to the image, where diffractors are represented by
small-scale perturbations which can be located very close to each other—
each individual diffraction, as it is recorded in the time domain, has a unique
shape which makes it distinguishable from other contributions.
In line with the observation that the shape of the diffraction response only
contains propagation effects and no information on the internal geometry of

A Receiver position (km) B Receiver position (km)


)

)
m

m
(k

(k
0 1 2 0 1 2
n

n
tio

tio

2 2
si

si
po

po

1.0
ce

ce

1 1
ur

ur
So

So

0 0
0.8 0.8 0.8

0.6
1.1 1.1
Time (s)

0.4

1.4 1.4 0.2

0
1.7 1.7
Data Coherence

C Receiver position (km) D Receiver position (km)


)
m
(k

0 1 2 0 1 2
on
ti

2 2
si
po

2.0
ce

1 1
ur
So

0 50
0.8 1.5

1.1 0 1.0
Time (s)

0.5
1.4 –50

0
1.7
Receiver emergence angle (deg) Event tag
Fig. 28 Prestack response of a two-diffraction experiment (A) and the corresponding
by-products of coherence analysis. Also displayed are (B) the detected coherence
and (C) the estimated wavefront emergence angle at the receiver. Very similar to an
image segmentation task, the local coherence and continuity arguments can help to
automatically detect, distinguish and tag individual diffractions (shown in (D), Bauer,
Schwarz, Werner, & Gajewski, 2019).
An introduction to seismic diffraction 47

the scatterer, it was recently suggested that local coherence and smoothness
arguments could be used to efficiently distinguish and tag individual diffrac-
tions directly in the data domain (Bauer et al., 2019). Alongside the prestack
data cube, Fig. 28 shows the detected coherence, the estimated emergence
angle, and the aforementioned event tag assigned following the strategy
described by Bauer et al. (2019). Confirming previous observations, the sym-
metries contained in the diffracted response clearly express themselves in the
estimated angle fields (Fig. 28C). The automated estimation of the event tags
allows for a monogamous application of Eq. (13), which leads to an accurate
reconstruction for the multidiffractor case (Fig. 29). So depending on the

Receiver position (km) Receiver position (km)


)

A B

)
m

m
(k

(k
0 1 2 0 1 2
n

n
tio

tio
2
si

2
si
po

po
ce

ce
1 1
ur

ur
So

So

0 0
0.8 –0.6

–0.4
Time shift (s)

1.1
Time (s)

–0.2

1.4 0

0.2
1.7
Data Cross-correlation of CS gathers

C Receiver position (km) D Receiver position (km)


)

)
m

m
(k

(k

0 1 2 0 1 2
n

on
tio

ti

2 2
si

si
po

po
e

1 1
c

rc
ur

u
So

So

0 0
0.8 0.8

1.1 1.1
Time (s)

1.4 1.4

1.7 1.7
Reconstruction (without tagging) Reconstruction (with tagging)
Fig. 29 Comparison of the interferometric reconstruction of the full prestack diffraction
response out of one single common-source (CS) gather (C) without and (D) with the use
of the automated tagging strategy suggested by Bauer et al. (2019). The spurious con-
tributions that arise from inter-event correlations and convolutions become more
numerous, the more diffracted events are recorded (compare the cross-correlograms
(B) with the actual prestack response (A)).
48 Benjamin Schwarz

objective and the available data configurations, inherent prestack symmetries


connected to the uniform radiation of diffraction either justify imaging based
on rapidly reduced acquisitions or data volumes, or help to stabilize and addi-
tionally inform multicoverage imaging and inversion. Strongly related to
these findings, the next subsection briefly elaborates on how a similar corre-
lation strategy can help to illuminate processes of multiple scattering.

4.3 Deciphering the coda


The coda of seismic events can be a very lively place in which interesting
patterns and systematic behavior can be observed. In particular in passive-
source seismology, coda waves, owing to the fact that they encode multiple
scattering, arrive later and spend significantly more time in the investigated
medium than the primary energy of an earthquake. Based on the concept of
stationary phase Snieder (2004) demonstrated that interferometric relations
can be employed for Green’s function retrieval in the seemingly diffuse coda
wavefield, thereby providing important evidence on how ambient noise
might (partly) be generated. Despite their overall amplitude weakness these
multiply scattered contributions, by means of coda wave interferometry,
were demonstrated to provide unique sensitivity with respect to structural
changes that occur in the course of an earthquake (Snieder, Gr^et,
Douma, & Scales, 2002). In the following I will briefly discuss, how the
autocorrelation of active or passive trace gathers may help in recovering
inter-scatterer traveltimes from the coda. As was demonstrated before, it
is a special property of point diffraction that radiation occurs uniformly
resulting in a somewhat wasteful redundancy of information in prestack
records. A similar redundancy can be observed, when multiple scattering
is considered.
Figs. 30–32 show passive seismic data that were modeled using the
pseudo-spectral technique by Tessmer (2011), which is capable of honoring
multiple scattering. In Fig. 30, for comparison, the two simple cases of a
smooth background medium without perturbations and a single diffractor
are considered. In the respective autocorrelation, as expected, most of the
energy is concentrated in the zero-lag regime, whereas nonzero lag times
exist owing to the correlation of the scattered with the primary event.
The more interesting case of multiple scattering can only be observed, if
more than one scatterer is present. As the simplest possible example, the
autocorrelation was performed for two scatterers embedded in a smooth
background medium (see Fig. 31).
An introduction to seismic diffraction 49

A Trace number Trace number


100 200 300 100 200 300
–0.8

2 –0.4

Time shift (s)


Time (s)

3 0.4

0.8
Data (smooth background) Autocorrelation (smooth background)

B Trace number Trace number


100 200 300 100 200 300
–0.8

2 –0.4
Time shift (s)
Time (s)

3 0.4

0.8
Data (one diffractor) Autocorrelation (one diffractor)
Fig. 30 Recorded passive synthetic wavefield and its autocorrelation for (A) a smooth
background medium without scattering structures embedded, and (B) when a single
diffractor is present. In both cases, the correlogram is dominated by the autocorrelation
of the primary event.

Whereas the autocorrelated wavefield is considerably richer, perfectly


horizontal features with nonzero time lag appear which correspond to the
trace-by-trace correlation of different orders of scattering. Very similar to
the multicoverage case, for multiple scattering the remaining diffractor acts
like a secondary source leading to the same, merely time-delayed observable
moveout. To be able to better decipher the noticeably more complicated cor-
relogram the primary event was separated before autocorrelation (Fig. 31A
and B). Fig. 31C and Fig. 31D show the separate contributions resulting from
correlation with the response of the first and the second scatterer, respectively.
50 Benjamin Schwarz

A Trace number B Trace number


100 200 300 100 200 300
–0.8

2 –0.4

Time shift (s)


Time (s)

3 0.4

0.8
Data (two diffractors) Autocorrelation (two diffractors)
C Trace number D Trace number
100 200 300 100 200 300
–0.8 –0.8

–0.4 –0.4
Time shift (s)

Time shift (s)

0 0

0.4 0.4

0.8 0.8
Autocorrelation (diffractor 1) Autocorrelation (diffractor 2)
Fig. 31 Displayed are (A) the simulated response of two scatterers, (B) the corresponding
autocorrelation, as well as the separated contributions of (C) diffractor 1 and (D) diffractor
2. As contributions with high energy typically dominate the correlogram, the primary
event was separated before autocorrelation (compare Fig. 30). Red lines indicate the time
shifts caused by multiple scattering at the left and the right diffractor, respectively.

Two distinct time lags of τ1  0.5 s and τ2  0.3 s can be observed (indicated
by the red lines). As can be deduced from Fig. 32, where the case of four dif-
fractors is considered, the autocorrelation’s structure becomes increasingly
complicated, when more scatterers are involved.
To better appreciate the implications of these systematic observations in
the autocorrelation, a conceptual sketch of the considered multiple scatter-
ing process (for two and four diffractors) and its controlled-source analog is
displayed in Fig. 33. As was recently suggested by other authors, the fact that
the moveout connected to a point diffraction is always the same can be used
to systematically investigate the order in which observed scattering occurred
An introduction to seismic diffraction 51

Trace number Trace number


100 200 300 100 200 300
–0.8

2 –0.4

Time shift (s)


Time (s)

3 0.4

0.8
Data (four diffractors) Autocorrelation (four diffractors)
Fig. 32 Even for the relatively low number of only four point diffractors the
autocorrelogram becomes noticeably more complicated and a systematic extraction
of the multiple-scattering time lags likely demands wavefield separation or, in analogy
to the full-prestack wavefield reconstruction example suggested in the previous subsec-
tion, the systematic identification and tagging of different contributions to suppress
undesired spurious correlations (e.g., Bauer et al., 2019).

A B C

Fig. 33 Schematic illustration of the multiple scattering process for (a) a passive source
and two diffractors, (b) a controlled source at the surface and the same two diffractors,
and (c) the case of four diffractors, for which an increasingly complicated network of
correlations arises.

(Dylan Mikesell, Van Wijk, Blum, Snieder, & Sato, 2012; L€ oer, Meles, &
Curtis, 2015; Meles & Curtis, 2014). Here, I seek to demonstrate that with
the help of wavefield separation and tagging, under certain conditions, inter-
scatterer traveltimes can be extracted interferometrically from the coda by
means of simple autocorrelation. The conceptual sketch in Fig. 33 suggests
that the correlation of primary and secondary scattering at diffractors A and
B yields the respective lag times τA and τB. The redundant leg Rr
52 Benjamin Schwarz

(corresponding to the invariant wavefront curvature originating from the last


scatterer) is eliminated in the process,
τA ¼ tB + tBA  tA , (14a)
τB ¼ tA + tAB  tB : (14b)
Due to reciprocity we have tBA ¼ tAB for the inter-scatterer propagation
time which can be written as the arithmetic mean of the observable individ-
ual time lags (for diffractor A and diffractor B)
τA + τB
tAB ¼ : (15)
2
In Fig. 33 the distances Ri, for a constant velocity or an effective replace-
ment medium, correspond to wavefront radii excited by the actual or
by secondary sources. Despite its simplicity, Eq. (15) constitutes an exact
interferometric correspondence of inter-scatterer traveltimes and observ-
ables, provided point diffraction is considered and the extraction of the
time lags τA and τB from the autocorrelation was successful. In contrast
to cross-correlation, physically feasible time lags are always positive, which
rightfully implies a positive absolute inter-scatterer time tAB. Like before,
the presented findings strongly suggest that in many useful ways, diffracting
structures can indeed be viewed as passive sources providing structure and
order to the seemingly chaotic coda observed in highly scattering media.
Provided that these signals, despite their weakness, can be reliably
extracted, the knowledge of inter-scatterer times might help in addition-
ally stabilizing and constraining, for example, wavefront tomographic
inversion in depth (Bauer et al., 2017). Again, owing to their uniform
radiation, the above considerations likewise apply to passive and controlled-
source investigations.

5. Conclusions and outlook


The aim of this chapter was to provide a conceptual and intuitive
introduction to the process of seismic diffraction, the resulting wavefield’s
unique properties and how they can be utilized for imaging and inversion.
Historically, the roots of diffraction and the underlying theoretical frame-
work can be traced back to pioneering works in optics, where the concept
(illustrated by the famous slit experiments of Young) can be viewed as the
ultimate manifestation of a wave process. Although a fully rigorous treatment
An introduction to seismic diffraction 53

might only be possible in a full-waveform picture of seismic wave propaga-


tion, powerful far-field approximations exist and can successfully be utilized.
This introductory text has also sought to convince the reader that a proper
understanding and utilization of the process of diffraction demands the treat-
ment of fields, rather than isolated waveforms as is still often the case in the
context of sparse acquisitions.
In that light I briefly reviewed how simple wavefield-based imaging
techniques such as Kirchhoff migration intrinsically capture and naturally
honor diffraction phenomena, resulting in highly resolved images of
small-scale subsurface discontinuities. It was likewise discussed, how insuffi-
cient spatial wavefield sampling can compromise diffraction imaging—a still
prevailing challenge in near-surface geophysical prospecting. A variety of
controlled synthetic and field data examples suggest that insightful, highly
resolved and complementary images can be gained using different data con-
figurations, shedding new light, for example, on past erosional processes and
fault systems, which are normally hidden or hard to distinguish in conven-
tional seismic images. Due to the fact that essentially all conventional migra-
tion schemes can be used, the successful separation of diffractions from
the more dominant reflected wavefield can be considered a major challenge.
Following analogies in optics, I briefly described a simple strategy for
wavefield separation that can be easily implemented and was employed for
all diffraction imaging examples.
Despite the fact that the first studies concerned with seismic diffraction
occurring at crustal faults and rugged unconformities date back at least to the
1950s, its practical utilization for imaging and inversion can still be consid-
ered a niche discipline. While this in part can be explained by the overall
faintness of the diffracted wavefield and the fact that strong interference with
reflections often leaves it obscured and largely invisible in the prestack data,
it may also be argued that its overall potential is still often underestimated.
Through a simple sequence of cross-correlation and convolution it was
demonstrated that the full multicoverage prestack diffraction response is
encoded in and can be reconstructed with a single shot record, suggesting
a variety of interesting applications.
In particular when academic objectives are concerned, target depths are
typically significant, while measurement campaigns are often limited to very
small source-receiver distances and, consequently, suffer from drastically
reduced illumination, which renders the successful inversion of laterally
resolved velocity structure practically impossible. Following from the
54 Benjamin Schwarz

revealed and discussed symmetries and redundancies, diffracting structures


seem ideally suited to act as highly illuminating secondary sources in the
subsurface and in addition to recent advances in diffraction imaging
and inversion, major advances in the processing of reduced data configu-
rations arising from these poor-man acquisitions are expected in the near
future.
Simple controlled simulations likewise suggested that there is order in
the coda and that multiple scattering and its systematic study by means of
autocorrelation can lead to the interferometric extraction of inter-
scatterer traveltimes, which could, for example, be utilized for linking
individual measurements and efficiently constraining tomographic inver-
sion in depth. Owing to the multidirectional radiation of diffractions,
joint method development and the linking of observations in earthquake
seismology and controlled-source exploration might eventually enable
an integrated view of the Earth’s crust (compare, e.g., Zhu, 2018). While
previous studies often suffered from poor data quality, the rise of dense
seismic acquisition strategies across the scales, as, for example, provided
by fiber-optic strain sensing (Daley et al., 2013; Jousset et al., 2018;
Lindsey et al., 2017), promise an exciting future for exploring faults,
fractures, and unconformities with the weak signatures of seismic
diffraction.

Appendix. Kirchhoff’s diffraction integral


Building on the notion of the Huygens–Fresnel principle of interfer-
ence, Hagedoorn (1954) was the first to develop a systematic strategy to focus
and reposition seismic energy across seismograms to arrive at so-called migrated
sections. Despite the success this systematic treatment had, the rapid increase
of the amount of data recorded, fuelled by the digital revolution, a more quan-
titative and computational approach to migration was in demand. Building on
Kirchhoff’s diffraction theory in optics, Schneider (1978) introduced an
approximate integral approach to the repositioning of seismic energy, called
Kirchhoff migration, which still, up to the present day, is routinely employed in
everyday practice (e.g., Etgen et al., 2009). Owing to its persistent importance,
its formal beauty, and because it engraines the process of diffraction at its heart,
in the following, I will briefly rederive the Kirchhoff integral Eq. (1) from first
principles, i.e., as an approximate solution to the wave equation. To maintain
comparability, I will largely follow the derivations provided by Schneider
(1978), Wiggins (1984), and Born and Wolf (2013). For a more complete
account, the reader is referred to these publications.
An introduction to seismic diffraction 55

A time-dependent scalar wavefield D, which may, for example, corre-


spond to the pressure, obeys the scalar wave equation

1 ∂2 D
r2 D  ¼ 4πQ (A.1)
v2 ∂t2
where r2 is the three-dimensional Laplacian, t is time, v is the wave velocity,
and Q encodes the spatial and temporal characteristics of the sources con-
tained in the investigated volume. Following up on the ideas of Huygens
and Fresnel, Kirchhoff sought a mathematically sound quantitative theory
of wave propagation and image reconstruction, in which the process of dif-
fraction, expressed as secondary, elementary wavefields, forms the central
ingredient. He showed that the Huygens–Fresnel principle may be explicitly
formulated by utilizing Green’s theorem, according to which the wavefield
within a closed volume can be expressed in terms of the sum of a volume
integral encoding the source terms as well as initial conditions and an integral
over the surface enclosing this volume (e.g., Born & Wolf, 2013). For the
problem of seismic back-scattering, in good approximation, the homoge-
neous wave equation (Q ¼ 0) can be investigated. Correspondingly, the
only nonvanishing contribution to the spatially and temporally varying
wavefield constitutes in an integral over the closed surface S
Z Z  
1 ∂Dðr, tÞ ∂Gðr, t j r0 , t0 Þ
Dðr0 ,t0 Þ ¼ Gðr, t j r0 , t0 Þ  Dðr, tÞ dS dt:
4π ∂n ∂n
(A.2)
In Eq. (A.2) the time and space coordinates (r, t) 2 S indicate measurements
on the closed surface, whereas (r0, t0) correspond to their counterparts within
the enclosed volume. Differentiation ∂/∂n is performed with respect to the
surface normal. This representation can be viewed as a form of holography,
because measurements of D on a closed surface, i.e., a two-dimensional man-
ifold allow for the full reconstruction of the wavefield (in time and space)
within the enclosed volume (compare, e.g., Maynard, Williams, & Lee,
1985). In a slightly different form, it is commonly also referred to as the Helm-
holtz-Kirchhoff theorem (Born & Wolf, 2013). A two-dimensional sketch of the
configuration can be found in Fig. A.1A. For consistency and in contrast to
other authors, here, the image or location and time of diffraction are denoted
by the subscript 0, whereas the spatial and temporal coordinates corresponding
to observations on the surface carry no subscript. This should be kept in mind
when other works are consulted.
In vertical seismic profiling, measurements are mostly confined to the
Earth’s surface, which locally, often in good approximation, can be assumed
56 Benjamin Schwarz

A B

Fig. A.1 Illustration of (A) the holographic representation theorem and (B) the config-
uration underlying the derivation of the Kirchhoff integral. In principle, closed-surface
data acquisition would permit the full and accurate reconstruction of wavefields in the
interior. Kirchhoff migration assumes single-sided illumination and is a quantitative
manifestation of the Huygens–Fresnel principle, in which any wavefield can be thought
of as a superposition of elementary point diffractions (compare Figs. 2 and 3).

to be planar (Fig. A.1B). Consequently, in seismic investigations, the exact


correspondence expressed by Green’s theorem Eq. (A.2) does not hold and
approximations are commonly introduced. One of these approximations is
the so-called Sommerfeld radiation condition, which requires that the energy
radiated by sources must scatter to infinity and that no energy may be radi-
ated back from infinity. In practice, this means that contributions from the
distant hemisphere illustrated in Fig. A.1B can be ignored (Schneider, 1978).
If this condition is fulfilled, single-sided illumination, as is typical in seismic
problems, is deemed sufficient for image reconstruction (Born & Wolf,
2013; Claerbout, 1971; Wiggins, 1984).
In the following, we assume that the Sommerfeld radiation condition
holds and, for convenience, reparameterize the location vectors in terms
of depth with respect to the registration surface, i.e., we have r ¼ (x, z)
and r0 ¼ (x0, z0). The two-component vectors x and x0 describe the lateral
position on the surface and within the investigated subsurface volume,
respectively. As a result, in practice, the closed surface S in Eq. (A.2) is rep-
laced by a finite surface aperture Sx and normal derivatives correspond to
derivatives in the z direction (compare Fig. A.1B). As normal wavefield
derivatives ∂D=∂n are commonly not available, an appropriate choice of
the Green’s function can suppress these contributions by vanishing on the
measurement surface. According to the original derivation by Schneider
(1978), the Green’s function of free space (placed at x) subtracted by its
reflection in the plane meets this requirement. With R ¼ j(x0, z0)  (x, z)j
and Rr ¼ j(x0, z0)  (x, z)j we have

δðt  t0  R=vÞ δðt  t0  Rr =vÞ


G¼  , (A.3)
R Rr
where δ is Dirac’s delta function. The radii R and Rr correspond to the dis-
tance between secondary source and the receiver (located on the registration
An introduction to seismic diffraction 57

surface Sx) and its reflected counterpart, respectively. Insertion of (A.3) into
the surface integral representation of the wavefield (Eq. A.2), after simplifi-
cation, then leads to
Z Z  
1 ∂ δðt  t0  R=vÞ
Dðx0 , z0 , t0 Þ ¼ Dðx, tÞ dSx dt: (A.4)
2π ∂n R
The wavefield Dðx, tÞ ¼ Dðx, 0,tÞ now corresponds to the actual densely
spaced seismogram recordings acquired in the field. Despite the assumption
of an acquisition plane Sx (Schneider, 1978), this result is formally equivalent
to the generalized solution Wiggins (1984) provided for mildly undulating
surfaces and represents one form of the Kirchhoff integral.
For practical purposes, the time integration in Eq. (A.4) can be carried
out explicitly. First, however, the normal differentiation, which in the
assumed setup is a differentiation in the vertical (z) direction, can be per-
formed. As a result, one arrives at
Z
1
Dðx0 , z0 , t0 Þ ¼ ðI1 + I2 Þ dSx , (A.5)

with
Z
1 ∂R δ½t  ðt0 + R=vÞ
I1 ¼ Dðx, tÞ dt,
Z vR∂n  t  ðt0 + R=vÞ
∂ 1
I2 ¼ Dðx,tÞ δ½t  ðt0 + R=vÞ dt:
∂n R
These constituent integrals can be solved in a straightforward manner
through integration by parts and by making use of some of the unique prop-
erties of Dirac’s delta function δ. After some minor simplifications, the time
integration then leads to
Z    
1 ∂ 1 1 ∂R ∂Dðx, tÞ
Dðx0 ,z0 ,t0 Þ ¼ Dðx,tÞ  dSx , (A.6)
2π ∂n R vR ∂n ∂t
where similar to optics t ¼ t0 + R/v is the advanced time, i.e., the time the field
will advance from the present time t0. Time advancement captures effects of
wave propagation through the second term R/v and, therefore, can be viewed
as a process of extrapolation. This reflects once more the formal correspon-
dence of diffraction and the excitation of elementary (secondary) sources.
Eq. (A.6) is in perfect correspondence with the classic results found by
Schneider (1978) and Wiggins (1984) in the context of seismic migration.
It is interesting to note that exactly the same result can be gained, if the gen-
eralized Kirchhoff formula, also valid for nonmonochromatic light, is
58 Benjamin Schwarz

considered instead of Green’s theorem (see, e.g., Born & Wolf, 2013).
A more generic form of the Kirchhoff integral can be gained, if the time
integration is not performed explicitly. Then, Eq. (A.5) can be written as
ZZ    
1 ∂ 1 1 ∂R ∂Dðx, tÞ
Dðx0 , z0 ,t0 Þ ¼ Dðx, tÞ  δ½tðt0 +R=vÞdSx dt:
2π ∂n R vR ∂n ∂t
(A.7)
For a homogeneous background, R/v denotes the one-way diffraction
traveltime tdiff, which is the traveltime from the diffractor to the considered
measurement location on the surface (compare Fig. 1). As, owing to reci-
procity, this position can be both, a source (s) or a receiver (g) location,
the more practical two-way analog reads
Rs + Rg
tdiff ¼ : (A.8)
v
In the frame of coherence analysis, it was shown by various authors that this
so-called double-square-root equation, in the wavefront picture, lends the basis
for powerful traveltime approximations that permit improved data enhance-
ment, attribute determination and time-domain imaging, when reflection at
highly curved features or diffractions are considered (e.g., Fomel &
Kazinnik, 2013; Landa et al., 2010; Schwarz et al., 2014). In homogeneous
media, the distances Rs and Rg correspond to the wavefront radii as they are
observed at the source and the receiver location and Eq. (A.8) provides exact
traveltimes (compare Eq. 8). For the heterogeneous case (see Fig. 1), approx-
imate traveltimes can be gained by considering an effective overburden or
optical (straight-ray) projections (Schwarz & Gajewski, 2017b, 2017c).
The first term in Eq. (A.7) has a 1/t dependence and therefore decays
more rapidly when the scattered far-field is concerned (Wiggins, 1984).
Exploiting this observation, common implementations of Kirchhoff migra-
tion, therefore, only consider the second contribution containing the tem-
poral derivative of the wavefield ∂D=∂t. Following the image principle
introduced by Claerbout (1971), setting t0 ¼ 0 lets one arrive at the
so-called Rayleigh–Sommerfeld approximation of the Kirchhoff integral
adapted to the problem of seismic migration
Z Z
∂Dðx, tÞ
I ðξÞ ¼ w δðt  tdiff Þ dt dx, (A.9)
∂t
which is in exact correspondence with Eq. (1) provided at the beginning of
this chapter. Following modern convention, dSx is now replaced by dx and,
for convenience and generality, the factor of the second term in Eq. (A.7) is
An introduction to seismic diffraction 59

abbreviated with w, as it represents a weight that is applied to the temporal


derivative of the wavefield during summation. Like in (1), I now denotes
the amplitude of the reconstruction at the image point ξ.
Owing to its generic formulation, Eq. (A.9) captures Kirchhoff migration
in common-source gathers, common-receiver gathers (i.e., prestack data) or
the poststack (zero-offset) domain. To account for heterogeneity, in practice,
the weight function w and the diffraction traveltimes tdiff are forward-
calculated for a given velocity distribution either using analytical expressions
in the case of time migration or by employing ray tracing or eikonal solvers in
the depth domain. After its initial introduction to the geophysical community,
extensions of Eq. (A.9), by means of employing more sophisticated weights,
are capable to deliver migrated images with fully preserved amplitudes for
complicated media (e.g., Bleistein, 1987; Schleicher et al., 1993a).
As was discussed by Schneider (1978), the Kirchhoff integral represents a
three-dimensional convolution of the observed wavefield with a space-time
operator related to the point-source solution to the scalar wave equation. In
that sense, it can indeed be viewed as a quantitative manifestation of the
Huygens–Fresnel principle of interference. It is interesting to see that detect-
able isolated diffractions, at least from a kinematic perspective, share striking
similarities with passive sources and Green’s functions. Provided the time of
scattering and the scattering location can be determined, for example, by means
of source localization techniques, they would represent what Dirk Gajewski
once called natural Green’s functions that could be used for velocity-independent
seismic imaging.

Acknowledgments
This chapter builds on and summarizes work that has been carried out in parts at the Institute of
Geophysics of the University of Hamburg and at the Department of Earth Sciences at the
University of Oxford. I am indebted to the respective working groups as a whole, but
would specifically like to thank Alexander Bauer, Dirk Gajewski, Leon Diekmann, Jonas
Preine, Sergius Dell, Karin Sigloch, and Tarje Nissen-Meyer for crucial support and the
many important discussions that substantially shaped my current perception of the presented
matter. Further I am grateful to Christian H€ ubscher, TGS, BP, and SMAART JV for the
permission to share the respective complex synthetic and field data examples. Encouraging
feedback during my visit at BP in Sunbury and the weeks preceding it, specifically by Paula
Koelemeijer, Alistair Crosby, Jeffrey Winterbourne, Joseph Dellinger, and Jean Virieux, is
highly appreciated. The preparation of the manuscript and major strands of the underlying
research were crucially supported through fellowships of the German Research Foundation
(DFG, grants SCHW 1870/1-1 and SCHW 1870/2-1) and by Geo.X, the Research
Network for Geosciences in Berlin and Potsdam. Thoughtful comments by an anonymous
reviewer greatly helped to improve presentation and style of this chapter. Last but not least,
I am very grateful to the volume’s editor Cedric Schmelzbach for inviting me to contribute
and the truly great support during the preparation and editing stages.
60 Benjamin Schwarz

References
Abma, R., Sun, J., & Bernitsas, N. (1999). Antialiasing methods in Kirchhoff migration.
Geophysics, 64(6), 1783–1792.
Aki, K., & Richards, P. G. (2002). Quantitative Seismology. University Science Books.
Araya-Polo, M., Dahlke, T., Frogner, C., Zhang, C., Poggio, T., & Hohl, D. (2017). Auto-
mated fault detection without seismic processing. The Leading Edge, 36(3), 208–214.
Bahorich, M. S., & Farmer, S. L. (1995). 3-D seismic discontinuity for faults and stratigra-
phic features: The coherence cube. In SEG technical program expanded abstracts 1995
(pp. 93–96): Society of Exploration Geophysicists.
Bakhtiari Rad, P., Schwarz, B., Gajewski, D., & Vanelle, C. (2018). Common-reflection-
surface-based prestack diffraction separation and imaging. Geophysics, 83(1), S47–S55.
Bansal, R., & Imhof, M. G. (2005). Diffraction enhancement in prestack seismic data.
Geophysics, 70(3), V73–V79.
Bauer, A., Schwarz, B., & Gajewski, D. (2016). Enhancement of prestack diffraction data and
attributes using a traveltime decomposition approach. Studia Geophysica et Geodaetica,
60(3), 471–486.
Bauer, A., Schwarz, B., & Gajewski, D. (2017). Utilizing diffractions in wavefront tomog-
raphy. Geophysics, 82(2), R65–R73.
Bauer, A., Schwarz, B., Werner, T., & Gajewski, D. (2019). Unsupervised event identifica-
tion and tagging for diffraction focusing. Geophysical Journal International, 217(3),
2165–2176.
Baykulov, M., & Gajewski, D. (2009). Prestack seismic data enhancement with partial
common-reflection-surface (CRS) stack. Geophysics, 74(3), V49–V58.
Berkovitch, A., Belfer, I., Hassin, Y., & Landa, E. (2009). Diffraction imaging by
multifocusing. Geophysics, 74(6), WCA75–WCA81.
Berryhill, J. R. (1977). Diffraction response for nonzero separation of source and receiver.
Geophysics, 42(6), 1158–1176.
Billette, F. J., & Brandsberg-Dahl, S. (2005). The 2004 BP velocity benchmark. In 67th
EAGE conference & exhibition.
Biondi, B. (2001). Kirchhoff imaging beyond aliasing. Geophysics, 66(2), 654–666.
Bleistein, N. (1987). On the imaging of reflectors in the earth. Geophysics, 52(7), 931–942.
Born, M., & Wolf, E. (2013). Principles of optics: Electromagnetic theory of propagation, interference
and diffraction of light. Elsevier.
Buchen, P. W., & Haddon, R. A. W. (1980). Diffraction of a plane pulse by thin arbitrarily
shaped obstacles. The Journal of the Acoustical Society of America, 68(1), 309–313.
Cecchini, R., & Pelosi, G. (1990). Diffraction: The first recorded observation. IEEE
Antennas and Propagation Magazine, 32(2), 27–30.
Chopra, S., & Marfurt, K. J. (2007). Seismic attributes for prospect identification and
reservoir characterization. Society of Exploration Geophysicists.
Claerbout, J. F. (1971). Toward a unified theory of reflector mapping. Geophysics, 36(3),
467–481.
Claerbout, J. F. (1985). Fundamentals of geophysical data processing. Citeseer.
Dafni, R., & Symes, W. W. (2017). Diffraction imaging by prestack reverse-time migration
in the dip-angle domain. Geophysical Prospecting, 65(S1), 295–316.
Daley, T. M., Freifeld, B. M., Ajo-Franklin, J., Dou, S., Pevzner, R., Shulakova, V., …
Henninges, J. (2013). Field testing of fiber-optic distributed acoustic sensing (DAS)
for subsurface seismic monitoring. The Leading Edge, 32(6), 699–706.
Daniel, J., Moreau, S., & Nicol, R. (2003). Further investigations of high-order ambisonics and
wavefield synthesis for holophonic sound imaging. In Audio engineering society convention 114.
Decker, L., Janson, X., & Fomel, S. (2014). Carbonate reservoir characterization using
seismic diffraction imaging. Interpretation, 3(1), SF21–SF30.
Dell, S., & Gajewski, D. (2011). Common-reflection-surface-based workflow for diffraction
imaging. Geophysics, 76(5), S187–S195.
An introduction to seismic diffraction 61

Dell, S., Hoelker, A., & Gajewski, D. (2018). Using seismic diffractions for assessment of
tectonic overprint and fault interpretation. Geophysics, 84(1), IM1–IM9.
Diekmann, L. (2018). Source localization and joint velocity model building using wavefront attributes.
University of Hamburg.
Diekmann, L., Schwarz, B., Bauer, A., & Gajewski, D. (2018). Source localisation and joint
velocity model building. In 80th EAGE conference and exhibition 2018.
Duveneck, E. (2004). Velocity model estimation with data-derived wavefront attributes.
Geophysics, 69(1), 265–274.
Dylan Mikesell, T., Van Wijk, K., Blum, T. E., Snieder, R., & Sato, H. (2012). Analyzing
the coda from correlating scattered surface waves. The Journal of the Acoustical Society of
America, 131(3), EL275–EL281.
Etgen, J., Gray, S. H., & Zhang, Y. (2009). An overview of depth imaging in exploration
geophysics. Geophysics, 74(6), WCA5–WCA17.
Fomel, S. (2002). Applications of plane-wave destruction filters. Geophysics, 67(6), 1946–1960.
Fomel, S., & Kazinnik, R. (2013). Non-hyperbolic common reflection surface. Geophysical
Prospecting, 61(1), 21–27.
Fomel, S., Landa, E., & Taner, M. T. (2007). Poststack velocity analysis by separation and
imaging of seismic diffractions. Geophysics, 72(6), U89–U94.
Gelchinsky, B., Berkovitch, A., & Keydar, S. (1999a). Multifocusing homeomorphic imag-
ing: Part 1. Basic concepts and formulas. Journal of applied geophysics, 42(3), 229–242.
Gelchinsky, B., Berkovitch, A., & Keydar, S. (1999b). Multifocusing homeomorphic imaging:
Part 2. Multifold data set and multifocusing. Journal of Applied Geophysics, 42(3), 243–260.
Gradmann, S., H€ ubscher, C., Ben-Avraham, Z., Gajewski, D., & Netzeband, G. (2005). Salt
tectonics off northern Israel. Marine and Petroleum Geology, 22(5), 597–611.
Grasmueck, M., Weger, R., & Horstmeyer, H. (2005). Full-resolution 3D GPR imaging.
Geophysics, 70(1), K12–K19.
Gray, S. H. (2013). Spatial sampling, migration aliasing, and migrated amplitudes. Geophysics,
78(3), S157–S164.
Greenhalgh, S., & Manukyan, E. (2013). Seismic reflection for hardrock mineral exploration:
Lessons from numerical modeling. Journal of Environmental and Engineering Geophysics,
18(4), 281–296.
Hagedoorn, J. G. (1954). A process of seismic reflection interpretation. Geophysical
Prospecting, 2(2), 85–127.
Heincke, B., Green, A. G., van der Kruk, J., & Willenberg, H. (2006). Semblance-based
topographic migration (SBTM): a method for identifying fracture zones in 3D georadar
data. Near Surface Geophysics, 4(2), 79–88.
Hennenfent, G., & Herrmann, F. J. (2008). Simply denoise: Wavefield reconstruction via
jittered undersampling. Geophysics, 73(3), V19–V28.
Hilterman, F. J. (1970). Three-dimensional seismic modeling. Geophysics, 35(6), 1020–1037.
Hilterman, F. J. (1975). Amplitudes of seismic waves—A quick look. Geophysics, 40(5), 745–762.
H€ ocht, G., Ricarte, P., Bergler, S., & Landa, E. (2009). Operator-oriented CRS interpola-
tion. Geophysical Prospecting, 57(6), 957–979.
Hubral, P. (1975). Locating a diffractor below plane layers of constant interval velocity and
varying dip. Geophysical Prospecting, 23(2), 313–322.
Hubral, P. (1977). Time migration—Some ray theoretical aspects. Geophysical Prospecting,
25(4), 738–745.
H€ ubscher, C., Ruhnau, M., & Nomikou, P. (2015). Volcano-tectonic evolution of the poly-
genetic Kolumbo submarine volcano/Santorini (Aegean Sea). Journal of Volcanology and
Geothermal Research, 291, 101–111.
Ishii, M., Shearer, P. M., Houston, H., & Vidale, J. E. (2005). Extent, duration and speed of the
2004 Sumatra–Andaman earthquake imaged by the Hi-Net array. Nature, 435(7044), 933.
J€ager, R., Mann, J., H€ ocht, G., & Hubral, P. (2001). Common-reflection-surface stack:
Image and attributes. Geophysics, 66(1), 97–109.
62 Benjamin Schwarz

Jousset, P., Reinsch, T., Ryberg, T., Blanck, H., Clarke, A., Aghayev, R., …
Krawczyk, C. M. (2018). Dynamic strain determination using fibre-optic cables allows
imaging of seismological and structural features. Nature Communications, 9(1), 2509.
Kanasewich, E. R., & Phadke, S. M. (1988). Imaging discontinuities on seismic sections.
Geophysics, 53(3), 334–345.
Keller, J. B. (1962). Geometrical theory of diffraction. Journal of the Optical Society of America,
52(2), 116–130.
Khaidukov, V., Landa, E., & Moser, T. J. (2004). Diffraction imaging by focusing-
defocusing: An outlook on seismic superresolution. Geophysics, 69(6), 1478–1490.
Klem-Musatov, K. (1994). Theory of seismic diffractions. Society of Exploration Geophysicists.
Klem-Musatov, K., Hoeber, H. C., Moser, T. J., & Pelissier, M. A. (2016a). Classical and
Modern Diffraction Theory. Society of Exploration Geophysicists. Tulsa.
Klem-Musatov, K., Hoeber, H. C., Moser, T. J., & Pelissier, M. A. (2016b). Seismic
diffraction. Society of Exploration Geophysicists. Tulsa.
Klem-Musatov, K. D., & Aizenberg, A. M. (1984). The ray method and the theory of edge
waves. Geophysical Journal International, 79(1), 35–50.
Klem-Musatov, K. D., & Aizenberg, A. M. (1985). Seismic modelling by methods of the
theory of edge waves. Journal of Geophysics, 57(2), 90–105.
Klokov, A., & Fomel, S. (2012). Separation and imaging of seismic diffractions using
migrated dip-angle gathers. Geophysics, 77(6), S131–S143.
Krey, T. (1952). The significance of diffraction in the investigation of faults. Geophysics,
17(4), 843–858.
Krijgsman, W., Hilgen, F. J., Raffi, I., Sierro, F. J., & Wilson, D. S. (1999). Chronology,
causes and progression of the Messinian salinity crisis. Nature, 400(6745), 652.
Kunz, B. F. J. (1960). Diffraction problems in fault interpretation. Geophysical Prospecting,
8(3), 381–388.
Landa, E. (2012). Seismic diffraction: Where’s the value? In SEG technical program expanded
abstracts 2012 (pp. 1–4): Society of Exploration Geophysicists.
Landa, E., & Keydar, S. (1998). Seismic monitoring of diffraction images for detection of
local heterogeneities. Geophysics, 63(3), 1093–1100.
Landa, E., Keydar, S., & Moser, T. J. (2010). Multifocusing revisited-inhomogeneous media
and curved interfaces. Geophysical Prospecting, 58(6), 925–938.
Landa, E., Shtivelman, V., & Gelchinsky, B. (1987). A method for detection of diffracted
waves on common-offset sections. Geophysical Prospecting, 35(4), 359–373.
Lindsey, N. J., Martin, E. R., Dreger, D. S., Freifeld, B., Cole, S., James, S. R., …
Ajo-Franklin, J. B. (2017). Fiber-optic network observations of earthquake wavefields.
Geophysical Research Letters, 44(23), 11–792.
Liu, J., & Marfurt, K. J. (2007). Instantaneous spectral attributes to detect channels.
Geophysics, 72(2), 23–P31.
L€
oer, K., Meles, G. A., & Curtis, A. (2015). Automatic identification of multiply diffracted
waves and their ordered scattering paths. The Journal of the Acoustical Society of America,
137(4), 1834–1845.
Lyrintzis, A. S. (1994). The use of Kirchhoff’s method in computational aeroacoustics. Journal
of Fluids Engineering, 116(4), 665–676.
Marfurt, K. J., Kirlin, R. L., Farmer, S. L., & Bahorich, M. S. (1998). 3-D seismic attributes
using a semblance-based coherency algorithm. Geophysics, 63(4), 1150–1165.
Maynard, J. D., Williams, E. G., & Lee, Y. (1985). Nearfield acoustic holography: I. Theory
of generalized holography and the development of NAH. The Journal of the Acoustical
Society of America, 78(4), 1395–1413.
Mayne, W. H. (1962). Common reflection point horizontal data stacking techniques.
Geophysics, 27(6), 927–938.
Meles, G. A., & Curtis, A. (2014). Fingerprinting ordered diffractions in multiply diffracted
waves. Geophysical Journal International, 198(3), 1701–1713.
An introduction to seismic diffraction 63

Meyer, C. F. (1934). The diffraction of light, X-rays, and material particles. The University of
Chicago Press.
Moser, T. J., & Howard, C. B. (2008). Diffraction imaging in depth. Geophysical Prospecting,
56(5), 627–641.
M€ uller, C. (2000). On the nature of scattering from isolated perturbations in elastic media and the
consequences for processing of seismic data (Unpublished doctoral dissertation). Christian-
Albrechts Universit€at Kiel.
Neidell, N. S., & Taner, M. T. (1971). Semblance and other coherency measures for
multichannel data. Geophysics, 36(3), 482–497.
Netzeband, G. L., Gohl, K., H€ ubscher, C., Ben-Avraham, Z., Dehghani, G. A.,
Gajewski, D., & Liersch, P. (2006). The Levantine Basin—Crustal structure and origin.
Tectonophysics, 418(3–4), 167–188.
Newman, P. (1990). Amplitude and phase properties of a digital migration process. First
Break, 8(11), 397–403.
Nomikou, P., Druitt, T. H., H€ ubscher, C., Mather, T. A., Paulatto, M., Kalnins, L. M., …
Parks, M. M. (2016). Post-eruptive flooding of Santorini caldera and implications for
tsunami generation. Nature Communications, 7, 13332.
Nomikou, P., H€ ubscher, C., Ruhnau, M., & Bejelou, K. (2016). Tectono-stratigraphic
evolution through successive extensional events of the Anydros Basin, hosting Kolumbo
volcanic field at the Aegean Sea, Greece. Tectonophysics, 671, 202–217.
Reiche, S., H€ ubscher, C., & Beitz, M. (2014). Fault-controlled evaporite deformation in the
Levant Basin, Eastern Mediterranean. Marine Geology, 354, 53–68.
Rost, S., & Thomas, C. (2002). Array seismology: Methods and applications. Reviews of
Geophysics, 40(3), 2.
Santos, L. A., Mansur, W. J., & McMechan, G. A. (2012). Tomography of diffraction-based
focusing operators. Geophysics, 77(5), R217–R225.
Sava, P. C., Biondi, B., & Etgen, J. (2005). Wave-equation migration velocity analysis by
focusing diffractions and reflections. Geophysics, 70(3), U19–U27.
Schilt, F. S., Kaufman, S., & Long, G. H. (1981). A three-dimensional study of seismic
diffraction patterns from deep basement sources. Geophysics, 46(12), 1673–1683.
Schleicher, J., Tygel, M., & Hubral, P. (1993a). 3-D true-amplitude finite-offset migration.
Geophysics, 58(8), 1112–1126.
Schleicher, J., Tygel, M., & Hubral, P. (1993b). Parabolic and hyperbolic paraxial two-point
traveltimes in 3D media. Geophysical Prospecting, 41(4), 495–513.
Schneider, W. A. (1978). Integral formulation for migration in two and three dimensions.
Geophysics, 43(1), 49–76.
Schoepp, A., Labonte, S., & Landa, E. (2014). Multifocusing 3D diffraction imaging for
detection of fractured zones in mudstone reservoirs: Case history. Interpretation, 3(1),
SF31–SF42.
Schwarz, B. (2019). Coherent wavefield subtraction for diffraction separation. Geophysics,
84(3), V157–V168.
Schwarz, B., Bauer, A., & Gajewski, D. (2016). Passive seismic source localization via
common-reflection-surface attributes. Studia Geophysica et Geodaetica, 60(3), 531–546.
Schwarz, B., & Gajewski, D. (2017a). Accessing the diffracted wavefield by coherent subtrac-
tion. Geophysical Journal International, 211(1), 45–49.
Schwarz, B., & Gajewski, D. (2017b). A generalized view on normal moveout. Geophysics,
82(5), V335–V349.
Schwarz, B., & Gajewski, D. (2017c). The two faces of NMO. The Leading Edge, 36(6),
512–517.
Schwarz, B., Vanelle, C., Gajewski, D., & Kashtan, B. (2014). Curvatures and inhomogene-
ities: An improved common-reflection-surface approach. Geophysics, 79(5), S231–S240.
Silvestrov, I., Baina, R., & Landa, E. (2016). Poststack diffraction imaging using reverse-time
migration. Geophysical Prospecting, 64(1), 129–142.
64 Benjamin Schwarz

Snieder, R. (2004). Extracting the Greens function from the correlation of coda waves:
A derivation based on stationary phase. Physical Review E, 69(4), 046610.
Snieder, R., Gr^et, A., Douma, H., & Scales, J. (2002). Coda wave interferometry for
estimating nonlinear behavior in seismic velocity. Science, 295(5563), 2253–2255.
Sturzu, I., Popovici, A., Pelissier, M., Wolak, J., & Moser, T. (2014). Diffraction imaging of
the Eagle Ford shale. first break, 32(11), 49–59.
Taner, M. T., & Koehler, F. (1969). Velocity spectra: Digital computer derivation applica-
tions of velocity functions. Geophysics, 34(6), 859–881.
Tessmer, E. (2011). Using the rapid expansion method for accurate time-stepping in model-
ing and reverse-time migration. Geophysics, 76(4), S177–S185.
Trorey, A. (1977). Diffractions for arbitrary source-receiver locations. Geophysics, 42(6),
1177–1182.
Trorey, A. W. (1970). A simple theory for seismic diffractions. Geophysics, 35(5), 762–784.
Tsingas, C., El Marhfoul, B., Satti, S., & Dajani, A. (2011). Diffraction imaging as an
interpretation tool. First Break, 29(12), 57–61.
Tyiasning, S., Merzlikin, D., Cooke, D., & Fomel, S. (2016). A comparison of diffraction
imaging to incoherence and curvature. The Leading Edge, 35(1), 86–89.
Virieux, J., & Operto, S. (2009). An overview of full-waveform inversion in exploration
geophysics. Geophysics, 74(6), WCC1–WCC26.
Walda, J., & Gajewski, D. (2017). Determination of wavefront attributes by differential
evolution in the presence of conflicting dips. Geophysics, 82(4), V229–V239.
Walda, J., Schwarz, B., & Gajewski, D. (2017). A competitive comparison of multiparameter
stacking operators. Geophysics, 82(4), V275–V283.
Wiggins, J. W. (1984). Kirchhoff integral extrapolation and migration of nonplanar data.
Geophysics, 49(8), 1239–1248.
Wiggins, R. A. (1978). Minimum entropy deconvolution. Geoexploration, 16(1-2), 21–35.
Wu, X. (2017). Directional structure-tensor-based coherence to detect seismic faults and
channels. Geophysics, 82(2), A13–A17.
Wu, X., & Fomel, S. (2018). Automatic fault interpretation with optimal surface voting.
Geophysics, 83(5), 1–52.
Wu, X., & Hale, D. (2016). 3D seismic image processing for faults. Geophysics, 81(2),
IM1–IM11.
Xie, Y., & Gajewski, D. (2017). 5-D interpolation with wave-front attributes. Geophysical
Journal International, 211(2), 897–919.
Xie, Y., & Gajewski, D. (2018). 3D wavefront attribute determination and conflicting dip
processing. Geophysics, 83(6), V325–V343.
€ (2001). Seismic data analysis: Processing, inversion, and interpretation of seismic
Yilmaz, O.
data. Society of Exploration Geophysicists.
Yin, J., & Nakata, N. (2017). Diffraction imaging with geometric-mean reverse time
migration. In SEG technical program expanded abstracts 2017 (pp. 974–979): Society of
Exploration Geophysicists.
Zhang, Y., Bergler, S., & Hubral, P. (2001). Common-reflection-surface (CRS) stack for
common offset. Geophysical Prospecting, 49(6), 709–718.
Zhu, T. (2018). Passive seismic imaging of subwavelength natural fractures: Theory and 2-D
synthetic and ultrasonic data tests. Geophysical Journal International, 216(3), 1831–1841.
Zhuge, X., Yarovoy, A. G., Savelyev, T., & Ligthart, L. (2010). Modified Kirchhoff
migration for UWB MIMO array-based radar imaging. IEEE Transactions on Geoscience
and Remote Sensing, 48(6), 2692–2703.
Zwartjes, P. M., & Sacchi, M. D. (2006). Fourier reconstruction of nonuniformly sampled,
aliased seismic data. Geophysics, 72(1), V21–V32.

You might also like