You are on page 1of 18

Whither seismic stratigraphy?

Bruce S. Hart
1
Abstract
Here, I provide an historical summary of seismic stratigraphy and suggest some potential avenues for future
collaborative work between sedimentary geologists and geophysicists. Stratigraphic interpretations based on
reflection geometry- or shape-based approaches have been used to reconstruct depositional histories and to
make qualitative and (sometimes) quantitative predictions of rock physical properties since at least the mid-
1970s. This is the seismic stratigraphy that is usually practiced by geology-focused interpreters. First applied to
2D seismic data, interest in seismic stratigraphy was reinvigorated by the development of seismic geomorphol-
ogy on 3D volumes. This type of reflection geometry/shape-based interpretation strategy is a fairly mature sci-
ence that includes seismic sequence analysis, seismic facies analysis, reflection character analysis, and seismic
geomorphology. Rock property predictions based on seismic stratigraphic interpretations usually are qualita-
tive, and reflection geometries commonly may permit more than one interpretation. Two geophysics-based ap-
proaches, practiced for nearly the same length of time as seismic stratigraphy, have yet to gain widespread
adoption by geologic interpreters even though they have much potential application. The first is the use of
seismic attributes for feature detection, i.e., helping interpreters to identify stratigraphic bodies that are
not readily detected in conventional amplitude displays. The second involves rock property (lithology, porosity,
etc.) predictions from various inversion methods or seismic attribute analyses. Stratigraphers can help quality
check the results and learn about relationships between depositional features and lithologic properties of interest.
Stratigraphers also can contribute to a better seismic analysis by helping to define the effects of stratigraphy
(e.g., laminations, porosity, bedding) on rock properties and seismic responses. These and other seismic-related
pursuits would benefit from enhanced collaboration between sedimentary geologists and geophysicists.
Introduction
Seismic stratigraphy is an approach to seismic inter-
pretation that is based on principles of stratigraphy. The
science of seismic stratigraphy was first formalized in a
series of papers published in AAPG Memoir 26 in 1977,
although it has older roots (Cross and Lessenger, 1988).
Papers (e.g., Mitchum et al., 1977) published in the
AAPG volume showed how reflection terminations,
continuity, and other qualitative geometry-based
(mostly) analyses of reflections could be used to help
to infer depositional histories and predict lithology in
undrilled areas. This seismic-based approach became
an invaluable tool in the petroleum industrys explo-
ration and development efforts and subsequently
spread to other applied and fundamental geosciences
pursuits. Seismic stratigraphic analyses based on 2D
seismic data were later expanded to application on
3D data sets.
Seismic stratigraphy helped to spawn the develop-
ment of sequence stratigraphy, a science that is now
routinely applied to stratigraphic analyses even if seis-
mic data are not available. Catuneanu (2006) and Miall
(2010) summarize these techniques and discuss the
historical development of sequence stratigraphy.
Nevertheless, seismic stratigraphy and sequence stra-
tigraphy are so genetically linked that many geoscient-
ists consider the two to be essentially synonymous.
In this paper and elsewhere (Hart, 2011), I argue for a
somewhat distinct definition of seismic stratigraphy.
Textbooks in sedimentary geology define lithostratigra-
phy as the study of stratigraphic units that are defined
on the basis of their lithology, biostratigraphy as the
characterization and correlation of sedimentary depos-
its based on their fossil content, chemostratigraphy as
the use of inorganic chemistry as a correlation tool, and
so on. In that spirit, seismic stratigraphy should be de-
fined as the study of stratigraphic units that are defined
on the basis of their seismic characteristics. This broad
definition is consistent with the definition of Cross and
Lessenger (1988) who defined seismic stratigraphy as
1
Statoil, Houston, Texas, USA. E-mail: BHART@statoil.com.
Manuscript received by the Editor 23 April 2013; published online 8 August 2013. This paper appears in INTERPRETATION, Vol. 1, No. 1
(August 2013); p. SA3SA20, 21 FIGS.
http://dx.doi.org/10.1190/INT-2013-0049.1. 2013 Society of Exploration Geophysicists and American Association of Petroleum Geologists. All rights reserved.
t
Special section: Interpreting stratigraphy from geophysical data
Interpretation / August 2013 SA3
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
the science of interpreting or modeling stratigraphy,
sedimentary facies, and geologic history from seismic
reflection data.
As defined here, seismic stratigraphy can be an end
in its own right (e.g., the study of how different strati-
graphic features manifest themselves seismically) or it
can be one part of a multidisciplinary toolkit that is in-
tegrated for other applied or fundamental purposes
(e.g., basin analysis, sequence stratigraphy, reservoir
modeling, geotechnical studies). Seismic stratigraphy
can be used to study depositional basins in their en-
tirety (perhaps from 10
4
to >10
5
km
2
in extent), individ-
ual hydrocarbon reservoirs (from <10 to 10
2
km
2
in
extent), for geotechnical site investigations (1 km
2
in extent) and for other purposes. Seismic stratigraphic
principles and techniques have even been adapted for
stratigraphic analyses of ground-penetrating radar data
sets that commonly cover areas much smaller than
1 km
2
(e.g., Jol and Bristow, 2003).
AAPG Memoir 26 is best known in geologic circles
for papers that showed how stratigraphic principles
could be applied to seismic interpretation, and that ap-
proach is summarized in the first part of this paper.
However, it is noteworthy that other papers in the col-
lection focused on rock properties, seismic modeling,
and complex trace attributes (Gregory, 1977; Meckel
and Nath, 1977; Taner and Sheriff, 1977). These and
other geophysics- and rock physics-based approaches
subsequently failed to gain traction with geology-
focused interpreters, although they are areas where
stratigraphers and geophysicists share much common
interest. Subequent sections of this paper suggest ways
that geophysics-based methods can be applied to help
with stratigraphic interpretations and how stratigra-
phers can help geophysicists to better understand rela-
tionships between geology and seismic responses
2
.
Seismic stratigraphy: Analyses based on reflection
geometries
As described in the Introduction, papers published in
AAPG Memoir 26 presented techniques for making
stratigraphy-based interpretations of seismic data.
The initial emphasis in seismic stratigraphy was on
the use of reflection geometries (amplitudes were
treated descriptively) as seen in 2D seismic profiles
for exploration purposes. The approach became widely
accepted, and subsequent work helped refine and
promote the science (e.g., Shell Oil Company, 1987;
Cross and Lessenger, 1988; Bertram and Milton, 1996;
Snedden and Sarg, 2008). Seismic stratigraphy was
originally developed for the analysis of 2D seismic data
because 3D seismic volumes (and associated visualiza-
tion technologies) did not become widely available until
the 1980s and 1990s.
As originally defined, there were three distinct but
related aspects of seismic stratigraphy:
1) Seismic sequence analysis uses reflection termina-
tions (e.g., downlap, onlap, erosional truncation)
on 2D data, or vertical transects through 3D vol-
umes, to define key stratigraphic surfaces (e.g.,
unconformities, maximum flooding surfaces). These
surfaces are then used to assign relative strati-
graphic ages to portions of the seismic data (e.g.,
Unit A is older than Unit B) and to help define other
aspects of the depositional history (e.g., history of
relative sea-level change). Integration of this unca-
librated stratigraphic framework with biostrati-
graphic data from wells gives the seismically
defined packages absolute chronostratigraphic sig-
nificance (e.g., Vail et al., 1984). Seismic sequence
analysis is the branch of seismic stratigraphy that
eventually gave birth to sequence stratigraphy. Seis-
mic sequence analysis may not lead to inferences
about changes global (or even relative) sea level
or other driving mechanisms if the data sets cover
too small an area (e.g., Hart et al., 2007), if the data
sets are from basinal settings where factors other
than sea level control sediment accumulation, or
if other data types are not available to calibrate
the seismic-based stratigraphy. It should also be
noted that, although seismic stratigraphic analyses
were originally purported to be useful for identifying
and correlating global (eustatic) changes in sea level
from basin to basin, criticisms of the methods used
(e.g., seismic and biostratigraphic resolution prob-
lems) have led many, if not most, stratigraphers
to shy away from using seismic surfaces as interre-
gional correlation tools (e.g., Miall, 2010).
2) Seismic facies analysis. This approach uses reflec-
tion geometries and amplitudes as seen on vertical
transects (e.g., 2D seismic data) to define seismic
facies that are linked by the interpreter to specific
stratigraphic bodies (e.g., submarine fan lobes, mass
transport complexes, leveed channel complexes)
that can be used to make qualitative lithology pre-
dictions away from existing well control (e.g.,
Prather et al., 1998; Colombera et al., 2012). The pre-
dictions can be made quantitative, at least in a prob-
abilistic way, if analog databases of properties are
available (e.g., net:gross ratios of frontal splays in
deepwater fan systems). Calibration of seismic fa-
cies with well control enhances confidence in the
interpretation because seismic facies are nonuni-
que. For example, a chaotic seismic facies can
be associated with reef cores in carbonate settings,
mass-transport deposits in deepwater clastic set-
tings, and other lithofacies types.
2
For stratigraphers and geophysicists to collaborate more effec-
tively, they will need to become more familiar with each others lan-
guage and paradigms. To facilitate this task, Hart (2011) describes
geophysical terms and workflows in ways that should be understand-
able by geologists, and geologic concepts in ways that should be
understandable by geophysicists. SEGs Wiki site (http://wiki.seg
.org) provides additional information about geophysical terms,
whereas the Society for Sedimentary Geologys Strata site (http://
www.sepmstrata.org) provides descriptions and definitions of stratig-
raphy terms and concepts.
SA4 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
3) Reflection character analysis focuses on lateral var-
iations in the character of reflections, or groups of
reflections, to predict lateral variations in the stra-
tigraphy (lithology, porosity, thickness, etc.). Seis-
mic modeling can play an important role in this
pursuit. Modeling of acoustic responses (i.e., reflec-
tions associated with compressional waves and den-
sity) of stratigraphic features has become less
commonplace, with increased emphasis on model-
ing of elastic responses (reflections associated with
compressional waves, shear waves, and density) be-
cause the latter provide additional constraints on
lithology and physical property predictions.
Seismic stratigraphy offers significant advantages
over the more traditional lithostratigraphic interpreta-
tion of basins, as illustrated in Figure 1. Part (a) shows
an uninterpreted 2D seismic line from the North Slope
of Alaska. The image shows a series of clinoforms that
represent the infill of a Cretaceous foreland basin. In
part (b), some of the more continuous reflections have
been identified (picked). These reflections represent
seismic surfaces in the sense of Bertram and Milton
(1996), i.e., they are laterally continuous reflections
against which other reflections terminate. One of the
fundamental tenets of seismic stratigraphy is that,
within the resolution of the seismic data, reflections
represent depositional timelines. Although there are ex-
ceptions (e.g., Tipper, 1993; Zeng and Kerans, 2003),
this has proven to be a useful starting point for most
seismic stratigraphic interpretations. The seismic surfa-
ces (seismic sequence boundaries) shown in Figure 1b
can therefore be used to divide the seismic transect into
units of relative geologic age using the Law of Superpo-
sition; the clinoforms at left represent sedimentary
deposits that are younger than those at right. Biostrati-
graphic data from wells, if any, along the seismic tran-
sect could be used to assign more rigorous ages.
Although not illustrated here, reflection terminations
(onlap, downlap, etc.; Figure 2) could be identified on
the image in Figure 1 and used to designate some of the
seismic surfaces as sequence boundaries or maximum
flooding surfaces
3
. Instead, I have chosen to illustrate
how seismic facies analysis of reflection geometries,
amplitudes, and other characteristics of the seismic
profiles can be integrated with other data sets (wireline
logs, outcrop analogs, etc., not shown) to interpret
probable depositional environments within the seismic
sequences (Figure 1c). Note that even this simple
Figure 1. Seismic reflections approximate timelines and cross lithologic boundaries within the limits of seismic resolution. This
example shows a progradational Cretaceous succession in a 2D seismic profile collected by the United States Geological Survey
from the North Slope of Alaska. (a) Uninterpreted seismic profile. Note the vertical scale, in meters, has been approximated using
local well control. (b) Profile showing seismic surfaces, continuous seismic reflections that can be traced over most of the length
of the profile. Note the well-developed clinoform geometries with topset, foreset, and bottomset portions to the reflections. The
seismic surfaces are dashed in places in the lower foreset portions where strata have been disrupted by approximately syn-dep-
ositional deformation. The blue dots represent the location of the paleo shelf break, the break in slope between the relatively gently
dipping topset reflections and the more steeply dipping foreset reflections. Changes in trajectory of the shelf break (upward,
downward, outward) are used to distinguish whether the system is aggrading, prograding, retrograding, or degrading. (c) Profile
showing depositional settings represented by the different portions of the clinoforms. Note that the seismic surfaces cut across
several different types of depositional setting and so must cross lithological boundaries. (d) Lithostratigraphic interpretation of the
profile, wherein the stratigraphic names are assigned on the basis of lithology. The Torok Formation corresponds to marine shales
and deep water clastics, whereas the Nanushuk Group consists of shallow-marine and nonmarine clastics (e.g., Houseknecht and
Schenk, 2004). The seismic surfaces cut across lithostratigraphic formation boundaries and so are better for defining depositional
histories than the diachronous lithostratigraphic units. From Hart (2011). Reprinted by permission of the AAPG, whose permission
is required for further use.
3
The SEPMs stratigraphy Web site (http://www.sepmstrata.org)
has mapping and sequence stratigraphy exercises that are based
on this same public-domain Alaska seismic data set.
Interpretation / August 2013 SA5
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
seismic stratigraphic interpretation has more predictive
power than the lithostratigraphic interpretation shown
in Figure 1d. Integrating the seismic surfaces with the
seismic facies analysis is important because it would
allow the interpreter to evaluate changes in depositio-
nal processes from one seismic sequence to the next.
From a reservoir perspective, it is also important be-
cause the rocks associated with these seismic surfa-
ces can act as barriers or baffles to lateral and
vertical fluid flow; the lateral continuity of the toe-of-
slope facies (potentially sandy reservoirs) suggested
by Figure 1c is likely to be more apparent than real.
The advent of 3D seismic technology led to a fourth
branch of seismic stratigraphy called seismic geomor-
phology (e.g., Posamentier, 2004). Once 3D data sets
became commercially available, it quickly became ap-
parent that plan-view images derived from 3D seismic
volumes could be used to detect and map stratigraphic
features (e.g., Brown et al., 1981). Time slices, horizon
slices, proportional slices, and other types of plan-view
images now are routinely used to generate paleogeo-
graphic images from 3D cubes (e.g., Brown, 2004;
Weimer and Slatt, 2004). These images can be very use-
ful for fundamental studies of depositional systems
(e.g., Figure 3) and for qualitative rock property predic-
tions based on relationships between depositional ele-
ments and properties of interest (see seismic facies
discussion above).
From a seismic sequence perspective, the ability to
view 3D seismic data from many angles is particularly
significant because reflection terminations and configu-
rations are most easily observed from specific angles
with respect to the stratigraphic bodies being imaged.
For example, Figure 4 shows mutually orthogonal slices
through a 3D seismic cube that are viewed from differ-
ent angles. The data set images a Tertiary shelf-phase
Figure 2. Two different representations of stratal termina-
tions that might be visible in seismic data, well-log cross sec-
tions, or sometimes exceptional outcrops. (a) Reflection
terminations as defined by Exxon workers in AAPG Memoir
26 (Mitchum et al., 1977). Reprinted by permission of the
AAPG, whose permission is required for further use. (b) Re-
flection terminations as defined by BP in 1996 (Redrawn and
used with permission from Bertram and Milton, 1996).
Figure 3. Seismic amplitude map of the Cadotte Member, a
Cretaceous-age clastic strandplain deposit from Alberta, Can-
ada. (a) Uninterpreted map showing prominent amplitude
trends that strike approximately eastwest, with other trends
more northsouth in orientation. Note the acquisition geometry
is oblique to these trends, suggesting that the amplitudes are not
acquisition artifacts. (b) Interpreted map. The eastwest-striking
amplitude lineations are interpreted to indicate strandplain ori-
entation (i.e., they represent paleoshorelines), whereas the ap-
proximately northsouth trends are related to: (1) Cretaceous
channels that cut through the Cadotte shoreline, (2) Cretaceous
channels in the stratigraphic unit immediately above the Cadotte
(i.e., the Paddy Member) that interfere with the Cadotte ampli-
tudes, or (3) poor-data areas beneath a modern river floodplain.
Seismic modeling, log-based stratigraphic interpretations, and
comparison of the amplitude map to modern surficial features
were used to constrain the amplitude interpretation. See McCul-
lagh and Hart (2010) for further details. Reprinted by permission
of the AAPG, whose permission is required for further use.
SA6 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
delta. Reflection terminations showing toplap, down-
lap, and erosional truncation are more clearly seen in
some orientations than in others. Horizontal slices
through clinoforms allow true progradation directions
to be determined. The 3D seismic visualization software
allows the viewer to pick these or other orientations,
whereas an interpreter working with 2D seismic data
would be constrained to view the data in the orientation
in which they were collected.
The 3D seismic data sets should also be used by in-
terpreters to help make the conceptual link between the
expression of seismic facies seen on vertical transects
and in plan (map) views. This exercise can help inter-
preters to more confidently interpret 2D data. The 3D
seismic interpretation packages also permit inter-
preters to detect and visualize stratigraphic features
in various ways, such as volume (opacity) rendering
and geobody detection (Figure 5). These types of ex-
tractions and visualizations need to be undertaken in
the context of a fit-for-purpose seismic stratigraphic
analysis.
Some of the large 2D and 3D (covering several
1000 km
2
) seismic data sets collected offshore provide
unparalled opportunities for visualizing entire deposi-
tional systems (e.g., Saller et al., 2004). Conversely,
small land surveys (covering areas of several km
2
to
several tens of km
2
) can be much smaller than the
sequences or systems tracts being imaged. In these
small data sets, it is commonly impossible to see the
reflection termination patterns that are necessary to
identify seismic-scale systems tracts. As such, they
need to be integrated with regional data sets consisting
of longer 2D seismic lines and/or log-derived correla-
tions (Hart et al., 2007).
Figure 4. The seismic expression of stratigraphic features in
a 3D seismic cube depends on which way the data volume is
sliced. (a) Vertical transect and (b) time-slice images through
a prograding deltaic system. Compare the seismic expression
of the clinoforms and the incised valley from one image to the
other. The clinoforms show toplap and downlap in the vertical
transect, but the time slice shows them as lineations that can
be used to determine the shoreline orientation. The incision in
the vertical transect is seen to have a meandering geometry in
the time slice. Area of timeslice covers 96 km
2
(38 square
miles). From Hart (2011). Reprinted by permission of the
AAPG, whose permission is required for further use.
Figure 5. A simple example of geobody detection. (a) A dis-
continuous trough (red) is observed in a vertical transect and
used as a seed point by the software to track the feature
through the 3Ddata set, only a portion of which is shown here.
The software traces the body as a series of connected voxels
through the data set (using user-defined thresholds) and, in
this case, identifies a channel feature (b and c). From Hart
(2011). Reprinted by permission of the AAPG, whose permis-
sion is required for further use.
Interpretation / August 2013 SA7
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Seismic stratigraphy was originally conceived using,
and applied to, compressional wave seismic data. How-
ever, the methods of seismic stratigraphy also can be
applied to multicomponent data, a subdiscipline re-
ferred to as elastic wavefield seismic stratigraphy
by Hardage and Aluka (2006).
Shear-wave seismic data can sometimes detect
stratigraphic features that are invisible to compres-
sional waves. For example, Figure 6 shows two differ-
ent time slices through some Cretaceous clastics of the
Western Canada Sedimentary Basin in Alberta. The
image on the left shows a conventional seismic image
obtained by recording (and then processing) P-wave
reflections. A channel corresponding to the producing
wells (black dots) is, at best, possibly visible in the data.
The converted-wave data
4
(right side of Figure 6) more
clearly delineate the extent of the channel because the
channel-filling deposits respond to shear waves more
clearly than the compressional waves. Despite the ap-
parent utility of shear-wave seismic data for seismic
stratigraphic analyses, a variety of logistical and cost
reasons prevent them from being more widely em-
ployed (e.g., Harris and OBrien, 2008).
Because of seismic resolution limits and the ambigu-
ity associated with seismic facies (i.e., any one seismic
facies can be associated with several types of strati-
graphic deposits), seismic stratigraphic analyses should
be integrated with wireline logs, core, and other data
types wherever possible. Figure 7 shows an example
of this approach. In this Cretaceous clastics example,
log-based correlations were ambiguous because of
the highly channelized nature of the deposits. Con-
versely, some important stratigraphic surfaces could
not be defined in the seismic data because of resolution
issues and, perhaps, imaging problems in this P-wave
data set. Integration of the log and seismic data, and
a modern analog, reduced the ambiguity in the seismic
stratigraphic correlations and led to more confident
lithology predictions in interwell areas.
These geometry-based approaches tend to be most
commonly applied by interpreters with geologic
backgrounds. Vertical transects (e.g., 2D seismic data)
continue to be used in exploration settings to help de-
fine lithologies and depositional histories in areas lack-
ing well control (e.g., Bachtel et al., 2004; Gregersen and
Skaarup, 2007). Diagnostic combinations of seismic
facies and reflection terminations have been advanced
for a variety of different depositional settings
(e.g., Handford and Loucks, 1993; Weimer and Slatt,
2004; Figure 8). However, the approach has some
pitfalls, including: (1) nonuniqueness of the seismic re-
sponse due to resolution problems, interference effects
Figure 6. In some cases, stratigraphic features that are invisible to P-waves can be detected using S-waves, as shown by these
images from the Cretaceous section of Alberta. (a) Time slice through P-P data (conventional seismic data P-wave down and
P-wave up). The black dots running from top to bottom toward the left side of the image indicate the location of oil wells that
produce from channel sands. There is only a faint indication from this P-wave image that a channel might be present. (b) A cor-
responding time slice through a P-S data volume (i.e., mode-converted S-waves recorded). This image much more clearly indicates
the presence of a channel connecting the productive oil wells. From Margrave et al. (1998).
4
Converted-wave seismic data generally involve using a compres-
sional-wave source to generate downgoing seismic energy but record-
ing upgoing shear waves that were produced by mode conversion at
stratigraphic boundaries. See Margrave et al. (1998) for more details.
SA8 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
or other phenomena (Figure 9); (2) nonuniqueness
of the genetic relationships between depositional proc-
ess, external form/seismic facies and lithology, (3) non-
uniqueness of the relationship between depositional
facies (e.g., turbidite sheets) and physical properties
of interest (e.g., porosity and permeability are
affected by the degree/type of diagenesis, which is
not readily predicted from depositional systems ap-
proaches), and (4) nonquantitative output. Finally,
although some slices/attribute extractions show geo-
morphic features that are readily identifiable as depo-
sitional features, in some cases, these views show
ambiguous patterns that can be interpreted in several
different ways.
Seismic attributes and feature detection
Not all stratigraphic features of interest are readily
apparent in amplitude data. Various seismic attributes
have shown to be useful for detecting stratigraphic fea-
tures in the data in the same way that some attributes
are useful for identifying structural features such as
faults. Unfortunately, attribute-based analyses are not
Figure 7. Integration of a vertical transect, stratal slice, wireline logs, and suspected modern analog to define meandering fluvial
point bar deposits. (a) Seismic transect showing key seismic stratigraphic surfaces mapped through the integration of seismic and
log data. (b) The shingled reflections in the yellow oval correspond to shalier-upward successions in gamma-ray logs (c). A slice
through the data at this level (d) shows crescentic amplitude patterns that are similar in scale and planform morphology to scroll
bars of a modern meandering river system (e). See Sarzalejo and Hart (2006) for fuller discussion.
Interpretation / August 2013 SA9
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
typically used by geology-based interpreters (i.e., see
papers published in the sedimentary geology or petro-
leum geology literature). I explore the application of
seismic attributes to seismic stratigraphy in this section.
I defer discussion of inversion results to a later section,
although these derived volumes are sometimes referred
to as physically significant attributes.
Several definitions of seismic attributes exist,
but generally they can be thought of as quantitative
measures derived from the seismic data or interpreta-
tions. They include simple amplitude extractions (e.g.,
Figure 3, complex-trace attributes, other mathematical
manipulations of the seismic trace (e.g., integration and
derivative of the seismic trace; Figure 10), and mea-
sures derived from interpretations (e.g., curvature;
Figure 11). Summaries and examples of seismic attrib-
utes are presented by Brown (2004), Chopra and Mar-
furt (2007), Hart (2011), and others. Meta-attributes are
Figure 9. As demonstrated in this example,
geometry-based seismic stratigraphic inter-
pretations can be ambiguous because of seis-
mic resolution problems, interference effects,
or other factors. Top row shows a simple geo-
logic model composed of layers having differ-
ent acoustic properties. Image at top right
shows how the stratigraphy can be subdivided
into three packages: (1) a lowermost unit con-
sisting of folded/dipping layers (yellow), (2) a
middle unit consisting of a divergent basin fill
(green), and (3) an undisturbed upper unit
consisting of horizontally layered strata
(blue). An unconformity (dashed red line) sep-
arates the uppermost unit from the underlying
two units. The middle row shows the stratig-
raphy as imaged using a 75 Hz Ricker wavelet.
The reflections in this model generally show
the structural/stratigraphic geometries of the
model, and the image allows the three strati-
graphic units to be identified (middle right).
However, reflections appear to converge in
the middle unit (might be interpreted as strati-
graphic pinchouts?) and there appears to be
some subtle relief associated with the uncon-
formity (might be interpreted as differential
erosion of harder/softer layers?). The Lower-
most row shows the stratigraphy as imaged
using a 25 Hz Ricker wavelet. In this image,
it would be possible to interpret the presence
of a sequence-bounding unconformity (red
dashed line, lower right) that has truncation
below and onlap above (reflection polarity
would need to be considered). Only two
stratigraphic units might be inferred from this
interpretation. Based on seismic modeling
presented in Hart (2000).
Figure 8. Seismic lapout geometries and
facies common to carbonate platform depos-
its. Labeled features are: (1) karst-related
truncations, (2) shelf mounds, (3) land-
ward migrating clinoforms (rimmed shelves),
(4) bioherms (rimmed shelves), (5) steep
depositional slopes (>angle of repose),
(6) downlapping clinoforms at toe-of-slope,
(7) alternating downlap/onlap, (8) conver-
gence of clinoform reflections, (9) shelf edge
incision, and (10) incision within sequences.
Redrawn from Handford and Loucks (1993).
Reprinted by permission of the AAPG, whose
permission is required for further use.
SA10 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Figure 10. Poststack attributes used for detection of stratigraphic features based on simple seismic modeling. (a) Geologic model
of a prograding succession of clinoforms. High-velocity sands (yellow) grade down into low-velocity shale (brown and gray).
(b) Predicted amplitude response (variable density display with wiggle overlay) of the clinoforms using a 60 Hz Ricker wavelet.
(c) Integrated trace (relative acoustic impedance) attribute derived from the amplitude data shown in (b). The sands correspond to
relatively high impedance (green) and the shales to low impedance (red). Wiggle trace overlay shows the original seismic am-
plitudes. The clinoform geometry is apparent. (d) Predicted amplitude response (variable density display with wiggle overlay) of
the clinoforms using a 30 Hz Ricker wavelet. Subtle changes in amplitude are produced by the changes in sand thickness, but
the clinoform geometry is not clearly visible. (e) Integrated trace (relative acoustic impedance) attribute derived from the am-
plitude data shown in (d). The upper sandy portion of the succession corresponds to relatively high impedance (green) and the
shales to low impedance (red), but the clinoform geometry is not apparent. Wiggle trace overlay shows the original seismic
amplitudes. (f) Second derivative attribute derived from the amplitude data shown in (d). This attribute is capturing subtle changes
in waveform that suggest the presence of the clinoforms. From Hart (2008b).
Interpretation / August 2013 SA11
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
attributes derived by the combination of more than one
attribute (de Rooij and Tingdahl, 2002; Figure 12).
Qualitative analyses of reflection amplitude and fre-
quency, in addition to geometric analyses, originally
formed part of seismic facies analyses (Mitchum et al.,
1977). Taner and Sheriff (1977) are generally credited
with introducing the seismic stratigraphy community
to complex trace attributes. Other workers sub-
sequently incorporated complex trace attributes into
seismic stratigraphic analyses (e.g., Fontaine et al.,
1987) but published examples are few. Similarly, rela-
tively few attempts have been made to define seismic
attributes that capture reflection geometries such as
parallelism, convergence, or downlap (e.g., Barnes,
2000; van Hoek et al., 2010).
The seismic response is frequency-dependent.
Although some analyses have examined changes in re-
flection configuration in vertical transects as a function
of frequency (e.g., Zeng, 2013), most published analyses
have focused on map view slices through 3D volumes.
Spectral decomposition methods break the seismic sig-
nal down into its component frequencies. The resultant
images commonly bring out stratigraphic (or structural)
features that are not readily apparent in the original
broadband data (Castagna and Sun, 2006; Chopra
and Marfurt, 2007). Different approaches are possible
for analyzing the results, but one common approach
is to optically stack different frequency bands (e.g.,
Figure 13). Frequency-dependent tuning observed in
these analyses can be used to predict the thickness
of stratigraphic features.
Automated seismic facies analyses use a variety of
computer-based techniques to characterize seismic
trace shape. Thereafter, it is assumed (or hoped) that
each facies can be related to lateral variations in lithol-
ogy, rock properties, and/or fluid content of the strati-
graphic bodies being imaged. Several different
approaches have been applied to this task, including
artificial intelligence-based methods such as neural
networks (Colou et al., 2003; Figure 14). Originally
applied to a narrow time window defined with respect
to a single reflection, 3D applications of seismic facies
analysis also have been developed (e.g., Farzadi and
Hesthammer, 2007; Gao, 2007). These automated seis-
mic facies can be very useful; however, knowledge of
depositional systems is required to: (1) help define geo-
logically meaningful windows for the facies analyses,
and (2) ensure that the stratigraphic features inter-
preted to be revealed by the analyses make geo-
logic sense.
References presented in this section about seismic
attributes are generally from the geophysical literature.
Despite the potential applications to sedimentary geol-
ogy, seismic attribute studies have yet to become a
mainstream part of that disciplines toolkit.
Physical properties prediction
While indirect methods for predicting physical prop-
erties from seismic data (i.e., exploiting relationships
Figure 11. Example of horizon curvature for stratigraphic
analysis. (a) Time-structure map of the top of a leveed-channel
complex. (b) Subtle structural and stratigraphic features are
emphasized when dip curvature is overlain over the surface
and directional lighting is applied. (c) Expanded and rotated
view of the dip curvature overlay showing some fine-scale
morphological features that are emphasized by curvature
visualization. Note the changes in sharpness of the channel
margin (green and blue) along its length, and the presence
of an incised inner channel (shown by red hues indicating neg-
ative curvature) in the upper portion of this image. Repro-
duced with permission from Hart and Sagan (2007).
SA12 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
between seismic facies, depositional processes, and
lithology) were being developed and exploited by
sedimentary geologists and geophysicists, other geo-
physicists were developing ways of making direct,
quantitative predictions of rock properties from seismic
data.
Seismic inversion methods are used
to transform seismic data into quantita-
tive predictions of rock properties (e.g.,
acoustic impedance, lithology, porosity,
fluid saturation). Several different ap-
proaches are available for this purpose,
with one of the principal distinctions
being whether the inversion is under-
taken before or after the seismic data
have been stacked. Prestack methods
are used to define elastic properties of
the strata being imaged. Poststack meth-
ods are generally used to predict acous-
tic impedance. Most inversion methods
are deterministic; however, stochastic
inversion methods (e.g., incorporating
elements of geostatistics; e.g., Rowbo-
tham et al., 2003) also have been devel-
oped. Rock physics analyses and
modeling are needed components of in-
version studies. Selected summaries of
seismic inversion methods include those
of Veeken and DaSilva (2004), Sen
(2006), and Bosch et al. (2010).
There are several interrelated rea-
sons for inverting seismic data to acous-
tic impedance data prior to commencing
stratigraphic analyses (Latimer et al.,
2000). Acoustic impedance is a rock
property that can typically be directly
related to lithology, porosity, and
pore-filling fluids. Second, seismic data
that have been inverted to acoustic
impedance show layer properties rather
than interface properties. Reflections in
conventional seismic data correspond
to interfaces where there are changes
in acoustic impedance. The reflections
do not image the layers themselves.
After inversion, the seismic image
should show the acoustic properties of
the layers themselves. As such, the data
more closely resemble a geologic cross
section. Finally, most inversion methods
incorporate methods to remove the
blurring effect of the seismic wavelet,
thereby improving seismic resolution
and minimizing tuning effects. Thin
beds, not visible in the original seismic
data, can become apparent in the in-
verted seismic data.
Figure 15 compares a seismic image
with a model-based acoustic-impedance
inversion result from the Western Canada Sedimentary
Basin where Cretaceous siliciclastic deposits (fluvial,
estuarine, and shallow-marine deposits) unconform-
ably overlie a succession of Devonian carbonates.
The contact between those two packages (black bar
Figure 12. Sweetness attribute applied to a deepwater clastics data set. Sweet-
ness is a meta-attribute derived by combining instantaneous amplitude and in-
stantaneous frequency. In this example, modified from Hart (2008a), the image at
left (a) shows an inclined slice through a submarine fan system as seen in the
original amplitude display. The image at right (b) shows an identical slice
through a sweetness attribute volume, corendered with semblance to highlight
channel margins. High sweetness values (yellow/orange) are associated with
sands of at least two crosscutting channel systems. These channels are not
readily apparent in the original amplitude volume (a). Reprinted by permission
of the AAPG, whose permission is required for further use.
Figure 13. Spectral decomposition example from Moser et al. (2004). Ampli-
tude variations for three different frequencies are displayed in red, green, and
blue then overlain. The resultant image shows stratigraphic features that are
not readily visible in the original broadband seismic data. No scale bar was pro-
vided by the original authors, but it seems reasonable to assume that the width of
the image is probably 35 km across (meandering channels, such as those seen in
the image, are typically several tens of meters wide).
Interpretation / August 2013 SA13
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
near base of each well display) is not readily apparent in
the original seismic data. The unconformity becomes
readily apparent after the inversion. Clearly the imped-
ance version of the data would be much more useful for
seismic stratigraphic analyses than the original ampli-
tude volume.
Prestack inversion methods solve for P-impedance,
S-impedance, and density. These products can then
be manipulated in various ways to predict elastic prop-
erties (Youngs modulus, Poissons ratio, V
P
/V
S
ratio,
etc.) that can be related to lithology, porosity, fluid
saturation, etc. Given the current interest in unconven-
tional reservoirs (e.g., shale plays), there is much
interest in using inversion products to predict areas
of better reservoir properties (porosity, hydrocarbon
saturation) and to design hydraulic fracture treatments
(e.g., Close et al., 2012; Figure 16).
Although seismic inversion products have become
widely used in the petroleum industry, there are several
ambiguities and limitations that need to be kept in mind.
These problems include: (1) dependence of the result on
the low-frequency model, (2) wavelet estimation, (3)
nonuniqueness of the inversion solution, and (4) non-
uniqueness of the relationship between elastic/acoustic
properties and lithologic properties of interest to seis-
mic stratigraphers. Geophysicists have fully embraced
Figure 14. Sample automated seismic facies classification
results from Marroquin et al. (2009). (a) Devonian pinnacle
reefs and (b) a probable Jurassic tidal channel (siliciclastic).
Although different seismic facies (and lithologies) are present
at the two stratigraphic levels, the software has used the same
color palette for both classification exercises.
Figure 15. Example showing how a simple model-based
acoustic impedance inversion was used to help with a strati-
graphic interpretation (i.e., mapping a significant unconform-
ity). (a) The original seismic amplitude data. Well control
indicates the presence of a significant unconformity near
the base of the wells that separates Cretaceous clastics from
underlying Devonian carbonates (black line and star). The
unconformity does not correspond to a prominent reflection
in the data. (b) Model-based inversion result showing the
same profile presented in (a). The unconformity is clearly vis-
ible as an upward transition between high-impedance carbon-
ates (purple) and the overlying clastics (blue, red, yellow,
green). A high-impedance layer (purple) above the uncon-
formity corresponds to carbonate-cemented sandstone. Wire-
line logs in both cases are sonic logs. From Hart (2011).
Reprinted by permission of the AAPG, whose permission is
required for further use.
Figure 16. Prestack elastic inversion-based prediction of the
distribution of brittle and ductile rocks in the Cretaceous
Eagle Ford Formation. Map area covers approximately
80 km
2
. Modified from Treadgold et al. (2011).
SA14 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
inversion methods for predicting rock properties (includ-
ing fluid content) but the sedimentary geology commu-
nity has been slow to adopt these methods for studying
depositional systems (but see Contreras and Latimer,
2010). Sedimentary geologists could play a key role in
inversion studies by helping to define which results
are geologically/stratigraphically possible.
Another approach to physical property prediction
was termed the data-driven methodology by Schultz
et al. (1994). This approach exploits empirically derived
relationships between seismic attributes and log-de-
rived properties of interest to predict those properties
away from well control (e.g., Hampson et al., 2001).
Tebo and Hart (2005) and Sagan and Hart (2006) dem-
onstrate how this approach could be used to study
porosity development in two different carbonate set-
tings, and Sarzalejo Silva and Hart (2013) show how
the approach could be used to evaluate a siliciclastic
heavy oil reservoir. Although once relatively widely
used, the popularity of this empirical approach has
waned as geophysicists attention has turned to focus
on rock physics-based inversion approaches described
previously. The sedimentary geology community did
not embrace attribute-based property predictions.
Stratigraphic controls on seismic response
Despite much overlap in topics of interest, there has
been relatively little collaboration between the sedi-
mentary geology community, rock physicists, and geo-
physicists in terms of understanding the relationships
between depositional/diagenetic processes and seismic
responses. This, despite the fact that sediment prov-
enance (i.e., the source of the sediment), depositional
processes, and diagenesis are the only controls on min-
eralogy, porosity, fabric, and other properties that, in
turn, are fundamental controls on density, elastic
moduli, anisotropy, etc.
Rock physicists have developed many different
mathematical treatments for predicting properties of in-
terests based on theoretical mixes of grain sizes, poros-
ity, mineralogy, etc., in siliciclastic successions (e.g.,
Mavko et al., 1998; Avseth et al., 2005). However, there
are few studies that specifically link depositional proc-
esses (e.g., debris flows, tractive sediment transport) or
Figure 17. Two stacked coarsening upward successions
(parasequences) exposed in Cretaceous clastics of the Book
Cliffs, Utah. Each succession is shaley at the base with sand-
stone beds generally becoming thicker and more abundant
upward. People at lower right for scale. Note that these para-
sequences would be below the resolution of most petroleum
industry seismic data sets.
Figure 18. Outcrop photo of the Eagle Ford
Formation, west of Del Rio, Texas, from Hart
et al. (this issue). The formation is not a litho-
logically homogeneous unit. Instead, different
lithologies of different physical properties
(porosity, Poissons ratio, Youngs modulus,
etc.) are arranged into laminae, beds, bedsets
(shown), and members, all of which are
shorter than a seismic wavelet. Do maps of
seismically derived physical properties, such
as the one shown in Figure 16, represent aver-
age properties over the entire interval or are
they the product of some other phenomenon,
such as changes in stratigraphic stacking pat-
terns that affect the AVO response (e.g., Fig-
ure 20) upon which the physical property
predictions are based?
Interpretation / August 2013 SA15
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
diagenetic processes (e.g., effects of chlorite rims on
porosity development) to an appropriate physical
model (but see, for example, Avseth et al. [2005],
p. 8390). Eberli et al. (2003) and Weger et al. (2009)
examined relationships between carbonate porosity
types (a function of depositional and diagenetic proc-
esses) and rock physical properties.
Closer integration between sedimentary geologists
and rock physicists might allow more realistic modeling
of mineral textures and would almost certainly help to
ensure that appropriate rock physics models are se-
lected for property prediction in seismic modeling stud-
ies, i.e., certain mineral textures (e.g., clay-supported
sandstones) are more likely to develop in some deposi-
tional facies than others. For example, Hart et al. (this
issue) state that mathematical/conceptual models de-
veloped for clay-rich shales are inappropriate for
source-rock reservoirs (shale plays) because the latter
are commonly clay poor. In fine-grained systems
(shales), there has been considerable work undertaken
to examine the origins and effects of anisotropy at
the particle scale (e.g., Sayers, 2005; Day-Stirrat et al.,
2010). Layering imparts a vertical axis of symmetry,
making sedimentary rocks vertically transverse
isotropic media
5
. The anisotropic parameters are de-
rived from measurements on core plugs and need to
be upscaled mathematically (e.g., Backus, 1962; Berry-
man, 2008) to predict seismic-scale properties that
cannot readily be measured otherwise. Sedimentary
geologists could help to define realistic stratigraphic
model parameters for this type of upscaling work.
Sedimentary rocks are typically bedded, with
collections of beds forming bedsets or parasequences
(Figures 17 and 18). Seismic modeling has been used
successfully to predict seismic responses (amplitude
and other attributes) for various types of stratigraphic
successions (e.g., Meckel and Nath, 1977; Hart and
Chen, 2004; Figure 19), typically with a focus on predict-
ing poststack character. However, variations in strati-
graphic layering also are known to affect amplitude
variation with offset responses that would be visible in
prestack analyses (Figure 20). The effects of stratigraphic
Figure 19. Simple 1D acoustic models showing the predicted
seismic response of various stratigraphic stacking patterns
(blocky, fining upward, coarsening upward) as a function
of thickness. Each block shows an acoustic impedance profile
(increasing to right) and the calculated seismic response using
a simple Ricker wavelet. From Hart (2008b).
Figure 20. Amplitude-variation-with-offset (AVO) responses
for various stratigraphic stacking patterns. Although the same
change in physical properties is present in each case, the AVO
responses are markedly different. Modified from Lindsay and
Van Koughnet (2001).
Figure 21. Example of how seismic geometries can help
guide log correlations. In this simple case, gamma ray logs
from a series of wells (top row) can be correlated in at least
two different ways (middle row). A seismic transect from this
area (bottom) shows clinoforms that would support the
shingled correlation option of the middle row. Modified
from Hart (2011) and reprinted by permission of the AAPG,
whose permission is required for further use.
5
I.e., properties such as lithology, porosity, Poissons ratio, and
others change more rapidly from bed to bed, rather than along beds.
SA16 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
variability on AVO responses, or physical property pre-
dictions (e.g., elastic inversion) derived from AVO
analyses, are seldom documented explicitly.
Conclusions
The field of seismic stratigraphy is a mature science
that has proven to be useful in a variety of exploration,
development, and fundamental studies in sedimentary
geology. The four main elements of seismic stratigraphy
are seismic sequence analysis, seismic facies analysis,
reflection character analysis, and seismic geomorphol-
ogy. The first three were developed for use on 2D seis-
mic data and can be transferred to 3D volumes for use
on vertical transects. Seismic geomorphology evolved
from map-view analyses of 3D seismic volumes and
has since expanded to incorporate a variety of visuali-
zation and analysis techniques. Unfortunately, a variety
of factors can lead to nonunique interpretations, and
physical property predictions are typically qualitative
or, at best, probabilistic in nature. Perhaps for these
reasons, the physics-based quantitative output of seis-
mic inversion methods has become favored for rock
property prediction in the petroleum industry. On the
other hand, seismic data availability and quality prob-
lems preclude the application of inversion methods
in all settings.
The field of seismic stratigraphy, as practiced by
sedimentary geologists, could benefit from more rou-
tine integration of seismic attribute studies and seis-
mic-based physical property predictions. To do so,
sedimentary geologists will need to learn the physics
behind these methods. The sedimentary geology com-
munity has embraced physics before, such as applica-
tions in sediment transport and bedform development
(e.g., Middleton and Southard, 1984) or basin analysis
(e.g., Allen and Allen, 2005), and perhaps will again.
I have tried to highlight a variety of geoscience sub-
disciplines where the methods and interests of sedimen-
tary geologists and geophysicists overlap. Here, I
present several reasons why enhanced collaboration
between these two groups could be mutually advan-
tageous.

Enhanced fundamental understanding of deposi-


tional systems. As described above, the sedimen-
tary geology community has been very slow to
use/accept seismic analyses that are based on
(qualitative) attribute studies or (quantitative)
physical property predictions. This is unfortunate
because these analyses can provide clear images
of stratigraphic features and quantitative mea-
sures that would be very useful for fundamental
studies of depositional systems. Miall (2002),
for example, derived a variety of quantitative mor-
phologic parameters from meandering fluvial
systems imaged in time slices through a 3D
seismic amplitude volume. Wood and Mize-Span-
sky (2009) did likewise for a deepwater leveed-
channel system. Enhanced seismic stratigraphic
analyses, for example using volume visualization
on attribute volumes, might provide new informa-
tion about other depositional systems that cannot
be adequately imaged using amplitude displays.

Better predictive capabilities in exploration set-


tings. The high-quality 3D seismic data needed
to apply prestack elastic inversion methods are
not available everywhere. The results of any given
inversion study can be used to provide analog
data for risk assessment in other areas but should
be employed only if appropriate analogs can be
identified. For example, it would be inappropriate
to use data from a sand-prone submarine fan sys-
tem to assess risk in a mud-prone system. Sedi-
mentary geologists play an important role in
selecting appropriate analogs.

Improved geology-based interpretations. Geo-


logic interpretations based on subsurface data
(wireline logs, core) have an inherent degree of
ambiguity because of data availability (or lack
thereof) and, typically, data that can be inter-
preted in more than one way. Seismic data can
provide interwell information that can be critical
for assembling a meaningful geologic story (e.g.,
Figure 21). For example, log-based interpreta-
tions are problematic for some deep shale ba-
sins that are essentially undrilled. Seismic data
also can provide useful analog dimensional and
morphology data for features that cannot be accu-
rately mapped with available subsurface control.

Improved geophysics-based interpretations. Geo-


physics-based rock property predictions (e.g., in-
version) are likewise associated with ambiguity
because of reasons such as data quality, data
availability, assumptions made during the inver-
sion (e.g., wavelet estimation), etc. Stratigraphic
analyses of seismic-based rock property predic-
tions should always be undertaken to ensure that
those predictions are geologically reasonable. For
example, it would be appropriate to ask whether
the predicted distribution of brittle and ductile
rocks shown in Figure 16 makes geologic sense,
given what is known about Eagle Ford stratigra-
phy (e.g., Figure 18) and depositional processes.

Subseismic prediction of rock properties. Many


reservoirs contain internal baffles, thief zones,
variations in porosity, or other stratigraphic
heterogeneities that can have important eco-
nomic impacts but are below seismic resolution
(e.g., Figures 10, 17, and 18). Seismic data can pro-
vide the structural and stratigraphic context in
these cases, but stratigraphic knowledge plays
an important role in predicting the presence
and distribution of these features.
References
Shell Oil Company, 1987, Atlas of seismic stratigraphy, in
Bally, A. W. ed., AAPG Studies in Geology 27, 1571.
Interpretation / August 2013 SA17
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Allen, P. A., and J. R. Allen, 2005, Basin analysis: Principles
and application: Blackwell.
Avseth, P., T. Mukerji, and G. Mavko, 2005, Quantitative
seismic interpretation: Cambridge University Press.
Bachtel, S. L., R. D. Kissling, D. Martono, S. P. Rahardjanto,
P. A. Dunn, and B. A. MacDonald, 2004, Seismic strati-
graphic evolution of the MiocenePliocene Segitiga
Platform, East Natuna Sea, Indonesia: The origin,
growth, and demise of an isolated carbonate platform,
in G. P. Eberli, J. L. Masaferro, and J. R. Sarg, eds., Seis-
mic imaging of carbonate reservoirs and systems: AAPG
Memoir 81, 309328.
Backus, G. E., 1962, Long-wave elastic anisotropy
produced by horizontal layering: Journal of Geophysical
Research, 67, 44274440, doi: 10.1029/JZ067i011p04427.
Barnes, A. E., 2000, Attributes for automating seismic
facies analysis: 70th Annual International Meeting,
SEG, Expanded Abstracts, 553556.
Berryman, J. G., 2008, Exact seismic velocities for
transversely isotropic media and extended Thomsen
formulas for stronger anisotropies: Geophysics, 73,
no. 1, D1D10, doi: 10.1190/1.2813433.
Bertram, G. T., and N. J. Milton, 1996, Seismic stratigraphy,
in D. Emery, and K. J. Meyers, eds., Sequence stratig-
raphy: Blackwell, 4560.
Bosch, M., T. Mukerji, and E. F. Gonzalez, 2010,
Seismic inversion for reservoir properties combining
statistical rock physics and geostatistics: A review: Geo-
physics, 75, no. 5, 75A16575A176, doi: 10.1190/1
.3478209.
Brown, A. R., 2004, Interpretation of 3-D seismic data, 6th
ed.: AAPG Memoir 42.
Brown, A. R., C. G. Dahm, and R. J. Graebner, 1981, A
stratigraphic case history using three-dimensional
seismic data in the Gulf of Thailand: Geophysical Pro-
specting, 29, 327349, doi: 10.1111/j.1365-2478.1981
.tb01017.x.
Castagna, J. P., and S. Sun, 2006, Comparison of spectral
decomposition methods: First Break, 24, 7579.
Catuneanu, O., 2006, Principles of sequence stratigraphy:
Elsevier.
Chopra, S., and K. J. Marfurt, 2007, Seismic attributes for
prospect identification and reservoir characterization:
SEG Geophysical Development Series 11.
Close, D., M. Perez, B. Goodway, and G. Purdue, 2012, In-
tegrated workflows for shale gas and case study results
for the Horn River Basin, British Columbia, Canada:
The Leading Edge, 31, 556569, doi: 10.1190/
tle31050556.1.
Colou, T., M. Poupon, and K. Azbel, 2003, Unsupervised
seismic facies classification: A review and comparison
of techniques and implementation: The Leading Edge,
22, 942953, doi: 10.1190/1.1623635.
Colombera, L., F. Felletti, N. P. Mountney, and W. D.
McCaffrey, 2012, A database approach for constraining
stochastic simulations of the sedimentary heterogeneity
of fluvial reservoirs: AAPG Bulletin, 96, 21432166, doi:
10.1306/04211211179.
Contreras, A. J., and R. B. Latimer, 2010, Acoustic imped-
ance as a sequence stratigraphic tool in structurally
complex deepwater settings: The Leading Edge, 29,
10721082, doi: 10.1190/1.3485768.
Cross, T. A., and M. A. Lessenger, 1988, Seismic stratigra-
phy: Annual Review of Earth and Planetary Sciences,
16, 319354, doi: 10.1146/annurev.ea.16.050188.001535.
Day-Stirrat, R. J., S. P. Dutton, K. L. Milliken, R. G. Loucks,
A. C. Aplin, S. Hillier, and B. A. van der Pluijm, 2010,
Fabric anisotropy induced by primary depositional var-
iations in the silt: Clay ratio in two fine-grained slope fan
complexes: Texas Gulf Coast and northern North Sea:
Sedimentary Geology, 226, 4253, doi: 10.1016/j.sedgeo
.2010.02.007.
de Rooij, M., and K. Tingdahl, 2002, Meta-attributes the
key to multivolume, multiattribute interpretation: The
Leading Edge, 21, 10501053, doi: 10.1190/1.1518445.
Eberli, G. P., G. T. Baechle, F. S. Anselmetti, and M. L.
Incze, 2003, Factors controlling elastic properties in
carbonate sediments and rocks: The Leading Edge,
22, 654660, doi: 10.1190/1.1599691.
Farzadi, P., and J. Hesthammer, 2007, Diagnosis of the
Upper Cretaceous palaeokarst and turbidite systems
from the Iranian Persian Gulf using volume-based multi-
ple seismic attribute analysis and pattern recognition:
Petroleum Geoscience, 13, 227240, doi: 10.1144/
1354-079306-710.
Fontaine, J. M., R. Cussey, J. Lacaze, R. Lanaud, and L.
Yapaudjian, 1987, Seismic interpretation of carbon-
ate depositional environments: AAPG Bulletin, 71,
281297.
Gao, D., 2007, Application of three-dimensional seismic
texture analysis with special reference to deep-marine
facies discrimination and interpretation: Offshore An-
gola, west Africa: AAPG Bulletin, 91, 16651683, doi:
10.1306/08020706101.
Gregersen, U., and N. Skaarup, 2007, A mid-Cretaceous
prograding sedimentary complex in the Sisimiut Basin,
offshore west Greenland Stratigraphy and hydrocar-
bon potential: Marine and Petroleum Geology, 24, 15
28, doi: 10.1016/j.marpetgeo.2006.10.005.
Gregory, A. R., 1977, Aspects of rock physics from
laboratory and log data that are important to seismic
interpretation: in C. E. Payton, ed., Seismic stratigraphy
application to hydrocarbon exploration: AAPG
Memoir 26, 1546.
Hampson, B., J. Schuelke, and J. Quirein, 2001, Use of
multi-attribute transforms to predict log properties from
seismic data: Geophysics, 66, 220236, doi: 10.1190/1
.1444899.
Handford, C. R., and R. G. Loucks, 1993, Carbonate depo-
sitional sequences and systems tracts responses of
carbonate platforms to relative sea-level changes, in R.
G. Loucks, and J. F. Sarg, eds., Carbonate sequence
SA18 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
stratigraphy, recent developments and applications,
AAPG Memoir 57, 341.
Hardage, B. A., and I. J. Aluka, 2006, Elastic wavefield
seismic stratigraphy: AAPG Search and Discovery
Article 40183, http://www.searchanddiscovery.com/
documents/2006/06005hardage_gc/imagim/hardage.pdf,
accessed 15 March 2013.
Harris, R., and J. OBrien, 2008, Imaging tight gas sand-
stones in the East Texas Basin: First Break, 26, 5768.
Hart, B. S., 2000, 3-D seismic interpretation: A primer for
geologists: SEPM Short Course Notes, 48.
Hart, B. S., 2008a, Channel detection in 3-D seismic data
using sweetness: AAPG Bulletin, 92, 733742.
Hart, B. S., 2008b, Stratigraphically significant attributes:
The Leading Edge, 27, 320324, doi: 10.1190/1.2896621.
Hart, B. S., 2011, Introduction to seismic interpretation:
AAPG.
Hart, B. S., and M. A. Chen, 2004, Understanding seismic
attributes through forward modeling: The Leading
Edge, 23, 834841, doi: 10.1190/1.1803492.
Hart, B. S., J. H. S. Macquaker, and K. G. Taylor, this vol-
ume, Mudstone (shale) depositional and diagenetic
processes: Implications for seismic analyses of
source-rock reservoirs: Interpretation (This issue).
Hart, B. S., and J. A. Sagan, 2007, Curvature for visualiza-
tion of seismic geomorphology, in R. J. Davies, H. W.
Posamentier, L. J. Wood, and J. A. Cartwright, eds.,
Seismic geomorphology: Applications to hydrocarbon
exploration and production: Geological Society of Lon-
don Special Publication 277, 139149.
Hart, B. S., S. Sarzalejo, and T. McCullagh, 2007, Seismic
stratigraphy and small 3-D seismic surveys: The Leading
Edge, 26, 876881, doi: 10.1190/1.2756867.
Houseknecht, D. W., and C. J. Schenk, 2004, Sedimentol-
ogy and sequence stratigraphy of the cretaceous na-
nushuk, seabee, and tuluvak formations exposed on
umiat mountain, north-central Alaska: U.S. Geological
Survey Professional Paper, 1709-B, p. 18.
Jol, H. M., and C. S. Bristow, 2003, GPR in sediments: A
good practice guide, in C. S. Bristow, and H. M. Jol,
eds., Ground penetrating radar in sediments: Geological
Society of London, Special Publication 211, 927.
Latimer, R. B., R. Davison, and P. van Riel, 2000, An inter-
preters guide to understanding and working with seis-
mic-derived acoustic impedance data: The Leading
Edge, 19, 242256.
Lindsay, R., and R. Van Koughnet, 2001, Sequential backus
averaging: Upscaling well logs to seismic wavelengths:
The Leading Edge, 20, 188191, doi: 10.1190/1.1438908.
Margrave, G. F., D. C. Lawton, and R. R. Stewart, 1998, In-
terpreting channels sands with 3C-3D seismic data: The
Leading Edge, 17, 509513, doi: 10.1190/1.1438000.
Marroquin, I. D., B. S. Hart, and J. J. Brault, 2009, A visual
mining-based methodology to conduct seismic facies
analysis, part 2: Application to 3-D seismic data: Geo-
physics, 74, no. 1, P13P23, doi: 10.1190/1.3046456.
Mavko, G., T. Mukerji, and J. Dvorkin, 2009, The rock
physics handbook: Cambridge University Press.
McCullagh, T., and B. S. Hart, 2010, Stratigraphic controls
on production from a basin-centered gas system: Lower
Cretaceous Cadotte Member, Deep Basin Alberta, Can-
ada: AAPG Bulletin, 94, 293315, doi: 10.1306/
08260908137.
Meckel, L. D., and A. K. Nath, 1977, Geologic considera-
tions for stratigraphic modeling and interpretation, in
C. E. Payton, ed., Seismic stratigraphy Application
to hydrocarbon exploration: AAPG Memoir 26,
417438.
Miall, A. D., 2002, Architecture and sequence stratigraphy
of pleistocene fluvial systems in the Malay Basin, based
on seismic time-slice analysis: AAPG Bulletin, 86, 1201
1216.
Miall, A. D., 2010, The geology of stratigraphic sequences,
2nd Edition, Springer.
Middleton, G. V., and J. B. Southard, 1984, Mechanics of
sediment movement: SEPM Short Course Notes, 3.
Mitchum, R. M., P. R. Vail, Jr., and J. B. Sangree, 1977,
Stratigraphic interpretation of seismic reflection pat-
terns in depositional sequences, in C. E. Payton, ed.,
Seismic stratigraphy applications to hydrocarbon
exploration: AAPG Memoir 26, 117133.
Moser, J. R., F. F. Strijbos, and J. P. Castagna, 2004, Spec-
tral variance in comparison with conventional spectral
decomposition attributes: 74th Annual International
Meeting, SEG, Expanded Abstracts, 441444.
Posamentier, H. W., 2004, Seismic geomorphology: Imag-
ing elements of depositional systems from shelf to deep
basin using 3D seismic data: Implications for explora-
tion and development: Geological Society London
Memoirs, 29, 1124, doi: 10.1144/GSL.MEM.2004.029
.01.02.
Prather, B. E., J. R. Booth, G. S. Steffens, and P. A. Craig,
1998, Classification, lithologic calibration, and strati-
graphic succession of seismic facies of intraslope
basins, deep-water Gulf of Mexico: AAPG Bulletin,
82, 701728.
Rowbotham, P. S., D. Marion, P. Lamy, E. Insalaco, P. A.
Swaby, and Y. Boisseau, 2003, Multidisciplinary sto-
chastic impedance inversion: Integrating geological
understanding and capturing reservoir uncertainty:
Petroleum Geoscience, 9, 287294, doi: 10.1144/1354-
079302-490.
Sagan, J. A., and B. S. Hart, 2006, Three-dimensional
seismic-based definition of fault-related porosity
development: TrentonBlack river interval, Saybrook,
Ohio: AAPG Bulletin, 90, 17631785, doi: 10.1306/
07190605027.
Saller, A. H., J. T. Noah, A. P. Ruzuar, and R. Schneider,
2004, Linked lowstand delta to basin-floor fan deposi-
tion, offshore Indonesia: An analog for deep-water res-
ervoir systems: AAPG Bulletin, 88, 2146, doi: 10.1306/
09030303003.
Interpretation / August 2013 SA19
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Sarzalejo, S., and B. S. Hart, 2006, Stratigraphy and litho-
logic heterogeneity in the Mannville group (southeast
Saskatchewan) defined by integrating 3-D seismic
and log data: Bulletin of Canadian Petroleum Geology,
54, 138151, doi: 10.2113/gscpgbull.54.2.138.
Sarzalejo Silva, S. E., and B. S. Hart, 2013, Advanced seis-
mic-stratigraphic imaging of depositional elements in a
lower cretaceous (Mannville) heavy oil reservoir, west-
central Saskatchewan, Canada, in F. J. Hein, D. Leckie,
S. Larter, and J. Suter, eds., Heavy-oil and
oil-sand petroleum systems in Alberta and beyond:
AAPG Studies in Geology, 64, 359372.
Sayers, C. M., 2005, Seismic anisotropy of shales: Geo-
physical Prospecting, 53, 667676, doi: 10.1111/j.1365-
2478.2005.00495.x.
Schultz, P. S., S. Ronen, M. Hattori, and C. Corbett, 1994,
Seismic-guided estimation of log properties: Part 1: A
data-driven interpretation methodology: The Leading
Edge, 13, 305310, doi: 10.1190/1.1437020.
Sen, M., 2006, Seismic inversion: SPE.
Snedden, J. W., and J. F. Sarg, 2008, Seismic stratigraphy
a primer on methodology: AAPG Search and Discovery,
40270, accessed 8 July 2008.
Taner, M. T., and R. E. Sheriff, 1977, Application of ampli-
tude, frequency and other attributes to stratigraphic and
hydrocarbon determination, in C. E. Payton, ed., Seis-
mic stratigraphy: Applications to hydrocarbon explora-
tion: AAPG Memoir 26, 391327.
Tebo, J. M., and B. S. Hart, 2005, Use of volume-based 3-D
seismic attribute analysis to characterize physical prop-
erty distribution: A case study to delineate reservoir
heterogeneity at the Appleton field, SW Alabama: Jour-
nal of Sedimentary Research, 75, 723735, doi: 10.2110/
jsr.2005.058.
Tipper, J. C., 1993, Do seismic reflections necessarily have
chronostratigraphic significance?: Geological Maga-
zine, 130, 4755, doi: 10.1017/S0016756800023712.
Treadgold, G., B. Campbell, B. McLain, S. Sinclair, and D.
Nicklin, 2011, Eagle Ford shale prospecting with 3D
seismic data within a tectonic and depositional system
framework: The Leading Edge, 30, 4853, doi: 10.1190/1
.3535432.
Vail, P. R., J. Hardenbol, and R. G. Todd, 1984, Jurassic
unconformities, chronostratigraphy and sea-level
changes from seismic stratigraphy and biostratigraphy,
in J. S. Schlee, ed., Interregional unconformities and
hydrocarbon accumulation, AAPG Memoir 36,
129144.
van Hoek, T., S. Gesbert, and J. Pickens, 2010, Geometric
attributes for seismic stratigraphic interpretation: The
Leading Edge, 29, 10561065, doi: 10.1190/1.3485766.
Veeken, P. C. H., and M. DaSilva, 2004, Seismic inversion
and some of their constraints: First Break, 22, 4770.
Weger, R. J., G. P. Eberli, G. T. Baechle, J. L. Massaferro,
and Y.-F. Sun, 2009, Quantification of pore structure
and its effect on sonic velocity and permeability in car-
bonates: AAPG Bulletin, 93, 12971317, doi: 10.1306/
05270909001.
Weimer, P., and R. M. Slatt, 2004, Petroleum systems of
deepwater settings: SEG Distinguished Instructor
Series no. 7.
Wood, L. J., and K. L. Mize-Spansky, 2009, Quantitative
seismic geomorphology of a quaternary leveed-channel
system, offshore eastern Trinidad and Tobago, north-
eastern South America: AAPG Bulletin, 93, 101125,
doi: 10.1306/08140807094.
Zeng, H., 2013, Frequency-dependent seismic-stratigraphic
and facies interpretation: AAPG Bulletin, 97, 201221,
doi: 10.1306/06011212029.
Zeng, H., and C. Kerans, 2003, Seismic frequency control
on carbonate seismic stratigraphy: A case study of
the Kingdom Abo sequence, west Texas: AAPG Bulletin,
87, 273293, doi: 10.1306/08270201023.
SA20 Interpretation / August 2013
D
o
w
n
l
o
a
d
e
d

1
0
/
1
4
/
1
3

t
o

1
8
6
.
1
1
2
.
2
4
3
.
2
3
8
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/

You might also like