You are on page 1of 48

Sequence Stratigraphy > Intro Sequence Stratigraphy

Preamble
Stratigraphy is the science of understanding the variations in the successively layered
character of rocks and their composition. These rocks may be sedimentary, volcanic,
metamorphic or igneous. The layering of sedimentary rocks is expressed as sets of simple to
complex sedimentary geometries, and a wide variety of different sedimentary facies.
Sequence stratigraphy, a branch of sedimentary stratigraphy, deals with the order, or
sequence, in which depositionally related stratal successions (time-rock) units were laid down
in the available space or accommodation. The chronostratigraphy of sedimentary rocks tracks
changes their character through geologic time. These changes may be shown in graphical
form as either geologic cross sections and/or as chronostratigraphic correlation charts or
Wheeler (1958, & 1964) diagrams. This is distinct from their geochronology or geologic age.
The discipline of sequence stratigraphy provides a tool used interpret the depositional origin
and predict the heterogeneity, extent and character of the lithofacies. This tool combines:

 The framework of major depositional and erosional surfaces bounding these


successions of strata
 The geometry that successive contemporaneous strata have following their
accumulation

This framework is most commonly interpreted to have been generated during


changes in relative sea level when surfaces formed during the associated deposition
and erosion. These surfaces, and the geometries of the sediments they envelope,
are combined to interpret the depositional setting of clastic and carbonate sediments
sediments, be they continental, marginal marine, basin margins and down-slope
settings of basins. The interpretation is better, and predictions of local and regional
stratigraphy more accurate, when the sequence stratigraphic framework
integrates accommodation successions with understanding of:
 Steno's Laws of sediment accumulation
 Walther's Law of the vertical and lateral equivalence of sediments
 The chronology of the succession of strata forming the geologic section
 Sedimentary structures

The major problem with sequence stratigraphy is that the definition, terminology and
interpretation of the surfaces of sequence stratigraphy is complex and often contentious (Neal
& Abrue, 2009). Terminology often involves conceptual depositional models, sea level,
and/or age duration, mixing interpretation with observation, applying different genetic names
to the same surfaces or deposits that depend on the model employed. In the hope of
circumnavigating this, the site places a heavy dependence on the explanation of terminology
linked pop-up boxes whose contents is intended to clarify the understanding and use of this
discipline of stratigraphy. In the hope of circumnavigating this, the site places a heavy
dependence on the explanation of terminology linked pop-up boxes whose contents is
intended to clarify the understanding and use of this discipline of stratigraphy.

Bounding Surfaces of Sequence Stratigraphy


The key to using sequence stratigraphy as a tool for interpreting the sedimentary
section are the major bounding and subdividing surfaces (see the banner image
above). These surfaces are commonly generated by the changes in relative sea
level. From the moment the oceans were first generated, their water volume and
distribution across the globe has varied. This would have resulted in transgressions that
caused the shore and the near-shore being flooded so transgressive surfaces (TS)
formed. When the rate of sea level rise reached its most rapid change, the rate of
sediment accumulating seaward of the shore slowed while from the onset of the
Phanerozoic the pelagic and benthic organic matter continued to accumulate. These
organics sequestered radioactive elements in the water column. The result is that
from the Phanerozoic on the sediments have a strong radioactive signal on gamma
logs with matching condensed sections of fossils that accumulated on a surface or in
a thin zone which is known as the maximum flooding surface (mfs). In contrast, a
drop in sea level may cause the shore and the near-shore to be eroded, forming
sequence boundaries (SB).

The Protocol Of Sequence Stratigraphic Analysis


Once interpretative analysis has identified the subdividing surfaces listed above, the
successions of layered stratal geometries of discrete contemporaneous sediment filling the
accommodation are examined. These stratal geometries of accommodation
successions commonly either:

 Prograde and aggrade (PA)


 Retrograde (R)
 Aggrade to prograde to degrade (APD)

The characteristic geometries of accommodation successions are


interpreted to differentiate the systems tracts and sequences listed
below. First the interpreter reverses the order of deposition by back-
stripping the contemporaneous stratal geometries from youngest to
oldest, These are then reassembled in the order of successive
accumulation. This reassembly uses the template of subdividing surfaces, lithofacies
geometry, and fauna to interpret the evolving character of depositional setting and
build conceptual models. Reassembly tracks the evolution of the sedimentary
system, its hydrodynamic setting, and accommodation (see Basin Clastic Sequence
Hierarchies for such a reassembly).
The back-stripping and analysis is aided by a variety of contemporaneous elements
subdivided by the surfaces and their hierarchy from low frequency to high frequency.
These include:

 Sequences
 Systems tract
 Parasequences and/or cycles
 Bedsets
 Beds

As these sediments are reassembled, the genetic character of the sequences,


systems tracts, parasequences, and beds will be seen as products of changes in
accommodation. A limit to this analytical strategy is often the extent of ones
understanding of the inferred depositional setting. The advantage of the strategy is it
considers new questions, leading to more realistic interpretations and enhanced
predictions of lithofacies heterogeneity's. Thus the sequence stratigraphic framework
is used to analyze and explain how sedimentary rocks acquire their layered
character, lithology, texture, faunal associations and other properties. These
properties in turn can be used to explain how the mechanisms of sediment
accumulation, erosion and inter-related processes produced the current
configuration of these rocks.
The sequence stratigraphic approach recommended on this web site for the
interpretation of sedimentary rocks contrasts with:

 Lithostratigraphic analysis which maps lithofacies independent of subdividing


external and internal boundaries
 Allostratigraphic analysis that identifies and includes as time markers
bounding discontinuities including erosion surfaces, marine flooding surfaces,
and markers that include tuffs, tempestites, and/or turbidite boundaries etc.,
but considers these as independent of any model of base level change

Analyses based on sequence stratigraphy apply allostratigraphic models to interpret


the depositional origin of these sedimentary strata but in contrast assumes, though
this is not always stated, an implicit connection of the contermporaneous strata to
relative sea level or base level change.

Niels Steno and Johannes Walther


Steno and Walther's contributions to sedimentary interpretation and so sequence
stratigraphy were profound. Steno established the order and the way sediments
were laid down. His principle of superposition recognized that older sedimentary
layers underlie the newer layers. His principle of original horizontality recorded how
sediments are deposited to fill a basal irregular surface enclosed above by a smooth
surface. His principle of lateral continuity proposed contemporaneous layers of
sediment are continuous till they pinch out, or a barrier prevents their further spread
during deposition, or subsequent abrupt changes in the landscape break up the
sediment layers. Walther then recognized that as depositional settings change their
lateral position and fill accommodation, so the sedimentary facies of adjacent
depositional settings succeed one another as a vertical sequence.

Stacking Patterns and Geometries


To establish the depositional setting of the sedimentary section, sequence
stratigraphy uses the geometric arrangement of sedimentary fill, particularly the
vertical succession of sedimentary facies geometries and their enveloping surfaces
known as stacking satterns. The geometries and so stacking patterns of un-
cemented carbonates and clastics are similar. This is because both respond to
changes in base level, both can be subdivided by similar surfaces and both respond
to wave and current movement similarly and may be transported.

Never the less major differences in the sequence stratigraphy of the two sediments
exist. All clastics are transported to their depositional resting place while carbonates
are produced and accumulate "in situ". Rates of carbonate production are linked to
photosynthesis, so are depth dependent with rates greatest close to the air/sea
interface. This means that carbonate facies and their fabrics are often used as
indicators of sea level position. Additionally rates of carbonate accumulation often
have a biochemical and physicochemical origin that is influenced by the chemistry of
the water from which they are precipitated. Stacking patterns of both sediments are
expressed by geometric bodies that may be:

 Unconfined by topography
 Confined within eroded topography.

Stacking patterns for both clastics and carbonates that are the product of physical
accommodation vary between:

 Unconfined sheets that:


o Prograde (step seaward) as
o Retrograde (step landward)
o Aggrade (build vertically)

 Unconfined prograding carbonate sheets that are the product of physical


accommodation are further subdivided below into:
o Low angle ramps of fine shallow-water carbonate that in deeper-water
pass to gravels
o Homoclinal ramps of fine shallow-water carbonate
o Distally steepened ramps of shallow-water grain-dominated carbonate

 Unconfined carbonate platform sheet geometries formed with ecological


accommodation form
o Flat-topped open shelves with moderate shallow-water ecological
accommodation
o Reef-rimmed platform with highest shallow-water ecological
accommodation
o Massive steep to cliffed margins with maximum shallow-water
ecological accommodation

Confined bodies represented by fill of Incised topography include

 Subaerial incised valleys


 Submarine incised valleys

Channel fill and stacking of confining valleys, unconfined lobes and sheets may be
expressed as:

 Organized bodies
 Disorganized bodies
 Multi-storied
 Amalgamated

Other Stratigraphic Tools Utilized with Sequence Stratigraphy


Prediction and interpretation improves not only when sequence stratigraphy is
coupled to the Laws of Steno and Walther but when tied to indicators of deposition
and time. Indicators of depositional setting include:

 Ichnofacies and fossils


 Sedimentary structures
 Volcanics
 Storm layers or tempestites
 Sequence stratigraphic boundaries

Chronostratigraphic markers include:

 Fossils
 Magneto-stratigraphic
 Radioactive markers or gamma ray log signalRadioactive markers or gamma
ray log signal
 Radiometric markers

Terminology
Though the linkage between the sequence stratigraphy and the other sub-disciplines
of stratigraphy can be ‘fuzzy’ these links are important to prediction and
interpretation. A key problem to strengthing theses links is not only that the
terminology of sequence stratigraphy carries connotations related to the
interpretation of the surfaces used to interpret the stratigraphic section but also a
consideration of sedimentology and chronostratigraphy. How the terminology is
defined and used and/or fits preconceived classifications is tied to the character of
the data and stratigraphic techniques used. In the end it is up to the user to consider
their data, and the goals of their interpretations. They should be able to explain their
choice of terms and then make their interpretation!

Use of the “Over Simplification” of time as it relates to sequence stratigraphy


As is explained in the pages that follow, using the above approach geologists infer
the processes responsible for that sedimentary rock and so interpret its origin. The
sedimentary layering of a stratigraphic section has a vast array of dimensional
hierarchies. These range from units millimeters thick that might be formed over
seconds to thousands of feet thick and formed of millions of years. As much of the
literature related to these surfaces indicates, it should be recognized that whatever
the dimension of a layer is and whatever the time involved in its deposition, each
layer is assumed to be contemporaneous though it is bounded by surfaces that
transgress time (Wheeler, (1958); Middleton, (1973); Vail et al (1977); Galloway,
(1989); Catuneanu, et al, (1998); Schwarzacher, (2000); Catuneanu, (2002); Embry,
(2002); Cross, and Lessenger, (submitted)). This means an interpretation of the
depositional setting for a section cut by these diachronous surfaces contravenes
Walther's Law. Most interpreters accept and take into consideration that the layered
units bounded by these surfaces formed at different times, and recognize that the
subdividing surfaces are of a higher order frequency than the time envelope of the
parasequence being considered. In other words the situation is simplified when the
surfaces are taken to represent instances in time between which assumed to be
contemporaneous sediments accumulated continuously over some unit of time. Thus
the surfaces of the layers transgress time and the sediments filling between these
surfaces also transgress time while being continuously reworked through a series of
geological events.

Figure - Hierarchy of sedimentary structures:


A - Flaser structures from an intertidal flat setting in which the individual
components probably accumulated over minutes but the whole section may
represent tidal cycles over months (Bar Scale - 2 cm)
B - Cross bedded Ordovician carbonates from a beach or nearshore shallow shoal
setting that probably represent accumulation and reworking over several years
C - Flat bedded Mississippian downslope siliciclastic fan deposits; each bed of
which may have accumulated over a period of hours, though the whole section
encompasses potentially hundreds of thousands of years (bar scale - 1m)
Thus it should be recognized that in sedimentary interpretation the application of
Steno's principles and Walther's Law provide powerful and useful simplifications that
assume the sediments packaged by surfaces accumulated within discrete moments
of time, and are considered contemperaneous. If one thinks about this, these
simplifications don't contravene logic (which is literally Fuzzy) and aid in the
interpretation of the sedimentary section. The above discussion provides a general
introduction to the subdivision of the sedimentary section by the surfaces listed
above and their relationship to base level change. For for a more complete and
thorough discussion of this topic you should read Catuneanu (2002).

Introduction to the Web Site


This Web Site explains:

1. How to make sequence stratigraphic interpretations of sedimentary sections:

 Subdivide of these sections into sequences, parasequences &/or their


associated systems tracts
 Determine their depositional setting
 Characterize and predict of the extent of their lithofacies, particularly
when associated with hydrocarbon reservoirs, and aquifers.

2. The use of:

 2-D and 3-D seismic sections


o Well log data
o Outcrops

to identify and correlate surfaces of:

 Erosion and non-deposition (sequence boundaries [SB])


 Transgressive surfaces [TS]
 Maximum flooding surfaces [mfs]

3. How the above surfaces have time significance and establish:

 a relative time framework for the sedimentary succession


 the inter-relationship of the depositional settings and their lateral
correlation
 a compartmentalization of hydrocarbon reservoirs

In summary this web site explains how "Sequence Stratigraphy" can be used to
study sedimentary rock relationships within a time-stratigraphic framework of
repetitive, genetically related contemporaneous strata bounded by surfaces of
erosion or non-deposition, or their correlative conformity
(Posamentier et al., 1988; Van Wagoner et al., 1988).

Sequence stratigraphy of depositional systems


This site provides an overview of both modern and ancient depositional systems in
terms of their sequence stratigraphy and their character. These systems include:
 Clastic Systems
o Marine: Barrier island Coasts; Deltaic systems; deepwater fans;
Deepwater basins
o Continental: Glacial; Aeolian; Alluvial Fans; Braided Streams; Coarse
and fine grained fluvial systems; Lacustrian
 carbonate Systems: Inner carbonate shelf; Outer carbonate Shelf and
Margins; Deepwater carbonates

Using the sidebar menu you can select topics in sequence stratigraphy and access
exercises related to this. You should be able to learn how to subdivide the
sedimentary section into packages defined by bounding unconformities and internal
surfaces. You will be able to see how sequences, parasequences and their
associated systems tracts are the products of changes in relative sea level and rates
of sedimentation. The various forms of sequence stratigraphic analyses outlined
include the use of seismic cross-sections, well logs and outcrop studies of
sedimentary rocks to infer changes of relative sea level and rates of sedimentation.
You will be shown how to construct chronostratigrapic diagrams and also be show
how to predict facies geometries and build depositional models using a variety of
techniques!

Later in the section on the Basics of Sequence Stratigraphy you will be introduced
to the details of how systems tracts respond to changing base level. However as a
preview you can trace clastic systems tract evolution through time in the linked
movie!

Friday, April 24, 2015

Sequence Stratigraphy > Basics of Sequence Stratigraphy

Click here for more Streamed Web Based Lectures!

The streamed video chalk board lecture below explains the ideal Vail et al 1977 sequence,
and considers this in terms of the component systems tracts, and their gross geometries as
products of changes in relative sea level (base level) and the evolving accommodation' space.
You are introduced to sea level and base level, sequence, concepts of sequence stratigraphy
from the perspective of evolving sediment geometry through time, and relative sea level
changes. The geological setting described in the lecture is hypothetical, and represents the
first of a series of other hypothetical and real geological examples of clastic and carbonate
sequences that are described in the pages that follow.
Downloads Formats:
Real Player
MP4 (zipped)
AVI (zipped)

The movie makes the following points:

The lecture seen in the movie involves a geologic model that makes the following
assumptions::

 Sea level position varied


 Subsidence was constant
 Sediment supply was constant

The sequence is divided by surfaces systems tracts. Each systems tract is represented by a
collection of the sediments of the associated sedimentary depositional systems that were
active during the different phases of base level change. Thus systems tract sediments can be
considered as sedimentary units that were deposited synchronously and can be mapped as
being enclosed by continuous surfaces that extend from sub-aerial and to sub-aqueous
settings.

The systems tracts defined in order of deposition to form the ideal sequence are:

 Early phase lowstand systems tract


 Late phase lowstand systems tract
 Transgressive systems tract
 Highstand systems tract

Early phase lowstand systems tract is associated with:

 Falling stage of relative sea level induced by eustasy falling rapidly and/or
tectonic uplift outpacing the rate of change in sea level position
 Fluvial incision up dip with formation of an unconformity or sequence
boundary and the focus of sediment input at the shoreline
 Forced regressions induced by the lack of accommodation producing stacking
patterns of downward stepping prograding clinoforms over the condensed
section formed during the previous transgressive and highstand systems
tracts
 Slope instability caused by the rapid deposition of sediment from the fluvial
systems
 Basin floor fans formed from sediment transported from the shelf margin when
this fails under the weight of the rapid sediment accumulation associated with
the forced regression
 Shelf margin and slope fans form when rates of sedimentation slows and
slope instability is reduced so sediment is not displaced so far downslope
 Onlap of sediments onto the prograding clinoforms below the shelf break
 The lower boundaries of the early phase lowstand systems tract are the updip
unconformity and the top of the downdip condensed section. These surfaces
form by different mechanisms and have different time significance
 The top of the downdip condensed section immediately underlies the
downlapping prograding clinoforms of the forced regression
 The top of the early phase lowstand systems tract in theory is marked by an
initial onlap onto the often eroded surface of the prograding clinoforms of the
forced regression

Late phase lowstand systems tract is associated with:

 A slow relative sea level rise is induced when eustasy begins to rise slowly
and/or tectonic uplift slows
 Sediment is now outpaced by an increase in accommodation and in response
the sediment begins to onlap onto the basin margin
 River profiles stabilize
 Valleys backfill
 Prograding lowstand clinoforms form and are capped by topset layers that
onlap, aggrade, become thicker upward and landward

Transgressive systems tract is associated with:

 A rapid relative sea level rise above the shelf margin occurs when eustasy
begins to rise rapidly, exceeding the effects of any tectonic uplift
 Condensed sequences are often composed of sediment layers rich in the
tests of fauna that are no longer masked by sediment accumulation because
sedimentation rates are very slow in response to the greater area of sea floor
exposed to sedimentation
 Ravinement erosion surface formed when the transgressing sea reworks
either the prior sequence boundary or the sediments that may have collected
during the forced regression that may have followed the generation of that
sequence boundary.
 Maximum flooding surface forms when the last fine-grained widespread
transgressive sediment collects before the high stand builds out over it.

Highstand systems tract is associated with:


 Slow rise of relative sea level followed by a slow fall; essentially a still stand of
base level when the slower rate eustatic change balances that of tectonic
motion
 Sediment outpacing loss of accommodation
 River profiles stabilize
 River valleys are dispersed laterally in a position landward of the shelf margin.
 Prograding highstand clinoforms develop capped by aggrading topsets that
become thinner upward.

To view carbonates and clastics with different inputs we refer you to the Quick Time movies
on our page.

Wednesday, January 14, 201

Stratigraphic framework and sedimentary systems

Figure: Information from detailed geological analysis populates a sequence stratigraphic framework
so leading to accurate interpretations of depositional setting and predictions of lithofacies geometries
in unknown portions of a basin.

Introduction
Lyell’s premise in 1830, 1832 & 1833 that “the present is the key to the past” is
fundamental to sequence stratigraphy's recognition that the sedimentary record of
the earth’s crust is the product of uniform and common physical processes that
interacted with sediments as they accumulated. This means that the processes
responsible for known portions of a geological section can be determined by careful
description and analysis focused on component lithofacies, fabric and geometry.
The origins of these sediments are then be interpreted by comparison with
observations of similar features in modern sedimentary systems and their processes,
results of flume experiments, computer simulations that re-create the fabrics and
geometries seen in the field and laboratory, and the body of known geological
information geologists have amassed in the geological literature. These
interpretations are then integrated with the sequence stratigraphic framework of
erosional and depositional surfaces that enclose and subdivide the section. This
template extends the interpretation of the depositional setting and predictions of
lithofacies’ geometries away from the known areas. It also aids prediction of
sedimentary rocks likely to contain both hydrocarbon and water resources and what
their characteristic fabrics might be.

Figure: A sequence stratigraphic framework leads to interpretations of depositional setting and


predictions of lithofacies geometries.

This portion of the web site is based on the geologic literature, in particular
Catuneanu, et al, (2011) and its summary of a collective understanding of sequence
stratigraphy. This section defines and explains the origins of:

 The sequence stratigraphic framework of enveloping erosional and


depositional surface boundaries
 The physical processes that generated the gross sediment geometric end
members:
o Sequences
o Systems tracts
o Parasequences.

The relationship of the different depositional systems, with their different genetically
related stratigraphic elements is then described in terms of the above sequence
stratigraphic geometric end members. Further consideration is given to the stratal
stacking patterns of different depositional systems combined to define trends in
geometric character and systems tracts. The text explains how the framework of the
elements of each depositional setting has common hierarchies that enable
reconstruction of the sedimentary section and the prediction of lithofacies and
paleogeography away from control points. The sources of information in this text are
referenced, though if inadvertent omissions occur, it is likely the information was
thought to be axiomatic.

Bounding surfaces and architectural elements


The interpretation of the depositional setting of sedimentary strata is enhanced
through understanding the origin of the character of the bounding surfaces to the
sedimentary geometries of the component lithofacies. The bounding erosional and
depositional surfaces of sedimentary geometries have hierarchical order. For
instance partings with a high frequency include those of that subdivide shales, while
bedding planes, the features commonly used to interpret the origin of the
sedimentary section, are of lower frequency. In contrast lower frequency bounding
surfaces include unconformities (Hutton, 1788), whose stratigraphic importance
increases where they subdivide sedimentary rocks of varying ages.
Brookfield (1977) was among the first known geologists to apply hierarchical order to
surface boundaries subdividing sedimentary rocks, using his perspective of surfaces
identified by Stokes (1968) in aeolian sediments. Brookfield (1977) identified:

 First order boundaries that cut across underlying aeolian sediments when the
migration of “draa” dunes occurred.
 Second order surfaces related to migration of transverse dunes,
 Third order boundaries that enclose groups of laminae interpreted to be the
products of local events within the depositional cycle.

Allen (1983) describing fluviatile systems extended this by recognizing that bounding
surfaces may be non- erosional or erosional. Using this he identified four surfaces:-

 Concordant -non-erosional (normal bedding)


 Discordant- non-erosional reactivation surface
 Concordant - erosional
 Discordant - erosional contacts

Studying braided streams, Allen (1983) used these surfaces to associate at least
eight geometric shapes with specific lithologies and fabrics that he named
"architectural elements”. Miall (1985) utilized this concept of depositional
architectural elements to further classify and communicate information on the
character and origins other fluvial depositional systems.

The application of the concepts of architectural elements is now widely used for most
depositional systems. For example Pickering et al (1998) subdivided deeper water
sedimentary bodies, recognizing a hierarchy of enveloping boundaries that
genetically related discrete stratigraphic “architectural elements", "bodies", or "units"
or “groups”. Others, including Sprague et al 2008, used a top-down hierarchical
classification of "architectural elements" for deep-water settings that starts at a
sedimentary basin scale. Sub-divided downward these form a series of broad
elements includes the larger stacked channel complexes, in turn subdivided ever
downward to an ultimate subdivision of laminae or even the individual sand grain.
This top down classification is used to provide a framework of the basin to its
interrelated broader larger scale "architectural elements" and their tie to the smaller
scale "architectural elements". This can be inverted from small to large equally
effectively.
A sequence stratigraphic analysis will iteratively use mixes and matches of a top
down classification with a bottom-up classification. This interactive approach uses
the general to guide an understanding of the specific and vice versa.

Below are further examples of the hierarchical architectures of different sedimentary


systems. Click on the thumb nail and right click the shadow box image to see full
size image! This will work on most of the images on site.

Beach Barrier Deep Water Carbonate


System System Systems

Figure: Sequence stratigraphic framework of Permian Section exposed in Guadalupe Mountains


(Tinker, 1998)
Surfaces or Boundaries of the Sequence Stratigraphic Framework
As indicated sequence stratigraphy uses a framework of surfaces or boundaries that
define "sequences", "systems tracts", and parasequences. These boundaries
include:

 Sequence Boundary (SB)


The sequence boundary (SB) envelopes a sequence extending down dip to a
correlative conformity (Figure …) (Mitchum, 1977). This surface demarks the
boundary between a highstand systems tract (HST) and falling stage systems
tract (FSST) (Catuneanu, et al, 2011).

 Basal surface of falling stage systems tract (FSST)


This was defined by Hunt and Tucker (1992) as the surface that underlies the
marine sedimentary wedge that builds seaward during a forced regression of
the shoreline. This basal surface of falling stage systems tract (FSST) (Plint
and Nummedal, 2000) is also known as the early lowstand systems tract
(ELST) (Posamentier and Allen, 1999) and overlies the transgressive systems
tract (Catuneanu, et al, 2011).
Transgressive surface (TS)

 This surface is the first significant marine-flooding surface in a sequence. In


most siliciclastic and some carbonate successions it occurs when the rate of
creation of accommodation space is greater than the rate of sediment supply.
Transgressive surface (TS) lies over the Lowstand System Tract (LST) and
beneath the transgressive systems tract (TST) (Catuneanu, et al, 2011).
 Maximum flooding surface (mfs)
Seismically an mfs is often expressed as a downlap surface and is a zone
where slow rates of deposition result in thin and fine grained pelagic-
hemipelagic sediments forming a condensed section (Mitchum, 1977). The
mfs marks the bounding surface between coarsening and/or fining upward
cycles. This commonly widespread zone is often characterized by the
presence of radioactive and often organic rich shales, glauconite,
hardgrounds and is composed of thin bedded concentrations of fauna
(condensed sections) with high abundance and diversity. This surface marks
the time of maximum flooding or transgression of the shelf and lies beneath
the highstand systems tract (Catuneanu, et al, 2011).

The subdividing "surfaces" of the sequence stratigraphic framework that envelopes


and encloses discrete geometric bodies of sediment establishes their order of
accumulation from oldest to youngest. Interpretation is conducted by dis-assembling
(backstripped) sedimentary bodies and then reassembling them in order in which
they formed. The depositional setting is determined through the iterative reassembly
and a consideration of the origin of the subdividing surfaces, geometry, lithofacies
and fauna and evolving character. ‘‘Each stratal unit is defined and identified only by
physical relationships of the strata, including lateral continuity and geometry of the
surfaces bounding the units, vertical stacking patterns, and lateral geometry of the
strata within the units"(Van Wagoner et al., 1990).

In addition the process of interpreting the origins of these surfaces, the depositional
setting and gross sedimentary geometry of the rocks enclosed within the sequence
stratigraphic framework involves Niels Steno’s Laws of Superposition and Walther's
Law. The latter proposes that a vertical succession of sedimentary facies likely
accumulated in adjacent depositional settings whether within a parasequence,
system tract or a sequence. Paradoxically the surfaces used to subdivide
stratigraphic sections are diachronous but in the process of interpretation this is
oversimplified and the diachronous character of the surfaces is essentially ignored.
For instance Holbrook and Bhattacharya (2012) indicate sub-aerial unconformities in
fluvial systems meet these criteria but suggest these boundaries can still be used to
bound systems despite intense diachroneity. The results of the use of the sequence
stratigraphic methodology is that interpretation of depositional setting and a
prediction of gross sedimentary geometry are confirmed in the field and with
subsurface data.

Without exception all these surfaces, and in many cases the zones inferred to
contain them and the sediment they enclose, transgress time, in other words are
diachronous and in some cases may not even have the regional extent proposed for
them, and so may be miss-correlated. Also the sequence stratigraphic surfaces
have become largely conceptual surfaces imposed upon tangible rocks (Helland-
Hansen, and Martinsen, 1996; Catuneanu, 2006; Embry et al., 2007; Miall, 2004;
Holbrook and Bhattacharya, 2012). For instance Holbrook and Bhattacharya (2012)
point out that the subaerial unconformity is more often than not is a conceptual
surface and is assumed to be an approximate time barrier that includes the defining
traits of originating as a ‘subaerial erosional surface’ preserved as an ‘unconformity
that separates younger from older strata with significant hiatus.

As a result all the sequence stratigraphic surfaces, (SB, BSTSST, TS, and mfs) often
violate Walther's Law, since they record shifts in facies deposition during
transgression and regression and/or rates of change of accommodation, particularly
at basin margins and along strike (Catuneanu, 2006). It is argued here that while
most of these sequence surfaces do not exactly fit their defining characteristics they
can be mapped and bound enclosing facies that accumulated over a generally short
time. The conceptual character of a surface is likely to be more so with increasing
hierarchical rank (Catuneanu, et al, 2011). Despite these caveats sequence
stratigraphic surfaces are useful for general "fuzzy" oversimple correlation. As
interpretive tools they are commonly used and are often referred to in the
stratigraphic literature. To conclude the definition of these surfaces is oversimplified
and form discrete boundaries that can be traced beyond the scale of a single valley
or comparable local depositional system, and used to make accurate facies
predictions.
Contention arises from the nomenclature of each of the sequence stratigraphic
surfaces and the bodies they contain. This argument is based not so much on the
constantly changing nomenclature as the developing understanding of sedimentary
systems and their interpretation. However it is unfortunate that, though changes in
nomenclature are well intentioned, these changes often add further to the confusion
to a scientific methodology already weighed down with complex multi-word and
multi-syllable terminology.

Sequence Stratigraphic Units


The gross sediment geometric end members are represented by sequences,
systems tracts, and parasequences are a hierarchy of stratigraphic packages or
units of similar sediment strata whose geometries are of increasingly higher
frequency, and are related to changes in the space or accommodation available for
sediment fill; accommodation driven by changes in eustasy and tectonics (Jervey,
1988). It is shown how the geometric hierarchy is expressed in these packages by
the subdividing and enveloping surfaces found in sedimentary sections. These
bodies and their lithofacies are keys to determining and interpreting the depositional
setting of the sedimentary sections that contain these bodies. It is contended that if
depositional systems are described in terms of the geometric hierarchy of their
lithofacies and elements this leads to a better understanding of the depositional
origins of similar sedimentary bodies in the rock record.

Figure: accommodation is "the space available for potential sediment accumulation" driven by relative
sea level (Jervey, 1998). Curray, (1964), Posamentier & Allen, (1999), Coe et al (2002), and
Catuneanu (2002) suggest rates of sedimentation are a co-equal control of accommodation.
Sequence
The sequences of the sedimentary record are generated by cycles of change in
accommodation and/or sediment supply that also form similar sequence stratigraphic
surfaces through geologic time. cycles may be symmetrical or asymmetrical, and
may or may not contain the systems tracts of a fully developed sequence. A function
of scale, sequences and their bounding surfaces may have different hierarchical
orders recording a series of geological events, and processes in sedimentary rocks
that form a relatively conformable succession of genetically related strata. Their
upper surfaces and bases are bounded by unconformities and their correlative
conformities (Vail, et al., 1977). A sequence is formed by a succession of genetically
linked deposition systems (systems tracts) interpreted to have accumulated between
eustatic-fall inflection points (Posamentier, et al., 1988). The sequences and
enclosed system tracts are subdivided and/or bounded by a variety of "key" surfaces
that bound or envelop them. As described above these include sequence
boundaries (SB), the basal surfaces of falling stage systems tracts (BSFSST),
transgressive surfaces (TS) and a maximum flooding surfaces (mfs). These
erosional and depositional surfaces mark changes in depositional regime or
"thresholds" across that boundary.

Sloss et al., (1949, and 1963) originally defined a 'sequence' as an unconformity-


bounded stratigraphic unit. Mitchum (1977) modified this to define a sequence as “a
relatively conformable succession of genetically related strata bounded by
unconformities or their correlative conformities.

Through the 80's and 90's sequences were defined from several perspectives
Catuneanu (2011):
 Depositional sequences, bounded by subaerial unconformities and their
marine correlative conformities (e. g., Vail 1987; Posamentier et al. 1988; Van
Wagoner et al. 1988, 1990; Vail et al. 1991; Hunt and Tucker 1992)
 Genetic stratigraphic sequences, bounded by maximum flooding surfaces
(Galloway 1989)
 Transgressive-Regressive (T-R) sequences, also referred to as T-R cycles,
bounded by maximum regressive surfaces (Johnson and Murphy 1984;
Johnson et al. 1985). The T-R sequence was subsequently redefined by
Embry and Johannessen (1992) as a unit bounded by composite surfaces
that include the subaerial unconformity and the marine portion of the
maximum regressive surface.

Catuneanu (2009 and 2011) felt that the various types of sequence should be
encompassed by the definition. They redefined a sequence as “a succession of
strata deposited during a full cycle of change in accommodation or sediment supply”.
The definition is generic, model-independent, and embraces the sequences listed
above that may develop at any spatial or temporal scale. The requirement that a
sequence coincide with a full stratigraphic cycle means that a sequence can be
distinguished from component systems tracts. Existing sequence stratigraphic
schemes incorporate a full cycle of change in accommodation or sediment supply
with a beginning and the end of one cycle manifest by the same kind of event. This is
the onset of a relative sea-level fall; the end of relative sea-level fall; the end of
regression; or the end of transgression. In contrast, the boundaries of any systems
tract correspond to different 'events’ within a relative sea-level cycle. The definition
of a sequence is updated to be the fundamental statal unit of sequence stratigraphy
(Catuneanu et al., 2011). As with Vail, et al., (1977) they see this as represented by
a relatively conformable succession of genetically related strata bounded by surfaces
but extend this to correspond to a full cycle of base-level changes or shoreline shifts
depending on the sequence model being employed.
The Posamentier et al.'s 1988 original interpretation was that sediments
accumulated during the falling stage of sea level cycle and this was where the
sequence boundary should fall. Hunt & Tucker, (1992), 1995) discuss the role of
forced-regressions and where the sequence boundary should be placed with respect
to sea level position. Hunt believes that the position of the sequence boundary
should be placed at the lowest position reached by sea level. A number of geologists
support this contention. One of these is Pomar (1991) who recognizes that within the
Late Miocene reefal platform of Mallorca, the sequence boundary and the downlap
surface are both coeval and formed during the falling stage of sea level. Both
surfaces bound the offlapping systems tract and merge landward in the erosion
surface and, basinward, in the condensed interval. Note the correlative conformity
on the top of the basin floor fan as suggested by Vail, 1987, versus the Hunt and
Tucker, 1992 & 1995, models.

Systems tracts
A systems tract is a subdivision of a sequence independent of spatial and temporal
scales representing “a linkage of contemporaneous depositional systems” (Brown
and Fisher, 1977). It consists of a relatively conformable succession of genetically
related strata bounded by conformable or unconformable sequence stratigraphic
surfaces with an internal architecture that varies from a succession of facies that
include high-frequency cycles driven by orbital forcing to a parasequence set or a set
of higher frequency cycles. Systems tracts are interpreted on the basis of stratal
stacking patterns, position within the sequence, and types of bounding surface (Van
Wagoner et al., 1987, 1988, 1990; Posamentier et al. 1988; Van Wagoner 1995;
Posamentier and Allen 1999). Systems tracts may be either shoreline-related, where
their origin can be linked to particular types of shoreline trajectory, or shoreline-
independent, where a genetic link to coeval shorelines cannot be determined
(Catuneanu, 2011).

Shoreline-Related Systems Tracts


Shoreline-related systems tracts are depositional systems that are often tied to
shoreline trajectory, be this a forced regression, normal regression, or transgression,
and are commonly interpreted to form during specific phases of the relative sea-level
cycle (Posamentier et al. 1988; Hunt and Tucker 1992; Posamentier and Allen 1999;
Catuneanu 2006; Catuneanu et al. 2009; Catuneanu et al. 2011). These systems
tracts may have different scales, and are defined by distinct stratal stacking patterns
(Figure…). Forced regressive deposits include 'early lowstand’, 'late highstand’,
'forced-regressive wedge', and 'falling-stage'. Normal regressive deposits include
'late lowstand’ and 'lowstand’, 'early highstand’ and 'highstand’ systems tracts.
Transgressive systems tract is composed of regressive stratal stacking patterns
comprise. Five of these systems tracts are described below.

Falling-Stage systems tract (FSST)


The FSST is formed by forced regressive deposits that accumulated after the onset
of a relative sea-level fall and before the start of the next relative sea-level rise. The
FSST lies directly on the sequence boundary sensu Posamentier and Allen (1999)
and is capped by the overlying lowstand systems tract (LST) sediments. Hunt and
Tucker (1992) differ with this placing the sequence boundary above the FSST, where
this boundary marks the termination of one cycle of deposition and the start of
another. Depending on the gradient of the depositional profile, the rate of sediment
supply, and the rate of relative sea-level fall, a variety of 'attached' or 'detached'
parasequence stacking patterns can be produced (Posamentier and Morris, 2000).
Catuneanu (2011) explain that the terminology applied to this systems tract varies
from 'forced regressive wedge' (Hunt and Tucker 1992) to 'falling sea-level'
(Nummedal 1992) and 'falling-stage' (Ainsworth 1994). The simpler 'falling-stage' has
been generally adopted by more recent work (e. g., Plint and Nummedal 2000;
Catuneanu 2006). This systems tract has also been termed the early lowstand
systems tract (Posamentier et al. 1988; Posamentier and Allen, 1999). The fall in
relative sea level is evidenced by the erosion of the subaerially exposed sediment
surface updip of the coastline at the end of forced regression, and the formation of a
diachronous subaerial unconformity that caps the highstand systems tract (HST).
The subaerial unconformity may be onlapped by fluvial deposits that belong to the
lowstand or the transgressive systems tracts. The subaerial unconformity may also
be reworked by a time-transgressive marine ravinement surface overlain by a
sediment lag.
Lowstand Systems Tract (LST)
The LST is formed by sediments that accumulate after the onset of relative sea-level
rise, during normal regression, on top of the FSST corresponding to an updip
subaerial unconformity. stacking patterns of clinoforms may forestep, and aggrade,
particularly in siliciclastic systems, thicken downdip, with a topset of fluvial, coastal
plain and/or delta plain deposits. LST sediments often fill or partially infill incised
valleys that were cut into the underlying HST and other earlier deposits, during the
forced regression. This systems tract has also been termed the late lowstand
systems tract (Posamentier et al. 1988; Posamentier and Allen 1999) or the
Lowstand Prograding Wedge systems tract (Hunt and Tucker 1992). In earlier
papers the 'shelf-margin systems tract' was recognized as the lowermost systems
tract associated with a 'type 2 'sequence boundary (Posamentier et al. 1988). With
the abandonment of the distinction between types 1 and 2 sequence boundaries, this
term is now redundant (Posamentier and Allen 1999; Catuneanu 2006); these
deposits are now considered to be part of the LST.

Transgressive Systems Tract (TST)


The TST is formed by sediments that accumulated from the onset of transgression
until the time of maximum transgression of the coast, just prior to the renewed
regression of the HST. The TST lies directly on the maximum regressive surface
formed at the end of regression (also termed a transgressive surface). A
transgressive systems tract is overlain by the maximum flooding surface (MFS)
formed when marine sediments reach their most landward position. stacking patterns
exhibit backstepping, onlapping, retrogradational clinoforms that, particularly in
siliciclastic systems, thicken landward. In cases where there is a high sediment
supply the parasequences may be aggradational.

Highstand Systems Tract (HST)


The HST includes the progradational deposits that form when sediment
accumulation rates exceed the rate of increase in accommodation during the late
stages of relative sea-level rise (Fig. 2). The HST lies directly on the mfs formed
when marine sediments reached their most landward position. This systems tract is
capped by the subaerial unconformity and its correlative conformity sensu
Posamentier and Allen (1999). stacking patterns exhibit prograding and aggrading
clinoforms that commonly thin downdip, capped by a topset of fluvial, coastal plain
and/or delta plain deposits.

Regressive System Tract (RST)


The RST lies above a TST and is overlain by the initial transgressive surface of the
overlying TST. The complete sequence is known as a Transgressive-Regressive (T-
R) sequence (Johnson and Murphy 1984; Embry and Johannessen 1992). The
sediments of this systems tract include the HST, FSST and LST systems tracts
defined above. There are cases where the data available are not sufficient to
differentiate between HST, FSST and HST systems tracts. In such cases the usage
of the regressive systems tract is justified. However, where permitted by data, the
differentiation between the three types of regressive deposits (highstand, falling-
stage, lowstand) is recommended because they refer to different stratal stacking
patterns; are characterized by different sediment dispersal patterns within the basin;
and consequently are associated with different petroleum plays. The last aspect
relates to one of the most significant applications of sequence stratigraphy, which is
to increase the resolution of stratigraphic frameworks that can optimize petroleum
exploration and production development.

Shoreline-Independent Systems Tracts


Shoreline-independent systems tracts are stratigraphic units that form the
subdivisions of sequences in areas where sedimentation processes are unrelated to
shoreline shifts. These systems tracts are defined by specific stratal stacking
patterns that can be recognized and correlated regionally, without reference to
shoreline trajectories (Figs. 9–12). In upstream-controlled fluvial settings, fluvial
accommodation may change independently of changes in accommodation at the
nearest shoreline and create sequences and component low- and high-
accommodation systems tracts (e. g., Shanley and McCabe 1994; Boyd et al. 2000).
Shoreline- independent systems tracts may also be mapped in deep-water settings
controlled by sub-basin tectonism (e. g., Fiduk et al. 1999), but no nomenclature has
been proposed for these situations. The timing of shore-line-independent sequences
and systems tracts is commonly offset relative to that of shoreline-controlled
sequence stratigraphic units and bounding surfaces (e. g., Blum and Tornqvist
2000).

Parasequence
A relatively conformable succession of genetically related beds or bedsets (within a
parasequence set) bounded by marine flooding surfaces or their correlative surfaces
(Van Wagoner, 1985). Patterns of the stacking of parasequence sets are used in
conjunction with boundaries and their position within a sequence to define systems
tracts (Van Wagoner et al., 1988). Thus a parasequence is commonly identified and
separated from other parasequences by flooding surfaces and is often characterized
by a cycle of sediment that either coarsens or fines upward. Thus the flooding
surfaces are usually identified by abrupt and correlatable changes of the grain size of
the sediments on either side of that flooding surface.
This change in grain size is often caused by the abrupt changes in energy that are
associated with the waves or currents of the sea transgressing across the sediment
interface. These abrupt changes in grain size that bound a parasequence can be
identified in well logs, outcrop and seismic and used to identify a parasequence
cycle. Examples of these grain size changes can be seen in the parasequences of
tidal flats, beaches, and deltas.

A parasequence in its original definition (Van Wagoner et al. 1988, 1990) is an


upward-shallowing succession of facies bounded by marine flooding surfaces. A
marine flooding surface is a lithological discontinuity across which there is an abrupt
shift of facies that commonly indicates an abrupt increase in water depth. The
concept was originally defined, and is commonly applied, within the context of
siliciclastic coastal to shallow-water settings, where parasequences correspond to
individual prograding sediment bodies.
In carbonate settings, a parasequence corresponds to a succession of facies
commonly containing a lag deposit or thin deepening interval followed by a thicker
shallowing-upward part, as for example in peritidal cycles. In contrast to sequences
and systems tracts, which may potentially be mapped across an entire sedimentary
basin from fluvial into the deep-water setting, parasequences are geographically
restricted to the coastal to shallow-water areas where marine flooding surfaces may
form (Posamentier and Allen 1999). In the case of carbonate settings, peritidal
cycles can in some cases be correlated into slope and basinal facies (e. g., Tinker
1998, Chen and Tucker, 2003).
Figure: Hierarchy of cyclicity. Each stratigraphic element is a component of the subsequent lower-
order element. Specific interpretations from McKittrick Canyon were used to construct the sequence
stratigraphic framework from Tinker (1998).
For this reason, it has been proposed that a parasequence be expanded to include
all regional meter-scale cycles, whether or not they are bounded by flooding surfaces
(Spence and Tucker 2007; Tucker and Garland 2010). However, following the
principle that a sequence stratigraphic unit is defined by specific bounding surfaces,
many practitioners favor restricting the concept of parasequence to a unit bounded
by marine flooding surfaces, in agreement with the original definition of Van
Wagoner et al. (1988, and 1990).

Scale and stacking patterns


As seen above in the diagram from Tinker (1998) parasequences are commonly
nested within larger scale (higher rank) sequences and systems tracts. However,
scale is not sufficient to differentiate parasequences from sequences. For example,
high-frequency sequences controlled by orbital forcing may develop at scales
comparable to, or even smaller than, those of many parasequences (e. g., Strasser
et al. 1999; Fielding et al. 2008; Tucker et al. 2009). As such, even cycles as thin as
a meter can sometimes be referred to as sequences and be described and
interpreted in terms of sequence stratigraphic surfaces and systems tracts (e. g.,
Posamentier et al. 1992a; Strasser et al. 1999; Tucker et al. 2009).

We recommend the use of the sequence stratigraphic methodology to the analysis of


any small, meter-scale cycles, as long as they display depositional trends that afford
the recognition of systems tracts and diagnostic bounding surfaces. Parasequences
consist of normal regressive, transgressive and forced regressive types of deposit,
and display various stacking patterns.

Parasequences may be stacked in an upstepping succession, in which case they


consist of normal regressive and transgressive deposits that accumulate during a
period of positive accommodation in response to variations in the rates of
accommodation and/or sediment supply. Upstepping parasequences may either be
forestepping or backstepping (see the figure below).

Parasequences may also be stacked in a downstepping succession, in which case


they consist primarily of forced regressive deposits that accumulate during a period
of overall negative accommodation. However, negative accommodation does not
occur during the time of formation of the parasequence boundary. The pattern of
stacking of parasequences defines longer term normal regressions, forced
regressions or transgressions, which correspond to shoreline-related systems tracts
of higher hierarchical rank

Parasequence set
This is often formed by a succession of genetically related parasequences that have
a distinctive stacking pattern that in many cases is bounded by major marine-
flooding surfaces and their correlative surfaces (AAPG Methods in Exploration 7,
1990). These include aggradational parasequence sets, progradational
parasequence sets, and retrogradational parasequence sets. Patterns of the
stacking of parasequence sets are used in conjunction with boundaries and their
position within a sequence to define systems tracts (Van Wagoner et al., 1988).

Figure: High frequency clastic parasequence sets from the Bookcliffs. Note hierarchy of sedimentary
structures and associated seaward to landward depositional systems (after Coe et al, 2003).
a) Upper foreshore planar-cross bedded sandstone of wave swash zone overlying trough-cross
stratified sandstone zone of breaking waves.
b) Burrowed sandstone of the middle shorface.
c) Offshore transition zone.
d) Upper foreshore planar-cross bedded sandstone of wave swash zone.
e) Upper shoreface sandstone of wave swash zone to offshore transition zone between storm wave
base & fairweather base.
Steps for extracting information from a seismic section to build a chronostratigraphic chart:
1. Carefully interpret the pseudo-seismic section by identifying and marking where reflector
terminations intersect seismic surfaces. Identify the type of reflector terminations (onlap,
downlap, toplap, and/or truncation). Reflectors are interpreted to bound contemperanous
sediment bodies.

2.On the chart first identify and number the interpreted seismic reflectors in order of
deposition, starting from oldest (numbered 1) to youngest (top reflector) as in the linked
movie at the top of the page.

3. Transfer the numbered reflectors to a time-scale:

a. The horizontal line matches the length of the pseudo seismic section or distance, with SP
referring to "Shot Points" numbered 10 through 240. The vertical axis on the lower part of the
diagram represents an arbitrary time line. The numbered time intervals, 1 through 30, are
assumed to be of equal duration and bound contemporaneous sedimentary bodies, or
chronosomes.

b. Transfer the horizontal dimension of the interpreted reflectors, starting with oldest
(numbered here as 1), to the bottom of the time chart. Draw this to match the horizontal
length of the equivalent reflector.

c.The introductory chapters described systems tracts and their bounding surfaces (sequence
boundaries (S.B.), the first transgressive surface, ravinement surfaces, condensed sections,
and maximum flooding surfaces). Mark up these properties as related to that reflector (type,
geometry, facies, systems tract info, SB type, etc., within the chronosome).

Continue this process in order of deposition for all remaining reflectors on the seismic cross
section.

4. The void space on the chart now represents areas of non-deposition, erosion, or
condensation of the sedimentary section with thicknesses below the resolution of the seismic.

5. Read Miall (2004), and Wheeler (1958 and 1964) to gain further incite into the application
of this chart and the significance of time in sequence stratigraphy.

Chronostratigraphic
exercise cross-section
illustrated above but
here in 3D perspective
with detail of clastic
response to varying sea
level. Evolving Coastal Margin.

Click for Baum's solution:


Chronostratigraphic-chart

Stratigraphic+Chronostratigraphic
chart

References
Use of well logs for sequence stratigraphic interpretation of the subsurface

The following pages are focused on the use of well logs to build sequences stratigraphic
models of depositional systems. For this process well logs are primarily used to establish the
grain size and the lithology of the penetrated sediment.

The best sequence stratigraphic models of the sedimentary fill of basins are provided by a
combination of seismic data, well logs and cores and outcrop studies in conjunction with
biostratigraphy. The cores and well logs and outcrop studies provide access to a detailed
vertical resolution of sedimentary sections while seismic and outcrop studies provide the
lateral continuity to the sequence stratigraphic framework and the biostratigraphy provides
the time constraints. All these different sequence stratigraphic techniques can be used
independently of each other to produce accurate interpretations of the depositional histories
of the sedimentary fill of a basin but the best models come from a mix of all three.

How well logs are used to interpret sequence stratigraphy


In the use of sequence stratigraphy the interpretations are best determined when well logs are
tied to biostratigraphic markers. Using these two in combination one can:
 Identify, match and tie sequence stratigraphic surfaces
 Interpret the stacking patterns of the vertical sedimentary sequences

The character of electric logs of wells that penetrate clastics often reflect changes in grains
size and so are easier to use in this process, while the logs of wells that penetrate carbonates
often should be calibrated with cores, since carbonates are more susceptible to diagenesis and
their change in character may be affected by more than changes in grain size. The sections
that follow initally focus on the well log response to shallow water clastics and then move on
to their response to shallow water carbonates.

At the start of an interpretation of


sequence stratigraphy using well logs
one must first identify the
predominant of sequence stratigraphic
surfaces. The MOST important of
these surfaces, and the FIRST that
should be identified when using logs,
are maximum flooding surfaces (mfs)
and transgressive surfaces (TS). These
coincide and are correlated with
radioactive shales (use of the gamma
log) that are interpreted to have been
deposited across relatively flat
surfaces. Once the mfs and TS are
established and tied, then the sequence
boundaries (SB) of both carbonate and clastic sedimentary systems are identified. These will
tend to lie directly beneath the sand sized sediment fill of depressions on eroded and incised
surfaces and over the prograding clinoforms of high stand systems tracts (HST).

Well logs used to interpret shallow water clastic sequence stratigraphy


In both clastics and carbonates the second and often co-incident step in the interpretation of
well logs and cores is the use of parasequence stacking patterns (the vertical occurrence of
repeated cycles of coarsening or fining upwards sediment) of to identify the lowstand systems
tracts (LST), parasequence cyclic stacking patterns are commonly identified on the basis of
variations in grain size and when these fine upwards are indicated by triangles whose apex is
up while those that coarsen upwards are indicated by inverted triangles whose apex is down.
The example presented in the above diagram catches the essence of the response of shallow
water clastics responding to a fall and rise in base level and the incision and fill of the shelf
margin during this cycle.

The repeated stacking patterns for LST cycles are: -

 cyclic fill of incised depressions that tend to fine upward.


 cyclic sand to shale bodies of basin floor fans that tend to fine and thin
upward.
 cyclic sand to shale bodies of shelf margin clinoforms that tend to coarsen
and thicken upward.

The repeated stacking patterns for TST cycles are: -

 Regressive cyclic shale to sand bodies of that tend to coarsen and thin
upward.

You are advised to link to the exercises and movies on the clastic well logs of the Guarico
basin on this site to learn more about these techniques.
Well logs used to interpret shallow water carbonate sequence stratigraphy
The above examples are focused on near shore shelf clastics. The same approach
should be taken for carbonates, with the second co-incident step in the interpretation
of carbonate well logs and cores being to identify parasequence stacking patterns
(the vertical occurrence of repeated cycles of coarsening or fining upwards
sediment) of to identify the lowstand systems tracts (LST), transgressive systems
tracts (TST) and highstand systems tracts (HST) that are enveloped by the mfs, TS
and SB. As in clastics, carbonate parasequence and cycle stacking patterns are
commonly identified on the basis of variations in grain size and when these fine
upwards are indicated by triangles whose apex is up while those that coarsen
upwards are indicated by inverted triangles whose apex is down. For more
information on this topic you should click here to link to the carbonate sequence
stratrigraphy page of this site.

Value of sequence stratigraphy built from well logs


Interpretations that use well logs for sequence stratigraphy should be tied to
biostratigraphy so they can be used to correlate and analyze sedimentary rocks from
the perspective of geologic time. Well logs lend themselves to the detailed
reconstructions of paleogeography and the generation of high frequency
stratigraphic models that predict the distribution of sedimentary facies, particularly
those associated with aquifers, sediment bound ore bodies, and hydrocarbon
reservoirs, their source rocks and seals.
Mond
Well Log Interpretations > Intro Well Log Seq Strat > Well Log Suites

(source: Emery, 1996)


Resistivity logs
This log measures the bulk resistivity (the reciprocal of conductivity) of the formation.
Resistivity is defined as the degree to which a substance resists the flow of electric
current. Resistivity is a function of porosity and pore fluid in a rock. Porous rock
containing conductive fluid (such as saline water) will have low resistivity. A non-
porous rock or hydrocarbon-bearing formation has high resistivity. This log is very
useful for determining the type of fluids in formations and is frequently used as an
indicator of formation lithology.

Spontaneous potential (SP) logs


This log measures the electrical current that occurs naturally in boreholes as a result
of salinity differences between the formation water and the borehole mud filtrate
(formation and surface). These logs are used as indicators of permeable beds
(including determining permeable sands and impermeable shales) or for locating bed
boundaries. The SP log was one of the first tools to be used to distinguish shale from
sand in clastic sequences (zero matches pure shale while high SP values match
sand).

Gamma ray (GR) logs


This log records the radioactivity of a formation. Shales (or clay-minerals) commonly
have a relatively high gamma radioactive response, and consequently gamma ray
logs are taken as good measures for grain size (and subsequently inferred
depositional energy). Thus coarse-grain sand, which contains little mud, will have
low gamma ray value, while a fine mud will have a high gamma ray value. The
values range of gamma ray is measured in API (American Petroleum Institute) units
and range from very few units (in anhydrite) to over 200 API units in shales.
Gamma ray logs are one of the most commonly used logs for sequence stratigraphic
analysis. As explained above the variation in GR log character is interpreted to be a
proxy for grain size and so enabled these logs to play a major role in the sequence
stratigraphy analysis.The rationale behind this interpretation is that variations in the
GR are related to the presence of organic matter that sequesters radioactive
minerals responsible for the GR signal.In this case it is possible to:1) identify the
portions of the section which had higher accumulations of organics and 2) those for
which this has been reduced by winnowing and oxidizing effects! In turn, the
presence or absence of organic matter is assumed to be a proxy for the winnowing
and oxidizing effects of shallow marine waters. The result is that abrupt changes in
the GR log response are interpreted to be related to sharp lithological breaks
associated with unconformities and sequence boundaries. The principle shapes that
GR log signal makes frequently used for interpreting the depositional setting of
sedimentary cycles.

Neutron logs
This log measures the porosity of a formation, indicating in its response the quantity
of hydrogen present in the formation. The log is calibrated to limestone. The linear
limestone porosity units are calibrated using the API Neutron pit in 19% porosity,
water-filled limestone is defined as 1000 API units. This log is useful in measuring
lithology (usually in combination with density log).

Density logs
This log is a measure of the formation's bulk density and is mostly used as a porosity
measure. Different lithologies can also be determined using density log based on
returned density value. For example, pure quartz will have a bulk density (g/cm-3) up
to 2.65, coal 1.2-1.8, halite 2.05, limestone up to 2.75, dolomite up to 2.87, anhydrite
2.98.
Density is mostly commonly used in conjunction with neutron logs to determine
lithology of formation (density-neutron suites such as Schlumberger FDC-CNL suite).

Sonic (acoustic) Logs


This log measures of the speed of sound in the formation, and is related to both the
porosity and lithology of the rock being measured. Thus, if the lithology of a
formation is know, this log can be used to determine its porosity. shales have lower
velocity (higher transit time) than sandstone of same porosity, making this log a good
indicator of grain size.
sonic log values (in ms/ft) for some rock types are: sandstone 51-56, limestone 47.5,
dolomite 43.5, anhydrite 50, halite 67.

NB: whenever possible, one should use core data to confirm or augment well log analysis.
This because core provides direct access to the character of the rock penetrated by the well
whereas well logs are used to infer the character of the rocks penetrated. Care in establishing
and matching the depths of the core and well logs is essential in the correlation of well logs
and core.

Gamma Ray Response as a Proxy for Grain Size


Variation in GR log character is interpreted to be a proxy for grain size and so enable
these logs to play a major role in the sequence stratigraphy analysis. The rationale
behind this interpretation is that variations in the GR are related to the presence of
organic matter that sequesters radioactive minerals responsible for the GR signal. In
this case it is possible to: 1) identify the portions of the section which had higher
accumulations of organics and 2) those for which this has been reduced by
winnowing and oxidizing effects! In turn, the presence or absence of organic matter
is assumed to be a proxy for the winnowing and oxidizing effects of shallow marine
waters.

The result is that abrupt changes in the GR log response were interpreted to be
related to sharp lithological breaks associated with unconformities and sequence
boundaries (Krassay, 1998). The principle GR log shapes were frequently used for
interpreting the depositional setting of sedimentary cycles.

 Cleaning-up trend (funnel shape): a gradual upward decrease in gamma


response. In shallow marine settings, this trend reflects a change from shale-
rich into sand-rich lithology and upward increase in depositional energy with
shallowing-upward and coarsening. In deep marine settings, this trend reflects
an increase in the sand contents of turbidite bodies. This trend also may
indicate gradual change from clastic to carbonate deposition.
 Dirtying-up trend (bell shape): a gradual upward increase in gamma
response: This trend may reflect upward fining (eg: a lithology change from
sand to shale) or upward fining of sand beds in a thinly interbedded sand-
shale unit. This trend usually implies a decrease in depositional energy. In a
non-marine setting, fining upward is predominant within meandering or tidal
channel deposits with an upward decrease in fluid velocity within a channel
(coarser sediments at base of channel). In a shallow-marine setting, this trend
usually reflects an upward deepening and a decrease in depositional energy
(shoreline retreat). In deep-marine settings, this trend reflects waning of
submarine fans (reducing of sand contents).
 Boxcar trend (cylindrical or block shape): with low gamma and sharp
boundaries and no internal change: this trend is predominant in fluvial channel
sands, turbidites (typically with greater range of thickness), and aeolian
sands. Evaporites also can have a cylindrical gamma trend.

Additionally two in-between trends can be recognized:

 Bow trend (symmetrical or barrel): with gradual decrease then gradual


increase in gamma response: this is usually the result of progradation and
retrogration of clastic sediments.
 Irregular trend (lack of character): this trend represents aggradation of
shales or silts and can occur in other settings.

Curve character can be smooth, complex or serrated (sawlike) with contacts can be
sudden or gradual.

The figure below summarizes the log response a variety of different clastic
depositional systems that Malcolm Rider (1996) assembled in his excellent book on
Well Logs. His diagram has been added to and slightly been modified for this web
site.
Well Log Interpretations > Intro Well Log Seq Strat > Stacking Patterns

Well Logs: stacking patterns of clastic parasequences


After Van Wagoner et al., 1990
Progradational parasequence sets
As in figure below each successively younger parasequence is deposited farther
basinward when the rate of deposition is greater than the rate of accommodation
(regression).

Retrogradational parasequence sets


In figure below each successively younger parasequence is deposited farther
landward in backstepping pattern when the rate of deposition is less than the rate of
accommodation and a transgression of sea occurs.
Aggradational parasequence sets
Note that below each successively younger parasequence is deposited above one another with
no significant lateral shifts when the rate of accommodation approximates the rate of
deposition.

Progradational parasequence sets of the lowstand systems tract


In figure below stacking patterns and trajectories of progressively younger parasequences are
deposited farther basinward when the rate of deposition is greater than the rate of
accommodation (regression) during a fall in base level. Click on image below to access a
larger version of this diagram and view grain size distribution and clastic geometries within
deepwater fans.
Mid Slope Channels Toe of Slope fan Proximal Fan

Distal Fan Mid Fan

Retrogradational parasequence sets of the transgressive systems tract


stacking patterns and trajectories of progressively younger parasequences which are
deposited farther landward when the rate of deposition is less than the rate of accommodation
(regression) during a rise in base level. Click on image below to access a larger version of
this diagram and view grain size distribution and clastic geometries within deepwater fans.

Aggradational progradational parasequence sets of the highstand systems tract


stacking patterns and trajectories of progressively younger parasequences which are
deposited farther seaward when the rate of deposition exceeds the rate of accommodation
(progradation) during a still stand in base level. Click on image below to access a larger
version of this diagram and view grain size distribution and clastic geometries within
deepwater fans.
This page is the first step of a seismic stratigraphy interpretation. Its objective is to
define the genetic reflection packages by the surfaces that envelope seismic
sequences and systems tracts. These bounding discontinuities are identified on the
basis of reflection termination patterns and their continuity.
Boundaries are defined on a seismic line by identifying the termination of seismic
reflectors at the discontinuity surfaces. These terminations occur:

 Below a discontinuity and the definition of the upper sequence boundary.


Examples of this include:
o Toplap: termination of strata against an overlying surface, representing
the result of non-deposition and/or minor erosion.
o Truncation: this implies the deposition of strata and their subsequent
tilting and removal along an unconformity surface. This termination is
the most reliable top-discordant criterion of a sequence boundary.
Such truncation can also be caused by termination against erosional
surface, as for instance a channel.

 Above a discontinuity and the definition of the lower sequence boundary:


o Onlap: A base-discordant relationship in which initially horizontal strata
progressively terminate against an initially inclined surface, or in which
initially inclined strata terminate progressively updip against a surface
of greater initial inclination.
o Downlap: a relationship in which seismic reflections of inclined strata
terminate downdip against an inclined or horizontal surface. Examples
of downlap surfaces include a top basin floor fan surface, a top slope
fan surface, and a maximum flooding surface.

Note: If onlap cannot be distinguished from downlap because of subsequence


deformation, the term baselap is used.
Recommended procedures for performing seismic sequence analysis include:

 Identifying the unconformities in the area of interest. Unconformities are


recognized as surfaces onto which reflectors converge.
 Mark these terminations with arrows.
 Draw the unconformity surface between the onlapping and downlapping
reflections above; and the truncating and toplapping reflections below.
 Extend the unconformity surface over the complete section. If the boundary
becomes conformable, trace its position across the section by visually
correlating the reflections.
 Continue identifying the unconformities on all the remaining seismic sections
for the basin.
 Make sure the interpretation ties correctly among all the lines.
 Identify the type of unconformity:
o Sequence boundary: this is characterized by regional onlap above and
truncation below.
o Downlap surface: this is characterized by regional downlap.

Recommended color codes:

 Red: Reflection patterns and reflection terminations.


 Green: Downlap surfaces
 Blue: Transgressive surfaces
 Other colors: Sequence boundaries

If using only black and white:

 Thin solid lines: Reflection patterns


 Thicker solid lines: Sequence boundaries
 Dashed lines: Downlap surfaces
 Dotted lines: Transgressive surfaces

This hierarchy of surfaces now provides a framework to the reflectors. Be as objective as


possible when identifying the discontinuity surfaces of the section. Where possible base your
observations more on the geometric relationships of the reflectors than on an interpretation of
their origin.
Your next step in the workflow is to interpret the origin of this framework. A suggested
approach is to assume that the framework is result of the generation of repeated successions
of accommodation and sediment fill (the accommodation succession of Neal and Abreu,
2009). These seismic units of the accommodation successions vary in magnitude and duration
and can be interpreted to be the products of sedimentary packages that accumulated on a
depositional profile. Successions can be seen to consist of component partial succession sets
that sequentially prograde to aggrade, retrograde, and aggrade to prograde to degrade.
Neal and Abreu, 2009 propose that when interpreting seismic it is assumed that deposition
has responded to accommodation successions that range across time scales of 10–105 ka and
have stacking patterns that are products of vectors of accommodation rate (δA)/sedimentation
rate (δS). Thus:

1. δA/δS < 1 and increasing = Progradation to aggradation stacking (PA or lowstand).


2. δA/δS < 1 and decreasing = Aggradation to progradation to degradation stacking
(APD or highstand).
3. δA/δS > 1 = Retrogradation stacking (R or transgressive).

The interpretation of depositional sequence from a relatively conformable unit will be


dependent on data resolution and the lateral extent of the coverage. Stacking patterns will be
seen to be repeated across a range of durations and magnitudes accommodation succession.
The depositional settings of these accommodation successions can be related to systems
tracts, depositional sequences, sequence sets, composite sequences, composite sequence sets,
and megasequences and used to describe and interpret a basin’s depositional fill. Thus the
framework of the accommodation succession provides a simple, objective observation based,
predictive, independent of time or sea-level terminology, flexible for all scales of data and
subsequent data-resolution improvement, and provides a guide to incorporate the new
observations and make predictions of previously unrecognized complexity elsewhere.
From the perspective of interpretation each full succession should be considered to consist of
component partial succession sets that are sequentially, lowstand—progradation to
aggradational; transgressive—retrogradation; and highstand—aggradation to progradation to
degradation. To paraphrase Neal and Abreu, (2009) the “terms highstand and lowstand as
originally defined to label systems tracts relative to a shelf edge, and with an implied
relationship between sea level and systems tracts, have been the root of confusion. Like
propose that these terms be used in the strict sense of the original definition, because their
meaning has been lost when applied to the many depositional settings and high-resolution
data sets to which the concepts of sequence stratigraphy are now applied". As Neal and
Abreu, (2009) propose this concept of accommodation succession stacking should be used in
the interpretation of stratigraphic data within a hierarchal framework of depositional
sequences, sequence sets, and composite sequences. As they point out this approach allows
an interpreter to accurately categorize observations, provide a basis for predictions away from
control points, and develop a framework that allows revisions as higher-resolution data
become available.

References
Boggs, S. Jr., 2001, Principles of Sedimentology and stratigraphy, 3rd Ed., Prentice-
Hall, Inc., New Jersey, 726 p.
Neal, J., and Abreu, V., 2009, Sequence stratigraphy hierarchy and the accommodation
succession method, Geology, v. 37, p. 779-782
Vail, P. R., 1987, Seismic stratigraphy interpretation procedure, in Bally, A.W. (ed.),
Atlas of seismic stratigraphy: AAPG Studies in Geology No. 27, Vol. 1, p. 1-10.
Saturday

You might also like