You are on page 1of 64

3

Sedimentology

CONTENTS
SEDIMENTOLOGY
3.1
INTRODUCTION
3.1.1 General Introduction
3.1.2 Definitions
3.1.3 Objectives
3.2
CLASSIFICATION OF SEDIMENTARY
ROCKS
3.3
SEDIMENT TEXTURE
3.3.1 Introduction
3.3.2 Texture in Granular Sediments
3.3.2.1 Grain Size
3.3.2.2 Sorting
3.3.2.3 Grain Shape
3.3.2.4 Fabric
3.3.2.5 Grain Morphology and Surface Texture
3.3.2.6 Textural Maturity
3.4
SANDS AND SANDSTONES
3.5
POROSITY AND PERMEABILITY
3.5.1 Definitions
3.5.2 Porosity Types
3.5.3 Controls on Porosity and Permeability
3.5.3.1 Grain Size
3.5.3.2 Sorting
3.5.3.3 Grain Shape
3.5.3.4 Packing
3.5.3.5 Fabric
3.5.3.6 Grain Morphology and Surface Texture
3.5.3.7 Diagenesis(e.g.Compaction,Cementation)
3.6
TRANSPORT AND DEPOSITION OF
SEDIMENTS
3.6.1 Physical Processes of Transportation
3.6.2 Transport by Fluids
3.6.2.1 Bedforms and Sedimentary Structures
3.6.2.2 Current-Generated Bedforms and
Sedimentary Structures (water currents)
3.6.2.3 Wave-Generated Bedforms and
Sedimentary Structures
3.6.2.4 Wind-Generated Bedforms and
Sedimentary Structures
3.6.2.5 Sediment Gravity Flows
3.7
OTHER SEDIMENTARY STRUCTURES
3.7.1 Introduction
3.7.2 Further Discussion of Primary
Sedimentary Structures
3.7.3 Secondary Sedimentary Structures
3.7.3.1 Erosional sedimentary structures

3.8

3.9

3.10

3.11

3.7.3.2 Deformational Sedimentary Structures


3.7.3.3 Biogenic Sedimentary Structures - Trace
Fossils
FACIES AND FACIES SEQUENCES
3.8.1 Facies - Definitions
3.8.2 Facies Associations and Facies Sequences
3.8.3 Graphical Sedimentary Logs
3.8.4 Controls on the Nature and Distributions
of Facies
3.8.5 Genetic Units
SELECTED CLASTIC DEPOSITIONAL
ENVIRONMENTS
3.9.1 Aeolian Environments
3.9.2 Fluvial Environments
3.9.3 Coastal and Shallow Marine
Environments
3.9.4 Deep Marine Clastic Environments
CARBONATE SEDIMENTS
3.10.1 Introduction
3.10.2 Carbonate Grains
3.10.3 Minerals Present
3.10.3 Classification of Carbonate Rocks
3.10.4 Carbonate Depositional Environments
3.10.5 Porosity in Carbonate Rocks
DIAGENESIS
3.11.1 Definition
3.11.2 Clastic Diagenesis
3.11.3 Carbonate Diagenesis
3.11.4 Impact of Diagenesis on Porosity and
Permeability

Sedimentology

SEDIMENTOLOGY
3.1 Introduction
3.1.1 General Introduction
The great majority of hydrocarbon reserves worldwide occur in sedimentary rocks.
It is therefore vitally important to understand the nature and distribution of sediments
as potential hydrocarbon source rocks and reservoirs. Two main groups of sedimentary
rocks are of major importance as reservoirs, namely siltstones and sandstones
(clastic sediments) and limestones and dolomites (carbonates). Although carbonate
rocks form the main reservoirs in certain parts of the world (e.g. in the Middle East,
where a high proportion of the worlds giant oilfields are reservoired in carbonates),
clastic rocks form the most significant reservoirs throughout most of the world. This
chapter will therefore concentrate on the sedimentology of clastic sediments, with a
relatively brief discussion of carbonates.

3.1.2 Definitions
Sedimentology ...is concerned with the composition and genesis of sediments and
sedimentary rocks ......including the nature and composition of the constituent
particles
The relevance of sedimentology to the petroleum industry is summarised by the
following quotation;
....the reconstruction of depositional environments in clastic sequences provides the
optimum framework for describing and predicting reservoir development and reservoir quality distribution on both regional (-exploration) and field (-production)
scales.

3.1.3 Objectives
The objectives of this chapter are to discuss

the classification of sedimentary rocks


the important elements of texture in granular sediments and their influence on
porosity and permeability
the nature of sediment transport
sedimentary structures
facies analysis and genetic sedimentary units
the nature of reservoir sandbodies from the main clastic depositional environments
carbonate depositional environments
clastic and carboniferous diagenesis and their influence on reservoir quality

3.2 CLASSIFICATION OF SEDIMENTARY ROCKS


Sedimentary rocks are formed by physical, chemical and biological processes and can
be classified on the basis of the dominant process or processes responsible for their
formation. Although five classes can be identified (Table 1), we will be concerned

Department of Petroleum Engineering, Heriot-Watt University

Class
Clastic
(also known as
siliciclastic

Carbonate

Organic

Chemical

Volcaniclastic

Main lithologies

Dominant process(es)

conglomerate 1
breccia 1
sandstone 1
siltstone (1)
mudrock 2,3

Physical and chemical.


Weathering and erosion of
existing rocks, transport and
deposition of the weathering
products.
Biogenic and biochemical.
Formation by plants or
animals of carbonate (mainly
calcium carbonate)
skeletons, or organicallyinfluenced precipitation.
Diagenetic alteration to dolomite

1,2

limestone
dolomite 1

Biogenic.
Fixing of carbon or
phosphatic compounds by
plants or animals.
Accumulation of dead plant
or animal material.

phosphate
coal 2
oil shale 2,3

evaporites 3
ironstones

Chemical.
Mainly direct precipitation.
Volcanic, physical.
Eruption of volcanic
material, transport and
deposition by volcanic or
other processes.

ignimbrites
tuffs
volcaniclastic sandstones
etc.

3.3 SEDIMENT TEXTURE


3.3.1 Introduction
Texture is the general term used to describe the size, shape and arrangement of grains,
matrix and cement in a sedimentary rock. It is of importance to us because sedimentary
texture is the single most important control on reservoir properties (e.g. porosity and
permeability).
In this chapter, the term texture is used mainly to describe the grains and matrix. The
texture of a sediment reflects both the available sediment and its mode of transport and
deposition. The majority of clastic sediments contain laminae on a scale of mm to cm,
which will have subtly different textures. Variations in porosity and permeability
between laminae can exercise a strong influence on fluid flow, especially in the case
of two-phase flow (i.e. where two fluids, for example oil and water, are involved).

3.3.2 Texture in Granular Sediments


The main textural components of granular rocks include:

grain size

grain sorting

packing

sediment fabric

grain morphology

grain surface texture


4

Table 1
Classification of
sedimentary rocks. Rocks
marked 1 may form
reservoirs, those marked 2
may act as source rocks,
and those marked 3 may
form seals (Modified from
Tucker, 1981)

Sedimentology

3.3.2.1 Grain Size


Clastic sediments are defined on the basis of their mean grain size, as shown on Table
2. The majority of naturally-occurring sediments have an approximately log-normal
grain size distribution. Grain size can be measured either in millimetres, or in phi ()
units () = -log2(d), where d = grain diameter in millimetres). Because of the lognormality of sediments, the use of phi units allows normal statistical measures to be
calculated. Care should be taken, however, to avoid confusion with the symbol used
for porosity (see section 3.5.1).

a
mm

b.
m

mm

1/16

clay
1/256

~4

very fine

silt
1/16

62.5

1/8

sand
2

2000

-1

fine
1/4

1/2

granule
medium
4

Table 2
Definition of grain size of
granular sediments. a. clay
to boulder grade; b.
subdivisions of sands and
sandstones

-2
pebble

64

-6

coarse
1

cobble
256

-8

very coarse
2

-1

boulder

3.3.2.2 Sorting
The sorting of a sediment quantifies how well a depositional process has concentrated
(sorted) grains of a given size. It is generally measured as the standard deviation (SD)
of the grain size (in phi units). The sorting of a sediment is generally described
verbally, according to defined ranges of standard deviation (Table 3).

Table 3
Verbal description of
sorting

SD (in phi units)

Description

less than 0.35


0.35-0.5
0.5-0.71
0.71-1.0
1.0-2.0
greater than 2.0

very well sorted


well sorted
moderately well sorted
moderately sorted
poorly sorted
very poorly sorted

Department of Petroleum Engineering, Heriot-Watt University

The sorting of a sediment is most commonly estimated by comparison with images of


circles of known ranges of diameters (Figure 1).

Well Sorted = 0.35

Moderately Poorly Sorted = 1.0

Moderately Well Sorted = 0.5

Very Poorly Sorted = 2.00

Figure 1
Graphical illustration of
sorting (modified from
Pettigrew et al, 1973)

3.3.2.3 Grain Shape


Grain shape is described in terms of:

aspect ratio
grain sphericity - approximation to a sphere
grain roundness - curvature of the corners

Aspect ratio is the ratio of the diamter of the grain measured in different directions.
The three dimensional shape of the grain can be classified in terms of of the ratios of
their long, intermediate and short diameters (figure 2)

Figure 2
Grain shape and sphericity

Sedimentology

It should be noted that roundness is the extent to which the corners of a grain have been
smoothed off, not the approximation to a spherical shape; well-rounded grains can
have shapes which are far from spherical (see Figure 3.3). Grain shape depends both
on the mineralogy of the grains and the degree and energy of transportation (e.g. desert
and beach sediment is generally well sorted and rounded).

Low Sphericity

High Sphericity

0
Very Angular
1
Angular
2
Sub Angular
3
Sub Rounded
4

Figure 3
Grain roundness, shown for
grains of low and high
sphericity (modified from
Pettigrew et al, 1973)

Rounded
5
Well Rounded
6

3.3.2.4 Fabric
The term fabric, when applied to granular sediments, refers to the orientation and
packing of grains and the nature of their contacts.
Packing
Packing is the term used to describe the three-dimensional arrangement of grains in
a sediment. In naturally-occurring sediments, the grains are somewhat randomly
arranged, but their packing can be compared to idealised packing arrangements, such
as cubic close packing (in which the grains are arranged in a rectilinear grid) and
hexagonal or rhombohedral close packing (in which grains are arranged at angles of
60o and 120o). Of these two packing arrangements (Figure 4), the rhombohedral
packing is more efficient, leading to a lower porosity (see Section 3.5).

Department of Petroleum Engineering, Heriot-Watt University

(A)
Cubic packing
(48% porosity)

(B)
Rhombohedral packing
(26% porosity)

(C)
Grain supported
fabric

(E)
Preferred orientation
of grains

(F)
Point contacts

(G)
Concavo-convex
contacts

yy
;;
;;
yy
;;
yy
(D)
Matrix supported
fabric

(H)
Sutured contacts

Matrix and clast support


Many sediments contain, between their grains or clasts, a matrix of finer grained
material. In sands and sandstones, this matrix is likely to be of silt or clay grade,
whereas in pebbly or bouldery sediments and conglomerates, the matrix will be of
sand grade. In sediments with a high proportion of matrix, the larger clasts may not
be in contact with each other, in which case they are described as matrix-supported (Figure 4 D).
Orientation
Non-spherical grains may be deposited with a preferred orientation. Flat grains
commonly lie with their short axis sub-vertical and elongate grains may be arranged
with their long axis either parallel to or perpendicular to the palaeocurrent, depending
on the exact process of deposition. In some situations, flat clasts may be arranged so
that they dip in the upcurrent direction, a fabric known as imbrication (see Figure 4E).
Grain contacts
Immediately after deposition, most grains in a clast-supported sediment will have
point contacts with other grains. It should be noted that on 2D sections (e.g.
microscope thin sections of sandstones or outcrop sections of conglomerates) not all
grains will appear to be in contact; in these cases, the grains will probably be in 3D
contact in front of or behind the 2D section. During compaction of a sediment,
deformation and dissolution of grains will lead to the grain contacts becoming longer
and as compaction continues, concavo-convex and sutured contacts may result
(Figure 3.4).

3.3.2.5 Grain Morphology and Surface Texture


The morphology and surface texture of grains will reflect both the mineralogy and the
transport of the sediment. Grains derived from the weathering of crystalline (igneous
and metamorphic) rocks commonly consist of single crystals, and their shape will
reflect the mineralogy. During transport, the grains will undergo a certain amount of
8

Figure 4.
Grain fabric in sediments;
packing, grain contacts,
orientation of grains and
grain-matrix relationships
(modified from Tucker,
1981). Note that 3.4E
shows imbrication in
response to a current from
left to right

Sedimentology

rounding, which will be influenced both by the mineralogy and the energy and
duration of transportation. Grains which have undergone significant transport,
particularly in high-energy environments, will tend to have smooth surfaces, which
will have an influence on the flow of fluids through the pore system (see Section 3.5).

3.3.2.6 Textural Maturity


As has been described above, grains will tend to increase in roundness during
transport, and there will also be a tendency for the sorting of the sediment to improve.
Sediments consisting of well sorted, well rounded grains are described as texturally
mature. This should not be confused with mineralogical maturity, which is a measure
of the ratio of chemically and physically stable grains to unstable grains. The term
maturity is also used for the thermal maturity of hydrocarbon source rocks.

3.4 SANDS AND SANDSTONES


Sands are defined (Section 3.3.2.1) as sediments with a mean grain size between
0.0625 and 2 mm which, on compaction and cementation will become sandstones.
Sandstones form the bulk of clastic hydrocarbon reservoirs, as they commonly have
high porosities and permeabilities.
Sandstones are classified on the basis of their composition (mineralogical content)
and texture (matrix content). The most common grains in sandstones are quartz,
feldspar and fragments of older rocks. These rock fragments may include fragments
of igneous, metamorphic and older sedimentary rocks. It should be noted that the
quartz and feldspar grains will also be derived from older rocks, but as they consist
of a single crystal they are not considered to be rock fragments. Other grains present
may include micas (muscovite and biotite) and heavy minerals, which are often metal
ores or semi-precious minerals. Several classification schemes have been developed
which represent the grain composition of sands and sandstones on a triangular plot,
with quartz, feldspar and rock fragments as the three corners (front triangle of Figure
5). On these plots, sediments with a high quartz component will plot close to the quartz
corner, whilst those composed mainly of quartz and feldspar will plot close to the
quartz-feldspar line. The triangular plot is divided into fields defined by the proportion
of grains present, and sediments plotting within a certain field are given a name
reflecting their mineralogy. Thus, for example, quartz arenites plot close to the quartz
corner, whilst lithic arenites contain less quartz and have a higher proportion of rock
fragments than feldspar.

Department of Petroleum Engineering, Heriot-Watt University

QA
SAA
SLA
AA
LAA
LA
FLA
LSA

=
=
=
=
=
=
=
=

Quartz Arenite
Subarkosic Arenite
Sublithic Arenite
Arkosic Arenite
Lithic Arkosic Arenite
Lithic Arenite
Feldspathic Lithic Arenite
Lithic Subarkosic Arenite

QW
FW
LW

=
=
=

Quartz Wacke
Feldspathic Wacke
Lithic Wacke

Mudrocks

Wackes

75
Arenites

QW

QUARTZ

25

25

FW
LW

trix

15

50

ing
reas

ma
ent

s
rain
e(i. . G

< 30

m)

perc

Inc

FELDSP
AR
50

ROCK
F

RAGME

NTS

Quartz + Polyquartz + Chert


QA 5

QA
SAA SLA

SAA SLA 25
LSA

AA
LAA

LA

AA

LA
LAA

AA

Feldspar

Rock
Fragments

b (Pettigrew et al)

FLA

Feldspar
10

10

Unstable, or Labile,
Rock Fragment

c (Mc Bride)

The position on the triangular plot reflects the mineralogical maturity of a sediment.
As quartz is more stable at atmospheric temperature and pressure than feldspar and
rock fragments, continuing chemical weathering and physical transport will tend to
decrease the proportion of the unstable grains, leading to a quartz-rich, mature
sediment. The issue of maturity highlights a problem with several of the classification
systems (modified from Pettigrew et al, 1973), which include quartz-rich
multicrystalline clasts such as chert, polycrystalline quartz and metamorphosed
sandstone or siltstone as rock fragments. Sediments rich in these clast types would
therefore be termed lithic arenites, a name which implies mineralogical immaturity.
However, these quartz-rich clasts will be almost as stable as monocrystalline quartz,
so the sediment itself is mineralogically mature. This fact is recognised by the
classification scheme of McBride (1963), which includes chert, polycrystalline quartz
etc. with quartz, and plots only the unstable (labile) rock fragments at the rock
fragments corner (Figure 5b).

10

Figure 5
Classification of sands and
sandstones according to
grain and matrix
composition (modified from
Pettigrew et al, 1973)
Triangles b and c show
alternative classification
schemes for the matrix free sandstones
("arenites")

Sedimentology

The classification described so far only takes account of the grains, but we know that
many sandstones contain a finer-grained matrix. This can be taken into account if the
triangular plot is extended into a triangular prism, with the long axis representing the
proportion of matrix (Figure 5). Sands and sandstones with less than 15% matrix are
called arenites (front triangle on Figure 5) and those with more than 15% matrix are
wackes (beyond second triangle on Figure 5). Sediments with over 75% muddy matrix
(i.e. less than 25% grains) are known as mudstones or mudrocks. The subdivision of
the triangles becomes simpler with increasing proportion of matrix.

3.5 POROSITY AND PERMEABILITY


3.5.1 Definitions
Total porosity () is defined as the volume of void (pore) space within a rock,
expressed as a fraction or percentage of the total rock volume. It is a measure of a
rocks fluid storage capacity.
The effective porosity of a rock is defined as the ratio of the interconnected pore
volume to the bulk volume
Microporosity (m) consists of pores less than 0.5 microns in size, whereas pores
greater than 0.5 microns form macroporosity (M)
The permeability of a rock is a measure of its capacity to transmit a fluid under a
potential gradient (pressure drop). The unit of permeability is the Darcy, which is
defined by Darcys Law (see Figure 6). The millidarcy (1/1000th Darcy) is generally
used in core analysis.
P1 <

> P2

<

Figure 6
Diagram illustrating
Darcys Law

>

Q = K.P.A
.L
Q
P
A

L
K

=
=
=
=
=
=

Rate of flow (cc / sec)


Pressure differential (atmospheres)
Area (cm2)
Fluid viscosity (centipoise)
Length (cm)
Permeability (Darcies)

3.5.2 Porosity Types


Primary porosity consists of pore space that results from primary depositional texture
(e.g. spaces between grains, or within fossils).
Secondary porosity is pore space generated by post-depositional processes (e.g.
dissolution of grains or cement, fracturing etc.)
Department of Petroleum Engineering, Heriot-Watt University

11

3.5.3 Controls on Porosity and Permeability


The porosity and permeability of the sedimentary rock depend on both the original
texture of a sediment and its diagenetic history. In many cases, despite a complex
diagenetic history, clastic sediments still retain a strong fingerprint from their original
facies-controlled texture (Figure 7). The main controls on porosity and permeability
are outlined in the following sections. Sections 3.5.3.1 to 3.5.3.6 are concerned with
textures, which control the depositional porosity and permeability (i.e. the porosity
and permeability immediately after deposition of the sediment) but will also influence
the final primary porosity and permeability (Figure 7). Section 3.5.3.7 discusses the
effects of diagenesis on final porosity and permeability. Diagenesis will effect both
the primary and secondary porosity.

1000

Aeolian

Permeability (mD)

100

10

Fluvial

Flash Flood/ Alluvial Fan

0.1
5

10

15

20

Porosity (%)

3.5.3.1 Grain Size


In theory, porosity is independent of grain size, as it is merely a measure of the
proportion of pore space in the rock, not the size of the pores. In practice, however,
porosity tends to increase with decreasing grain size for two reasons. Finer grains,
especially clays, tend to have less regular shapes than coarser grains, and so are often
less efficiently packed. Also, fine sediments are commonly better sorted than coarser
sediments. Both of these factors result in higher porosities. For example, clays can
have primary porosities of 50%-85% and fine sand can have 48% porosity whereas the
primary porosity of coarse sand rarely exceeds 40%.
Permeability decreases with decreasing grain size because the size of pores and pore
throats will also be smaller, leading to increased grain surface drag effects (Figure 8).

3.5.3.2 Sorting
For a given grain shape, porosity and permeability decrease with decreasing sorting
(Figure 8). This is due to the fact that, in poorly sorted sediments, smaller grains can
accommodate themselves between the larger ones, leading to a reduction both in the
percentage of pore space and the size of pores

12

Figure 7
Porosity and permeability
as a function of
depositional environment
within a fluvio-aeolian
system

Sedimentology

Figure 8
Depositional porosity as a
function of grain size and
sorting (after Beard and
Weyl)

3.5.3.3 Grain Shape


The more unequidimensional the grain shape, the greater the porosity (see also section
3.5.3.1). As permeability is a vector, rather than scalar property, grain shape will affect
the anisotropy of the permeability. The more unequidimensional the grains, the more
anisotropic the permeability tensor.

3.5.3.4 Packing
The closer the packing, the lower the porosity and permeability.

3.5.3.5 Fabric
Rock fabric will have the greatest influence on porosity and permeability when the
grains are non spherical (i.e. are either disc-like or rod-like). In these cases, the
porosity and permeability of the sediment will decrease with increased alignment of
the grains.

3.5.3.6 Grain Morphology and Surface Texture


The smoother the grain surface, the higher the permeability

3.5.3.7 Diagenesis (e.g. Compaction, Cementation)


Diagenesis (see Section 3.11.2) is the totality of physical and chemical processes
which occur after deposition of a sediment and during burial and which turn the
sediment into a sedimentary rock. The majority of these processes, including
compaction, cementation and the precipitation of authigenic clays, tend to reduce
porosity and permeability, but others, such as grain or cement dissolution, may
increase porosity and permeability. In general, porosity reduces exponentially with
burial depth (Figure 9a), but burial duration also an important criterion. Sediments that
have spent a long time at great depths will tend to have lower porosities and
permeabilities than those which have been rapidly buried (Figure 9b).
Department of Petroleum Engineering, Heriot-Watt University

13

B
Porosity (percent)
0
0

Porosity (percent)

40

30

Young Sands

Maximum Depth of Burial (m)

Burial Depth

Old Sands
Mesozoic

Tertiary

Palaeozoic

6000

Tertiary
Mesozoic
Palaeozoic

3.6 TRANSPORT AND DEPOSITION OF SEDIMENTS


3.6.1 Physical Processes of Transportation
Sediments are generally transported by one of three basic processes
by mass movement. This is mainly gravity-driven, and so the processes are
commonly referred to as sediment gravity flows
by fluids (water and air)
by glaciers
In the following sections, we will be dealing mainly with movement by fluids and
sediment gravity flows. These discussions will concentrate mainly on clastic sediments,
but it should be noted that many carbonate sediments consist of grains which behave
in similar ways to clastic grains. For example, in many parts of the tropics, beaches
and shallow marine environments, which in temperate zones are dominated by quartz
sand, contain mainly carbonate sands.

3.6.2 Transport by Fluids


The main types of fluid motion by which grains are transported are currents (of water
or air) and waves. In both of these, sediment is transported either in suspension
(through the effects of turbulence) or as bed load, by rolling or bouncing (saltating).

3.6.2.1 Bedforms and Sedimentary Structures


During transport by currents or waves, granular sediments generally form a range of
distinctive shapes, or bedforms, on the sediment surface. Continued sediment transport,
bedform migration and sediment deposition will lead to the development of sedimentary
14

Figure 9
Changes of porosity with
burial depth and burial
duration (modified from
North, 1985)
A. Exponential loss of
porosity with burial depth
for "typical" sandstones
B. loss of porosity with
maximum burial depth
(may not be the same as
present depth) for
sandstones of different ages

Sedimentology

structures within the sediment pile. It is important to differentiate between bedforms


and sedimentary structures; bedforms are the features on the sediment surface, whilst
sedimentary structures are the features within the sediment which are commonly
preserved in the rock record.
Structures produced by fluid flow are known as primary sedimentary structures, but
structures may also be formed by organisms (e.g. by burrowing) or by deformation.
These are known as secondary sedimentary structures. The transport of sediment and
the formation of bedforms are discussed in the remainder of Section 3.6, primary and
secondary sedimentary structures are discussed in section 3.7 and the importance of
sedimentary structures in environmental interpretation is described in section 3.8.

3.6.2.2Current-GeneratedBedformsandSedimentaryStructures(WaterCurrents)
Bedforms formed by unidirectional water currents have been extensively studied,
both in the laboratory and in nature, and the relationship between sediment grain size,
current velocity (measured in a number of different ways) and the bedforms produced
has been established. This has led to the development of bedform stability diagrams,
which show the bedforms which occur under different conditions (Figure 10).

Upper Regime
(Rapid flow)

40
Antidunes

Transition

10

Plane Bed
B

Stream Power, v

Figure 10
Schematic representation of
various bedforms and their
relationship to grain size
and stream power. Based
on Simons et al. 1965 and
Allen 1968a. Plan views A
and B show the change in
shape of ripples (A) and
megaripples (B) as stream
power increases
(palaeocurrent on these
plan views is from bottom to
top) 1) straight-crested 2)
undulatory 3) lingoid 4)
lunate. More recent flume
experiments show that the
megaripple field pinches
out at 0.1mm grain size

Lower Regime
(Tranquil flow)

0.1
Megaripples

Plane Bed

0.01

0.625 1.25 0.2.


0

very fine
fine sand
sand

Ripples

04

0.6

0.8

1.0 mm

coarse
sand

medium
sand
Median Fall Diameter

Department of Petroleum Engineering, Heriot-Watt University

15

For medium sand (0.25-0.5mm), as the current velocity or stream power increases, the
first bedforms to form are current ripples. As the stream power increased, larger scale
structures, known as megaripples or dunes, form and are replaced at even higher
stream powers by a flat bed, (upper stage plane bed). For both small ripples and
megaripples, ripple crests tend to become more curved and discontinuous (three
dimensional) with increasing stream power.
Both ripples and megaripples have a distinctive form in cross-section (Figures 11 and
12). They have a relatively low slope on their up-current stoss side and are steeper
on their downcurrent lee side. As a current passes over the ripple, it detaches from
the sediment surface near the crest and forms a separation eddy downstream of the
ripple (Figure 12A). Idealised path lines of sediment grains are shown. (Modified
after Jopling, 1967) In the case of ripples, grains roll or saltate up the stoss side and
periodically avalanche down the lee side. The dip of the lee side is thus controlled by
the angle of repose, the maximum slope at which grains of a given grain size and
sorting can rest without slope failure. In the zone of back-flow, some sediment is
caught in the backflow eddy and is deposited at the toe of the lee slope. As the ripple
migrates, successive positions of the lee side are marked by inclined foresets, which
can be seen within the body of the bedform (figure 11). These foresets are either planar
or concave-upwards.

yyyyyyy
;;;;;;;
;;;;;;;
yyyyyyy
yyyy
;;;;
Sloss Side Laminae

Fore Set Laminae

Bottom Set Laminae

Zone of
No Diffusion
Zone of
Mixing

Line of Zero Velocity


Zone of Backflow

16

Figure 11
Profile and internal
structure of a welldeveloped ripple. The
geometry of a megaripple/
dune will be essentially
similar

Figure 12
Flow pattern and
sedimentation processes
over a ripple. A. Velocity
distribution and flow
separation on the lee side of
the ripple (modified after
Jopling 1963, 1967) B.
Flow pattern and
sedimentation processes
(modified after Jopling,
1967)

Sedimentology

As a ripple train migrates downcurrent each ripple trough will erode the next ripple
downcurrent. For net deposition to occur, the ripple troughs must climb relative to the
sediment surface in a downcurrent direction. In this case, sets of cross lamination
bounded by erosive surfaces result, (Figure 13). It should be noted that the term cross
lamination applies to structures generated by ripples, and so sets are less than 4cm
thick (generally 1-3cm).

Figure 13
Experimentally produced
climbing-ripple crosslamination seen in vertical
profile parallel with flow.
The increasing angle of
climb from top to bottom is
caused by the increasing
rate of net vertical
deposition relative to the
speed of advance of the
ripples (after J.R.L. Allen
1972)

As shown on Figure 10 and discussed above, as current energy or stream power


increases, ripples are replaced by larger scale bedforms. The terminology of these
bedforms is very confused, with different researchers referring to them as megaripples,
dunes, large scale ripples or sandwaves (Aslley, 1991). Although dune is probably the
most commonly used term, the term megaripple is preferred here, because of the
Department of Petroleum Engineering, Heriot-Watt University

17

desert connotation of the term dune (see Section 3.6.2.4). As described above,
megaripples have a similar general form to ripples, but they are a distinct bedform
type. If the size of naturally occurring ripple and megaripple bedforms are plotted on
histograms, they are found to form two distinct size populations. In addition,
megaripples tend to wave lower height/wavelength ratios. Their wavelengths range
from 0.6m to hundreds of metres and their heights from 0.05m to ~10.00m, but they
are most commonly 0.1-1m or 2m high, with wavelengths of 1m to 20m. Experimental
and field observations show correlation between bedform height and the depth of flow
and superimposed hierarchies of ripples and megaripples or small and large megaripples
may co-exist.
As discussed above, both straight-crested and curved-crested megaripples occur
(Figure 14). The straight crested megaripples generally occur at lower current
velocities than the curved-crested megaripples. This is because, as current velocity
increases, the strength and localisation of separation eddies in the lee side of the
bedforms becomes greater, leading to increased and localised erosion of the trough.
This leads to localised embayments on the crestline, and eventually to the occurrence
of discrete concave-downcurrent (lunate) bedforms (Figure 14b). Because their
cross-sectional shape perpendicular to the crest does not vary much along the
crestline, straight-crested megaripples are sometimes referred to as 2-dimensional,
whereas lunate megaripples are referred to as 3-dimensional.

Figure 14
Block diagrams showing (a)
straight-crested (2dimensional) and (b)
lunate (3-dimensional)
megaripples and the
sedimentary structures they
produce ((a) tabular cross
bedding and (b) trough
cross bedding)

As the bedforms migrate and climb, straight-crested (2D) bedforms produce tabular
cross bedding. In sections parallel to the palaeocurrent, this consists of near-parallel
set boundaries separating inclined foresets. These foresets may be either planar or
curved. The curved foresets are concave-upwards, and are sometimes referred to as
tangential or asymptotic. On sections perpendicular to the palaeocurrent, the foresets
appear to be almost parallel to the set boundaries (Figure 14a). Lunate, 3D, megaripples
produce trough cross bedding. On sections parallel to the palaeocurrent, trough cross
bedding looks similar to tabular cross bedding, although the set boundaries are less
parallel and the sets tend to be slightly shorter. Foresets are always curved. On sections
perpendicular to the palaeocurrent, the set boundaries are strongly concave-upwards
and the foresets are almost parallel to the boundaries (Figure 14b).

18

Sedimentology

Cross-bed sets are typically decimetres thick. Trough sets are commonly 10-50cm
thick, 1-2m wide perpendicular to flow and 5-10m long (parallel to flow). Tabular sets
are generally more laterally extensive for a given set thickness than trough sets.
As the current velocity or stream power is increased, megaripple bedforms become
lower and flatter and are eventually replaced by a plane bed on which there is intense
sediment transport, with most of the grains are moving most of the time. This feature
is known as upper-stage plane bedding or upper-phase plane bedding. The rapid flow
over the bed produces vortices with their axes parallel to the flow, and these act to align
the sand grains and form subtle ridges parallel to the flow. Sandstones containing
upper-phase plane bedding split readily parallel to bedding and bedding planes exhibit
subtle linear features, which reflects the grain alignment and which are parallel to the
palaeoflow direction. This structure is known as primary current lineation.
As the current velocity is increased still further, standing waves develop on the water
surface. With increasing current velocity, these may migrate a short distance upstream
before breaking. These standing waves and antidunes are mimicked on the
sediment surface by similar, in phase features with a more subdued relief. Because of
the very rapid movement of grains over such bedforms and their limited stability field,
antidune bedding is very rarely preserved, so it will not be discussed further here.

3.6.2.3 Wave Generated Bedforms and Sedimentary Structures


Waves commonly form in standing bodies of water, in response to wind shear over the
water surface. Individual water particles have an orbital motion, with the net effect of
these motions being to produce waves on the water surface which migrate in the wind
direction. The orbital radius decreases with depth, from a maximum at the water
surface (Figure 15). The depth at which the radius reaches zero is known as the wave
base; below this, the waves will have no effect.
If the sediment surface is above the wave base, the waves will impinge on the
sediment, which modifies the behaviour of both the waves and the sediment. The
orbital motion becomes elliptical and as the waves shoal (i.e. enter shallower water)
they become steeper, migrate more rapidly onshore and eventually break onto the
shore.

Department of Petroleum Engineering, Heriot-Watt University

19

Wavelength (L)

Direction of Wave Movement

Calm Water Level

D=L
2

Effective Wave Base (D)

Direction of Wave Movement

Figure 15
Diagrams to show the
orbital motion of open
waves (a), and the
ellipsoidal motion of
shoaling waves (b)

Sea Floor
B

The waves impinging on the sediment surface can produce wave-ripples to depths as
great as 200m. Wave-ripples are generally straight crested, and may be symmetrical
or asymmetrical in section. They vary greatly in size, with their size being dependent
on wave dimensions. Ripple wavelengths () are between 0.0009m and 2m and have
heights (H) between 0.003m and 0.25m. Wave ripples can be distinguished from
current ripples by lower ripple indices (/H) and crestal bifurcation.
The sedimentary structure produced by wave ripples is wave-ripple cross lamination.
It has a number of distinctive features (Figure 16) which can be used to differentiate
it from current ripple cross lamination.

Oscillation Ripples

Wave-Formed Current Ripples


?????-Laminae, Sometimes oppsed
Chevron Laminae
Draping

Planar laminations formed at high


applied bed shear stresses

Irregular, Undulattory
Set Boundaries

As waves shoal, and the shear stress on the sediment surface becomes greater, wave
ripples are replaced by planar beds (Figures 16 and 17)
20

Figure 16
Diagram showing some of
the distinctive features of
wave-ripple cross
lamination (from De Raaf
et al, 1997,)

Sedimentology

Plane Bed

100

Velocity (cm/s)

80
Wave Ripples
60
40
20
No Grain Movement
0

Figure 17
Bedform stability diagram
for wave-generated
structures (modified from
Allen, 1985)s

0.5

0.125 0.25

0.2
Very Fine
Fine

0.4

1.0

0.6
0.8
Grain Size (mm)

Medium Sand

Coarse Sand

1.0

1.2

Very Coarse

Sand

An additional sedimentary structure, which is found only in the rock record, is felt to
be generated by waves or combined waves and currents. Hummocky cross stratification
(HCS) was first described by Harms (1975). On bedding planes, it can be seen to
consist of low-relief mound-like hummocks separated by troughs (Figure 18). In
section, HCS sets are typically 10-15cm thick and include both concave-up and
convex-up laminae (in contrast with other forms of cross bedding, in which upwardconvex laminae are rare to absent). Set bases are erosive and produce low-angle
truncations.

Figure 18
Block diagram of
hummocky cross
stratification

Sharp

Based

Direc

Bed

tiona

l Sole

Mark

Sets up to 25cm Wave length 1-5m Height up to 40 cm

In ancient successions, hummocky cross stratification generally occurs in shallow


marine successions. When it occurs in thin, discrete sandstone beds, the beds
generally have sharp bases and sharp or gradational tops. Because of its inferred
shallow marine origin and common association with wave-generated structures,
hummocky cross stratification is interpreted as the product of complex waning
oscillatory currents related to storm activity.

Department of Petroleum Engineering, Heriot-Watt University

21

Thin hummocky cross-stratified beds often amalgamate to form thicker beds, for
example in middle shoreface environments. In this case, there is often pronounced
erosion between the sets, leading to erosion of many of the upward-convex laminae
in the upper parts of sets and therefore an increase in the proportion of upward-concave
laminae. This slightly modified type of hummocky cross stratification is sometimes
referred to as swaley cross stratification (scs).

3.6.2.4 Wind-Generated Bedforms and Sedimentary Structures

Grain Size of
20th percentile (mm)

In addition to the water-generated bedforms and sedimentary structures described in


Section 3.6.2.2, currents of wind are also capable of transporting sediment. Windgenerated bedforms include wind-ripples, dunes and compound dune-like bedforms
called draas. Their relative sizes are shown on Figure 19.

2.0
Ripple Field

Dune Field

1.0
0.7
0.5
0.3

Draa Field

0.2
(m) 0.01
(cm) 1

0.04
4

0.16

0.64

2.56

16

64

256

10

40

160

640

2560

(m)

Bedform Wavelength

Figure 19
Plot of sediment grainsize v.
wavelength for windgenerated bedforms
Modified from Wilson,
1972)

Aeolian ripples have wavelengths of 0.01m to 20.0m and heights of mms to


approximately 1m. Grains move mainly by saltation and the wavelength of these
impact ripples is controlled by the mean saltation jump length (Figure 20). Internal
lamination is generally poorly defined.

Descending Grains

(A)

"Shadow Zone"

"Shadow Zone"

Wind Direction

(B)
Figure 20
Saltation of sand grains
over wind ripples (based on
Bagnold, 1954)

22

Sedimentology

Aeolian dunes have diverse morphologies and are differentiated mainly by their
structure (Figure 21). Draa, or complex/compound dunes, are larger-scale topographic features with superimposed dune-scale bedforms. If the superimposed
bedforms are of the same type (but different scale) as the larger bedform is the latter
described as a compound. if the superimposed bedforms are of a different type, the
large bedform is complex.
Simple dunes include straight-crested transverse dunes and strongly lunate barchan
dunes (Figure 21). Star-shaped or stellate dunes have several arcuate slipfaces,
arranged in different directions and longitudinal seif dunes are elongated parallel to
the mean wind direction and may have slipfaces on both sides (Figure 21).

A. Lunate Dune or Barchan

Wind Direction

B. Transverse Dune

C. Stellate Dune

Figure 21
Aeolian dune and draa
morphology

Wind Direction

C. Seif or Longitudinal Dune

Aeolian dunes, like subaqueous dunes/megaripples, produce cross bedding. The


foresets include laminae produced by a number of different processes. On steeper
parts of the slipface, periodic avalanching of sand produces lobate grainflows
laminae, whilst fall-out of finer sediment from suspension produces finer grained
grainfall laminae. Wind ripples may also occur locally on the slipface, particularly
near its base and at the lateral fringes. Because the wind direction can change relatively
frequently during the migration of a dune, interval erosive surfaces known as
reactivation surfaces are common. Complex aeolian cross bedding and the hierarchy
of bounding surfaces are discussed in more detail in Section 3.9.1.

Department of Petroleum Engineering, Heriot-Watt University

23

In many deserts, the low-relief areas between dunes are, at least occasionally, close
to the water table and may therefore be damp. If sand is blown onto a damp surface
the grains tend to stick to the surface, producing a range of adhesion structures such
as adhesion warts and adhesion ripples

3.6.2.5 Sediment Gravity Flows


All the bedforms and sedimentary structures described above are produced when
sediment is entrained, transported and deposited by moving air or water. There is
another important group of sedimentary processes in which sediment is transported by
gravity acting directly on the sediment or sediment/water mix. These sediment gravity
flows can be divided into four main types:

debris flows
grain flows
fluidised/liquefied flows
turbidity flows

Because they are driven by gravity, these flows all transport sediment down slopes.
They differ in the process by which the shear strength of the sediment is reduced in
order for it to move.

Rock Fall

Olistholiths

Olistholith
Sliding

Slumping

Slide

Shear Planes

Slump

Shear Planes

Mass Flow,
e.g.
Debris Flow

Mass Flow,
or Debris Flow
Deposit

Suspension Mainly Due to


Fluid Turbulence

Turbitity
Current

24

Turbidite

Figure 22
Sediment gravity flow
processes and deposits
(modified from Rupke,
1978 after Kruit et al, 1975)

Sedimentology

Sediment on a slope will commonly fail by slumping (Figure 23). Failure occurs along
a curved plane and the sediment above this plane deforms as it moves down-slope. In
many cases, slumps will move only a short distance down-slope before stopping.
However, if the slope is sufficiently steep or the sediment sufficiently mobile, the
slump may continue to move down the slope, developing into a debris flow. The most
commonly observed debris flows are mudslides which, as the name suggests, consist
of assorted debris in a muddy matrix. However, more sandy debris flows also occur,
especially in sub-aqueous environments. The larger clasts in debris flows are
supported by the strength of the matrix and by their buoyancy. Debris flow deposits
are generally chaotic, although there may be a slight tendency for the largest clasts to
occur towards the top of the deposit.

Sediment Gravity Flows


Turbidity
Current

Fluidized
Sediment Flow

;yy;

Grain
Flow

Debris
Flow

Sediment Support Mechanism

Turbulence

Figure 23
Sediment gravity flow
processes (modified after
Rupke, 1978)

Distal
Turbidite

Proximal
Turbidite

Upward
Intergranular
Fluid Flow

Resedimented
Conglomerate

Grain
Interaction

Matrix
Strength

Pebbly
Some
"Fluxoturbidites" Mudstones

Deposit

In grain flows, the grains are kept in the flow, and prevented from being deposited, by
the exchange of kinetic energy between grains, with the grains effectively bouncing
off each other. Such flows can move down relatively steep slopes (>18o) and are
generally only a few cms thick. Their deposits are structureless, with sharp bases, and
commonly sharp tops and reverse grading may occur.

Department of Petroleum Engineering, Heriot-Watt University

25

Turbidity Current

Grain Flow

Rippled or flat top


Cross lamination,
Climbing ripples

Flat top
No grading ?
Massive,
grain orientation parallel
to flow

Laminated
Good grading
("Distribution grading)

Reverse grading near base?


Scours, injection structures

Flutes, tool marks on base

Fluidized / Liquified Flow

Debris Flow
Irregular top
(Large grains projecting)

Sand volcanoes or flat top


convulate lamination
fluid escape "Pipes"

Massive,
poor sorting random fabric

Dish structure?

Poor grading, if any ("Coarse tail")

Poor grading
("Coarse tail grading")

Basal zone of "Shearing" broad "Scours"


? Striations at base

? Grooves, flame and load


striations structures on base

The least well known of the four flow types discussed here are fluidised flows. They
occur most commonly when loosely-packed silt or sand deposits collapse. The grain
framework is no longer supportive, with the grains being held partly in suspension by
the escaping fluid. The minimal sediment strength allows fluidised flows to flow
rapidly down slopes as low as 2o or 3o. Deposition occurs by gradual freezing from the
bottom up, with little grain segregation. This leads to deposits with sharp bases and
tops, poor grading, local diffuse lamination and common fluid escape structures
(Figure 24; see Section 3.9.4).
The final gravity flows to be considered here are turbidity currents. As their name
suggests, they are composed of a mixture of sediment and water in which the sediment
is kept in suspension by the turbulence of the flow. The study of turbidity currents and
their deposits, turbidites, began in the early 1950s. The first turbidity currents to be
examined consisted of high-density suspensions of mud and sand. The flows typically
consist of a pronounced, highly turbulent head followed by a thinner body (Figure 22).
The turbulence of the head commonly causes erosion of the underlying sediment, and
in the more proximal, energetic parts of the flow, this may act to entrain more sediment
into the flow. As the flow progresses, mixing of the sediment-laden flow with the
ambient water, in both the head and tail, leads to dilution of the flow.
As the flow loses energy, either by dilution or by a decrease in slope, the sediment will
begin to be deposited. The coarser grains will tend to be deposited first, leading to the
common occurrence of graded bedding (Figure 24). The rapid deposition and lack of
traction leads to this interval generally being structureless. As deposition continues,
traction at the interface between the sediment and active flow may form upper phase
plane bedding, overlain by ripple cross lamination, climbing ripple lamination or
wavy lamination. In turbidites, this interval commonly exhibits deformation of the
structure, causing convolute lamination. This interval is commonly overlain by
26

Figure 24
Sediment gravity flow
deposits (from Rupke, 1978
after middleton and
Kempton, 1976)

Sedimentology

parallel lamination of uncertain origin. This sequence of structure within a single bed
forms the basis for the Bouma classification of classical turbidites, in which the
intervals are given the letters A to D (Figure 60). The fine material deposited between
turbidity currents is assigned to Bouma E. It should be noted that the full Bouma
sequence is rarely seen, with most turbidites only showing a subset of the subdivisions
(e.g. AB, ABC or BCD).
More recently, the occurrence of mud-poor turbidites has been recognised and these
are attributed to high-density turbidity currents in which the sediment load is
dominated by sand and silt, with little mud (Figure 61). There has been much
discussion over the origin of these beds, but it is likely that the final transport
mechanisms may have included grain flow and fluidised flow processes. Coarsegrained turbidites, the deposits of high-density turbidity currents, generally have a
structureless base, overlain by a faintly laminated interval, and with abundant
evidence of water escape in the upper part of the bed (Figure 61).

3.7 OTHER SEDIMENTARY STRUCTURES


3.7.1 Introduction
The sedimentary structures discussed in section 3.6, which are produced by the
physical processes of transport and deposition of the sediment, are known as primary
sedimentary structures. However, a number of processes, including water escape,
deformation of the sediment and disruption by organisms living in or on the sediment,
may occur after deposition. The structures produced are known as secondary sedimentary structures. This section includes a brief discussion of some primary sedimentary structures which were not discussed in Section 3.6, followed by a description of
some of the more important secondary structures.

3.7.2 Further Discussion of Primary Sedimentary Structures


As discussed above, primary sedimentary structures are produced by the movement
of sediment in response to fluid flow and reflect the nature of the bedforms which
produced them. The structures formed are related directly to the grain size of the
sediment and the fluid behaviour (e.g. current strength or wave energy. In addition to
telling us about the process of deposition, sedimentary structures may also tell us the
palaeocurrent or palaeowind direction or the direction of wave approach.
Many of the structures mentioned above have laminae on the scale of mms or cms.
The laminae we can see are the product of changes in texture, such as grain
composition or the size, shape, orientation, packing or sorting of grains. However,
larger scale bedding may also occur (e.g. individual turbidite beds or flood-generated
beds on a fluvial floodplain. In the description of sediments, it is important to have a
uniform description of bed thicknesses. Such a definition is given on Table 3. In
addition to bed thickness, it is important to note, when describing bedded sediments,
whether the base and top of the bed are sharp or gradational and, if the base is sharp,
whether it is erosional.

Department of Petroleum Engineering, Heriot-Watt University

27

cm

Description of bed thickness


very thick

Splitting of beds

mm

massive

100

Description of lamina thickness


very thick

30
thick

blocky

medium

slabby

30

thick
10

10

medium
1

thin

flaggy

thin
?0.5

very thin

laminated

very thin

3.7.3 Secondary Sedimentary Structures


There are three main types of secondary sedimentary structures, erosional, deformational
and biogenic.

3.7.3.1 Erosional Sedimentary structures


As discussed above, the currents and waves which transport sediment may often erode
the underlying sediment prior to deposition of their own sediment load. Such erosion
may occur over a wide area, leading to a planar erosive surface, or may be localised,
producing discrete erosive features such as grooves, flute marks and impact marks.
These erosive features are generally filled with sediment as the eroding current loses
power, so that they are generally preserved as positive features on the base of beds. In
addition to giving some indication of the power of the current, many of the features
(e.g. flutes and grooves) have distinctive shapes which will indicate the palaeocurrent
direction.

3.7.3.2 Deformational Sedimentary Structures


Recently-deposited sediment is often poorly packed, with a high water content and so
is relatively unstable. It is therefore liable to deformation of various kinds. If the
sediment is on a slope, it may begin to move down-slope, leading to the development
of slumps and syn-sedimentary folds. Stresses applied to recently-deposited beds, for
example by traction currents, may cause shear of the sediment, leading to overturned
foresets or convolute lamination. In an interbedded sequence, more efficiently packed
and therefore denser beds may overlie less dense beds, a situation which is inherently
unstable. In response to gravity, the base of a dense bed may bulge down into the
underlying bed, producing bulbous load casts at its base. Poorly-packed sediments
with a high water content may rearrange their packing, leading to water being expelled
vertically. This may produce a range of water escape structures, including upward
concave dishes and vertical pipes (see Figure 61).

3.7.3.3 Biogenic Sedimentary Structures - Trace Fossils


Many organisms live in or on sandy or muddy sediment. Some burrow into it to
provide a secure home, others eat their way through the sediment, extracting nutrients
and excreting the rest, whilst others move around on the sediment surface. This
activity leads to a range of biogenic trace fossils, including burrows and trails. If the
disturbance, bioturbation, is particularly intense, it may lead to destratification and
homogenisation of the sediment.

28

Table 3
Descriptive terms used for
bed thickness, splitting of
beds and lamina thickness

Sedimentology

Trace fossils come in a wide variety of forms reflecting both the range of organisms
which produced them and their mode of life; different organisms living a similar
lifestyle may produce very similar trace fossils. Like conventional body fossils,
trace fossils are formally classified into genera and species. They can give much useful
information about the environment of deposition, including sedimentation rates
(continuous or discontinuous, low or high rate?), substrate consistency, water depth
and energy of the environment (e.g. current activity and direction). They may
therefore aid the interpretation of the depositional environment.
In addition, because burrows may cut across laminae and bed boundaries, may be
filled by different sediment than the surrounding material and may homogenise
laminated or bedded sediment, bioturbation may have a pronounced influence on
reservoir quality. It may influence both the small-scale permeability kv/kh and larger,
reservoir-scale heterogeneity. In different situations it may either improve or reduce
reservoir quality.

3.8 FACIES AND FACIES SEQUENCES


The aim of reservoir sedimentology is to develop three-dimensional models of
reservoir variability. These models are developed from the recognition of the
depositional environments of the sediments, from which an understanding of reservoir character can be gained. It should be noted that, in most cases, reservoir models
have to be generated from one-dimensional data (i.e. well data).
The method used to interpret the environment of deposition of a sedimentary
succession is known as facies analysis. In the following sections, the term facies is
defined and discussed and the methodology of facies analysis briefly described.

3.8.1 Facies - Definitions


A facies is defined as a body of sediment with specified characteristics. These may
include lithology, primary sedimentary structures, secondary sedimentary structures
and diagenetic character. Depositional facies refers to a sedimentary rock that is
generally quite distinctive, having formed under certain conditions of deposition that
reflect a particular depositional process or setting. A depositional system is a threedimensional association of facies that are in some way genetically linked by sedimentary
processes and environments of deposition.
In addition to depositional facies, other types of facies can be defined. These include
seismic facies, based on a rocks seismic character and petrophysical facies, based on
a rocks petrophysical characteristics.
It should be noted that interpretation of a facies tells us only what processes were
responsible for the deposition of that facies; it does not tell us the environment of
deposition. For example, a medium grained, trough cross-bedded sandstone facies
was clearly deposited by the migration of a 3D megaripple or dune (see Section
3.6.2.2). However, such megaripples may occur in a wide range of environments,
including rivers, lakes, estuaries and shallow seas (in addition to deserts, in the case
of aeolian cross bedding) and examination of a single facies will not allow us to

Department of Petroleum Engineering, Heriot-Watt University

29

differentiate between these environments. To identify the environment of deposition ,


we need to consider the vertical and lateral association of facies.

3.8.2 Facies Associations and Facies Sequences


Different researchers have, historically, used different terms to describe both lateral
and vertical associations of facies. The term facies association can be used to refer to
both vertical and lateral associations of facies, though it is probably used more
frequently in the vertical sense. Throughout the 1960s and 1970s, the majority of
sedimentologists used the term sequence to describe vertical associations of facies.
With the advent of seismic sequence stratigraphy, the term was appropriated by the
seismic stratigraphers and re-defined. To the non-specialist, the term sequence now
has sequence-stratigraphic connotations, which were certainly not the intention of the
earlier workers. To avoid confusion, it is probably now necessary to use the term facies
succession to describe vertical associations of facies. It should be noted, however,
that any papers written before the 1980s are unlikely to make this distinction.
When vertical successions of sediments are examined, it becomes clear that sedimentary facies are often superimposed on top of one another in quite specific sequences.
In other words, the interrelationships of facies are not random, but conform to a limited
number of geological patterns. One of the basic tenets of sedimentology is Walthers
Law, which states: The various deposits of the same [environmental] area and,
similarly, the sum of the rocks of different [environmental] areas were formed beside
each other in time and space, but in crustal profile we can see them lying on top of each
other . . . it is a basic statement of far reaching significance that only those
[environmental] areas can be superimposed, primarily, that can be observed side by
side at the present time In other words, we can only see in a vertical succession those
sedimentary facies that were once side by side during deposition. It should be noted,
however, that this relationship only applies when there are no major breaks, either
stratigraphic or structural, in the sedimentary facies sequence.
In order to interpret the environment of deposition of an observed sedimentary
succession, it is usual to compare the observed vertical and lateral associations of
facies with an existing facies model. Facies models are general summaries of the threedimensional arrangement of sedimentary facies produced by a given depositional
system and are distilled from an analysis of facies relationships in both modern and
ancient examples of that depositional system.

3.8.3 Graphical Sedimentary Logs


In the last thirty years, it has become the norm to represent vertical facies sequences
by the means of graphical sedimentary logs. These logs represent the well depth or
outcrop height as the vertical axis and the sediment grain size as the horizontal axis
(Figure 25). The grain-size scale is as shown in Section 3.3.2.1 and uses the scale.
In academia, the grain size normally increases to the right, but many industrial
sedimentologists use a scale in which grain size increases towards the left (compare
Figures 25a and 25b). The logs show the grainsize and thickness of beds and the nature
of bed contacts.

30

Sedimentology

Figure 25
Examples of sedimentary
logs

It is common to have separate columns which show the lithology, by means of


standard lithological symbols, and the sedimentary structures. The symbols used to
illustrate the sedimentary structures should be designed to look as much like the
required structure as possible, though they must be idealised to a certain extent. There
is no single scheme of symbols, but most sedimentary logs can be read without
frequent references to the key. They have the advantage over verbal descriptions that
they can contain much more information of a more subtle nature and, to the trained
eye, can be understood almost instantaneously. The vertical scale of logs can be varied
to suit the requirements of a particular study. Core is most commonly logged at a scale
of 1:50, but core logs may also be drawn at 1:200 or redrawn to this scale. This allows
the core logs to be compared directly with 1:200 wireline logs. Any discrepancy
between the log depth and the drillers depth can then be identified and the
interpretation of the cored interval can be extrapolated to uncored intervals.

3.8.4 Controls on the Nature and Distributions of Facies


As discussed above, sedimentary facies reflect the process and facies associations
reflect the depositional environment. The behaviour of depositional systems and
therefore the distribution of sedimentary environments is influences by a number of
controls, both internal and external. The internal, or autocyclic, controls include
normal sedimentary process, such as fluvial channel migration or delta switching (see
Section 3.9.2 and 3.9.3). These may occur independently of other, external, controls.
External autocyclic, include sediment supply, tectonics, climate and sea level changes.

Department of Petroleum Engineering, Heriot-Watt University

31

Facies distributions may also reflect biological activity and water chemistry (especially in the case of carbonate rocks).

3.8.5 Genetic Units


To aid the description and modelling of reservoirs, it is useful to group facies or facies
associations into genetic units, which are also known as genetic sedimentary units or
architectural elements. These units are the fundamental building blocks of reservoirs
and can be used build reservoir models. It is a challenge for reservoir geologists and
geo-engineers to identify genetic units in the subsurface and to characterise their
shape, size and petrophysical properties. Models obtained from outcrop and subsurface examples of similar successions can be used to predict the spatial distribution of
genetic units and to model the interwell volume.

3.9 SELECTED CLASTIC DEPOSITIONAL ENVIRONMENTS


It is beyond the scope of this chapter to give a full description of all the common clastic
depositional environments. Instead, a few environments, namely aeolian, fluvial,
coastal and shallow marine and deep marine, have been selected. What follows in the
remainder of Section 3.9 should be treated as a brief introduction to these environments. Much more detail can be obtained from the recommended reading list.

;
y
y
;
;
y
y;y;y;y;y;y;y;;y;y;y;y
y;y;y;y;;y ;y;y;yy;y;y;y;
;
y
;
y
y; y;y; y; y;

30
10

Zone of Tropical

Low Pressure

10

SubTropical

30

;y

Important Mountain
and Plateau Areas

Dry Coastal Areas

Major Areas of Desert

Simplified pattern of
Prevailing Winds

Desert with Sand Dunes

32

High Pressure

30
10
0
10
30

Figure 26
Distribution of the worlds
major deserts in relation to
major atmospheric
circulation and topography
(after Glennie, 1970)

Sedimentology

3.9.1 Aeolian Environments


Aeolian, or wind-transported, sediments occur most commonly in desert environments. Deserts are defined as areas where potential evaporation and transpiration
exceed precipitation. Such areas are commonly located at low latitudes (Figure 26).
The pattern of prevailing winds moves to about 5o North of its mean position in July
and 5o South in January. This simplified pattern is further modified by the large land
masses, which heat up rapidly in summer and cool rapidly in winter. Sandy deserts are
dominated by large fields of dunes (ergs) surrounded by extra-erg areas. Within ergs,
draa are the main bedforms: (Figure 29).The draa are generally covered by smaller
scale dunes (see Section 3.6.2.4). The foresets of these dunes contain laminae
deposited by a number of different processes, or combinations of processes (see
Section 3.6.2.4). The central parts of the slipface are commonly dominated by
grainflow laminae being more common near the dune crest and wind-ripple laminae
present at the flanks and low on the slipface (Figure 27). As the grainflow laminae
generally have the best porosity and permeability, the best reservoir properties in
aeolian successions are commonly found in the dune core areas.

Grainfall Laminae

Cone-Shaped Grainflows

Figure 27
Distribution of different
types of lamination within
small aeolian dunes. A.
Relationship of topset and
different types of lee-side
laminae. B. Horizontal plan
and section (A-B) of crossbedding in a dune truncated
by wind deflation.
Simplified from an
exposure on Padre Island,
USA. (After Hunter, 1977)

Department of Petroleum Engineering, Heriot-Watt University

33

Aeolian dune sets and cosets are typically ms to 10s of m thick. It is rare (but not
impossible) for subaqueous cross bedding to reach these sizes, so large set size is often
taken as an indication of an aeolian origin. Aeolian deposition is episodic at a number
of scales and each phase of deposition is separated from the next by a period of erosion.
This results in the formation of a bounding surface;the temporal and spatial scale over
which they occur give rise to a heirachy of such surfaces (Figure 28). First order
surfaces are very extensive, low-angle features inferred to represent interdune
migration. Second order surfaces are commonly concave-up on sections parallel to
palaeowind and are interpreted as set boundaries due to superposition of bedforms,
whilst third order surfaces are discontinuities (reactivation surfaces) between bundles
of foresets within the same set (Figure 28)
Interdune
(first order)
surface

Superposition
(second order)
surface

Reactivation
(third order)
surface

Figure 28
First-, second- and thirdorder bounding surfaces in
idealised aeolian crossbedding. The second order
surfaces may be inclined
either up wind or down
wind depending on whether
or not they are
superimposed on a larger,
draa-scale form (based on
Brookfield, 1977)

The dunes and draa are separated by low-lying interdune areas. Within draa, interdune
areas between individual dunes are small and relatively short-lived, but interdune
areas between major dune areas or draas may be larger, more permanent features
(Figure 29).
Prevail

ing Win

Interdu

d Direc
ti

ne

on
Draa

Draa cross - bedding


Barchan cross - bedding

Small Barchans
Crecentic Dunes

Interdune Deposits

34

200 m

Figure 29
Reconstruction of draa,
dunes and interdune
environments in relation to
cross bedding and
bounding surfaces: cross
bedding not drawn to scale,
(after Clemmensen and
Abrahamsen, 1983)

Sedimentology

Depositional Conditions
DRY

DAMP

WET

Wind ripples
Aeolian dune cross strata
Lag grain surfaces
Deflation scours
Bioturbation structures
Plant root structures
Sand drift behind obstacles

Figure 30
DIstribution of sedimentary
structures within interune
sediments deposited under
different conditions. Both
modern examples and those
found in the Jurassic
Entrada Formation of the
western USA are indicated
(after Kocurek 1981a)

Adhesion laminae
Microtopography
Rain-impact ripples
Brecciated laminae
Adhesion ripples
Adhesion warts
Evaporite structures
Algal structures
Fenestral porosity
Contorted structures
Rill marks
Wavy laminae
Wrinkle marks
Channels
Small deltas
Water ripples
Subaqueous cross strata

Sedimentary Structures/Features
Modern InterDune Facies
Entrada InterDune Facies

The low-lying interdune areas are influenced by different processes than the dunes
themselves and so contain a different suit of sedimentary structures (Figure 30). Dry
interdune areas, where the water table and its associated capilliary fringe lie far below
the depositioned surface, are dominated by wind ripples and possibly small dunes.
Because they are often sediment-starved, winds blowing them will tend to be undersaturated with sediment and may be erosional. If the water table and its capilliary
fringe intersect the interdune surface, the interdune areas may be damp. Wind borne
grains will tend to stick this damp surface, leading to the formation of adhesion
structures. Wetter interdunes may contain moving or standing water, leading to the
formation of current ripples, wave ripples and other water-generated structures.
Increased organic activity may lead to the preservation of plant rootlets and animal
burrows. Deposition of fine material from suspension provides a muddy blanket
which, on drying, cracks to form typical polygonal desication features. as the water
evaporates, precipitation of evaporite minerals may occur.
Clearly, damp and wet interdune areas will be more extensive during periods of high
water table. These periods may be due to a number of controls, including a rise in sea
or lake level or increased rainfall. Whatever the origin, periods of wetting and
drying can be identified in many ancient aeolian successions. During wetting
periods, aeolian dunes become less active and may be eroded. Extensive interdune
areas may develop and, in more pronounced periods of wetting, fluvial conditions may
predominate. This leads to the development of extensive interdune or fluvial intervals
Department of Petroleum Engineering, Heriot-Watt University

35

overlying aeolian sediments (Figure 31). As conditions again become drier, rivers will
become less active, wet interdunes will become drier and large aeolian dunes will
again become more active. Such a drying trend within interdune facies is shown on
Figure 31).

Dune Foresets

Wind Ripples and / or


Small Aeolian Cross Strata

Adhesion Laminae

Adhesion Ripple
Pseudo-Cross-Strata
Algal Mat Structures
Fenestral Porosity
Water Ripples

Truncated dune Foresets

Figure 31
Drying-upwards sequence
of interdune deposits
showing a transition from a
wet to a dry interdune. Dry
interdune conditions are
terminated by the
encroachment of the next
dune. Present-day example.
Padre Island, USA (After
Kocurek, 1981a)

Both interdune and fluvial sediments have poorer reservoir quality than aeolian dune
sands, so that extensive interdune or fluvial intervals may form baffles to vertical flow
and therefore tend to compartmentalise aeolian reservoirs (Figure 32).

Figure 32
Distribution of crossbedding, bounding surfaces
and interdune deposits in
sections through the
Jurassic Entrada
Formation, Western USA
(after Kocurek, 1981b)

36

Sedimentology

3.9.2 Fluvial Environments


Fluvial, or river-deposited, sediments occur in a wide range of climates and tectonic
regimes. Rivers flow downhill from the source area towards a lake or the sea and their
form reflects a number of controls including climate (especially rainfall), slope and
the available sediment. The geomorphology and behaviour of rivers form a continuum
of types but it is convenient, for discussion of rivers to divide them into somewhat
arbitrary classes. The most common classification of river forms identifies four types
of channels (Figure 33).
Moderate to High Sinuosity

Single Channel

Low Sinuosity

Figure 33
Classification of fluvial
channels according to their
shape in plan. (based on
Miall, 1977)

Meandering

Multiple Channels

Straight

Braided

Anastomosing

Bar surfaces covered during flood stages

The two most common types, which will be discussed here, are meandering and
braided rivers. Meandering rivers have a single channel with a strongly sinuous form
(figure 34). Flow in the apical parts of the bends is helical, with surface flow moving
from the inner to outer bank and flow at the river bed having a component towards the
inner bank. The outer bank is eroded and sediment is deposited on the inner bank to
form a point bar. Continued erosion of the outer bank and deposition on the point bar
increases the amplitude of the meanders and produces relatively narrow necks on the
point bar. During a severe flood, the point bar neck may be breached, leading to a
shortening of the channel course and abandonment of the old meander loop.

Department of Petroleum Engineering, Heriot-Watt University

37

Point Bar
With Scroll Bars

Crevasse
Splay

Flood Plain
Fires

Older
Crevasse
Splay

Lateral Accretion
Sufaces

Older
Charred
Splay

Meandering channels may transport sandy or muddy sediment but, from a reservoir
point of view, we are interested mainly in the more sandy rivers. During periods of
flooding, the river may flood onto the surrounding low-lying land, the floodplain. The
river may either break through its banks, to form a temporary crevasse channel, or may
flood over the banks over a longer length. In either case, the flood waters will tend to
deposit their coarsest sediment close to the main river, producing thin beds which will
tend to become finer and thinner away from the river. Repeated floods over many years
will produce elevated ridges of sediment, known as levees, close to the channel. The
meandering river will continue to flow along its raised alluvial ridge until, following
a major breach of its banks, it will follow a new path across the lower-relief floodplain.
Such avulsion of the channel will abandon the old alluvial ridge downstream of the
point of avulsion. Thus, meandering river systems will tend to produce complex
meander-belt sandbodies separated by finer-grained floodplain sediments.
Unlike meandering rivers, which have only one active channel at any time, braided
rivers have a number of active channels separated by sandy or gravelly bars (Figure
35). Braided rivers tend to form on slightly steeper slopes, and where there is a high
proportion of sandy or gravelly sediment.

38

Figure 34
Block diagram showing the
three-dimensional form of a
meandering river (modified
after Miall, 1985)

Sedimentology

Floodplain

Sandbar
with superimposed
megaripples/dunes

Figure 35 Block diagram


showing the three dimensional form of a
braided river (after Miall,
1985)

Because of their easily-eroded sandy banks, individual channels in a braided system


will tend to migrate laterally and to shift their course frequently. The intervening bars
migrate both downstream and across the streams. Braided rivers therefore tend to
produce compound sandbodies consisting of a number of mutually-erosive channel
bodies. These multi-storey and multilateral bodies will be both thicker and wider than
the channel dimensions.
The other two types of channels, straight and anastomosing channels, are rarer and less
well described than meandering and braided rivers. Straight channels are single
channels of low sinuosity, and are characterised by side bars which are attached to
alternate sides of the channel. Straight channels produce single channel-fill sandbodies.
Anastomosing rivers, like braided rivers, consist of a number of active channels which
split and rejoin in a down-valley direction. In contrast to braided rivers, with their
active bars between channels, the individual channels of anastomosing rivers are
separated by larger, finer grained, more stable islands. These islands are commonly
low-lying, boggy and vegetated and the channels do not migrate much laterally. This
leads to the development of relatively narrow but thick multi-storey sandbodies
(Figure 36).

yyyy
;;;;
;;
yy
;;;;
yyyy
;;
yy
yyyyy
;;;;;
;;
yy
;;
;
yy
;
yyyyy
;;;;;
yy
;;
yyyyy
;;;;;

Vegetated Island

Figure 36
Block diagram showing the
three-dimensional form of
an anastomosing river
(after Miall, 1985)

Channel Sandstones

Department of Petroleum Engineering, Heriot-Watt University

39

The deposits of meandering and braided rivers contain a variety of scales and degrees
of heterogeneity. Classically, meandering channels produce erosive-based, upwardfining sandbodies (Figure 37). As well as the vertical variation in grain size and
sedimentary structures, the channel sandbodies may also contain inclined lateral
accretion units, which represent deposition on the point bar as it migrated across the
channel (Figure 37). In addition to these within-channel heterogeneities, the meander
belt sandbody will also consist of a complex of erosive-based channel sandstones and
channel abandonment facies.
Surface Current
Bottom Current

1.5 M
St

Sp

an

tb

Cu

Chut

Sr

Lateral accretion units

Figure 37
Diagrams showing the
development of an upwardfining trend in meandering
channel sandstones, and
laterally-accreted point bar
deposits. (after Miall, 1985)

Lateral accretion is less common in braided rivers, but downstream migration of bars
may lead to the development of downcurrent-dipping or downstream-accreted
elements (Figure 38).

Figure 38
Development of
downstream-accreted units.

In addition to channel forms and laterally-accreted and downstream-accreted units, a


number of other architectural elements have been identified in fluvial successions
(Figure 39). These have been classified by Miall (1985) and can be used both to
differentiate between deposits of the different channel types and as the basic building
blocks in reservoir modelling.
40

Sedimentology

Ch Channel

FI
Lateral Accretion

GB
Gravel Bar and Bed Form
SG
Sediment Gravity Flow

DA
Downstream Accretion

SB
Sand Bed Form

Figure 39
The eight basic
architectural elements in
fluvial deposits, (after Miall
1985). No vertical
exaggeration. Note the
variable scale.

LS
Laminated Sand
OF
Overbank Fines
0.5-5m

As has been shown for both meandering and braided systems, the sandbodies
produced are generally complex, so that fluvial reservoirs consist of channel belt
sandbodies rather than individual channel sandbodies. It should be noted that the
geometry of these sandbodies will be controlled by the stacking pattern and may have
little relationship to the geometry of the individual channel sandbodies (Figure 40).

Department of Petroleum Engineering, Heriot-Watt University

41

3.5

13
B

19
C

17.5
D

Latest Channel

Before leaving the fluvial system, it is relevant to consider briefly two other alluvial
environments. Where rivers leave the confines of a valley, they commonly form cones
of sediment known as alluvial fans (Figure 41). In addition to channelised flows,
alluvial fans are also influenced by sheet floods, which deposit sheet-like sandstones
and conglomerates. The sediment tends to become finer away from the fan apex and
migration of the fan produces an upward-coarsening trend. Fans commonly form at
the fault-influenced margins of mountain ranges and fault movements may cause
rejuvenation of the fans and the influx of coarser material. As the newly-uplifted
mountains are eroded, the sediment will become progressively finer, leading to the
development of upward-fining megasequences.

yy
;;
yy
;;
yyyy
;;;;
yyyy
;;;;
;
y
;;;
yyy
;;;;
y;
;;;
yyy
yyyyyy
;;

Figure 40
Diagram to show the lack
of relationship between the
geometry of an individual
active channel and the
geometry of resulting
channel-fill sand bodies
(After Miall, 1985).
Numbers above each
channel are the width/
thickness ratios of the sand
bodies. A,D Simple
Channels; B,E, F, broad
channel-fill complexes
formed by lateral channel
migration or switchng with
little contemporaneous
subsidence; C stacked
channel complex formed by
vertical aggradation.

Siltstones

Cross-bedded Sandstone

Siltstones and Sandstone

Conglomerate and Sandstone

km

42

Approx. Scale

5
km

10

Figure 41
Alluvial fans

Sedimentology

Since the development of sequence stratigraphy, attention has turned to incised


valleys caused by falls in relative sea level. As the relative sea level falls, the incised
valleys are largely bypassed by the sediment (Figure 42A) but, as the relative sea level
stabilises and begins to rise, sedimentation will begin in the valley, which will
continue into the highstand period. The fills of the incised valleys are very complex,
and include deltaic sediments as well as fluvial sediments (Figure 42B).

High

Shelf / Ramp

Non-Incised
Fluvial System

Incised Valley System

Low

Incised Valley
Time
Lowstand (Fan)
Systems Tract

Figure 42
Formation and fill of an
incised valley (modified
from Zaitin et al)

Low

High

Shoreline / Delta

Meandering
River System

Braided
River System

Non-Incised
Fluvial System

B
Time
Lowstand (Fan)
Systems Tract

3.9.3 Coastal and Shallow Marine Environments


There is a very wide range of clastic coastal environments. In the following section
we will describe the beach and shoreface environment, before a discussion of barrier
island and deltaic environments.
Beaches form the boundary between the shallow marine and terrestrial environment.
They are dominated by wave processes, but in most cases are also affected by tides.
The constituent parts of a beach profile are defined in terms of the tide marks and are
shown on Figure 43. The area above the high water mark is the backshore, and the area
between the high and low water marks is the foreshore. The shoreface extends from
the low water mark to the fairweather wave base.

Department of Petroleum Engineering, Heriot-Watt University

43

Waves begin to build up


Spilling Breakers
Shoaling Waves

Fairweather Wave Base

5-15m

L
2

Low
water
mark

Longshore Bars

Lower
Storm Wave Base

Middle

Upper
Foreshore Backshore

Shoreface

Offshore
Muddy Substrate

High
water
mark

Sandy Substrate

Skolithos

Figure 43
Definition of the beach
profile

Cruzlana
Ichnofacies
Zoophycos

Water is driven onto the beach by waves, and then returns to the sea as localised
currents. The beach profile is therefore influenced by both waves and currents. As the
waves break on the shore, they produce rapid, shallow currents which flow up the
beach before flowing back into the sea. These swash and backwash currents form the
seaward-dipping plane beds which characterise most foreshores. Below the low water
mark, the dominant processes on the shoreface depend on a number of factors,
including the wave energy. Fairweather waves will tend to produce a mixture of wavegenerated bedforms such as wave-ripples and bedforms, including megaripples,
produced by wave-driven currents. When a shoreline is dominated by storm waves,
the dominant bedform may be hummocks.
As storm-generated currents flow offshore, they transport sediment into deeper water,
often as bottom-hugging currents similar to turbidity currents (Figure 44). After this
sediment has been deposited, it may be reworked by the storm waves themselves.
Storm-Surge Ebb
Storm surge tide briefly
stores sediment-laden
waters in lagoon

Storm Winds

<6m
As storm abates, storm
surge ebb currents emerge from washover
channels and flow seawards

Fairweather Wave Base (FWB)


Storm Wave Base (SWB)
HCS Sands
Turbidity Current

Geostropic Currents

Storm Winds

Surface wind-forced current


Seaward-directed geostropic
current, interacting with wave
orbital motions to deposit HCS beds

(FWB)
(SWB)
HCS Sands

Wind

l
Coasta ing
well
Down

Surfac

Core F

low

Botto

44

Figure 44
Alternative mechanisms of
storm surge ebb currents
and wind-forced or
geostrophic currents for the
generation of storm
deposits. (After Walker
1979, Morton 1981 and
Swift, Figueiredo et al,
1983 Elliott Reading II
Figure 7.6 1985)

Sedimentology

In general, the energy is greatest, and the sediment coarsest, on the higher parts of the
beach profile. In deeper water, discrete storm-generated beds may be separated by
mudstones whereas, in the more proximal parts of the beach, the sand beds coallesce
to form sand-dominated intervals. As a beach progrades offshore, it produces an
upward-coarsening facies sequence (Figure 45).

Coastal Plain
Coal/Backshore
Beach/Foreshore
Ridge and Runnel/
Rip Channels

Middle Shoreface.
Swaley Cross - Stratification

Figure 45
Typical upward-coarsening
facies sequences produced
by shoreline progradation.
Progradation of the
Shoreline produces a
gradationally-based
succession passing from
outer through inner shelf
deposits into sanddominated shoreface and
beach sediments

Lower Shoreface Inner Shelf Transition.


Hummocky Cross Stratification

Mid - Shelf.
Bioturbated
Sandy Siltstone

Outer - Shelf.
Bioturbated
Mudstone

Whilst some beaches are joined directly to the main land area, others are separated
from it by an area of standing water known as a lagoon (Figure 46). Because of the
mixing of fresh water and sea water, lagoonal waters are brackish, that is, of
intermediate salinity. The beach ridge on the seaward side of a lagoon is termed a
barrier island. At each high tide, water passes through the barrier island into the lagoon
via tidal inlets. In areas with high tidal ranges, these tidal inlets are closely spaced. Ebb
and flood tidal deltas may form on the seaward and landward side of the inlets, as the
confined flow expands laterally and loses its power.

Department of Petroleum Engineering, Heriot-Watt University

45

yyyyy
;;;;;
;;;
yyy
;;
yy
;
y
;;;;;
yyyyy
;;
yy
;;;
yyy
;;
yy
;;
yy
;
y
yy
;;
yy
;;
;;;;;
yyyyy
;
y
;;
yy
;;;
yyy
y
;
;;
yy
;;
yy
;
y
;;
yy
yy
;;
;;;
yyy
;;;;
yyyy
yy
;;
;;;; yy
yyyy
;;
Washover

yy
;;
;;
yy
Tidal Flat

Dunes
Beach

Flood Tidal
Delta

Marsh

Figure 46
Block diagram illustrating
the various
subenvironments in a
transgressing barrier-island
system

Ma

in
Tid
(In al C
let ha
)
nn

Secondary Tidal
Channel

Ebb Tidal Delta

el

Tidal inlets migrate rapidly along the barrier island, eroding the upper shoreface and
foreshore deposits. The inlet deposits are similar to those of fluvial channels, with
erosive bases and upward-fining trends (Figure 47). Migration of the inlets can lead
to the upper parts of barrier island successions being dominated by tidal inlet deposits.
Barrier islands and beaches will tend to produce linear sandbodies oreiented parallel
to the coastline. If the beach or barrier island migrates seaward, it will produce more
sheet-like bodies.
Longshore Drift
Beach and Dune Ridge

Direction of Channel Migration


Sea Level
Foreshore
and
Shoreface
Deposits

Channel Deposits

Erosion Surface

4000

3000

2000

1000

15

Metres

y;;
;
yy
;;
yy
yy
;;
yyy
;;;
Time Lines

30

Metres

Where a major river reaches a standing body of water, such as a lake or the sea, the
basinal processes will attempt to rework the sediment supplied by the river. If the river
supplies sediment faster than it can be reworked by the basinal processes, the shoreline
will project locally into the basin, forming a delta. On the subaerial part of the delta
plain, the river will split into two or more smaller distributary chanels, which may
themselves split into still smaller channels,causing sediment to be supplied to many
points along the delta front (Figure 49).
The form of the delta will depend on a wide range of parameters, including climate,
tectonic setting, sediment supply and energy of the receiving basin. A commonlyused classification of deltas uses a triangular diagram to compare the relative
importance of fluvial, tidal and wave processes (Figure 48).
46

Figure 47
Generalised cross-section
parallel to shoreline
illustrating the development
of a barrier-inlet sand body
by lateral inlet migration.
(Modified from Hoyt and
Henry, 1965).

Sedimentology

ng
Elo
te

4
Fluvial
Dominated

ba

ate
ng
Elo

ate

SEDIMENT
INPUT

Lo

Mahakam

Nile

Cu
s

pa

Classification of deltas in
terms of river, wave and
tide influence, simplified
from Galloway (1975)

Ord
Klang-Langat

Sao Fransisco

Wave Energy Flux

e
rin
tua

Figure 48

Tide
Dominated

Rhone

Es

te

Niger

Wave
Dominated

Tide Energy Flux

MISSISSIPPI LOBES
6 Modern, 4 St. Bernard, 5 Lafourche

Fluvial-dominated deltas will supply more sediment to the coastline than can be
reworked by the basinal processes. The resultant data will therefore form a pronounced protuberance of the shoreline. Depending on the depth of water into which
the delta is prograding, and the degree of reworking, the delta may be either lobate or
elongate (Figure 49a and b ). In the case of tidally-influenced deltas, the tidal
processes will tend to produce a radial pattern of distributary channels which become
broader towards the basin (Figure 49d). In wave-dominated deltas, a high proportion
of the sediment supplied to the river mouth will be reworked into beach ridges on
either side. The resulting delta will, therefore, often cause only a slight deflection of
the coastline (Figure 49c).

Figure 49
Delta models based on the
relative dominance of
fluvial, wave and tidal
processes (from Fisher et
al, 1969)
Department of Petroleum Engineering, Heriot-Watt University

47

It should be noted that the descriptions above are based on deltas dominated by a single
process (i.e. plotting near one of the corners of Figure 48). In reality, many deltas, for
example the Nile and Niger, reflect an interaction of two or more processes, and so plot
nearer the centre of the triangle (Figure 48).
At the mouths of the distributary channels on fluvial-dominated deltas, the sediment
laden river waters interact with the basinal water. The exact result depends on the
relative density of the two waters, the basinal energy and the water depth but, in
general, as the fluvial flow expands on leaving the confines of the levees, it loses power
and begins to deposit sediment. This leads to the development of a mouth bar, on
which the sediment fines away from the apex (Figure 50). Progradation of the delta
mouthbar therefore produces an upward-coarsening sequence dominated by currentgenerated structures (Figure 51).

yyyyy
;;;;;
;;;;
yyyy
;;;;;
yyyyy
;;;;
yyyy
;;;;;
yyyyy
;;;;
yyyy
;
y
yy y;
;;
;;;;;
yyyyy
;;;;;;
yyyyyy
;;;;
yyyy
y
;
;
y
;
y
;;
yy
;;;;;;
yyyyyy
;;;;
yyyy
;;;;;;
yyyyyy
y
;
y
;
;
y
;
y
;;
yy
;;;;;;
yyyyyy
y
;
y
;
yy
y
;
y
;
;;
;;;;
yyyy
;;;;;;
yyyyyy
;;;;;;
yyyyyy
;;
yy
;;;;
yyyy
;;;;;;
yyyyyy

Friction-Dominated River Mouth

Subaerial
Levee

L
S ub A q u e o u s

ev

ee

Channel

Coarsest
Sands

Middle Ground Bar

Finer
Sands

Interbedded
Sands and Silt

Silt and
Clay

Buoyancy-Dominated River Mouth


Subaqueous
Levee

Subaerial Levee

Channel
Subaerial Levee

Subaqueous
Levee

48

Bar Back

Bar Crest

Bar Front

Distal
Bar

Prodelta

Figure 50
Friction-dominated and
buoyancy-dominated river
mouth bars which develop
in shallow-water and deepwater areas respectively of
fluvial-dominated deltas,
for example in the east and
south of the modern
Mississippi delta (modified
after Wright, 1977)

Sedimentology

Figure 51
Typical vertical succession
in a fluvial-dominated delta
front (from Kelling &
George, 1971)

Progradation of the mouth bar produces elongate sandbodies, known as bar finger
sands, which form a radial pattern (Figure 52). It should be noted that the bar finger
sands have a considerably greater cross-sectional area than the distributary channel
which produced them (Figure 52b).
BRANCHING PATTERN

30

60

30

Delta Plain
Silty sands, silty clays

0
0

60

FO

60

RM

60

30
30

AT

Interbranch
areas widen
seaward

DE
EPW
AT E R

Marsh
Organic-rich silty clays

Natural Levee
Silty sands, silty clays

Fingers
narrow upstream

Figure 52
Bar finger sands of the
Mississippi delta as
described by Fisk (1961).
Sand contour thickness is
in metres

;;;;;;
yyyyyy
yyy
;;;
;
y
;
y
;
y
;;;;;;
yyyyyy
;;;;
yyyy
;;;
yyy
;
y
;
y
;
y
;;;;
yyyy
yyyyyy
;;;;;;
LENTICULAR CROSS SECTION

EL
N-D
MARGI

TA

Sparse to
abundant fauna

"Clean"
sand zone

Mud diapir

Delta Front
Clayey silts

Prodelta
Silty clays, clays

ne
Zo
Tr a n
s iti o n
Sa
n d s a n d S ilt s

Sparse
Fauna

Abundant
Fauna

In contrast, wave-dominated deltas consist of a series of beach ridges associated with


the distributary channels (Figure 53). Their vertical facies sequences will be very
similar to those produced by prograding beaches (described in the previous section),
differing mainly by the presence of associated distributary channels.

Department of Petroleum Engineering, Heriot-Watt University

49

Gulf Of Mexico

18 30'

le
xe
s

Active Distributary Mouth

Be a c h

g
Rid

m
Co

San Pedro
and
San Pablo River

N
Usumacinta River
92 30'

93
Grijalva River

10

20

km

When a delta is influenced by a combination of fluvial, wave and tide processes, it is


commonly fronted by a beach/barrier island complex, with extensive tidal flat, tidal
channel and estuarine facies behind it (Figure 54). More tide-dominated deltas may
consist mainly of estuarine facies and, in the rock record, may be difficult to
distinguish from estuaries.

Figure 53
An example of a wavedominated, high-destructive
delta, the Grijalva Delta.
(modified from Collinson,
1978, after Psuty, 1966)

Figure 54
An example of a mixed
fluvial, tide and wave
influenced delta, the Niger
Delta (from Oomkens,
1974)

50

Sedimentology

From a detailed study of the vertical succession of facies in preserved deltaic


successions, it is often possible to differentiate between the deposits of the different
delta types. This can be done either by the study of cores or simply by wireline log
interpretation (Figure 55). From this interpretation, it is possible to map the sandbody
distribution in the subsurface using a model derived from the interpretation (Figure 56).

Figure 55
Subsurface interpretation
of fluvial-dominated deltaic
reservoirs

Figure 56
Sandbody geometries of the
six delta types of Coleman
and Wright (1975) plotted
on the river-, wave- and
tide-dominated tripartite
classification of Galloway
(1975).

Department of Petroleum Engineering, Heriot-Watt University

51

Another important phenomenon frequently associated with deltaic deposits is the


formation of growth faults. Deltas commonly lead to the rapid deposition of sediment,
which thins, and therefore slopes, towards the basin. This may encourage the
development of early syn-sedimentary normal faults, which generally dip towards the
basin and are often curved, or listric, in form. These growth faults produced
localised topographic lows in their downthrown side, which may act as depocentres
for coarse grained sediment. Rollover of the beds on the downthrown side may also
lead to the development of anticlinal traps. Thus, the occurrence of growth faults can
control the distribution of both reservoir sandbodies and traps in deltaic environments.
In some deltas, such as the Niger, growth faults are one of the major controls on
hydrocarbon distribution (Figure 58).

Figure 57
Inter-relationships between
flow initiation, transport
and deposition of and by
sediment gravity flows.

Figure 58
Niger Delta growth fault
traps (from Elliot,1978,
after Weber and Daukoru,
1975)

52

Sedimentology

3.9.4 Deep Marine Clastic Environments


The physical processes, such as tides and waves, which dominate coastal and shallow
marine environments are generally absent or ineffective in the deep marine environment. For the majority of time, the deep oceans are low-energy environments in which
fine grained carbonate or clastic muds are able to accumulate. Coarse-grained
sediment is transported into these environments by a number of infrequent and shortlived processes, of which the sediment gravity flow processes (slumps, debris flows,
turbidity currents etc.) are dominant. Our discussion of the deep water environment
will, therefore, concentrate on the deposits of these processes.
Sediment gravity flows were discussed in Section 3.6.2.5. It should be noted that the
dominant process in a gravity flow may change with time and space as the flow
develops (Figure 60). In recent years, there has been much discussion in the literature
concerning the differentiation of the deposits of high-density turbidity currents and
sandy debris flows. It is likely that many of the deposits in question may have been
transported over long distances by turbidity currents but, immediately prior to
deposition, the lower parts of the currents took on the characteristics of a debris flow.
The deposits, which record only their mode of deposition, rather than their longdistance transport, therefore resemble debris flow deposits.

Figure 59
Inter-relationships between
flow initiation, transport
and deposition of and by
sediment gravity flows

The earliest deep marine sandstones to be studied in detail were turbidites. The facies
present in an idealised classical turbidite are summarised by the Bouma sequence
(Figure 60). For many years, this was thought to be the paradigm for all turbidite
sandstones. However, later work showed that many deep-water sandstones did not fit
this idealised model, so a number of new models for the deposits of single sediment
gravity flow were proposed (Figure 61).

Department of Petroleum Engineering, Heriot-Watt University

53

A
E (h)
E (t)

(D)
Figure 60
The Bouma sequence for a
classical turbidite (Bouma,
1962). Division A is
structureless; B is parallellaminated sand; C is
rippled and/or convoluted;
D consists of parallellaminated silt and mud.
The pelitic interval E is
partly of turbidite origin (t)
and partly hemipelagic (h).

C
B
A

Sole Marks

Turbidity Current

Grain Flow

Rippled or flat top


Cross lamination,
Climbing ripples

Flat top

Laminated

Massive,
grain orientation parallel
to flow

No grading ?

Good grading
("Distribution grading)

Reverse grading near base?


Scours, injection structures

Flutes, tool marks on base

Fluidized / Liquified Flow

Debris Flow
Irregular top
(Large grains projecting)

Sand volcanoes or flat top


convulate lamination
fluid escape "Pipes"

Massive,
poor sorting random fabric

Dish structure?

Poor grading, if any ("Coarse tail")

Poor grading
("Coarse tail grading")

Basal zone of "Shearing" broad "Scours"


? Striations at base

? Grooves, flame and load


striations structures on base

In many of the early outcrop studies, relatively little attention was paid to the vertical
trends within turbidite successions. However, careful examination of some turbidite
successions demonstrates the occurrence of facies sequences in which the thickness
and/or mean grain size of the sandstone beds increases or decreases upwards. These
trends were attributed to deposition on different parts of a submarine fan (Figure 62).
Submarine fans occur offshore from major river systems or off the continental shelf
in many parts of the world. Like deltas, they are sourced from a single point and contain
54

Figure 61
Facies models for
turbidites, debris flow
deposits (debrites) and
slump deposits.

Sedimentology

a system of distributary channels supplying lobes. The upward-thickening trends in


turbidite sequences are attributed to the progradation of fan lobes and the upwardthinning trends to deposition within a channel. It should be noted, however, that this
model is somewhat simplistic, and that the thickening or thinning trends may be very
subtle.

Figure 62
Single point-source
submarine fan in a sandrich system

In addition to point-sourced fan systems, multiple-sourced systems also occur. This


may cause the separate fan systems to overlap and coalesce, forming a more ramp-like
system (Figure 63). Away from the submarine fans or ramp systems, turbidites may
not be organised into upward-thickening or upward-thinning trends. Such successions
are commonly attributed to deposition on the basin plain.

Figure 63
Multiple-source submarine
ramp in a sand-rich system

Within deep marine clastic systems, a number of architectural elements can be


identified. These include channels, lobes and sheets, as well as slumps and slides. The
main reservoir targets are sand-rich channels and lobes, although the other architectural
elements may also form good reservoirs under certain conditions.
Department of Petroleum Engineering, Heriot-Watt University

55

As in the case of the other depositional systems described above, a detailed examination
of 1D well data (either wireline logs or core) from deep water clastic successions may
enable sedimentary trends to be identified and a three-dimensional model of the
system to be built (Figure 64).

Figure 64
Log signatures of a single
point-source mixed sandmud fan system

3.10 CARBONATE SEDIMENTS


3.10.1 Introduction
In many parts of the world, beaches are dominated by quartz sand and, for those of us
brought up in these areas, it is tempting to believe that this must be the case
everywhere. However, in other parts of the globe, carbonate sands are the norm. It
should be remembered that the natural sediments in marine environments are
carbonates; it is only when clastic material is brought into the area in sufficient
quantities to subdue carbonate-secreting organisms, and to dilute what carbonate
material is being formed, that clastic sediments result.
Carbonate sediments have many things in common with clastic sediments, but also a
number of important differences.
56

Sedimentology

Whereas the majority of clastic grains have their origin some distance from the site
of their deposition, many carbonate grains are formed at or very close to their eventual
site of deposition. A small proportion of these carbonate grains are precipitated
directly from marine or lacustrine water, but the majority are precipitated by, or with
the assistance of, plants or animals. Many carbonate rocks are composed almost
entirely of the broken shells of marine animals, which may have been transported only
a short distance.
Because the majority of clasts, and matrix, in carbonate sediments are composed
initially of various forms of calcium carbonate, they are more soluble than most clastic
rocks. The ions liberated by dissolution will then be available for precipitation as
carbonate cements. Carbonate sediments are therefore likely to undergo earlier
dissolution and/or cementation than clastic sediments.
Because of these differences, it is often forgotten that, under certain circumstances,
carbonate sediments behave in essentially the same way as clastic sediments. In those
areas where carbonates predominate, the carbonate grains respond to the physical
processes of the environment in the same way as any other type of sand. In currentdominated environments, carbonate grains may be transported in megaripples, and so
form cross bedding, whilst wave-ripples may form in wave-influenced carbonate
environments. Well-bedded deep marine carbonates commonly have all the characteristics of clastic turbidites.

3.10.2 Carbonate Grains


Bioclasts
As mentioned above, the majority of carbonate grains are precipitated by organisms,
as internal or external skeletons, or by biochemical processes mediated by organisms.
Clasts derived from skeletal material are known as bioclasts. These may consist of
whole skeletons, but are much more commonly broken fragments of skeletons. In the
marine environment, a wide range of organisms, including bivalves, brachiopods and
gastropods, produce external shells, whilst others, such as echinoids and crinoids
have internal calcite skeletons and some, like corals, produce frameworks in which
small colonial organisms can live. Some algae (e.g. Halimeda) also produce calcareous hard parts and these may be the dominant bioclasts in some environments.
On the death of these organisms, their hard parts become available as potential clasts.
In some cases, the shells remain in situ and unbroken , but they are more commonly
transported by waves or currents and become broken and/or rounded during transport.
The mineralogy and crystalline structure of skeletal material varies from organism to
organism, so that organic debris generally has a greater range of shapes and densities
than terrigenous clastic sands. This obviously has an influence on their hydrodynamic
behaviour and can lead to efficient sorting of the different grain types.
Ooids
Ooids are small (less than 2mm) near-spherical carbonate grains with a pronounced
concentric structure. They generally form in moderate to high energy shallow marine
environments. In the past, there has been much debate over their origin, but their
growth is generally accepted to be associated with algae, which bind the fine-gained
sediment of which they are composed, and may also have a role in the precipitation
Department of Petroleum Engineering, Heriot-Watt University

57

of the carbonate material. Larger, concentrically-structured algal nodules, known as


oncoids, also occur in some marine and lacustrine environments.
Intraclasts
Because carbonate sediments may be subject to earlier cementation than terrigenous
clastic sediments, it is common for carbonates to be lithified at, or close to, the
sediment surface. If erosion then occurs, instead of breaking up into individual grains,
the sediment may break into groups of grains lightly cemented together. These
compound clasts are known as intraclasts.

3.10.3 Minerals Present


As mentioned above, different organisms precipitate hard parts composed of different
minerals. The majority of shelly material is composed of calcium carbonate, but this
may occur in several forms. The most stable mineral, and that which occurs in most
old carbonate rocks, is calcite. However, many organisms, including bivalves,
gastropods and many corals, precipitate skeletons of a different form of calcium
carbonate, known as aragonite. Aragonite is less stable than calcite and is therefore
easily dissolved (see Section 3.11). Calcite may also occur as 2 different forms, high
Mg calcite and low Mg calcite. Of these two, High Mg calcite is less stable. Other
organisms may produce hard parts composed of phosphate minerals.
The other minerals common in carbonate rocks form by diagenesis (see Section 3.11)
and include iron-rich ferroan calcite, dolomite (calcium magnesium carbonate) and
siderite (iron carbonate).

3.10.3 Classification of Carbonate Rocks


In addition to the grain types, carbonate sediments may also contain carbonate mud,
or micrite. Micrite may form by the physical abrasion of clasts during transport, but
is also produced by the organic breakdown of clasts. Dead shelly material is
commonly colonised by organisms such as algae, which bore into the surface of the
shells and, in doing so, produce fine-grained micrite. This algal boring is often seen
in the form of dark algal rims on bioclasts.
The main components of carbonate sediments are therefore a range of grain types
including bioclasts, ooids and intraclasts, and micrite mud. An early classification
scheme for carbonates, proposed by Folk in 1973, divides carbonate rocks on the basis
of the clast types and the presence of matrix or cement. Carbonate rocks with micrite
matrix are given the suffix -micrite, whereas those with crystalline ("Spary") cement
are given the suffix -sparite. Thus a rock composed of bioclasts in a micrite matrix is
a bio-micrite. If it contains little or no matrix, but has a sparry cement, it is a bio-sparite.
If more than one clast type is present, the different clasts are listed in decreasing order
of abundance, for example oo- bio-sparite, bio- intra-micrite.
A later classification concentrates on the primary grain and matrix texture and ignores
cement (Dunham, 1962;). Carbonates consisting only of grains and cement, with no
matrix, are known as grainstones. Those with some matrix, but with a clast-supported
fabric, are packstones, whilst those with a matrix-supported fabric are wackestones.
Carbonate rocks consisting mainly of micrite, with less than 15% grains, are known
as mudstones. It should be noted that the mud in this case is carbonate material, in
58

Sedimentology

contrast with terrigenous mudstones, which are composed mainly of clay minerals. To
avoid confusion, it is sometimes better to refer to these sediments are carbonate
mudstones. In addition to these textural terms, further terms are needed to describe
sediments strongly influenced by growing organisms. Where a rigid framework is
built, for example on coral reefs, the resulting rock is a framestone. Where growing
animals or plants reduce the power of currents or waves, leading to the deposition of
sediment, the sediment is a bafflestone.

3.10.4 Carbonate Depositional Environments


As described above, carbonate depositional environments will not be discussed in as
much detail as the clastic environments and the discussion in the following sections
will concentrate on those features of relevance to the reservoir performance of the
rocks. The majority of carbonates are deposited in marine or lacustrine environments;
we will concentrate here on marine environments.
Shallow marine environments are sites of high biological diversity and productivity.
Particularly in areas of low clastic input, carbonate clasts, such as bioclasts and ooids,
and carbonate mud will form the dominant sediment. Carbonate sands will be washed
onto the beach, so that the coastal and shallow marine environments are dominated by
carbonate deposition. This is the case in many modern dry, low-latitude areas, such
as the Gulf coasts.
In the shallow marine waters, waves and tides rework the sandy sediment into shallow
bars or shoals. In general, the energy will be highest on the topographically higher,
shallower water parts of the shoals. The sediments are therefore coarser grained and
better sorted at the top of shoals and lateral migration of a shoal will produce an
upward-coarsening sequence, with mudstones and wackestones being replaced
upwards by coarser grained packstones and grainstones. Between the shoals, the
sediment is likely to be finer grained and dominated by mudstones and packstones.
The organisms living in this environment commonly occur in great abundance in
certain areas. A combination of the accumulation of skeletal debris and the baffling
effect of the living organisms can lead to the development of a mound of sediment
known as an organic buildup. These buildups commonly contain high concentrations
of coarse skeletal material, such as crinoids and rudist bivalves. If a framework of
organic material is built, for example by colonial corals, a reef results. Reefs are very
complex environments, which include both living and dead corals, as well as a large
number of other organisms which live on or near the reef. The reef may contain holes
and caves, and a talus slope of broken coral debris may form on the seaward side.
Shoals, buildups and reefs may all protect the area on their landward side from the
highest wave energy, allowing a low-energy lagoonal environment to develop. These
lagoons are generally dominated by fine-grained carbonates, but may contain local
shoals and buildups.
In tidal environments, extensive areas of carbonate tidal flats may develop. These
consist of horizontally bedded fine grained carbonates cut by tidal channels of various
sizes. The fills of the channels are generally coarser grained and cleaner than the tidal

Department of Petroleum Engineering, Heriot-Watt University

59

flat sediments. The topographically higher parts of the flats may be exposed at low
tide, and so be subjected to repeated wetting and drying. This can lead to the
development of a number of distinctive features such as mud cracks, tepee structures
and fenestrate fabrics.
Progradation of carbonate tidal flats or beaches leads to the development of extensive
areas of low-lying coastal plain or sabkha. Marine waters may soak into the sabkha
sediments and evaporate from surface of the sabkha, precipitating a range of evaporite
minerals and causing complex early diagenesis (see Section 3.11).

3.10.5 Porosity in Carbonate Rocks


In addition to the porosity types present in clastic sediments, carbonates also contain
a number of types of pores which occur exclusively in carbonates or are more
important in carbonates than in clastic sediments.
Vuggy pores
Because of the variable chemistry of many carbonate clasts, there is a likelihood of
partial or total dissolution of the less stable clasts. Dissolution of, for example, large
aragonite bioclasts can lead to the development of large secondary pores. These vuggy
pores are commonly isolated from each other within a lower-porosity matrix.
Intragranular pores
Many bioclasts, such as bivalves, are hollow or, in the case of corals or crinoids, have
a microporous texture. There is therefore a possibility of high proportions of
intragranular porosity in bioclast-rich carbonates.
Shelter porosity
Platy clasts, such as broken shells, may protect the area underneath them from
sediment falling from above. This area may therefore remain empty, as shelter
porosity, or be more loosely packed than the surrounding sediment.
Fenestrae
Shrinking associated with repeated wetting and drying can cause sedimentary laminae
to pull apart, producing irregular pores or fenestrae (from the Latin for window).

Intercrystalline porosity
Recrystallisation of carbonate rocks may cause changes in volume. For example,
dolomitisation of calcitic or aragonitic sediment involves a slight decrease in volume.
Thus the dolomite crystals may not fill the entire volume, allowing intercrystalline
porosity to exist between the dolomite crystals.
Fracture porosity
Fractures may occur in all lithologies, but they are particularly common in carbonate
rocks because the early cementation of many carbonates causes them to behave in a
brittle manner for most of their history. Fractures are also very important because the
permeability of many carbonates is low, even for relatively high porosities, so fluid
flow may depend on the presence of fractures. Also, many of the porosity types
described above (e.g. vuggy, fenestrate and intragranular porosity) commonly consist
of large but unconnected pores, which will only form an effective flow network if
fractures are present.
60

Sedimentology

3.11 DIAGENESIS
3.11.1 Definition
Diagenesis consists of the range of physical and chemical processes and changes
which turn a sediment into a rock. It may begin immediately after deposition and
continues during burial. The majority of diagenetic changes will tend to reduce the
porosity and permeability of a sediment, but some, such as dissolution, may increase
the porosity and/or permeability.
In the following section, we will be concerned only with those aspects of diagenesis
which impact on porosity and permeability and therefore on reservoir performance.

3.11.2 Clastic Diagenesis


As sediments are buried, they undergo compaction due to the weight of the overlying
sediment. Initial compaction may occur by a change in packing of the grains, but once
an efficient packing pattern has been established, further compaction can occur only
by grain deformation. The high pressure at grain contacts may cause local dissolution,
leading progressively to straight, concavo-convex and sutured grain contacts (Figure
3.4). Grains may also deform brittly, by breaking, or may deform ductily. Certain clast
types, such as mudstone intraclasts and lithic clasts, are more liable to deform ductily,
so sediments containing high proportions of these clasts are likely to lose a higher
proportion of their porosity by ductile deformation than, for example quartz-rich
sands.
Clastic sediments may be cemented by a wide range of mineral cements and clays. The
type and distribution of the cements will depend partly on the burial history of the
sediment. The majority of cements are precipitated from pore waters, so the type of
minerals precipitated, and their order, will reflect the changes in water chemistry with
time. At shallow burial depths, the pore waters will contain micro-organisms such as
aerobic and anaerobic bacteria which will influence the water chemistry. During
burial, the sediment may be surrounded by the same water for a long period of time,
in an essentially closed system, or circulation of water may constantly replenish the
water or replace the pore water with water of a different chemistry. A full study of the
geochemistry of diagenesis is beyond the scope of this chapter, but the following
paragraphs are a brief introduction to some of the more important points.
Immediately after deposition, the pore spaces in most sediments will be occupied by
either fresh or marine waters, depending on the environment of deposition. Fluvial or
aeolian sediments may be deposited above the water table, so will have little or no
water round the grains. However, sediment in this vadose zone may become wet
following periods of rainfall or may be affected by water rising from the water table
due to capillary action. Water may occur as thin rims round the grains, as menisci
between grains or as pendulous drops hanging from grains. As cements are only
precipitated were water is present, vadose zone cements may have a distinctive form,
either rimming grains, bridging the gap between grains at pore throats, or being thicker
on the lower side of grains. Continued burial of sediments in the vadose zone will
eventually move them below the water table.

Department of Petroleum Engineering, Heriot-Watt University

61

Cements precipitated below the water table can grow anywhere in the pore spaces, but
it is still common for cement crystals to be nucleated on sediment grains. Silica cement
may nucleate on quartz grains, where it grows with the same crystal orientation as the
host grain and is said to be in 'optical continuity'. These quartz overgrowths can often
be differentiated from the host grain by a thin dusty rim marking the original margin
of the grain. Feldspar overgrowths may also occur, though they are rarer than quartz
overgrowths.
Many sandstones are cemented by carbonate minerals, including calcite, dolomite and
siderite. The cements may rim the grains or may occur as small crystals in the
intergranular pores. Calcite and dolomite often form cements with crystals which are
significantly larger than the grains and so enclose a number of grains. These
poikilotopic cements are sometimes visible even in hand specimen, as the large
crystals sparkle with reflected light.
In addition to mineral cements, clays minerals may also be precipitated in the pore
spaces. These authigenic clays may have a variety shapes and relationships to the host
sediment. Most clay minerals form plate-like crystals, and several minerals, including
chlorite and illite may grow on grains as concentric or radial arrangements of plates.
The radial arrangement is more common. In addition to its platy fabric, illite also forms
more elongate crystals, and this fibrous or hairy illite commonly grows at the
margins of more platy illites. Other clays, such as kaolinite, tend to form denser
clusters of crystals, arranged like the pages of a book, in intergranular pores.
In the case of sands with high proportions of clay, detrital clays may increase in size
during burial by the addition of authigenic overgrowths or may be replaced by other
clay minerals. In general, the clay-rich matrix tends to become better crystallised with
increasing depth of burial. As the pressure and temperature increase during burial the
sediment may pass from the stability field for one clay mineral into that for another.
For example, smectite is generally replaced by illite with increasing depth, with the
most pronounced change occurring at depths of between 2.5k m and 3.5km. The illite
crystallinity increases with burial and can be used to give an estimate of the maximum
burial depth of a rock.
Thin section petrography and scanning electron microscopy are used to examine the
authigenic mineral fabrics and the relationship between different cements and clays.
For example, it maybe possible identify cements growing over other cement minerals
or clays. By a detailed study of the relationship of the different cements and authigenic
clays, it is possible to establish the order of diagenetic events. This diagenetic history
enables important information to be gained about both the burial history and the
evolution of the pore waters.

3.11.3 Carbonate Diagenesis


As mentioned above, carbonate sediments are prone to early diagenesis and may
undergo many phases of cementation during burial. Early cements, particularly those
precipitated in the vadose zone, may form rims round the grains. Both calcite and
aragonite may grow in this way and may show many of the characteristic features of
vadose cements described in Section 3.11.2.

62

Sedimentology

Many bioclasts and the majority of micrite are initially composed of aragonite, which
is an unstable mineral under normal burial conditions. If a clast is totally dissolved,
its original form may be preserved as a mould, either by the surrounding matrix or by
an early rim cement. The clast-shaped pore then behaves like any other pore and may
be later filled by cement. Alternatively, unstable grains may be gradually replaced by
another mineral, a process known as neomorphism. In this case, a ghost of the original
texture may be preserved.
Pore-filling calcite cement commonly develops a distinctive fabric. The initial pore
lining consists of a large number of small, blade-like crystals. As they continue to
grow into the pores, some crystals grow over their neighbours, reducing the number
of active crystals. The cement crystals therefore tend to increase in size towards the
centre of the pore.
Diagenetic dolomite may occur either as rhombic crystals in the pore spaces or as more
extensive pore-filling or poikilotopic cement. It may also replace the existing clasts
and matrix, producing a sedimentary rock composed entirely of dolomite. In some
dolomites, the original fabric of the sediment is entirely lost, but in others the faint
relict structure is visible.
In any dissolution or recrystallisation, the least stable minerals or grain types tend to
be dissolved or replaced first. It is fairly common for bioclasts of a certain type to be
entirely replaced by a diagenetic mineral, whilst others are either unaltered or are
replaced by a different mineral.
Although diagenesis of carbonates generally involves a smaller number of minerals
than clastic diagenesis, it often involves many phases of dissolution, replacement and
cementation. Many carbonate minerals can contain variable proportions of certain
ions. For example, the proportion of iron in ferroan calcite or ferroan dolomite may
vary, and it is often possible to identify distinctive growth zones of subtly different
chemistry within a single crystal. Other ions present only in trace amounts may also
vary in their proportions, and special techniques may be used to identify more subtle
zonation.
As in the case of clastic diagenesis, it is possible, by using a variety of techniques, to
identify diagenetic sequences in carbonates and to postulate a burial history and pore
water evolution. For example extensive dissolution may indicate the presence of
meteoric water, and so may suggest a period of uplift.

3.11.4 Impact of Diagenesis on Porosity and Permeability


The majority of diagenetic processes lead to a reduction of porosity and permeability,
as the sediment is compacted and authigenic minerals fill the pore spaces. It should
be noted, however, that the timing, mineralogy and fabric of cements and authigenic
clays can produce very different results.
Even quite low volumes of early cement may reduce the degree of compaction of a
sediment during later burial and may therefore increase its porosity at a given depth.
The shape and position of mineral cements may also have a significant impact on
permeability. For example, a certain volume of meniscus cements occurring near pore
Department of Petroleum Engineering, Heriot-Watt University

63

throats will have a greater impact on permeability than the same volume of cement
spread more evenly throughout the rock. In the same way, clay minerals have a
significantly greater impact on permeability than mineral cements, but the degree of
their influence will also depend on their fabric and location. Radial platy clays will
restrict flow more than tangential clays, as they extend further into the pore spaces. For
a given volume of clay, fibrous illite will have the greatest effect, particularly if it
occurs in pore throats. In contrast, kaolinite is often more patchilly developed, and
may have little or no effect on the majority of pores.
Dissolution of grains (for example feldspars in clastic rocks and aragonitic bioclasts
in carbonates) will produce secondary porosity. However, this increase in porosity
may have little impact on permeability if the dissolution pores are not well connected
to the existing pore system. Dolomitisation may also lead to an increase in porosity
and permeability, due to the reduction in volume it involves.
Before leaving the subject of diagenesis, it is relevant to look briefly at the impact of
man on the rock. Poor drilling or production methods may cause physical or chemical
changes in a rock which may be deleterious to its reservoir performance. For example,
production at too high a rate, particularly near the well bore, may cause clay minerals
to move, leading to the clogging of pore throats. Also, chemical techniques such as
acidisation, used to improve the permeability, may have the reverse effect if the acids
alter the clay minerals or cause them to move. It is vitally important, therefore, before
undertaking any programme of acidisation, to understand fully the diagenetic nature
of the rock, so that the reaction of the fluids introduced into the reservoir can be
estimated.

64

You might also like