Download as pdf or txt
Download as pdf or txt
You are on page 1of 79

THE OPEN UNIVERSITY

PETROLEUM GEOLOGY FOR PRODUCTION STAFF

Manual 1 An Introduction to Some Basic Geological Concepts


Manual 2 Reservoir Architecture I: depositional environments

1989 The Open University and Shell Internationale Petroleum Maatschappij BV

Produced for the Training Division of Shell Internationale Petroleum Maatschappij by the
Open University contract Training Unit

Production Team
Andrew Bell
Brent Cheshire
Kevin Gowans
Frank Jahn
Chris Wilson

(Author) consultant to the Open University


(Author and OU-Shell liaison) Shell Training
(Production Manager) contract Training Unit, Open University
(Author and OU-Shell liaison) Shell Training
(Author and Project Co-ordinator) Earth Sciences, Open University

ii

Manual 1 An Introduction to Some Basic Geological Concepts


Acknowledgements
The bulk of this manual has been adapted from Open University teaching texts. Sections
2, 3, 4, 6 and 7 are based on material from S101 Science: a Foundation Course, and
Section 5 from S238: The Earth's Physical resources. Section 7.6 is from S338:
Sedimentary Processes and Basin Analysis.
The following diagrams from sources other than the Open University or Shell have been
used from:
1.2
1.8
1.10
1.24

Basic Concepts of Petroleum Geology, IHRDC GL101


Trustees, British Museum (Natural History)
J. Lameyre (1975), Roches et Minraux - Les Materiaux, Editions Doin
Modified from Brown (1986), Jurrasic, in Glennie K (ed) Introduction to the
Petroleum Geology of the North Sea, Blackwell Scientific Publications
1.27
Nature, Vol 277,1979, MacMillan Journals Ltd.
1.29
D.H Tarling (ed), 1981, Economic Geology and Geotechnics, Blackwell
Scientific
1.30
J.M. Hunt, Geochemistry of Petroleum, American Association of Petroleum
Geologists
1.31
B.P. Tissot and D.H. Welte (1978), Petroleum Formation and Occurrence,
Springer-Verlag Inc
1.50
I. Gass et al (1972), Understanding the Earth, 2nd Edition, Artemis Press
1.56
B.C. Heezen, Deep Sea Floor, in S. Runcorn (ed), 1962, Continental Drift,
Academic Press
1.58a
Geological Museum (1973), The Story of the Earth
1.59
G. Balliot, Geology of the Continental Margins (trans. A. Scarth), Longman
1.60b&c Geological Museum (1973), The Story of the Earth
1.65
D. McKenzie (1978), "Some remarks.. .", Earth Plan. Sci. Lett, 40, Elsevier

iii

Manual 2 Reservoir Architecture: Depositional environments


Acknowledgements
Sections 1-7 of this manual are based on Block 4: Surface Processes of the Open
University Course S236 Geology. Section 8 is based on course notes prepared by the
Open University in 1984 to accompany a lecture and practical course for industry. The
original notes were written by Dr. Maurice Tucker, and subsequently published in P. J.
Brenchley and B. P. J. Williams (eds): Sedimentology: recent developments and applied
aspects, Special Publication of the Geological Society, 18, pp147-172, 1985.
The following diagrams from sources other than the Open University or Shell have been
used from:
2.5
2.13
2.14
2.21
2.22
2.32
2.46
2.47
2.51
2.52b
2.54
2.55a
2.56
2.59
2.62
2.63
2.64
2.66
2.68
2.70
2.71

F. Sawkins et al (1978), The Evolving Earth : A Text in Physical Geology,


Cromwell Collins and MacMillan Ltd
R.G. Walker, Shelf and Shallow Marine Sands, in R. Walker(Ed), 1984, Facies
Models, 2nd Edition, Geoscience Canada, Ainsworth Press
J.D. Collinson, Deserts, in H. Reading (Ed), 1986, Sedimentary Environments
and Facies, 2nd Edition, Blackwell Scientific
M. Tucker (1981), Sedimentary Petrology: an Introduction, Blackwell Scientific
redrawn from F.J.I.Pettijohn et al (1973), Sand and Sandstone, Springer-Verlag
M.W. Longman (1980), Carbonate digenetic textures from near surface digenetic
environments, Am. Assoc. Petrol. Geol, Bull 64
B.K. Lovell, H.J. Braackman and K.W. Rutter (1988), Oil bearing sediments of
Gondwana glaciation in Oman, Am. Assoc. Petrol. Geol, Bull 72
Adapted from H. Reading (Ed), 1980, Sedimentary Environments and Facies, 1st
Edition, Blackwell Scientific
M. Tucker (1981), Sedimentary Petrology: an Introduction, Blackwell Scientific
Adapted from H. Reading (Ed), 1980, Sedimentary Environments and Facies, 1st
Edition, Blackwell Scientific
A. Miall, Deltas, in R. Walker (Ed), 1984, Facies Models, 2nd Edition,
Geoscience Canada, Ainsworth Press
T. Elliot, Deltas, in H. Reading (Ed), 1986, Sedimentary Environments and
Facies, 2nd Edition, Blackwell Scientific
T. Elliot, Siliciclastic shorelines, in H. Reading (Ed), 1986, Sedimentary
Environments and Facies, 2nd Edition, Blackwell Scientific
T. F. Moslow and R.S. Tye (1985), Recognition and characterisation of
Holocene tidal inlet sequences, Marine Geology, 63, Elsevier Science
R.G. Walker, Shelf and Shallow Marine Sands, in R. Walker (Ed), 1984, Facies
Models, 2nd Edition, Geoscience Canada, Ainsworth Press
R.G. Walker, Shelf and Shallow Marine Sands, in R. Walker (Ed), 1984, Facies
Models, 2nd Edition, Geoscience Canada, Ainsworth Press
R.G. Walker, Shelf and Shallow Marine Sands, in R. Walker (Ed), 1984, Facies
Models, 2nd Edition, Geoscience Canada, Ainsworth Press
B.W. Sellwood, Shallow marine Carbonate Environments, in H.Reading (Ed),
1986, Sedimentary Environments and Facies, 2nd Edition, Blackwell Scientific
B.W. Sellwood, Shallow marine Carbonate Environments, in H.Reading (Ed),
1986, Sedimentary Environments and Facies, 2nd Edition, Blackwell Scientific
R. Till, Arid shoreline and evaporites, in H. Reading (Ed), 1980, Sedimentary
Environments and Facies, 1st Edition, Blackwell Scientific
Modified from E.A. Shinn (1983), Tidal flat environment, in Scholle et al (Eds),
Carbonate depositional environments, Am. Assoc. Petrol. Geol, Mem 3, Tulsa
iv

MANUAL 1
An Introduction to Some Basic Geological Concepts
Contents

SECTION 1 OVERVIEW .................................................................................................1


1.1 Introduction............................................................................................................. 1
1.2 The reservoir units .................................................................................................. 1
1.3 Seals........................................................................................................................ 3
1.4 Traps ....................................................................................................................... 3
1.5 Sources of hydrocarbons......................................................................................... 3
1.6 Maturation of source rocks ..................................................................................... 4
1.7 Hydrocarbon migration and entrapment ................................................................. 4
1.8 Geological time and rock units ............................................................................... 4
1.9 Hydrocarbon plays and sedimentary basins.............................................................6
SECTION 2 ROCKS AND THE ROCK CYCLE.............................................................8
2.1 Minerals .................................................................................................................. 8
2.2 Crystalline and fragmental rock textures .............................................................. 12
2.3 Igneous, metamorphic and sedimentary rocks...................................................... 13
2.4 Geological cycles...................................................................................................14
SECTION 3 SEDIMENTARY PROCESSES .................................................................17
3.1 Introduction........................................................................................................... 17
3.2 Physical and chemical weathering of rocks .......................................................... 17
3.3 The products of chemical weathering ................................................................... 18
3.4 The transport of weathered material ..................................................................... 20
3.5 The transport and erosion of sediment by wind and ice ....................................... 21
3.6 The transport and deposition of the soluble products of chemical weathering......22
SECTION 4 GEOLOGICAL TIME ................................................................................24
4.1 Introduction to geological time............................................................................. 24
4.2 Putting things in order .......................................................................................... 26
4.3 Evolution .............................................................................................................. 28
4.4 How the stratigraphic column was developed ...................................................... 30
4.5 Calibrating the stratigraphic column......................................................................34

SECTION 5 THE NATURE, ORIGIN, GENERATION, MIGRATION AND


ENTRAPMENT OF PETROLEUM................................................................................38
5.1 Nature of petroleum .............................................................................................. 38
5.2 Origin of petroleum .............................................................................................. 39
5.3 Generation of petroleum ....................................................................................... 40
5.4 Migration of petroleum......................................................................................... 43
5.5 Traps ..................................................................................................................... 43
5.5.1 Structural traps............................................................................................. 44
5.5.2 Stratigraphic traps........................................................................................ 46
5.5.3 Combination traps ........................................................................................47
5.6 Natural gas ............................................................................................................ 47
5.7 Solid petroleum and oil shale ................................................................................48
SECTION 6 THE EARTH'S STRUCTURE AND SURFACE FEATURES..................49
6.1 A simple model of the structure of the Earth........................................................ 49
6.2 Layers within the crust and mantle ....................................................................... 50
6.3 Elevations of the surfaces of continental and oceanic crust ................................. 52
6.4 Earth patterns .........................................................................................................53
SECTION 7 PLATE TECTONICS .................................................................................58
7.1 The model ............................................................................................................. 58
7.2 Continental fit and continental drift...................................................................... 59
7.3 Palaeomagnetism .................................................................................................. 61
7.4 Sea floor spreading ............................................................................................... 62
7.5 The model - more details ...................................................................................... 64
7.5.1 Constructive plate margins .......................................................................... 64
7.5.2 Destructive plate margins ............................................................................ 69
7.6 Basin formation and classification........................................................................ 70
7.6.1 Thermal processes ....................................................................................... 70
7.6.2 Stretching of the lithosphere........................................................................ 72

vi

SECTION 1
OVERVIEW
Study Comment
The purpose of this introductory chapter is to introduce some of the key concepts
necessary to understand how the geological sciences are used in the search for, and
production of hydrocarbons. The following topics will be introduced:
the three types of rocks and the rock cycle;
sedimentary rocks and processes;
geological time;
hydrocarbons;
the Earth's structure and surface features;
plate tectonics and sedimentary basins.
In order to give an insight into why petroleum engineers need to understand these
aspects of geology, this introductory section explores some of the features of two
North Sea hydrocarbon reservoirs. Many geological terms are used in this section,
which will be explained in more detail in later sections of the manual.
1.1 Introduction

Figure 1.1 shows the location of two North Sea reservoirs. In both cases the nature of the
reservoir unit on a microscopic scale is shown, together with its geometry in cross
section, and its setting within the top few kilometres of the Earth's crust. The labelling of
Figure 1.1 involves a considerable amount of geological jargon which is the subject of
this and succeeding manuals. We will examine it at successively larger scales.
1.2 The reservoir units

Both reservoir units consist of sand grains composed predominantly of the mineral
quartz, whose composition is pure silica (SiO2). The hydrocarbons are situated in the
spaces between the grains - that is the porosity. Unless the spaces between the grains are
connected together in some way, hydrocarbons cannot migrate into the reservoir, and of
course they could not be extracted either. This connectedness is termed permeability.
Virtually all hydrocarbon reservoir units are sedimentary rocks, which are composed of
grains derived from pre-existing rocks or as particles formed by organic or chemical
processes at the Earth's surface. They are the products of the recycling of rock debris
produced by processes of weathering, transport and deposition. After deposition, they
were buried by later episodes of sedimentation. During burial, the porosity and
permeability of rocks was modified, as the grains became packed more tightly together,
and new minerals were precipitated in the pore spaces.

FIGURE 1.1 The geologic


features and settings of a North
Sea oil field and a gas field at
different scales

Three sections through the fields are given at different scales. The sketches at a microscopic scale show the sand
grains within the reservoir and the pore spaces between them. The structural cross-section of the fields shows the
nature of the trap in which hydrocarbons accumulated. The regional structural cross-sections show the setting of the
fields within the Northern and Southern sub-basins. The terminology used in this figure is briefly introduced in
section 1, and explained in more depth in later sections. Manuals 2 and 3 examine the formation and subsequent
deformation of reservoir sands in greater depth.

On a larger scale, the geometry of the reservoir unit is important both in terms of
estimating what reserves are in place, and determining how to extract them. The
geometry of a reservoir will be determined by the kind of environment it was deposited
in, and whether it was deformed subsequently by Earth movements. In the two examples
illustrated, the Brent reservoir sands were deposited by water in a delta, and the
Indefatigable reservoir sands largely by wind in a desert. Knowledge of the nature of the
depositional environment of a reservoir sands helps to predict not only what their overall
geometry might be, but also the extent to which their porosity and permeability might
vary horizontally and vertically.
1.3 Seals

Unless reservoir units are capped by an impermeable rock unit, or seal, hydrocarbons
would not be retained in them. Again, knowledge of the depositional environment of seal
units aids the prediction of their continuity. In the case of the Brent reservoir units (there
are two principle sand units), the seal consists of impermeable shales (which were muddy
sediments when they were deposited). The Indefatigable reservoir sands are overlain by
minerals formed by the evaporation of seawater (anhydrite, CaSO4; halite, NaCl), which
are also impermeable.
1.4 Traps

Reservoir units capped by seals will not trap hydrocarbons unless a three dimensional
container' exists that is sealed in order to stop the relatively light hydrocarbons rising to
the surface. In the case of the two North Sea examples, the structures of the traps were
largely controlled by faulting. The main classes of traps are shown in Figure 1.2, which
also indicates the proportion of world oil occurring in each type. The nature of traps is
explained further in Section 5.5.
FIGURE 1.2 Major
types of oil traps and
percentage of world's
petroleum occurrence for
each. Most of the
reserves of the Middle
East occur in anticlinal
traps, which largely
accounts for the fact that
three quarters of world
oil occurs in this type of
trap.

1.5 Sources of hydrocarbons

Modern seismic techniques enable petroleum geologists to identify structures,


which could be potential traps in the subsurface relatively easily. But even if
porous, these are only likely to contain oil or gas if there are source rocks nearby.
Hydrocarbon source rocks are sediments containing large amounts of organic
matter, and so knowledge of the depositional environments of rock successions
enables the vertical and horizontal distribution of the source units to be estimated,

and hence their volume. Calculations can then be made concerning the amount of oil
and/or gas that might have been generated. In the case of the Indefatigable field, the
source rocks are coals lying beneath the field. In the Brent field, they are organic rich
shales, which not only overlie the reservoirs, but are also situated at lower intervals
across the faults (this situation of a source rock overlying the associated productive
reservoir is unusual).
1.6 Maturation of source rocks

Without being sure of the presence of the trinity of source rock, reservoir unit and seal,
there is no point embarking on a major exploration programme. Once satisfied that the
trinity does exist, the explorationist needs to be reasonably satisfied that the source rock
has generated oil. Source rocks need to be cooked to over 100C for significant amounts
of oil to be yielded. Temperature increases with depth within the Earth's crust (the
average rate - or geothermal gradient - is 30C km-1). This means that in most areas of
the world, the explorationist is looking for source rocks to have been buried to depths of
at least three kilometres for a significant period of geological time.
Calculations concerning the burial history of the Kimmeridge Clay source rock in the
vicinity of the Brent field show that it reached oil-generating depths relatively late (in
geological terms) during the history of the area - perhaps only about 30-40 million years
ago when Tertiary rocks were deposited. This was long after the trap was formed by the
faulting and tilting of the reservoir units, and the structures being draped by the
Kimmeridge Clay.
In the southern North Sea, peak gas generation occurred from the coals beneath the
Indefatigable reservoir unit at depths of around four kilometres; such depths were reached
150-100 million years ago as Jurassic and Cretaceous rocks were deposited.
1.7 Hydrocarbon migration and entrapment

Unless pathways exist for hydrocarbons to migrate out of the source rock after generation
and into traps, no economic reserves would exist. Again, knowledge of depositional
environments and the nature of deformation of rocks by Earth movements are applied by
explorationists to evaluate whether migration was likely.
The final piece of the jigsaw that has to be put in place is determining whether the trap
existed at the time hydrocarbons migrated out of the source rock. If a trap structure
formed later than migration, the hydrocarbons would be able to escape to the surface. For
this reason, it is important to know what the sequence of events was that contributed to
geological features of the kind illustrated on Figure 1.1. This involves working out the
geological history of the area being explored. Thus not only do petroleum geologists have
to produce three-dimensional models of reservoirs and their environs, they also have to
incorporate a fourth dimension-time.
1.8 Geological time and rock units

For reasons, which will become apparent in Section 4 of this manual, geologists do not
routinely give the ages in millions of years of rocks encountered during surface
mapping or drilling. This is because although it is usually possible to work out the

relative ages of different rocks from their geometrical relationship and content of fossils,
it is not possible to determine age in quantitative terms on a routine basis- and rarely
possible to do so accurately for sedimentary rocks. In the last century geologists agreed
on how to divide up geological time based on the changing assemblages of fossils in
rocks even before radiometric dating techniques were devised. This division is still used
today, and is shown on Figure 1.3(a). If cores or drill cuttings yield diagnostic fossils, it is
possible to assign an age to the sample. You will see that terms such as Cretaceous' and
Carboniferous' are used on Figure 1.1; these denote periods of geological time.
Local names are used locally to label individual rock units, and examples for the North
Sea shown in Figure 1.1 include the Brent Group, Coal Measures, Zechstein and Chalk.
Stratigraphy - the study of rock strata - is an essential ingredient in determining the
geological history of the Earth's crust. It involves establishing locally the sequence of
different rock types to produce a rock-stratigraphic or lithostratigraphic succession. The
ages of rocks within this succession can then be determined through their fossil content,
or by radiometric dating, and the lithostratigraphic units then assigned to the correct part
of the stratigraphic column shown to the left of Figure 1.3.
Figure 1.3 shows that hydrocarbons are not distributed evenly through the stratigraphic
succession of the North Sea because the distribution of source, reservoir and seal rocks in
space and time was not random. The distribution of hydrocarbon reserves on a world
scale is rather different, as shown on Figure 1.27 (page 40). This part of the Figure on the
right side of 1.27 shows the distribution of hydrocarbon fields by geologic age of the
reservoir unit, so it is showing the volumetric distribution of reserves. However, it shows
a general trend of more oil in younger rocks.
FIGURE 1.3
Distribution of
hydrocarbons through
geologic time. Current
estimates of UK
recoverable oil and gas
reserves.

The two cross sections across the North Sea shown in Figure 1.1 show that the rock
successions and structures present in the north and south of the North Sea are different.
There is no point going into too much detail at this stage, but the key points to note are as
follows:
Southern North Sea:
Source rock for gas: coals beneath Permian reservoir sands.
Seal: Permian evaporite mineral unit (the Zechstein), which has flowed to deform
overlying rocks.
Reservoir: desert sands of the Rotligende.
Trap structure: due to faulting during the late Jurassic and early Cretaceous.
Timing of gas generation: 100-150 million years ago during the late Jurassic and early
Cretaceous.
Northern North Sea:
Source rock for oil: organic rich shales of the Kimmeridge Clay.
Seal: various shale units.
Reservoir: delta and river sands of the Brent and Statfjiord Formations.
Traps: due to faulting that preceded deposition of the seal.
Timing of oil generation: 40-60 million years ago.
1.9 Hydrocarbon plays and sedimentary basins

The previous discussion implicitly introduced the concept of a hydrocarbon play, which
is a set of geological circumstances necessary for the accumulation of oil and gas. To
recapitulate, these are:
1) the trinity of source, reservoir and seal rocks;
2) the existence of a suitable trap structure;
3) burial of the source rock to mature (i.e. cook) the source rock;
4) a migration route for hydrocarbons to move from the source to the trap;
5) timing: maturation of the source rock must post-date the formation of the trap.
As knowledge of a play is improved by exploration, quantitative estimates of
hydrocarbons likely to have been trapped can be made in order to evaluate the likely
success of exploration drilling.

Examination of the map of the North Sea in Figure 1.1 shows that gas fields are
concentrated in an east-west zone across the southern part, whereas the oil fields in the
north form a narrower north-south zone. The geological history of these two fairways
must have resulted in a set of circumstances that combined to produce numerous
hydrocarbon plays. Likewise, the entire North Sea area is a large area of the Earth's crust
in which circumstances favorable to hydrocarbon generation and entrapment have
occurred, and may be termed a hydrocarbon province.
The most significant hydrocarbon provinces occur in the Earth's major sedimentary
basins (see Figure 1.4), which are regions of the crust, which have subsided to
accumulate thick sequences of sedimentary rocks. Oil and gas also occurs in some
present day and ancient mountain or orogenic belts, but it is virtually never found in the
cratons or shield areas shown on Figure 1.4, which consist of rocks older than one
thousand million years. The development of sedimentary basins is inextricably linked to
past movements of huge slabs, or plates of the Earth's crust.

FIGURE 1.4 Distribution of


some of the world's
sedimentary basins.
Orogenic fold belts are
areas of the crust that have
been intensively folded
during past compressional
episodes. Shield regions are
static areas of crust
exposing rocks much older
than 1000 million years;
they are often referred to as
cratons.

The rest of this manual is devoted to introducing, in more depth, the key geological
concepts necessary to understand the geological development of reservoirs.

SECTION 2
ROCKS AND THE ROCK CYCLE
Study Comment

The purpose of this section is to illustrate the nature of the three basic rock types
-namely igneous, metamorphic and sedimentary- and in particular to show how
sedimentary rocks are derived by recycling pre-existing rocks as part of the rock
cycle. The main characteristics of certain igneous rocks will also be outlined to aid
understanding of Sections 4 and 5 dealing with Geological Time' and The Earth's
structure and major surface features'. As rocks are aggregates of minerals, we begin
by summarising some of the key features of the latter.
2.1 Minerals

A few minerals are simply single elements, occurring in nature in the uncombined state.
Examples of such native elements are gold (Au), and carbon (C), which occurs as the
minerals diamond and graphite. Most minerals, however, are chemical compounds, and
they can be grouped into three main categories.
1

Minerals such as galena (PbS), pyrite (FeS2) and hematite (Fe2O3) have relatively
simple internal structures, such as the one shown in Figure 1.9. These structures
may be thought of as being made up principally of separate positive ions
(cations) such as Pb2+, Fe2+ and Fe3+, and negative ions (anions) such as S2-. This
group includes many minerals, which form cubic shaped crystals related to their
atomic structure as shown in Figure 1.5. A metallic lustre characterises this group
of minerals.
FIGURE 1.5 Atomic
structure of galena (PbS).
(a) To true relative scale,
showing the relative sizes of
Pb and S atoms and the way
they are packed together in
the structure. (b) An
expanded view of the
structure to show the cubic
arrangement of the planes of
Pb and S atoms.

Minerals such as calcite (CaCO3) and gypsum (CaSO4.2H2O) have slightly more
complex internal atomic arrangements, such as the one illustrated in Figure 1.6.
Here the minerals can be thought of as consisting of simple cations (Ca2+ in the
two examples) and of composite negative ions; these are complex anions, CO32and SO42- in the examples. (The H2O in gypsum is not an essential component of
the structure.) Nearly all the minerals in this group have a glassy lustre and the
great majority have rather low hardness - they can be scratched easily.
8

FIGURE 1.6
(a) The atomic
structure of the
mineral calcite
(calcium carbonate,
CaCO3). (b-d) Some
common crystal
habits of calcite.

The silicate minerals are by far the largest and most important group of minerals
and form the majority of rocks. In these minerals also there are simple cations,
such as Ca2+, Fe2+, Mg2+, K+, and so on, but the anions are more complex. The
basic building block of all silicate minerals is the silicate tetrahedron. It consists
of four oxygen atoms arranged symmetrically around a central silicon atom
(Figure 1.7). This basic unit has a formula of SiO44-, and is thus similar to other
complex anions (e.g. CO32) and can behave in the same way, combining with
cations (e.g. Fe2+ and M2+) to form minerals such as olivine (Mg,Fe)2SiO4.
FIGURE 1.7
(a) The tetrahedron,
the basic shape of the
SiO44- unit, to show
how the silicon and
oxygen atoms (ions)
are linked. (c) The
SiO44- tetrahedral
unit with atoms of
oxygen and silicon
shown to true relative
scale.

What sets the silicate tetrahedron apart from the other kinds of anion, however, is its
ability to polymerize. Silicate tetrahedra can polymerize in various ways (Figure 1.8) to
form chains, bands, sheets and framework structures by a process known as
oxygen-sharing. The giant composite anions of these structures are joined to each other
by a variety of simple cations to form the silicate minerals. The sole exception is quartz,
SiO2, in which all the oxygens are shared, to make a framework structure such as the one
shown in Figure 1.9. This structure is extremely resistant to mechanical breakdown, and
the mineral is also chemically very stable, so that it easily survives the processes of
weathering, transport and deposition.
9

FIGURE 1.8
Polymerization of SiO44units to form (a) chains,
(b) double chains or
bands and (c) sheets.
The oxygen and silicon
atoms (ions) are shown
true to relative scale in
(b) and (c), but the
silicon atoms cannot be
seen in (a).

FIGURE 1.9 (a) Quartz


(SiO2) atomic structure
consisting of silicon atoms
(black) and oxygen atoms
(brown) forming a
framework in which there
are no obvious planes of
weakness. (b) View of
quartz crystal; these may
range in size from 0.1 mm
across to tens of
centimetres.

The silicates encompass a wide and rich variety, from fibrous asbestos to
sheet-like mica, from soft talc (soapstone or talcum powder) to hard quartz, from
deep blue lapis lazuli (ultramarine) to deep red garnet. However, from the point of
view of petroleum geology, we need consider only the main groups of silicates; a
systematic summary of their principal features is provided in Figure 1.10. You
10

should use Figure 1.10 for reference; you do not need to learn the details, but it is
important that you perceive the control which atomic structure exerts upon the shape and
cleavage patterns (crystals tend to break relatively early along cleavage planes) of the
different silicate mineral groups shown in Figure 1.10. The mineral names listed in the
left-hand column of Figure 1.10 will be used from time to time in later sections of this
manual.

FIGURE 1.10 The


relationship between
the main types of
silicate structures and
the physical properties
of the corresponding
materials.
Note in particular the
influence of structure
on cleavage patterns.
You do not need to
learn the details of this
Figure.

11

2.2 Crystalline and fragmental rock textures

The texture of a rock describes the physical relationship between its constituent mineral
grains. Virtually all rocks are either crystalline or fragmental. Although extremely
useful, these terms are not altogether precise. It can be argued, for instance, that, since
almost all rocks (coal and natural glasses are notable exceptions) consist of minerals
which themselves are crystalline, then all rocks must be crystalline. There are difficulties
also with some kinds of limestone, which are major repositories of fossil remains of
ancient marine life forms. However, this very broad classification is used as a convenient
working introduction to rock textures.
1.

In crystalline rocks, the minerals have grown together to form aggregates of


crystals (Figure 1.11(a)).

2.

In fragmental rocks, minerals and rock particles have been transported and
deposited together, by water, wind or ice, and may therefore be more or less
rounded (Figure 1.11(b)).

FIGURE 1.11
Examples of textures
of (a) crystalline and
(b) fragmental rocks,
and the fracture
patterns displayed by
them.

Crystalline rocks tend to be harder and more compact than fragmental rocks. This is not
invariably so, but is often useful in preliminary examination or rocks, because rocks will
break along surfaces of internal weakness. In crystalline rocks, the minerals commonly
form strongly interlocking aggregates so that fracture usually occurs within the crystals,
along the planes of weakness in the structure called cleavage planes (Figure 1.11(a)).
Fragmental rocks fracture along boundaries between the more or less rounded grains
(Figure 1.11(b)). The result is that broken surfaces of many crystalline rocks show
numerous flat surfaces, which reflect and glint as they catch the light. This glinting effect
is generally less typical of fragmental rocks but there can be no hard or fast rules. Many
fragmental rocks fracture mainly between grains, but if the grains are very tightly held
together, then the rock will break mainly across the grains.

12

2.3 Igneous, metamorphic and sedimentary rocks

Igneous rocks contain minerals crystallised out from melts or magmas; most, but not all,
show crystalline textures.
Metamorphic rocks are formed by the alteration of pre-existing rocks due to heat and/or
pressure; most, but not all, show crystalline textures.
Sedimentary rocks are formed largely by the weathering, transport and deposition of
debris derived from pre-existing rocks, and so mostly show fragmental textures.
The fundamental control on the mineral composition of any rock is the nature of the
material from which it has formed. For an igneous rock, this is the magma, a liquid
mixture of various chemical elements. As the magma begins to cool, its composition
determines which minerals will crystallize out first. As progressively lower temperatures
are reached, other minerals will crystallize, while those minerals, which crystallized at the
higher temperatures, may cease to form. The temperature ranges over which the essential
minerals of igneous rocks can form during cooling of a magma are shown in Figure 1.12.
You can see that the ranges for several minerals do not overlap at all (muscovite and
pyroxene, or quartz and olivine, for example). Where there is a gap between the
temperature ranges of two minerals, those minerals are only exceptionally found together
in igneous rocks. Thus, olivine is never found with quartz and very rarely with alkali
feldspar. In contrast, minerals whose ranges of crystallization temperatures overlap are
commonly found together, so that olivine, pyroxene and Ca-rich plagioclase feldspar
form a predictable mineral association.

FIGURE 1.12 Ranges


of crystallization
temperatures for
common minerals of
igneous rocks and the
mineral associations of
gabbro, diorite and
granite.

13

Crystallization of minerals at different temperatures may lead to fractionation of an


original magma to produce a variety of rock types. Thus, crystals of olivine, forming at
high temperatures, sink to the bottom of magma chambers to form the rock peridotite,
which is considered to be similar in chemical composition to the Earth's mantle (see
Figure 1.39). At slightly lower temperatures, pyroxene and plagioclase feldspar will
crystallize from magmas. If the magma cools slowly, the crystals will grow relatively
large, resulting in the coarse grained rock termed gabbro, but rapid cooling at the Earth's
surface results in fine grained basalt being formed. At lower temperatures still, micas
(biotite and muscovite), alkali feldspars and quartz will crystallize. Slow cooling will
result in the formation of granite. Magmas of granitic composition that reach the Earth's
surface are extremely viscous (in contrast to basaltic magmas) and are likely to produce
highly explosive volcanic eruptions.
To summaries, slow cooling of magmas occurs at depths in excess of several kilometres
and results in the formation of coarse grained rocks such as peridotite, gabbro and
granite. Fast cooling at or near the Earth's surface results in the formation of fine grained
rocks such as basalt.
2.4 Geological cycles

Geological cycles act within the Earth on large and small scales, and over long and short
periods. They may affect large masses of rock, or individual elements; they may be
simple or complex, but they all involve the manufacture and re-processing of the Earth's
materials.
One important geological cycle is the water cycle, or hydrological cycle. In this cycle
water, in the form of rain, hail or snow, that falls on land is carried to the sea in rivers,
and then evaporates, condenses, and falls again. This is illustrated in
Figure 1.13.

FIGURE 1.13 The


hydrological cycle.
Movement of water
into the atmosphere by
evaporation from the
oceans (and to a minor
extent the land) is
matched by
precipitation as rain
and snow. Water is
returned to the ocean
via several `routes'.
Run-off from the land
carries the products of
rock weathering back
to the ocean.

Intimately linked to the hydrological cycle is the best-known cycle in geology, the
rock-cycle. This was first recognized by the pioneering Scottish geologist, James Hutton.
In a book published in 1785 called, rather immodestly, The Theory of the Earth with
Proof and Illustrations, he showed how:
14

1)

igneous rocks may be eroded to form sediments by weathering and erosion;

2)

the sediments may become compacted into rocks, and:

3)

a later mountain-building event may then expose these sedimentary rocks at the
Earth's surface, where they may be eroded away, thus forming a fresh generation
of sediments.

The parts of the cycle are illustrated in Figure 1.14. The burial and subsequent uplift of
rocks is labelled tectonic cycle' on this Figure.

FIGURE 1.14
Summary diagram of
the rock cycle.

A cycle of any kind requires an energy source to drive it. What are the energy sources
driving the hydrological and rock cycles?
The energy source driving the hydrological cycle is relatively easily identified as the heat
of the Sun. The rock cycle is more complex; the hydrological cycle is involved because
moving water plays a major part in the formation of sedimentary rocks. Equally
important, however, is the Earth's internal heat, which is responsible for driving plate
tectonic processes (see Section 7), which results in the formation of mountain belts,
which are subsequently eroded.
15

All the variety of rocks composing the Earth's crust have ultimately been derived from
peridotite of the Earth's mantle by successive episodes of fractionation involving partial
melting of the mantle (in essence, the reverse of fractionation due to early settling of
minerals formed in very hot magmas), crystal settling, and sedimentary processes at the
Earth's surface. Fractionation at the Earth's surface is discussed in the next section, and in
Section 7, the role of plate tectonics in the rock cycle will be outlined.

16

SECTION 3
SEDIMENTARY PROCESSES
Study Comment

In this section, that part of the rock cycle that operates at, or near, the Earth's
surface is explained. Such processes are responsible for the formation of
hydrocarbon source rocks, most reservoir units, and seals. A more detailed
discussion of types of sedimentary rocks and the architecture of reservoir units
produced by sedimentary processes is given in Manual 2.

3.1 Introduction

Igneous and metamorphic rocks are exposed to a totally different environment at the
surface of the Earth from the one in which their constituent minerals were formed. They
are therefore not in equilibrium with these conditions: temperature and pressure are both
much lower, and the environment is oxygen rich, and often wet. In response to such
environmental change, many silicate minerals undergo changes that may lead to their
total or partial breakdown and, consequently, to the destruction of the rocks themselves.
These breakdown processes are known collectively as weathering, and it is the products
of weathering that eventually accumulate to form sedimentary rocks. Sedimentary rocks
themselves may be subjected to fresh cycles of weathering.

3.2 Physical and chemical weathering of rocks

Two distinct types of weathering are distinguished: physical weathering (or mechanical
disaggregation) and chemical weathering (or chemical decomposition). Both processes,
either separately or together, lead to the formation of a wide range of sedimentary rocks.
One of the most important forms of physical weathering, which causes the break up of
large masses of rock into fragments, results from the repeated alternation of freezing and
thawing of rainwater in cracks and fissures. Water expands on freezing, with an increase
in volume of about 9%, and the resulting pressure increase on the sides of the fissures
causes them to split further apart. Large blocks of rocks fall under gravity from rock
faces, where they are shattered into smaller angular fragments, which accumulate as scree
slopes. This process is called frost shattering. Such shattering of rocks into smaller
fragments exposes a much larger surface area to rainwater run-off, and so that they
undergo chemical weathering. This attack is enhanced by the presence of dissolved
atmospheric gases such as carbon dioxide (and also sulphur dioxide) in rainwater, which
make it slightly acidic. Chemical reactions may take place between the rock minerals and
water, sometimes dissolving them completely and sometimes producing new mineral
types.

17

Figure 1.12 shows that the silicate minerals of igneous rocks crystallize in a sequence that
is controlled by their temperature of formation. The relative resistance of the mineral to
chemical weathering is shown below.

By examining Figure 1.12, it can be seen that the first-formed, high-temperature silicate
minerals are the first ones to break down, whereas the last-formed, low temperature
minerals are the most stable. Chemical weathering has virtually no effect on quartz at all.
This sequence of silicate mineral stability is not coincidental: it is directly related to the
chemical compositions and structures of the minerals themselves. In simple terms, silicon
ions bond strongly to oxygen, to form silicate anion complexes, mainly because they are
small and highly charged. However, the other cations in silicate minerals are both larger
and more weakly charged, so bonds between silicate anion complexes and the other
cations present are relatively weak. So, minerals with weak cation-silicate bonds, such as
olivine, pyroxenes, and amphiboles, are more easily broken down than those with many
silicon-oxygen bonds, such as quartz.
3.3 The products of chemical weathering

Three products result from chemical weathering.


1)

Substances that are soluble in water are carried away in solution by rainwater
and rivers, and eventually enter the sea.

2)

New minerals are formed when soluble substances are leached out of existing
minerals and their atomic structures partially collapse. An example of this
process is the formation of clay minerals through the weathering of feldspar and
micas in granite.

3)

Resistant minerals, such as quartz, are not chemically attacked at all.

The weathering products of average igneous rocks are shown in Figure 1.15, and more
details of their composition given in Table 1.1.
The new minerals and the resistant minerals are known as residual minerals. The most
common of the resistant minerals is quartz, but if chemical weathering is not too
prolonged, feldspar and micas will also resist chemical attack. Quartz is found in huge
volumes as a residual mineral in the form of sand grains on beaches all over the world
(although not all beaches are made up of quartz sand grains).

18

FIGURE 1.15 The


weathering products
of average igneous
rocks.

TABLE 1.1 A summary


of weathering processes
and products.

The breakdown of ferromagnesian minerals, such as olivines, allow magnesium to be


carried away in solution, but this usually leaves a residue of insoluble iron oxides and
hydroxides, which gives some weathered rock surfaces their rusty brown colour. Silica, in
the form of SiO44-, from the silicate minerals, is also very slightly soluble. When the
soluble cations are leached out of the more complex silicates (such as feldspar where the
silicate framework forms large molecules), the collapsed atomic structure usually
incorporates water to form a flakey, hydrated clay mineral.
Climate plays a very important part in weathering. For chemical weathering to be most
effective, annual rainfall should exceed about 1,000 mm and temperatures should be high
to speed the rate of the chemical reaction. Frost shattering can still be very effective at
quite low annual rainfalls, but obviously it requires the temperature to fluctuate above
and below freezing. For these reasons chemical weathering predominates in the hot,
humid low latitudes and physical weathering predominates in the colder high latitudes.
Not all chemical weathering takes place at the site of a rock outcrop being weathered. As
soon as rocks have become fragmented the fragments are moved under gravity, by water,
wind or glaciers, if climatic conditions are right. So both physical and chemical
breakdown continue during transportation of the fragments.

19

3.4 The transport of weathered material

It is fairly obvious that the soluble products of chemical weathering are removed by
rainwater and rivers. However, the residual products of chemical weathering, and the
larger fragments produced by physical weathering, are also removed by various
transporting media. The removal of weathered material is usually referred to as erosion,
and the combined processes of weathering and erosion that gradually wear down
mountainous landscapes are termed denudation. Wind and ice, as well as water, can be
important transporting media for rock fragments.
The transport of rock fragments and mineral grains by water involves either the bouncing
or rolling of the material along the bed of a river or the sea as bedload, or the transport of
the material within the water itself as suspended load. Coarser material tends to move as
bedload, and finer sediment is carried along within the water itself as suspended load.
The largest fragments may be deposited and remain immobile on the stream bottom.
Water current speeds and changes in speed are largely responsible for the selective
uptake and deposition of sediment gains, or the process of sorting a sediment, as well as
whether, overall, it is coarse or fine grained. The sorting is simply a measure of the range
of grain sizes about the average or mean grain size value for that deposit. A well sorted
sediment has grains that are very similar in size and that have a narrow range about the
mean, whereas a poorly sorted sediment has grains of various sizes with a wide range
about the mean. Well sorted sediments have a higher porosity than poorly sorted ones.
High-energy water environments result in the deposition of coarse-grained sediments, and
low-energy environments lead to the deposition of fine-grained sediments. There can be a
good deal of variation in the type of sedimentary rock within a small area. For example,
around headlands, where wave action is concentrated, large boulders and pebbles are
strewn on the beach, whereas in sheltered bays, sands are deposited. Similarly, on the
outside of a river bend (or meander), where the flow is fastest, coarse sands and grains
may be found on the river bed whereas on the inside of the bend, where the flow is
slowest, finer sand may be deposited.
The sorting of sedimentary material is a progressive process. The further a mass of
sedimentary material is transported, the greater is the chance that it will become separated
out into fractions of different sizes. Sometimes this separation is aided actively by the
way in which water moves. For example, the constant action of waves on a beach
winnows out the finer sands and silts and keeps them in suspension to be deposited later
where low energy conditions prevail, such as in a sheltered bay or the mouth of an
estuary.
The discussion so far has assumed that the particles being transported are all of the same
mineral type. But the mineral type is important because it determines the shape of the
sediment grains and their density. Flakey minerals, such as micas, are not deposited
according to the conditions shown in Figure 1.16 because their shape offers considerable
resistance to the water as they settle. This is why mica flakes tend to occur in sandstones
and siltstones that have been deposited in water under low-energy conditions. In fact,
Figure 1.16 is based on the behaviour of spheres of quartz, which are the most common
mineral grains transported.

20

FIGURE 1.16 Graph to


illustrate the relationship
between size of sediment
grains and the speed of
water current necessary to
pick them up (upper curve)
and deposit them (lower
curve). The zone between
the two lines represents
the field in which sediment
grains can be carried in
suspension by the current,
once they have been
picked up.

Particles that have been picked up and transported are subjected to mechanical abrasion,
and undergo changes in shape. It is fairly obvious that the longer particles are in motion,
the more eroded they will become, as edges and corners are worn away. It is perhaps less
obvious that large particles are much more rapidly rounded than small ones; this is
mainly because large particles have more momentum than smaller ones and, hence, can
do more damage to each other when they collide. The upshot of this is that rounded
particles are much more common in coarse-grained sediments deposited in a high-energy
environment than in fine-grained sediments deposited in a low-energy environment. For
example, the pebbles on a beach are almost always smoothly rounded, but the sand grains
are much more angular.
Fine grains, falling out of suspension like confetti, come to rest on the bed beneath the
water as very fine flat and even layers, or laminae, usually only a millimetre or so in
thickness. So, fine-grained, laminated sediments are characteristic of suspension deposits.
Coarser grained sediments, which move in the bedload prior to final deposition, do not
form such neat layers. Often they are built up by the motion of the water near the bed into
regularly spaced mounds and hollows known as bed forms, such as ripple marks on
beaches. A more detailed discussion of bedforms and the manner in which they produce
structures within reservoir units is given in Manual 2.
3.5 The transport and erosion of sediment by wind and ice

Air is a fluid, so that wind behaves in exactly the same way as water when
transporting sediment. The only difference is that the density of air and,
consequently, its viscosity is very much lower than that of water. Therefore,
moving air can move a much more restricted range of grain sizes than water. To
give an example, whereas it takes a water current speed of 0.08ms-1 to start a quartz
grain of 1 mm diameter rolling, it would take a wind speed of nearer 10 m s-1
21

(36 km h-1 or force 6 on the Beaufort wind scale) to do the same thing. Most transport by
air is as a kind of bedload sweeping across the ground, and only the very finest dusts are
lifted up into suspension. This means that, after re-deposition of the particles as wind
speeds gradually fall, the resulting wind-blown sediments are usually very well sorted.
The individual quartz grains are usually more rounded than those of water transported
sediments because the wind-blown grains collide without the cushioning effect of water,
and so have their corners and edges worn down much more rapidly.
By contrast, moving ice in the form of glaciers or ice sheets is totally indiscriminate
about what it transports. Anything from the fine clay-sized particles to vast boulders
several metres in diameter can be embedded in the ice. The mechanism of erosion is also
different. Whereas some particles and even sizeable fragments may be picked up by the
moving ice, many of the larger glacial boulders fall from the sides of a glaciated valley
onto the surface of a glacier. Subsequently, such boulders work down through the ice
because of the greater density of the rock material. Once embedded in the ice, the
fragments are not free to collide with each other, and most will not be deposited until the
ice melts (though large boulders may be deposited at the base of the glacier), often many
miles away from where the fragments were picked up. When the ice melts, all the
remaining fragments, from clay sized particles to boulders, are deposited at the same time
to form a long ridge of sediment, known as a moraine. Consequently, true glacial
sediments are very poorly sorted and the fragments are characteristically angular. The
larger fragments are also often faceted, owing to scraping of the entrained boulders at the
base of the glacier. This means that glacial sediments are not characterized by diagnostic
grain shapes or degrees of sorting that will give any clues about how far the material has
been transported.
3.6 The transport and deposition of the soluble products of chemical weathering

The most abundant soluble cations released from silicate minerals by chemical
weathering are Ca2+, Na+, K+ and Mg2+. These, along with some silica, in the form of
SiO44-, are carried away in rivers, eventually reaching the sea. Therefore, you might
expect that as weathered material in solution has been added to the oceans over many
millions of years, the sea should be becoming progressively more saline. This is not the
case because the dissolved salts are removed at the same rate as they are added, which
means that the composition of seawater has remained reasonably constant over at least
the past 1000 million years. Two major ways in which dissolved salts are removed are (i)
by the action of marine organisms and (ii) by direct chemical precipitation.
(i) The action of marine organisms
Many marine animals build shells composed of calcium carbonate by extracting
both Ca2+ and HCO3 from seawater to form the mineral calcite (CaCO3). After the
death of the organisms, their calcareous remains may accumulate on the sea floor.
Provided they are not greatly diluted by weathered sediment brought into the sea
from adjacent land areas, they may become cemented by chemically precipitated

22

calcite, in the form of a lime mud, to form limestone. Many limestones are formed on the
continental shelves in tropical and equatorial latitudes, where marine organisms thrive in
the warm, sunlit waters and there is a low input of continental sediment.
Some calcareous sediments from in the deeper ocean basins from the accumulation of
calcareous microfossils that are far too small to be seen with the naked eye. These
unconsolidated sediments are called calcareous oozes. When compacted and cemented
they form fine-grained chalky limestones.
Once exposed at the Earth's surface, limestones are very susceptible to chemical
weathering because they dissolve easily in rainwater that has been acidified by
atmospheric carbon dioxide. The chemical reaction that takes place is:
CaCO3
Calcite in
limestone

H2O
+
rainwater

CO2

from the
atmosphere

Ca2

2HCO3
soluble ions
in water

The weathering of limestones is a major source of both calcium and bicarbonate ions in
river waters and groundwaters, which explains why water in limestone regions is hard':
calcium carbonate is re-deposited when the water is boiled.
(ii) Direct chemical precipitation
Sediments formed from soluble cations by direct precipitation are comparatively rare, but
some are of great economic importance. The best known are salt deposits, or evaporites,
which form in isolated inland seas or in narrow, newly formed ocean basins where
circulation with the other masses of sea-water does not take place. Here, the evaporation
of the water may exceed the input by rivers, so that the water becomes saturated with
respect to ions such as Na+, K+, M2+, Ca2+, CI-, (SO4)2- and (HCO3)-, and salts of these,
such as halite (NaCl), gypsum (CaSO4.2H2O) and, of course, calcite (CaCO3), crystallize
out. Evaporites also form along arid shorelines like that present-day Persian Gulf. Here
they crystallize out within the sediments of the mud flats. High evaporative rates are
common in arid climates; thus, evaporites in rock sequences are a good indicator of
ancient arid climates.
A few limestones are precipitated directly from seawater but they are relatively rare. They
require warm shallow waters (CaCO3 is less soluble in warm water than cold water) that
are well agitated by waves or currents, and particles such as sand grains or shell
fragments act as nuclei for the precipitated CaCO3 to grow around. The resulting
calcareous grains are slightly well rounded and are termed ooids.

23

SECTION 4
GEOLOGICAL TIME
Study Comment

At the beginning of this manual, when studying Figure 1.1, you met terms such as
Cretaceous, Jurassic and Carboniferous, which were used to label' rock units in the
North Sea. Terms such as these are routinely used all over the world to identify
intervals of geological time.
The purpose of this section is to show the enormous scale of geological time, and how
it is divided up into four major Eras, each made up of shorter periods.
The way in which individual rock units can be assigned to specific geologic time
intervals will be explained, as will the methods used to assign a quantitative age to
rocks.
4.1 Introduction to geological time

The whole of geological time is divided into four major divisions called Eras (see Table
1.2). These Eras were defined and given their names from the general character of their
fossils long before geologists knew anything about the timescales involved, at a time
when the beginning of the Palaeozoic was thought to correspond roughly to the origin of
life.

TABLE 1.2 The


geological Eras.

These Eras have been arranged with the oldest at the bottom here and in Figure 1.17 to
form the Stratigraphic Column for the Earth. The Stratigraphic Column is made by
stacking up the rocks formed during each geological Era in their correct sequence, always
starting with the oldest at the bottom.
It is important that you should remember the names and approximate dates of the four
Eras, since they are in common use in any discussion of geological time.

24

As you can see from Figure 1.17, the last 570 million years is further subdivided into
eleven Periods. These, as you will see in Section 4.4, were defined long before any
radiometric ages were known. Radiometric ages will be explained later in Section 4.5.

FIGURE 1.17
Stratigraphic column for
the whole of geological
time. (a) To true scale:
note the length of the
Precambrian Era, about
4000 million years, or
nearly 90 percent of
geological time. (b)
Expanded scale since the
Cambrian, to show
details of the Periods in
the three most recent
Eras: Palaeozoic
(`ancient life'), Mesozoic
(`middle life') and
Cenozoic (`recent life').

Even before people had any accurate means of measuring time, they knew that there were
regular sequences in nature: day follows night, spring follows winter, and sometimes the
results of these daily or annual fluctuations are preserved in nature, for example, in
annual tree rings. Occasionally, it is possible to work out a chronology giving actual
dates, by counting annual events of the fairly recent past, starting from the present and
working back. But, apart from radioactive dating, most geological methods are purely
relative: they can only be used to work out the relative order in which things happened.
25

From what is known of the forces controlling the process that can now be observed on the
Earth's surface, namely the laws of physics and chemistry, there is no evidence that they
were any different in the past. It seems reasonable, therefore, to suppose that the present
is the key to the past. This common sense approach to the interpretation of past geological
events is often termed uniformitarianism. It may seem very obvious today, but it was a
revolutionary idea when it was first presented about 150 years ago.
There is one further implication of the principle that the present is the key to the past: it
can also be used in some cases as a key to the future. Some geological events happen on a
timescale that makes their prediction of vital importance to us, earthquakes and volcanoes
being the most dramatic examples. In many countries well away from active zones of the
Earth's crust, there is little chance of major damage by either of these in the foreseeable
future, but on a slightly longer timescale it is important to know whether a new Ice Age is
imminent.
4.2 Putting things in order

The ways in which geological events can be ordered in time can best be explained by
examples and analogy.
Consider first an example from the recent past. Various objects have been found in old
dumps near mining camps in North America, and seven examples of these are shown in
Figure 1.18. Their distribution in three boreholes through one dump is shown in Figure
1.19. The dates of manufacture of the bottles, cans and nails are known from their
makers' old records, and so a range chart can be worked out showing the duration of
manufacture of each item (Figure 1.18). Some of these articles can be used to give a
fairly exact age to a part of the dump because they were in use for only a few years, while
other were used over a long time and so are not so valuable for dating. Some groups of
articles were only in use together for a very short time, and so an assemblage of these
found together can tie down the age of the dump very precisely. For example, finding
items 3 and 4 together would indicate the date of 1900.
FIGURE 1.18 Range
chart of dates of
manufacture of cans,
bottles and nails
prepared from
historical records.
Each article is given
a number, and these
are used in Figure
1.19 to show where
the articles were
found in boreholes.

26

FIGURE 1.19 Rubbish


dump of old American
mining camp, to show
the distribution of cans,
bottles and nails in the
rubbish (stippled) tipped
on the original land
surface (black).
The rubbish found in the
three boreholes A, B and
C is shown below. The
numbers beside the
boreholes refer to the
articles shown in Figure
1.18 found at depth.

The next example goes back a little further in time to consider the tools of our early
ancestors, and the ways in which these tools were modified and improved as time went
by. Man appeared on the Earth about three million years ago, and Figure 1.20 illustrates
some of the primitive tools he invented. These tools can be used to date early remains
since the sequence in which they were developed is now known, through first the Stone
Age, to Bronze Age and then Iron Age.
FIGURE 1.20 Axes
made of various
materials: A, B, D of
stone, E of bronze,
and C of iron.

Arrange the axes shown in Figure 1.20 in order of decreasing age.


A, D and B are all stone axes, and so are older than those made of metal. A is the most
primitive, and therefore the oldest of all, with D being more skillfully fashioned, but not
set in a handle, as is B. E comes next, being made of bronze, with C, composed of iron,
being the most advanced tool.
Can you see the assumption behind this method? It is that the use of tools evolved' from
Stone to Bronze to Iron everywhere at about the same time (otherwise a stone axe from
one area might be the same age as a bronze one from elsewhere). On a broad scale this is
probably true, but consider present implements: people in industrialized countries and
contemporary primitive tribes are producing very different artifacts for future geologists
to discover!
Once you have placed these items in their correct relative order, you need further
information to find out when the tools were used, that is, some quantitative method of
measuring time is needed.
27

In a directly analogous way many sedimentary rocks can be dated by the fossils they
contain. Fossils can be formed when plant or animal remains become buried in
accumulating sediment (such as mud on the floor of a sea or lake) before they are
completely decomposed, and as the sediment hardens into rock, the traces of organic
material are preserved. Preservation can take a variety of forms: the hard parts of the
organism may remain more or less intact, while softer parts may form an impression or
cast which is later filled with sediment. Thus, insects completely preserved in every detail
in drops of amber (fossilized resin) and the footprints left by a dinosaur on an old land
surface can both be described as fossils'.
4.3 Evolution

Fossils are widely used to determine the relative ages of rocks, and so place them in their
correct order, that is, their correct stratigraphic sequence. But how was this sequence
originally worked out? As you will see later, earlier in the Nineteenth Century, William
Smith and Georges Cuvier discovered that overall there is a sequential order of fossils
through time; the biological explanation of this came later, particularly with the work on
evolution by Charles Darwin (1809-1882).
By careful collection, palaeontologists have discovered a discontinuous record of fossils
in the Stratigraphic Column from the end of the Precambrian to the present. The
stratigraphic sequence of the rocks was established first by mapping them in the field
and, from the relationships of rock strata to each other, deciding on their relative ages.
Having done this, it was then possible to work out the sequence of fossils in which fossils
appeared through geological time. Once this sequence of fossils had been established, the
fossils themselves could then be used to date other rocks.
The fossil record is not complete because very few of the animals and plants living at any
one time are preserved as fossils. Some marine organisms are fairly abundant in the rocks
in which they are found and the evolution of individual species can be traced in greater
detail. Sometimes a particular species is only found through a few metres of rock in one
Period, with its ancestors below, and its descendants above. In such cases a single fossil
can pinpoint the position of a rock in within, it is found very accurately within the
Stratigraphic Column.
Fossils can be used to determine the relative ages of rocks back to about 600 My (million
years) before the present, at which time abundant shelly fossils first appear in
sedimentary rocks. The evolutionary order of animals and plants can be used to establish
the relative ages of rocks because we assume that, in general terms, evolution proceeds
from simpler to more complex organisms; for example, the evolution of the vertebrates
followed the sequence:
fish

amphibia

reptiles

If we place this sequence in stratigraphic order we have:

28

mammals.

Before the Devonian Period the only vertebrates were fishes; by about the Cretaceous all
four groups had evolved.
Thus the oldest rocks in this sequence will contain only fossil fish and the youngest,
besides containing mammals, will also yield fossils of all four groups, which have
evolved, from the common ancestor, the fishes. The detailed evolution of the vertebrates,
to which all these animals belong, is much more complicated than this (Figure 1.21). For
example, during the Jurassic and Cretaceous Periods (the age of the dinosaurs) fish,
amphibia, dinosaurs and reptiles existed, but mammals and birds were barely significant.

FIGURE 1.21
Geological range of
the vertebrates to
show when the
various groups of
animals evolved
from each other.
The width of each
group indicates the
approximate number
of species living at
the time.

29

When the relative ages of strata from different localities have been established by the
sequence of the fossils they contain, the rocks can be arranged into a Stratigraphic
Column, but no actual ages can be given to rocks by the use of fossils alone; it is
therefore only a relative dating method.
In the oil industry, microfossils obtained from well cuttings are the usual method of
assigning a stratigraphic age to the rocks being drilled.
4.4 How the stratigraphic column was developed

The first attempt to recognize a sequence of historical events in sedimentary strata was
made by a Dane, Nicolauz Steno (1638-1687) in the mountains of western Italy. He
recognized that older rocks are overlain by younger rocks and so established the
principle of superposition. Steno also realized that strata that are normally deposited
slowly are laid down in a near-horizontal position, although later they may be folded and
even overturned.
There are often features present within a sedimentary rock that can be used to interpret
which way up the sediment was when it was laid down. These will be described in
Manual 2.
An Italian, Giovanni Arduino (1713-1795), prepared a simple stratigraphic column for
the rocks of northern Italy. Having studied how individual rock strata were related to
each other where he saw them exposed in the field, and having looked at the character of
the rocks themselves, he split his column into three:
Tertiary
Secondary
Primary

soft limestone with fossils, clays and sandstones;


hard limestones and mudstones with fossils;
severely folded metamorphic and igneous rocks without fossils.

Only the term Tertiary as survived from this first classification (as a Period name within
the Cenozoic (see Figure 1.17)). Arduino's Secondary very roughly corresponds to the
present Mesozoic and Palaeozoic Eras combined, and his Primary is roughly equivalent
to the present Precambrian Era. This was a beginning, but there was no certainty that
these three divisions could be applied elsewhere, because there were no clearly expressed
criteria to separate them, particularly the upper two divisions. There was no reason to
suppose that rocks of similar age elsewhere would look the same as these strata in Italy.
Early philosophers, such as the Greek, Herodotus, deduced accurately in the 4th century
B.C. that fossils were the remains of ancient sea creatures. He believed that this indicated
that the rocks in which they were found had been formed originally beneath the sea.
However, this theory, along with many other perceptive theories of the ancient Greeks,
became lost during the next 2000 years, overshadowed by religious beliefs. Even as late
as the seventeenth century, some scholars thought that the Earth was only 6000 years old.
The most plausible explanation offered for fossils was that they were either relics of
Noah's Flood, or the work of the Devil, who deliberately placed fossils in rocks to
deceive, mislead and perplex mankind.
A Frenchman, Georges Cuvier (1769-1832), was one of the first to describe
systematically the skeletal remains found preserved in rocks as fossils, to interpret
them in terms of the living organisms they represented, and to work out their
succession in Earth history. He studied the fossil plants and animals in the

30

Tertiary rocks of the Paris Basin, and concluded that older fossils differed more from
living creatures than did younger ones. He summarized his conclusions diagrammatically
by sequences of strata, such as those shown in Figure 1.22. He deduced from these that
some older forms of life had become extinct and that the extinct forms had been replaced
by newer forms.
But some of Cuvier's conclusions were not correct. For example, in his answer to the
question of how new species arose, Cuvier believed that each old species was wiped out
by a universal catastrophe followed by the special creation' of new species.

FIGURE 1.22
Diagrammatic
section of strata
(after Cuvier)
showing how
particular fossils are
associated with
particular strata.
Breaks within the
sequence where
erosion of the
underlying bed has
occurred before
subsequent beds
were laid down can
be seen in several
places, perhaps best
where strata with a
walking vertebrate
(shown by vertical
ruling) are overlain
by a horizontally
shaded bed.

Catastrophism is the name given to the idea that geological history can be explained by
a series of catastrophic events. Starting for the biblical idea of the Flood, Cuvier invoked
similar deluges' to explain each break in the series of fossils in his sequence of
sediments. Since he was not able to find intermediate species connecting the fossils at
different levels in the Stratigraphic Column, and since, moreover, there was often
evidence of a break in the deposition of the sediments themselves, marked by changes in
rock type that coincided with the faunal changes, it was logical to invoke a new deluge'
for each break. Geological history then was thought to be a whole series of deluges,
which killed off all life, each followed by special creation of a whole new fauna.
Cuvier noticed that very often the breaks in the sequences of strata and fossils were
marked by a horizon where there was evidence of erosion of the underlying beds
before deposition of the next layer, and that the beds immediately above contained
pebbles. Indeed this was powerful evidence for the deluge. Such breaks in the
Stratigraphic Column often do represent a long interval of time, during which
sedimentation ceased and erosion occurred because the area had risen above sea31

the formation of an unconformity is shown in Figure 1.23. The pebble bed that marks the
time when the area sank beneath the sea, and sedimentation began again is in effect a
fossil beach' formed as the sea gradually drowned the land. Recognition of
unconformities, and the missing strata they may represent, is crucial in working out the
geological history of an area. Several unconformities are shown in the cross sections of
the North Sea shown on Figure 1.1.
FIGURE 1.23
Unconformities. (a)
Stages in the formation
of an Unconformity (as
seen in cross section).
(i) Sediments A are laid
down under water on
pre-existing rocks. (ii)
The area is tilted, lifted
above sea-level, and
some of the original
sediment eroded away.
(iii) The area sinks
beneath the sea and a
second series of
sediments, B, laid down
on the old erosion
surface, which is now
called an unconformity.
(b) Cross-section of a
typical unconformity
(simplified) which
occurs in the Ingleton
area of northern
England.
The time interval
represented by the
unconformity can be
many tens or hundreds
of million years. In this
case the older strata, A,
were folded and eroded
before the younger
strata, B, were laid
down. Since strata B
were deposited there has
been a subsequent phase
of uplift and erosion.

In the last years of the eighteenth century an Englishman, William Smith (1769-1839), an
engineer and surveyor, who worked on canals, roads and drainage schemes all over
England, found that he could recognize distinctive beds within rocks such as the Chalk
on the North and South Downs, or within the coal-bearing strata in widely separated
coalfields. He noted that each group contained a particular assemblage of fossils quite
distinctive from those of the strata above and below.
Smith began to correlate apparently dissimilar sedimentary strata because they contained
similar fossils. Furthermore he found that there was the same succession of fossil
assemblages from older to younger beds in all parts of the country. He concluded that
each stage of this succession of fossils represented a particular span of geological history,
or a discrete length of time, and that rocks formed during that time would contain the
same fossils wherever they occurred geographically. This he called the principle of
faunal succession, and using it he was able to correlate widely separate outcrops of rock
by the fossils they contained. An outcrop of rock is a piece of the solid rock strata (not a
boulder), which is visible at the surface, so it crops out' through the soil and vegetation
cover.
32

The acceptance and application of William Smith's ideas led to the division of the
stratigraphic column into the periods shown in Table 1.3. Each period was defined using
a distinctive sequence of rocks at various locations in Europe.
TABLE 1.3 Origin of the
names of the Periods in the
stratigraphic column.

When reading geological accounts of reservoirs or oilfields, you should be aware that
three different sets of terminology may be used to describe the stratigraphy of a region,
namely lithostratigraphy, biostratigraphy and chronostratigraphy.
Lithostratigraphy (rock stratigraphy) is based on the description of rock units in terms
of their mineralogy, petrography, internal structure and fossil content. An example
lithostratigraphy from the Brent area of the North Sea is shown in Figure 1.24.
Lithostratigraphic units are defined using objective criteria, and are named after
geographic locations. There is a hierarchy of units:
Supergroup
Group
Formation
Member
Bed
Three of these - group, formation and member- are illustrated in Figure 1.24. Units of
formation status and up are mappable over significant areas of basins, but members and
beds have only localised distributions. Units are defined using type sections, where the
top and bottom of the unit can be seen. In offshore areas, or with units, which are, only
encountered in the subsurface, type well sections are defined for reference purposes.
Biostratigraphy encompasses the use of fossils to correlate rock successions in different
locations. Biozones are the units of biostratigraphy, and are based on the distribution of
fossils which evolved relatively rapidly, and which can be found in a wide range of
lithologies. The example shown in Figure 1.24 shows the ammonite biozones for part of
the middle Jurassic. Ammonites are used as the basis for biozonation of Mesozoic rocks,
but as suitable material is seldom recovered even in cores. the petroleum exploration
industry uses microfossils as an alternate basis for biozonation. At outcrop locations, the
ammonite and microfossil biostratigraphies can be integrated.
33

FIGURE 1.24 An
example of the use of
lithostratigraphic,
biostratigraphic and
chronostratigraphic
units for Jurassic rocks
in the Brent area of the
Northern North sea.
Further explanation in
text.

Chronostratigraphy is not the quantitative dating of rock successions. It aims to


produce an internationally agreed hierarchy of units, the boundaries of which are defined
by master points within an agreed outcrop. The hierarchy of units is:
System
Series
Stage
The Stage is the fundamental unit. It is defined by a given set of biozones in a particular
area, but in different parts of the world, different biozones may be used, due to the past
provincial distribution of organisms.
Attempts to calibrate the chronostratigraphic units began in the 1930's when sufficient
radiometric dates had been obtained. However, agreement about the dates of some
intervals of geological time has still not been reached. For example, the date of the
Jurassic-Cretaceous boundary ranges is defined by different time ranges from 135 to 144
million years ago!
4.5 Calibrating the stratigraphic column

Rocks that have minerals containing a radioactive isotope have a built-in clock' for
measuring their age. The principle is very simple. The rates of decay of all the common
radioactive isotopes are constant and are known from accurate laboratory measurements
on pure samples. If the amount of a radioactive isotope present in a material when it was
formed is known, then the age of that material is calculated from its present radioactivity,
using the known decay rate for that isotope.
Most geological ages are calculated from radioactive decay series where the original
parent isotope decays to give a stable daughter isotope. To calculate the age of a
mineral grain, it is necessary to find out how much of the parent isotope has
34

decayed since the mineral was formed. This is normally done by measuring the amount of
daughter isotope present, as well as the amount of parent isotope left, to get the parent:
daughter ratio.
There are several radioactive decay processes that have been used for geological dating,
and a selection are shown in Table 1.4:
TABLE 1.4 Commonly
used radioactive dating
processes.
The details of these
processes need not
concern you.

To date a rock sample, a mineral within it has to be found that contained atoms of a
radioactive isotope when it originally crystallized. Each parent atom eventually decays to
a daughter isotope, which is retained in the same crystal. Suppose we find in igneous rock
containing a uranium-rich mineral and that when it initially crystallized, it did not
incorporate any lead in its constituent crystals. As time passed, some of the uranium was
converted to lead by radioactive decay. To find out how old the rock is, samples of the
mineral can be dissolved to get the uranium and lead into a solution. The parent :
daughter ratios of the isotopes 23592U : 20782Pb and 23892U : 20682Pb can then be measured
with a mass spectrometer.
As an igneous magma cools and minerals containing a radioactive isotope crystallize, the
radiogenic clock' is started in the rock, so that at any time subsequently the age of
cooling can be determined by the methods just described. Although igneous rocks are
relatively rarely found interstratified with sedimentary rock that can be dated by fossils in
the Stratigraphic Column, when this happens the igneous ages provide crucial calibration
points, since they can give an exact age in millions of years to that part of the
Stratigraphic Column. The best examples are lavas, which clearly postdate the rocks
beneath and predate subsequent strata; in other words, their stratigraphic age
relationships are exactly the same as for the adjacent beds of sediment, to which the
provide an absolute' age.
But not all magma reaches the Earth's surface, and if magma solidifies by infilling
fissures or larger spaces in the crust the igneous rocks that result can be shown to be later
than the surrounding rocks because the igneous rocks may cut across pre-existing
structures such as bedding planes.
An igneous rock, which is intruded as a sheet along a bedding plane is called a
sill. Sills are often from a few metres to a few tens of metres in thickness and
spread over an area of many tens or hundreds of square kilometres at roughly the
same horizon in a sedimentary sequence (Figure 1.25(a)). A near vertical crack
filled with magma is known as a dyke (Figure 1.25(b)). Sills, dykes and lava
flows,

35

which can be radiometrically dated, are all used for bracketing the ages of sedimentary
rocks in which they are found. Sills and dykes yield a minimum age for the sediments
into which they are intruded. However, as lava flows are laid down on top of sediments,
and may be covered by them, they provide an exact age.
FIGURE 1.25 Various
types of igneous intrusions.
(a) Cross-section of sill.
The beds above and below
are likely to have been
baked by the intrusion.
(b) Cross-section of dyke.
The rocks on both sides of
the dyke may show effects of
baking by the intrusion.
(c) Block diagram of a
pluton.
The zone of contact
metamorphism around the
igneous rock may be quite
extensive.

In a similar way, large intrusions, which may represent many cubic kilometres of magma,
and which have crystallized at depth of several kilometres, can also be used to determine
a minimum age for the surrounding rocks (Figure 1.25(c)).

The relationships seen in the field between an igneous intrusion and the surrounding
sedimentary rocks are very important if the igneous rocks are to be used for dating
purposes. For example, a granite may cut across the bedding of the adjacent strata and
therefore be later than those strata (Figure 1.26(b)), or the contact between granite and
sediments may be an unconformity, in which case the granite is older than the overlying
sediments and does not metamorphose them at all (Figure 1.26(a)).
Each time a particular radiometric date is used to calibrate some part of the Stratigraphic
Column, a similar age can then be applied to rocks known to be of the same age because
they contain similar fossils. Moreover, in a uniform series of sediments, if radiometric
ages can be found for several horizons in a sequence, the age range of rocks between can
be estimated by interpolation. Each new radiometric date can then be used to refine this
dating procedure.
36

FIGURE 1.26
Diagrammatic sections
through plutonic
intrusions. In (a) the
intrusion was
emplaced after rock
series A, and eroded
before series B was
deposited. In (b), the
intrusion must postdate series A, but its
age cannot be pinpointed as precisely as
in (a).

Some unconformities, representing events than can be recognized over wide areas, can be
dated quite precisely, and many of the geological Periods in the Stratigraphic Column are
separated from the rocks above and below by such unconformities.
By the detailed correlation of igneous dates from all parts of the world, virtually any
horizon after the Precambrian in the Stratigraphic Column can now be allocated a date'.
However, in certain parts of the Stratigraphic Column, there are as yet very few
calibration points, so the age of some biostratigraphic divisions is still disputed.
However, the individual zone, defined by fossils, remains the unit of calibration for the
practicing geologist in the field. Marker horizons, often characterized by one or more
specific fossils, are used to establish the relative ages of individual outcrops. This is
basically the same method used by William Smith nearly 200 years ago.

37

SECTION 5
THE NATURE, ORIGIN, GENERATION, MIGRATION AND
ENTRAPMENT OF PETROLEUM
Study comment

In Sections 5.1-5.3 we consider the composition and processes of formation of


petroleum, which, after water, is the second most abundant fluid in the Earth's
crust. Economic accumulations of hydrocarbons form if migration (5.4) into a trap
(5.5) occurs. The last two sections discuss natural gas and oil shales.
5.1 Nature of petroleum

Petroleum is a general term for a mixture of various hydrocarbons, and the relative
amounts of these compounds in a given sample determine its properties. Petroleum exists
naturally in gaseous (natural gas), liquid (crude oil) and solid (asphalt) states. The
commonest elements in the compounds of petroleum are thus carbon and hydrogen, with
much smaller amounts of oxygen, nitrogen and sulphur (Table 1.5).

TABLE 1.5 Elemental


composition of typical
petroleum samples.

Petroleum normally occurs in sedimentary rocks deposited under marine conditions, and
its complex nature is evident from the fact that, so far, over 1200 different hydrocarbons
have been identified in crude oil. The majority of these compounds contain between one
and 40 carbon atoms, and the number of carbon atoms is the basis of a means of grouping
the components of petroleum. The seven main groups, or petroleum fractions, are gases
(one to three carbon atoms), gasoline (C4 to C10), kerosene (C11 to C13) diesel fuel (C14 to
C18), heavy gas oil (C19 to C25) lubricating oil (C26 to C40) and waxes (over C40).
So petroleum, like coal, is a carbon-rich material, and there are also similarities in their
processes of formation. There is one crucial difference between them, however. Coal is
solid and stays where it was formed; petroleum in its fluid forms can move away from its
source and accumulate in favourable structures elsewhere in the sedimentary sequence.

38

5.2 Origin of petroleum

There is no doubt that most of the compounds that have been identified in petroleum are
organic in origin. Ideally, the organic matter is deposited on the sea floor, where it must
be protected from scavengers that would consume and destroy it. This protection is
afforded by (i) anaerobic conditions, which arise because the bottom waters are free of
oxygen, and (ii) a relatively high rate of sedimentation of inorganic material, such as clay
or sand particles, to bury the organic matter.
Table 1.6 lists the annual production of organic carbon from each of the seven main types
of environment found on Earth. The continental shelf regions are more productive than
the open oceans because of the higher nutrient content of shallow water; there is an added
consideration because rivers transport large amounts of land-derived carbon into shallow
seas. So present-day marine sediments can contain organic matter from several sources.

Table 1.6 Annual


production of
organic carbon from
principle present day
environments.

However, land plants appeared 400 million years ago in the Devonian and the first rapid
development of these floras 345 million years ago gave rise to the first vast deposits of
coal in the Carboniferous. So the main source of organic material before the Devonian
must have been marine in origin. From the Precambrian until the Devonian, the largest
producer of organic matter was phytoplankton. These are microscopic marine algae,
which live in the upper layers of the oceans and on death sink in countless millions to the
sea floor. There the algae become part of the sediment and form organic-rich marine
shales. Figure 1.27 shows how the abundance of phytoplankton has changed with time.
There were two peaks of maximum production: the first in the Ordovician - Silurian
400-500 million years ago aid the other in the Jurassic - Cretaceous 65-200 million years
ago.
The two phytoplankton peaks tend to coincide with global rises in sea-level (Figure 1.27)
and it has been suggested that, during the major rises, seawater flooded into lowland
areas and into vast terrestrial depressions, where enormous phytoplankton populations
developed in nutrient-rich marine basins. On death, these algae would sink into
oxygen-free muds, which on burial formed beds of organic-rich material. These contain
0.5 per cent or more of organic matter and have the potential to generate petroleum; for
this reason they are termed source rocks. Source rocks are normally shales and contain
about 90 per cent of all the organic matter found in sediments; on average, shales contain
1.2 per cent organic matter as compared with limestones and sandstones, which have 0.56
and 0.05 per cent, respectively.

39

FIGURE 1.27 Major


changes in sea-level
and abundance of
phytoplankton for the
last 570 million years.
Phytoplankton are
microscopic floating
plants.
Also shown are the
source rocks for
petroleum and the
crude oil reserves in
relation to the
stratigraphic column.

As you can see from Figure 1.27, the greatest volumes of source rocks are of Jurassic Cretaceous age, and these are responsible for about 70 per cent of world resources of
petroleum. But how can the much lower level of Palaeozoic crude oil reserves be
reconciled with quite high phytoplankton production at that time? There is no great
difference in the abundance of phytoplankton in the two Eras, so the explanation must be
that the crude oil in the Palaeozoic rocks has been partly destroyed as a result of
mountain building events, which caused deep burial, heating, folding and erosion of the
rocks; the Mesozoic rocks have not been so severely affected by these destructive
geological processes mainly because they are much younger.
5.3 Generation of petroleum

The conversion of sedimentary organic matter into petroleum is termed maturation and
can be divided into three phases (Figure 1.28).

FIGURE 1.28 The


products of the
maturation of organic
matter as the
temperature increases
during diagenesis,
catagenesis, and
metagenesis: principal
zones of oil and gas
formation are shown.

40

Diagenesis: Immediately after deposition, biochemical decomposition of the


organic matter starts and produces biogenic methane. As the organic-rich
sediment is buried and subjected to slightly increased pressures and temperatures
the organic matter is converted into kerogen, which is the precursor of
petroleum.

Catagenesis: With higher temperatures (and pressures) kerogen is altered and


the majority of crude oil is formed. During this and the next phase of maturation
(metagenesis), the larger molecules break down into simpler molecules; this
process is called cracking, and is very similar to the process that takes place in
oil refineries.

Metagenesis: In the final stage of alteration of kerogen and crude oil, natural gas
mainly in the form of methane is produced and residual carbon is left in the
source rocks.

Although the main source of crude oil is phytoplankton, other types of organic matter
such as land-plant remains, bacteria and zooplankton (microscopic floating animals) are
present in marine sediments and can be converted into kerogen. Three kinds of kerogen
from different sources have been identified, each of which yields different products
during maturation. Table 1.7 shows how the hydrogen:carbon (H:C) and oxygen:carbon
(O:C) ratios differ in the three kerogen types.

TABLE 1.7 The


characteristics of the
three types of kerogen.

Although the maturation product is controlled by the composition of the original matter,
the maturation of kerogen into crude oil and natural gas is achieved mainly by an increase
in temperature, and therefore the formation of petroleum depends on the geothermal
gradient in the crust. Figure 1.30 shows the relative proportions of crude oil and gas
formed at increasing depth below the surface, for a region where the geothermal gradient
is about 3.5C per 100 metres. The maximum crude oil production is about 2700 metres
and the peak for gas is at about 3500 metres.
Petroleum is probably generated from kerogen in very small amounts at temperatures
below 50C with the peak production at about 100C ; when the temperature is raised to
above 150C even for a short time, crude oil itself will begin to break down or crack
(Figure 1.29). The first gas to be produced contains some C4-C10 compounds and is
termed wet gas. But as the temperature increases these break down to give C1-C8
gaseous compounds and the product is then called dry gas. So the most important factor
in the generation of petroleum is heat, which depends on the geothermal gradient in the
sedimentary basins. Such gradients vary from basin to basin. Source rocks that remain at
very shallow depths will not normally generate petroleum: deeper burial, accompanied by
an increase in temperature, is required.
41

FIGURE 1.29 (left)


The generation of
crude oil and natural
gas from kerogen in an
area with a geothermal
gradient of about
3.5C per 100 metres.
Note that during the
maturation process
some of the kerogen
remains unaltered as
an insoluble carbon
residue.

FIGURE 1.30 (right) A


model showing a
possible relationship
between the time after
burial of source rocks
and the temperature for
oil and gas formation.

The time factor is also important in the generation of petroleum. It is assumed that
different volumes of petroleum would be generated from similar source rocks if they
were subjected to the same temperature for different lengths of time (Figure 1.30).

FIGURE 1.31
Reconstruction of burial
histories of rocks from two
basins.
(a) The Jurassic Paris Basin.
Rocks were buried to a
depth of just below 2500
metres, whereas the much
younger rocks of the late
Tertiary Los Angeles Basin
were buried to a depth of
3250 metres.
(b) The principle zones of
oil formation (brown) in
relation to depth in the two
basins.
The estimated temperatures
at which oil begins to form
are also shown.

So much for a theoretical model. Let us now examine two basins with different
geological histories (Figure 1.31). In both the Paris and Los Angeles basins the
threshold, or beginning of petroleum generation, has been dependent on depth and
42

temperature, and duration of burial. In the Paris basin, the threshold was reached after 40
million years when the early Jurassic rocks were buried to a depth of 1400 metres where
the temperature was 60C. The source rocks of the Los Angeles basin were buried to
2500 metres, where the temperature was 115C, before the threshold was reached. It has
been estimated that the petroleum in the Paris basin was generated over a period of 120
Ma; in contrast the generation of the petroleum in late Tertiary Los Angeles rocks, which
were buried quickly to a much greater depth, took less than 10 Ma. Because the
geothermal gradient, and the depth and duration of burial, can vary considerably from
basin to basin, so can the time taken for petroleum to be generated.
5.4 Migration of petroleum

Although the generation of petroleum in source rocks is fairly well understood, its
migration into other rocks is not. Two stages have been recognized in the migration
process (Figure 1.32): primary migration, is the movement of the hydrocarbons within
and then out of the source rocks, and this is followed by secondary migration into and
within the rocks where it accumulates, which are called reservoir rocks.

FIGURE 1.32 Primary


migration of petroleum
out of deeply buried
source rocks that have
reached maturation
temperatures, followed
by secondary migration
into sandstone resevoir
rocks.

The source rocks are mainly organic-rich shales, which are very fine-grained and
impermeable. So how can the generated petroleum move out of the source rocks? One
current suggestion is that when kerogen is subjected to maturation temperatures, methane
is formed which creates so much internal pressure within the source rocks that
microfractures are formed which then permit the primary migration of the hydrocarbons.
Having escaped out of the source rocks, the petroleum appears to move with greater
freedom along joints, faults and bedding planes into a reservoir rock such as sandstone.
5.5 Traps

There are three types of trap. Structural traps are formed as a result of the deformation
of strata in the Earth's crust. Stratigraphic traps occur where there is a permeability
barrier caused by lateral and / or vertical variation in sedimentary rock types. Some large
petroleum traps have been formed by both structural and stratigraphic means and these
are known as combination traps. Seventy eight per cent of the world's crude oil
resources are held in structural traps, 13 per cent in stratigraphic traps and 9 per cent in
combination traps (Figure 1.2).

43

5.5.1 Structural traps

The anticlinal trap is the commonest form of petroleum accumulation, and these traps
account for the largest volume of the world's resources. Anticlinal traps are common in
regions where the crust has been subjected to horizontal compression and they can be
large structures. Figure 1.33 shows an anticline; in three dimensions the structure is a
dome, in which the rocks dip downwards from the centre in all directions.

FIGURE 1.33 An
idealized block
diagram showing a
dome structure in
folded sedimentary
strata which form
concentric ridges at
the surface.
Section A-B shows
an anticlinal
structure with oil
trapped at two levels.

Two separate traps are shown in Figure 1.33, and they are there because of the vertical
continuity of the dome.
The limiting factors that determine the volume of petroleum in an anticlinal trap include
the following:
1

The size of the structure: anticlines can be hundreds of kilometres long and
thousands of metres in height;

The volume of the source rocks that have undergone maturation after the
structure was formed;

The porosity of the reservoir rocks;

The thickness of the reservoir(s);

The amount of closure on the structure, that is the position of the spill-point
relative to the height of the arched reservoir that lies above this level.

The last point is illustrated in Figure 1.34. Petroleum migrates up-dip from the source
rocks because of its low density and gradually fills a trap. Any gas present will always lie
above the oil because it has a lower density. A stage is reached when oil can escape from
the trap at the spill-point; such a trap is said to have only a limited closure.

44

FIGURE 1.34 The


migration of oil and
gas into and out of
an anticlinal trap. (a)
During the early
phase of secondary
migration the lower
level of the oil in the
reservoir has not
reached the spillpoint. (b) Later, the
spill-point has been
reached and oil can
flow further along
the reservoir to
accumulate in higher
traps, or escape to
the surface.

Anticlinal structures can also be formed in non-compressional areas of the crust and may
result from a drape of sedimentary rocks over an uplifted or faulted older rock; an
example of this structure is provided by the Forties field in the North Sea. Other anticlinal
structures can be formed by salt plugs. These develop in regions underlain by evaporite
beds rich in halite (salt, NaCl). When it is deeply buried and under high pressure, salt
behaves as a plastic material within denser rocks. Its relatively low density means that it
tends to rise towards the surface in elongated cylindrical plugs or pillars of salt. Salt
plugs, which are commonly about 1 km in diameter, are impermeable to petroleum, and
in suitable areas the upward moving plug will pierce the overlying strata with the result
that traps are formed if porous reservoir sands are present (Figure 1.35). Because some
salt pillars can be up to 3000 metres in height, there may be many reservoirs and hence
many traps. Associated with the upward movement of salt are faults in the surrounding
rock, and traps can also be formed by these faults (see below). Above the salt plug is the
cap rock, a series of layers of anhydrite, gypsum and limestone, which may contain
cavities and show evidence of fracturing. In some instances the cap rock acts as a
reservoir for oil and gas. The exact origin of the cap rock is unknown although there are
several hypotheses for its formation.

FIGURE 1.35 A salt


plug and associated
oil traps.

Fault traps are formed when an inclined bed of reservoir rock is brought into
contact with impermeable strata by a fault. Some fault traps are illustrated in
Figure 1.36, where the reservoir rocks are beds of sandstone, above and below

45

which are thicker beds of impermeable shale that act as seals. Where the complete
thickness of the sandstone reservoir has been faulted against a shale there is said to be
unlimited closure; in other cases the closure is limited and there are spill-points.
FIGURE 1.36 Various
kinds of fault trap.

So far we have discussed traps that have been formed mainly as a result of folding or
faulting of rocks. The next category of traps normally occurs in relatively undeformed
sedimentary basins.
5.5.2 Stratigraphic traps

From modern analogues we know that a sandy bed deposited near shore can pass laterally
into muds deposited in deeper water. Similarly some sandstones are discontinuous within
shale deposits and appear like lenses or wedges in cross-section (Figure 1.37a; note that
gas oil and water are shown in separate layers and this is always the relationship because
of their relative densities). When wedge-shaped reservoirs are inclined, petroleum can be
found at the upper end in what are known as pinch-out traps (Figure 1.37b).

FIGURE 1.37 Two


stratigraphic traps.
(a) Lens-shaped
sandstone fillred with
gas, oil and water.
(b) Two pinch-out traps
containing gas and oil.

46

Reefs similar to modern coral reefs have existed in the geological past. Like modern reefs
the fossil reefs can be circular in outline and are composed of limestone made up of
corals and shell fragments, which are later dissolved to give secondary porosity. When
reefs become buried underneath impermeable sediment, such as mud, which later
becomes compacted to form shale, they form reef traps. Such traps are restricted to rocks
formed in former tropical and sub tropical climates and account for only 3 per cent of the
world resources of petroleum.
The final example we shall examine is the unconformity trap. A period of uplift and
erosion results in tilted beds below the unconformity overlain by nearly horizontal beds,
which can act as a seal. Major unconformities extend over large areas of the Earth and
thus might be expected to have trapped large accumulations of petroleum. So what is the
main reason why unconformity traps contain only 4 per cent of the world resources?
Large volumes of petroleum probably escaped when the older series of rocks were
uplifted, so although the younger strata above the unconformities form suitable seals,
there is no oil left to seal in.
5.5.3 Combination traps

Combination traps normally occur where reservoir rocks have been folded and
subsequently eroded and sealed by the deposition of shales on top of an unconformity;
note that the petroleum can have migrated into the trap only after the reservoir had been
sealed by the much younger rocks. Figure 1.38 illustrates such a combination trap of the
type found in the Prudhoe Bay field in northern Alaska. The trap is an east-west anticline
which was tilted westwards and eroded; the overlying shales were deposited much later
and only after this seal was deposited did burial become deep enough to generate the oil
from source rocks.
FIGURE 1.38 A
combination trap.
(a) The reaervoir rocks
are folded into a dome;
(b) tilting followed by
erosion exposes the
reservoir rocks; (c)
shale deposited on top
of the unconformity
acts as a seal and
creates a trap.

5.6 Natural gas

This Section may appear to be a digression at this stage, but it is important to discuss how
natural gas is related to other hydrocarbons. It is possible to trace a crude oil back to its
source by comparing its geochemical analysis with that of hydrocarbon extracts from a
particular source rock. But what of a gas source? If the gas is purely methane then it is
virtually impossible to determine its exact source and hence its origin.
There are different kinds of natural gas, with different compositions and modes of
formation. Methane is produced by bacteria in the very early stages of accumulation and
anaerobic decay; it occurs at low pressures, which are approximately atmospheric
pressures, and normally escapes into the atmosphere as marsh gas' or can be trapped at
shallow depths. Now let us consider the natural gas that accumulates in traps, with or
without crude oil.
47

Coal is an important source of natural gas, and it has been established that coals are able
to generate large enough volumes of methane for commercial accumulations to form.
Although small amounts of liquid hydrocarbons are generated during coal formation,
very few commercial oil accumulations have been traced to a coal origin. It has been
suggested that if liquid hydrocarbons are formed during the maturation of coal, only
limited primary migration is possible because oil is rapidly absorbed by coal.
Because coal is formed from terrestrial organic matter it has a strong similarity to type III
kerogen (Table 1.7), which yields gas rather than oil during maturation. If you refer back
to Figures 1.28 and 1.29 you can see that during the maturation of any kerogen, natural
gas and crude oil can form simultaneously. Natural gas dissolves in crude oil under
pressure, but comes out of solution as soon as the pressure is released (e.g. by a
borehole). In the later stages of maturation the crude oil can be converted into wet gas,
which is a mixture of oil and gas, and finally with increased temperature all the oil is
converted into dry gas.
5.7 Solid petroleum and oil shale

Crude oil that has migrated through a reservoir can sometimes reach the surface, where
the volatiles escape and the rock becomes impregnated with a hydrocarbon residue; such
a rock is known as an oil sand or tar sand. The Athabasca oil sands of Alberta, Canada,
are such a deposit, and they contain 600 x 109 barrels of asphaltic hydrocarbons. In this
case the oil first accumulated in Devonian reef traps from which it escaped and entered a
permeable Cretaceous sandstone, with a westerly dip, which was later exposed at the
surface. Because of this exposure, the oil lost most of its volatiles and became too viscous
to flow; the rock has to be heated to extract the oil. The Athabasca oil sands are mined by
open cast methods.
Pitch lakes are large deposits of asphalt formed from oil that has seeped to the surface,
accumulated in large depressions and become solid. Examples include deposits in
Trinidad, and the Bermudez Lake in Eastern Venezuela with an estimated reserve of 6.1
million tones.
A shale that contains 2.5 per cent or more of unaltered kerogen is termed an oil shale,
and when heated to about 500C it yields crude oil; because it has to be produced by
heating, the product is termed synthetic' crude oil. The energy required to heat the oil
shale to that temperature is about 1 kJ per gram of rock and the calorific value of kerogen
is about 40 kJg-1; therefore if the kerogen forms only 2.5 per cent of the rock its total
calorific value is taken up heating the shale. To be economic the lower limit is frequently
taken as 5 per cent kerogen in the USA, and this yields 25 litres of synthetic crude per
tonne of rock.
There are major economic problems in the extraction of synthetic crude oil, but it is still
being distilled in the Baltic region of the USSR and by the Chinese in Manchuria.
Oil shales have been deposited in lakes, shallow seas and lagoons from Precambrian to
Cenozoic times, and some of the largest deposits are located in the American states of
Utah, Wyoming and Colorado. It has been estimated that the oil shales could yield about
18000 x 109 barrels of oil, but because of the escalating costs, and a shortage of available
water for processing, most of the projects involved with the production of synthetic crude
oil were temporarily abandoned in the early 1980's.

48

SECTION 6
THE EARTH'S STRUCTURE AND SURFACE FEATURES
Study Comment

This section reviews briefly the structure of the Earth, the nature of its crust and the
major surface features, including topography, age distribution of rocks and
earthquake and volcanic activity. Space does not permit a detailed discussion of the
evidence on which the Earth's structure is based. A basic knowledge of the features
described in this section is necessary in order to understand the way in which plates
of relatively rigid crust interact with each other to produce geological structures,
including sedimentary basins and mountain belts.
6.1 A simple model of the structure of the Earth

The density of the whole Earth has been calculated to be 5.5 x 103 kg m-3, yet rocks that
constitute its crust, such as sedimentary rocks, granites and basalts, have densities
ranging between 2.5 x 103 to 3.5 x 103 kg m-3. Therefore, rocks towards the centre of the
Earth must have a larger density than that for the whole Earth.
Studies of the way in which seismic waves are refracted through the interior of the Earth
indicate that the Earth is composed of three layers: the crust, mantle and core (Figure
1.39).
FIGURE 1.39 A
simplified cut-away
section showing the
internal structure of the
Earth.

The Earth's crust varies in thickness, from about 6 km at its thinnest, to about 90 km at
its thickest and is, on average, 35 km. But on the scale of the whole Earth (radius about
6370 km) this is very small, so the crust is relatively very thin. As you know from
looking at the world around you, most of the top of the crust is fairly rigid and cold (for a
rock 50C - the maximum temperature of the Earth's surface - is cold, and the places
where molten rocks reach the surface are few in number). When crustal rocks are
subjected to strong and long-lasting forces, or when they are warm, rock beds may bend
and crumple to produce folds. But when pressures are exerted over short time-scales, rock
beds may fracture and break, such as when a bar of cold toffee breaks, or the substance
silly putty'.
49

Below the crust is a very thick, denser layer, which occupies nearly 70% of the Earth's
volume and is called the mantle. It is considered to be composed of peridotite. The
mantle itself is subdivided into a number of shells, which are described below. In the
centre of the Earth there is a core, which is even denser and, we believe, metallic rather
than rocky. In part this accounts for the increase in density towards the centre of the
Earth. The outer part of the core is liquid. Another important factor is that the Earth's
magnetic field appears to originate within the core; consequently the study of the Earth's
magnetism provides information about the core and its properties.
It might help you to get a feel for the internal structure of the Earth by thinking of it as a
spherical, unripe avocado pear. In this model, the spherical avocado pear model, the thin
skin of the avocado pear represents the crust, the flesh represents the mantle, and the
large, hard stone at the centre represents the core. Even the deepest borehole when scaled
down to the avocado pear model is only akin to scratching the skin of the fruit to find out
what is inside!
6.2 Layers within the crust and mantle

Detailed studies of compressional seismic wave travel times have produced the profile of
speed varying with depth shown in Figure 1.40. Notice that there is a zone from 400 km
to 1050 km depth across which the seismic-wave speed increases, with some distinct
steps, from 8 km s-1 to 11 km s-1. This zone of change is known as the transition zone
and its base separates the upper mantle from the lower mantle.
FIGURE 1.40 Graph of
compressional seismic
wave speed against depth
in the upper 1200km of the
Earth's mantle (continuous
line).
The dashed part of the line
represents the speed in the
outermost layer of the
Earth (the crust).

Experiments on peridotite have shown that under high pressure the internal structure of
the rock changes so that the constituent particles become more compacted. It is thought
that the three steps in the seismic-wave speed profile across the transition zone are due to
stepwise jumps in the density of peridotite without any change in composition. Such a
change in structure is known as a phase change. A more familiar example of such
structural differences is the contrast between diamonds, density 3.5 x 103 kg m-3, and
graphite (pencil lead), density 2 x 103 kg m-3. Diamond is a compressed form of
graphite that is produced naturally at depths of nearly 200 km within the Earth's mantle,
and then brought to the surface by volcanoes. If diamonds are subjected to very high
temperatures at atmospheric pressure they can be changed into graphite. (We do not
recommend
50

this experiment!) The changes from ice to liquid water and from water to steam are also
phase changes with no change in composition but these are also changes of state (from
solid to liquid and from liquid to gas).
Figure 1.40 also shows a dip in speed, which is shown at about 100 km depth. In 1926,
the seismologist Beno Gutenberg noted that seismic waves from earthquake foci at depths
between 50 and 250 km took longer to arrive than they should have done on the
assumption of a homogeneous upper mantle. He suggested that there was a layer between
these depths, which had a markedly lower compressional seismic wave speed than
expected, and this came to be known as the low-velocity layer though, strictly, it is a
low-speed layer. This layer is thought to be composed of a slush' of peridotite, in which
about five per cent of the rock is molten. It is probable that basaltic magmas erupted at
volcanoes originate from this zone. A Yugoslavian geologist, Andrija Mohorovicic
(pronounced mo-horovichick), studied seismograms for shallow-focus (less than 40 km
deep) continental earthquakes recorded over low distances up to 800 km from their
epicentres (the point on the Earth's surface above the site of the earthquake). The
surprising results showed that one set of wave arrivals was followed soon after by another
distinct set. Mohorovicic's idea is illustrated in Figure 1.41. One set, the direct' set of
waves passes through an upper region of fairly uniform, low wave speed. The other set
travels steeply downwards, is strongly refracted away from the normal at a boundary with
higher-speed material and, after traveling through this material for much of the distance,
it re-emerges to be refracted towards the normal and thence up to the receiver at the
surface. The boundary between the two layers is known as the Mohorovicic
discontinuity, or Moho for short. It varies in depth between 5 and 11 km beneath the
ocean floor but has an average depth of 35 km beneath continents, and is nearly 100 km
deep beneath high mountain ranges (Figure 1.42). The crust is defined as the layer of the
Earth above the Moho, and there are important differences between oceanic crust which
floors all the great ocean basins, and the continental crust beneath the continents and the
adjoining continental shelves at the edges of the oceans.

FIGURE 1.41 The


Mohorovicic
discontinuity (Moho).
The Figure shows how
both a direct seismic
wave and a seismic wave
refracted at the Moho
will be recorded at
distances up to several
hundred kilometres.

One of the results of the seismic investigation of oceanic crust has been to show that all
the deep ocean basins of the Earth are floored by rocks which, beneath a thin veneer of
sedimentary rocks, are all very similar, consisting of basalts and gabbros.

51

FIGURE 1.42
Schematic illustration
of the oceanic and
continental crust
showing the variations
in thickness in an eastwest cross-section
from the Pacific Ocean
to Africa.

The oceanic crust is very thin compared to the continental crust (Figure 1.42) and also
has a more homogeneous composition. Many of the lowland areas of the continents are
geologically young and expose sedimentary rocks at the surface. These overlie older,
more compacted sedimentary rocks together with igneous and metamorphic rocks. Where
these older and harder rocks are exposed at the surface, their resistance to weathering and
erosion by rain and snow results in mountainous landscapes. The dominant rocks in these
areas are granites, together with a wide variety of metamorphic rocks. These rocks are
actually typical of the deeper parts of the continental crust, but geological processes have
caused uplift of this crust and erosion has subsequently revealed these deeper rocks for
our inspection.
So the geological investigation of older uplifted parts of the continental crust which have
been eroded, together with seismic investigations and many other techniques, have
enabled us to arrive at a quite detailed picture of the nature of the continental crust. The
whole continental crust has an average composition which is quite close to that of
granite, and which is markedly different from the basaltic/gabbroic composition of the
oceanic crust.
6.3 Elevations of the surfaces of continental and oceanic crust

The originator of the continental drift hypothesis, which will be discussed in the next
section, was Alfred Wegener, a German meteorologist. Wegener was struck by the
distribution pattern of elevations of the surfaces of the continents and ocean basins. This
pattern is shown in Figure 1.43 which shows that there are, as Wegener described two
preferential levels for the world's surface which occur in alternation side by side and are
represented by the continents and ocean floors, respectively Wegener expressed surprise
that no previous workers had commented on this pattern. At the time, conventional
wisdom had it that different elevations of the earth's solid surface were caused by uplift
and subsidence of one original level (that is the top of a crust with uniform composition
laterally - Figure 1.43(c)) caused probably by forces resulting from the Earth's
contraction. But Wegener argued that if this were the case, the pattern of elevations
should be that shown in Figure 1.43(b) - a simple distribution with one peak, not two.
Wegener concluded that the double-peak pattern reflected two distinct layers in the
earth's crust, one more dense than the other: To put it in rather picturesque terms, the two
layers behave like open water and large ice floes (see Figure 1.43(d)). Wegener's
description and analogy are now believed to be correct. As outlined in the preceding
section, the ocean floors are underlain by comparatively denser crust, whereas the
continental crust is less dense. Indeed, Wegener's analogy of ice floating in water can be
taken further in the light of modern knowledge, for just as 90% of icebergs are
submerged beneath the sea, so the earth's high mountain belts have deep crustal roots (see
Figure 1.42).
52

FIGURE 1.43 (a) Bar


chart showing the
percentage of
distribution by elevation
of the Earth's solid
surface.
(b) Curve 1 represents
data presented in (a) as a
smoothed line; curve 2
shows the distribution
pattern that would be
expected if one original
level had been uplifted
and subsided (see c). (c)
A crustal model
involving the uplift and
depression of a crust of
uniform density: this
would result in curve 2
on (b). (d) Wegener's
crustal model to explain
the double peak
distribution pattern
(curve 1 in b).
Note that at the present
time, the world's oceans
lap over continental
crust, and for this reason
Wegener did not try to
match present-day
coastlines on either side
of the Atlantic.

Wegener used his model of the two-layered crust to further his argument for continental
drift in two ways. Firstly, as it was known that Scandinavia was rising at the rate of about
one metre per century consequent on the melting of a thick ice cap (which is further
evidence in favour of the crustal model shown in Figure 1.43(c)). Wegener suggested that
if such vertical movement were possible, then why not horizontal movements as well?
6.4 Earth patterns

Figure 1.44 is a cross section from the Pacific Ocean, across South America and the
South Atlantic Ocean to Africa showing examples of the major topographic features
developed over continental and oceanic crust. Figures 1.45-1.48 show the global
distribution of ocean ridges and trenches, ages of continental rocks, volcanoes and
earthquakes. There are relatively simple patterns and relationships to the discerned from
examining these maps. Using these maps, and your knowledge of the distribution of
mountain belts, answer the following questions before reading on.
FIGURE 1.44 Crosssection to show the
surface of the Earth's
crust between South
America and Africa.

53

FIGURE 1.45 Map showing the distribution of ages of rocks on the continents formed during the past phases of mountain building.
1 Rocks formed more than 1000 million years ago; these regions, termed cratons, have remained unaffected by Earth movements for a
considerable period, and they form the stable `nuclei' of the continents. Such regions exhibit a relatively subdued relief in contrast to
the rugged topography associated with `young ' mountain belts such as the Alps, Andes, Himalayas and Rockies (see 5 and 6 below).
2 Rocks formed 1000-600 million years ago.
3 Rocks formed 600-400 million years ago.
4 Rocks formed 350-250 million years ago.
5 Rocks formed 200-100 million years ago.
6 Rocks formed 60 million years ago to the present.
You will see from the above intervals of rocks that there were gaps between the main phases of mountain building.

FIGURE 1.46 Map showing the distribution of active subaerial volcanoes.


Black dots show the occurrence of effusive volcanoes, whose activity is dominated by outpourings of liquid lava. Blue dots
signify explosive activity, produced by the release of large amounts of gas as the lava nears the vent of the volcano; the
resultant rock types include volcanic ash and lava containing abundant gas bubbles.

54

FIGURE 1.47 Map showing the distribution of all shallow (less than 100km deep) earthquake foci recoded between 1961
and 1967.

FIGURE 1.48 Map showing the distribution of intermediate-focus (100-300km depth, black dots) and deep-focus (300-700km
depth, blue dots) earthquakes recorded between 1961 and 1967.

55

QUESTION 1
What is the difference in age of the rocks that occur at the surface of continents and
ocean basins?
QUESTION 2
What kind of volcanic activity characterizes ocean ridges and mountain belts?
QUESTION 3
What kind of earthquake activity (in terms of depth) occurs beneath of adjacent to:
a)
b)
c)
d)

ocean ridges;
ocean trenches;
mountain belts;
cratons (see caption to Figure 1.45 for explanation of this term) and areas
between ocean ridges and mountain belts/ocean trenches (comment on oceanic
and continental areas).

Spend about half an hour answering these questions before reading the answers.
Answers
1)

Oceanic rocks are relatively young (less than 180 million years: see caption to
Figure 1.45) whereas parts of the continents - the cratons (see Figure 1.45) are
extremely old in comparison (older than 1000 million years). This contrast
emphasizes the difference between continental and oceanic crust that was
stressed by Wegener, and undoubtedly points to a different origin for each.

2)

Ocean ridges are everywhere characterized by effusive volcanic activity, but


mountain belts show explosive volcanic activity. Note how the Pacific is ringed
by explosive activity - often termed the ring of fire'. Note too, how this activity
generally occurs on the landward side of ocean trenches (compare Figures 1.45
and 1.46).

3)
a)

Occurrences of earthquakes.
Ocean ridges are characterized by shallow earthquake activity. The ridge running
down the middle of the Atlantic is marked by a belt of shallow activity (see
Figures 1.45 and 1.47).

b)

Ocean trenches are always associated with intense earthquake activity as shown
by comparing Figure 1.45 with Figures 1.47 and 1.48. On a larger scale (not
possible to detect on these figures) the activity occurs on the side of the trenches
away from the ocean ridge. Moreover the earthquakes become deeper in this
direction; this relationship can be seen on Figure 1.46. The trenches off the
Andes (Figure 1.45) are flanked to the east by a belt of earthquakes at 0-100 km
(Figure 1.48) and 100-300 km depths (Figure 1.47), and further to the east,
deeper earthquakes occur (300-700 km; coloured dots on Figure 1.48). The same
relationship can be seen north of New Zealand, and in the Japanese area, but in
these cases, earthquake depths increase to the west, but still away from the
Pacific ocean ridges.

56

c)

The mountain belts shown on Figure 1.45 are all zones of frequent earthquake
activity (Figures 1.47 and 1.48).

d)

Cratons and both continental and oceanic areas between ocean ridges and
mountain belts and ocean trenches, are largely devoid of earthquake activity.
Indeed, examination of Figures 1.45 and 1.46 show that there are quite distinct
zones of volcanic and earthquake activity separated by large quiescent areas. The
significance of this distribution will become apparent as you read the next
section.

The patterns and relationships described above are manifestations of plate tectonics and
movements of material in the Earth's mantle, which are the subjects of the next section.

57

SECTION 7
PLATE TECTONICS
Study Comment

The purpose of this section is to introduce the concept of plate tectonics and show
some of the evidence that supports it. You will see that the concept explains many of
the features of the Earth's crust reviewed in the previous section. It also permits the
classification of sedimentary basins into eight types, and has predictive value in
enabling basins with hydrocarbon potential to be identified.
7.1 The model

The plate tectonic model (Figure 1.49) envisages plates of rigid lithosphere about 100 km
thick overlying a layer of relatively low strength. Mantle material rises beneath
constructive margins, and plunges back into the mantle at destructive margins. Along
a third type of margin, crustal plates slide past each other, forming conservative
margins. Figure 1.50 shows the distribution of the principal lithospheric plates, and types
of earthquake occurring along their boundaries. Figure 1.49 uses the terms lithosphere
and asthenosphere. The former term is used to describe the crust and top of the mantle,
which together behave as relatively rigid plates. The asthenosphere is a zone of plastic
deformation beneath the lithosphere. It permits both lateral movement of the plates, and
vertical crustal movements, which compensate changes in crustal thickness during, plate
movements, or loading and unloading when thick ice-sheets form or melt.

FIGURE 1.49 Diagram


showing the basic concept
of plate tectonics.

58

FIGURE 1.50 The distribution of lithospheric plates and the seismicity at their boundaries.
All the plate boundaries are regions of shallow seismicity; deeper-focus earthquake zones mark the sites of
destructive plate boundaries. Spreading rates at constructive plate boundaries are shown schematically by
the width between the parallel lines used to show them. The directions of plate movements are shown by
arrows, the length of which are proportional to the rate of movement. These plate motion directions have
been calculated by assuming that the African plate remains stationary. Minor plates are numbered as
follows: 1 Arabian: 2 Philippine: 3 Cocos: 4 Nazca: 5 Carribean.

The plate tectonic theory is so elegantly simple that it may seem surprising that before the
1970's, the majority of geologists believed that earth movements were largely vertical in
nature, and that the oceans were permanent features of the Earth. New evidence
supporting the continental drift hypothesis, and exploration of the ocean floors showed
that lateral crustal movements are extremely important.
7.2 Continental fit and continental drift.

The beginnings of plate tectonics are rooted in continental drift. The first modern
protagonist of drift was a German meteorologist, Alfred Wegener. His book "The origin
of the continents and oceans" was first published in 1915.
The evidence presented by Wegener in favour of the continental drift hypothesis can be
divided into several groups: the fit of continental masses, records of past climatic changes
(including glaciation), crustal structure (as described in Section 5.3), direct measurement
of increasing distance between continents, and the present day distribution of certain land
plants and animals. The last two categories of evidence will not be discussed here,
because they are not regarded as valid today - indeed it is only with the advent of
artificial satellite observations that direct measurements of intercontinental distances to
an accuracy of a centimetre or so have become possible.
59

Wegener cited three lines of evidence in favour of the continents bordering the Atlantic
having once been united. These concerned (i) the jigsaw fit of the outlines of the edges
of continental crustal areas (see Figure 1.51), which (ii) also enabled portions of old fold
or mountain belts to be nicely matched (i.e. the picture on the jigsaw pieces match up),
and (iii) the distribution of certain fossils to be explained without invoking the former
presence of land-bridges (Figure 1.52)

Wegener collected evidence concerning the distribution of rock types that indicated the
nature of the climate at the time they were formed. The formation of these rocks (and still
is today) interpreted using the uniformitarian approach. Thus coal is considered to form
from the accumulation of plant material that grew in very wet hot conditions, such as
those in present-day equatorial regions. Salt and gypsum are minerals that precipitate
from seawater when it is evaporated, usually, desert coastal areas. Desert sandstones
contain beautiful rounded and frosted sand grains, and can easily be recognized in the
rock record. Ice, on melting, leaves behind a characteristic deposit consisting of a
jumbled mass of boulders, pebbles and clay.
Note how, on Wegener's reconstructions (Figure 1.53), the occurrences of coal are on or
near the equator, and indications of hot, dry conditions flank this area to the north and
south, just as African equatorial forests are flanked today by the Sahara and Kalahari
Deserts. Without invoking continental drift, the south polar ice cap shown in Figure 1.53
(a) would have covered a huge area of the globe - from Australia, through India and
South Africa to South America.
60

FIGURE 1.51 (left)


Wegener's Atlantic
continent reconstruction.
Old mountain ranges
(the ages of which were
not known in Wegener's
time: they are now
known to be up to 600
million years old in
North America and
Europe, and about 200
million years in the
southern continents) are
shown in blue and form
continuous belts when
the continents are fitted
together. Note that
Wegener has not fitted
the coastlines, but the
rough positions of
continental margins.

FIGURE 1.52 (right)


Wegener argued that the
distribution of fossils of
Mesosaurus, a small
reptile that lived over
200 million years ago in
shallow brackish (a
mixture of salt and
freshwater) waters in
South America and
Africa, could best be
explained by involving
continental drift. The
`conventional'
explanation was that a
`land-bridge' connected
the two continents and
that this later subsided
beneath the Atlantic
Ocean.

FIGURE 1.53
Wgener's maps showing
climatic belts as they
may have existed during
the Carboniferous and
Permian periods of
geological time. (a)
Approximately 300
million years ago. (b)
Approximately 250
million years ago.
The climatic indicators
are: I, Ice; D, Desert
sandstone; S, Salt; G,
Gypsum; C, Coal;
stipple, postulated
arid/desert areas.

7.3 Palaeomagnetism

Some sedimentary and igneous rocks contain evidence of the orientation of the Earth's
magnetic field at the time of their formation (Figure 1.54). Measurements of the magnetic
properties of such rocks enable the location of past magnetic poles to be determined.
FIGURE 1.54 Fossil
magnets.
On the left magnetic
mineral grains settling
through water
accumulate on the
bottom as microscopic
compass needles, so that
the resultant sediment,
and later rock, contains
within it, a record of the
orientation of the Earth's
magnetic field. On the
right molten lava cools
and minerals solidify
from it. Past a critical
temperature they become
magnetized in the
direction of the Earth's
field then prevailing.

From the mid 1950's onwards the results of palaeomagnetic studies yielded evidence
that forced geophysicists to reconsider continental drift as a possibility, despite their
reservations that the physical properties of the earth's crust seemed to make it
impossible. Determinations of apparent palaeopole positions over the past 500
million years of earth history showed that the position of the magnetic poles had
apparently changed considerably during this time. Observations from rocks of the
same age from the same continent were found to be clustered together, and so the
average readings gave the palaeopole position. A line connecting successive
palaeopole positions is known as a polar wandering curve. As the earth's magnetic
field is considered over a period of time to have coincided with the earth's axis of
rotation, such curves either imply migration of the continents through latitudes, or a
61

shift in the position of the earth's rotational axis. Wanderings of the geographic poles,
that is a change in the position of the axis of the earth's rotation, was considered possible,
but this would produce magnetic polar wandering curves that are the same for each
continent. But the data showed that the polar wandering curves from different continents
(which are drawn without assuming the drift has occurred) could only be reconciled by
invoking continental drift (Figure 1.55).
FIGURE 1.55 Polar
wandering curve for
Europe and North
America, with the
present North pole
shown in the centre
of each map.

(a) The curve for Europe shows a steady shift of the palaeopole position (as measured with reference to the present location of Europe)
from near the Equator to the present North Pole. This observation can either be interpreted as being due to the latitudinal migration of
Europe through time, or being caused by changes in the position of the Earth's rotational axis. A single curve such as this is not
evidence of continental drift.
(b) When the curve for North America is added, a similar shift is revealed, but the curve is not the same shape. The main difference
occurs between 200 and 100 million years ago, suggesting that the two continents drifted apart during this period. The difference
between these two polar wandering curves can only be reconciled by invoking continental drift.
Polar wandering curves indicate movement only in a relative north-south sense, and provide no information about the longitudinal
position of past fits of slabs of continental crust. Nonetheless, they can be used to aid reconstructions of past fits of slabs of continental
crust

Thus the palaeomagnetic evidence confirmed Wegener's idea that continents had moved
considerable lateral distances. The fact that they migrate across climatic belts has
significance for petroleum exploration. For example, the gas fields of the southern North
Sea owe their origin to the region's northward drift. The coals that form the source of the
gas were laid down when the region was situated near the equator. As it migrated
northwards into an arid zone (akin to the Sahara today), desert sands were formed, which
later became reservoirs. When the area was flooded by the sea, it was still in an arid zone,
and so the thick evaporites of the Zechstein seal unit were formed.
7.4 Sea floor spreading

The shape of the ocean floors, which for a long time defied explanations in terms of
land-based geology, is also significant for continental drift and plate tectonics. The
existence of the Mid-Atlantic Ridge was known in the late nineteenth century as a result
of surveys conducted by cable laying ships, but Pacific Ocean ridges were only
discovered in the late 1950's. Similarly, the axial valley or rift zone of the ridges and the
incredible symmetry displayed by parts of them (as shown in Figure 1.56) are relatively
recent discoveries.
As shown in Figure 1.56(c) ocean ridges are areas of tension, and are regions of relatively
high heat flows compared to other crustal areas - just the phenomena to be expected over
rising convection currents in the mantle.
62

FIGURE 1.56
Topographic profile
(a) across the midocean ridge in the
North Atlantic, (b)
across an ocean ridge
in the Pacific and (c)
idealised sketch
showing the structure
of the axial rift zone.

Note the much more `rugged' topography of the Atlantic ridge and its marked symmetrical (or mirror image across the central rift)
appearance. The pacific example shows a degree of symmetry, but it is much less that that for the Atlantic. The Mid-Atlantic Ridge
develops a central rift valley that runs along its length; it is broadly comparable in topography to continental rift valleys, such as
the Rhine rift valley and those of East Africa. The pattern of geological faults that produce these rifts is shown on (c); such features
are produced by tension (pulling apart) of the crust. Note that (a) contains three gaps due to the incomplete nature of the survey on
which it is based.

During the 1950's and early 1960's, samples recovered from the ocean floors had yielded
ages never in excess of 180 million years, a remarkable contrast to the huge age range of
rocks shown on the continents. Thus the evidence mounted that ocean crust had different
origins to that underlying the continents; it was denser, it was younger, and its surface
formed a relatively simple symmetric pattern. With the collection of data concerning the
magnetic properties of the ocean floors, the differences between the two types of crust
became starker still.
Early magnetic surveys conducted in the 1950's showed that the central rift valley of the
Mid-Atlantic Ridge was marked by higher than normal magnetic values, and more
detailed work showed that ocean floors are characterized by a striped pattern of higher
and lower than average magnetic values quite unlike anything seen on continents (which
generally show curved or even blob-like patterns).
At the same time as the new data was being obtained from the oceans it was also
discovered that the polarity of the earth's magnetic field had switched periodically, so that
during periods of reversed polarity, a compass needle would point to the south magnetic
pole. The reversal timetable for the past few million years is shown on Figure 1.57. It was
also soon realised that it was the magnetic reversals that had caused the magnetic stripes
running parallel to ocean ridges. As newly formed ocean floor cooled, the magnetic
polarity at the time was frozen into the basalts. The formation of new ocean floor is
referred to as sea floor spreading.
The sea floor spreading model enabled the age of the ocean floor to be deduced, as the
magnetic stripes could be matched to the polarity timetable. The age of the ocean floor
shown on Figure 1.57 was verified by radiometric dating of ocean floor basalts and
palaeontological dating of the sediments immediately overlying them.

63

FIGURE 1.57 Map showing the age of basins, as determined by studies of magnetic stripes.
Blank areas are those in which there are insufficient data to support the interpretation of ocean-floor ages.

The fact that the ocean floor is continuously forming and none is older than 200 million
years implies that somewhere it is being destroyed or subducted. Overlying the sites of
destruction are ocean trenches and mountain belts.
7.5 The model - more details
7.5.1 Constructive plate margins

Oceanic crust covers about 70 percent of the Earth's surface and has a remarkably simple
structure (Figure 1.58(b)).
Figure 1.58 summarizes the main features of constructive plate boundaries. It must be
stressed that this is a hypothetical model, based both on investigations of ocean ridges
and on studies of slices of presumed oceanic crust that have been thrust up into
continental regions.
Figure 1.58(a) shows the ocean ridge sited over the rising limbs of two mantle convection
cells. Mantle material is shown rising under the ridge, filling the gap' as the two plates
move away from each other, and by so doing, actually grow. But what is the process by
which they grow? And how is the characteristic layered structure of oceanic crust
produced? Before speculating on the processes that might be operating beneath ridges, a
little more must be said about the nature of the layers.
Deep-sea drilling has only penetrated layer 1, and a little way into layer 2. Layer 1
consists of sedimentary rocks, formed either from the dead remains of organisms
living in the surface waters of the ocean, or from chemical precipitates on the sea
bottom, such as manganese nodules. Layer 1 may also contain volcanic ash that
has settled out through the ocean water, but is remarkable for its lack of

64

sedimentary particles derived from the continental crust. It thins towards the ridge crest
because sediment only starts to accumulate after oceanic crust was formed (that is layer 2
and below).
FIGURE 1.58 The
features and processes
characteristic of
constructive plate
boundaries. (a) Block
diagram of the crust
and the mantle beneath
a spreading axis (ocean
ridge). (b)
Enlargement of part of
(a) showing an
interpretation of the
crustal and mantle
processes operative
during sea-floor
spreading, and the
nature of the layers
produced.

Layer 2 has a fairly uniform thickness of just under 2km, and is exposed in the vicinity of
ocean ridges. Here, its surface is seen to consist of basalt pillow lavas (the pillow shape is
characteristic of lavas formed underwater), and deep-sea drilling confirms the view that
the whole of layer 2 is largely composed of such lavas.
Studies of rocks now exposed on continents, but believed to represent oceanic crust,
suggest that layer 3 can be divided into two parts. The upper part is probably formed
of vertical sheets of igneous rock, again of basaltic composition, known as dykes.
The structure of this layer is analogous to that of a pack of cards stood vertically,
each card representing a single dyke. It indicates that the crust was experiencing
extension. The upper boundary of the dykes with the lavas of layer 2 is not abrupt,
for some dykes penetrate the lavas. The lower half of layer 3 consists of gabbro, a
rock with the same chemical and mineral composition as basalt, but with a larger
crystal size. The larger crystal size of the gabbro is due to the fact that it cooled more
slowly than basalt. In the crustal environment we are discussing, this is hardly
surprising, as the gabbros must have formed several kilometres below the basaltic
lavas.
65

It is possible that the upper part of the gabbros of layer 3 is uniform in structure in
contrast to a layered base.
The base of layer 3 is of course marked by seismic discontinuity - the Moho. Layer 4 is
generally agreed to be composed of peridotite, which has a greater density than gabbro
or basalt. the top kilometre or so of layer 4 is thought to be composed of layered
peridotite.
The kind of rock sequence described above can be seen in continental regions in many
parts of the world. All these examples occur in regions interpreted as ancient
destructive plate margins in which slices of oceanic crust, instead of being subducted
downwards into the mantle, have been thrust upwards onto continental crust. What
kind of igneous process could form the rock sequence just described? Any hypothesis
must not only account for this sequence itself, but also the uniform thickness of layer
3, and of course for sea-floor spreading itself!
Figure 1.58(b) shows a large chamber filled with liquid rock of basaltic composition
(magma). The existence of this body of liquid is postulated both because seismic wave
velocities beneath ocean ridges are anomalous (lower than in normal mantle), and
because it is a model that can produce a layered structure.
The first crystals to appear in the magma are less dense than the surrounding liquid,
and so sinks to the bottom of the magma chamber. It is thought that these crystals
sinking to the bottom of the magma chamber account for the densest layer of peridotite
accumulating at the bottom of the chamber. Layering in both the peridotite and the
gabbros is produced by convection currents within the chamber, which at times may
prevent all but the densest crystals settling. During periods when there is no
convection, crystals of all densities will settle, and so layers of different density are
produced. The uniform gabbro is produced by steady crystallization on the walls of the
chamber, so that away from the region beneath the crest the magma is totally
solidified. Above the chamber, magma is constantly being forced upwards to form
dykes, which in turn feed' the outpourings of lava on the ocean floor. The dykes
continually shoulder apart' their predecessors, much as an extra playing card might be
forced up into the vertical stack of cards we imagined earlier. Whether this
shouldering apart' is the driving force behind sea-floor spreading, or whether the
dykes are filling spaces left as the ocean ridge region is pulled apart is still uncertain.
Finally, why are there ocean ridges over rising convection currents, and why is the
topography of these regions so rugged? The answer to both questions relates to the fact
that the region is one of high heat flow. Because this part of the crust is hotter, it is less
dense, and so it stands higher than the surrounding regions. As the newly formed
crustal material moves away from the spreading centre, it cools, and so its elevation
decreases, the drop in height being accommodated along faults running parallel to the
ocean ridge.
The initiation of a new axis of sea floor spreading involves the separation of a formerly
continuous slab of continental crust. The two new continental margins that develop as
the ocean widens accumulate large thicknesses of sediment in which hydrocarbons
may occur.

66

A generalized model for continental rifting and separation has been developed from
studies of the East African rift systems and the Red Sea (regions regarded as representing
continental rifting and the earliest phases of ocean opening, respectively) and the
continental margins of South Australia and the Atlantic Ocean. This model is illustrated
in Figure 1.59.
FIGURE 1.59 Key
stages in the rifting
of continental crust
and subsequent seafloor spreading.

Figure 1.59 shows the features that, taken together, can indicate past episodes of
continental rifting and the development of continental margins adjacent to opening
oceans. The following stages can be recognized:
Continental rift stage: rifting, comparable to the East African rift systems, forms a graben
into which rivers carry sands. Lake sediments, evaporites or even shallow marine
sediments may occur if the sea is able to invade the newly created fault-bounded trough.
As the water bodies, be they marine or fresh, are restricted in area, stagnation is likely,
causing oxygen-deficient conditions at the bottom, which result in the preservation of
organic matter. The resultant organic-rich sediments may be potential oil source-rocks.
Regional uplift may result in siliciclastic sediments being shed laterally away from the
rift zone. Volcanic rocks are associated with the rifting and uplift; they are generally
basaltic lavas and dykes.
Red Sea stage: this stage marks the development of an embryonic ocean along the
site of the earlier rift. Thus the assemblage of igneous rocks characteristic of ocean
crust is developed, namely sea-floor basaltic pillow lavas underlain by basaltic

67

dykes, beneath which occur gabbros and peridotites. As the embryonic ocean is narrow,
water circulation is still restricted (as in the previous stage) and so organic rich and/or
evaporitic sediments may accumulate. Rapid subsidence occurs as the rifted crustal
blocks slide oceanwards along spoon shaped faults (usually termed listric faults).
Sag stage: movement along the faults initiated at the continental rift stage ceases, and
the entire continental margin subsides (rather than subsidence being the result of
movement of fault blocks) as the crust cools with increasing distance from the
ridge-spreading axis as the ocean widens. The continental margin progrades laterally as
continental shelf and continental slope sediments build upwards and outwards.
Important indicators of initial rifting and separation of continents include the association
of basaltic igneous activity with the development of relatively narrow rift basins, which
are followed by a more widespread regional subsidence or sag. This is referred to as a
rift-to-sag pattern. This is not exclusively confined to continental margins; it is also
exhibited by intra-continental rifts that failed to develop as ocean spreading centres. The
North Sea (Figure 1.1) is a good example of a rift-to-sag basin.

FIGURE 1.60 The


features of the three
types of destructive
plate margin. (a) Eastwest sketch across the
Pacific Ocean showing
the location of
ocean/ocean (islandarc type) and
ocean/continent
(Andean type)
destructive plate
margins. (b) Block
diagram showing the
main features of an
ocean/continent
destructive margin.

68

7.5.2 Destructive plate margins

Figure 1.60 shows the main features of margins of this type. A great variety of complex
igneous, sedimentary and metamorphic processes occur on these margins, most of which
are beyond the scope of this manual.
Ocean/continent and ocean/ocean destructive margins have a number of features in
common:
1)

Subduction - that is the consumption of ocean crust in the mantle - of oceanic


crustal material is revealed by an inclined zone of earthquakes, which are
generally steeper beneath ocean/ocean margins;

2)

Explosive volcanic activity;

3)

Ocean trenches.

Friction between the two plates generates liquid rock (magma). Melting also occurs as the
down-going plate descends into hotter regions at depth. The resultant igneous activity is
largely responsible for building up the island arcs on previously existing oceanic crust.
Not surprisingly, basaltic rocks dominate the island-arc volcanic environments, but the
ocean-continent destructive margins are dominated by volcanic lavas intermediate in
composition between basalt and granite - these are known as andesites (named, of course,
after the Andes). This intermediate composition is characteristic of continental crust as a
whole; only its upper part has a granitic composition. However, andesites are also formed
beneath island arcs.
Ocean/continent margins are also characterized by the occurrence of large intrusive
masses of granite, produced by the melting of the plates along subduction zones. The
granitic magma, being lighter than the surrounding rocks, moves upwards in the crust,
and may reach within a few kilometres of the surface. There it solidifies to form huge
elongate masses of granite, which may later be exposed at the surface once the overlying
rock cover is removed. Continental rocks may be crumpled and metamorphosed at
destructive margins.
Sediments accumulate in the ocean trenches associated with the two types of destructive
margin, and their composition will be different in the two types of plate margin, because
only in margins of Andean type are fragments of continental rock swept into the trenches.
In both margins, the sediments accumulating in the trenches may eventually be carried
down the subduction zone, and be plastered' onto either the island arc, or continental
crust.
Sedimentation along active i.e. destructive, ocean margins is more complex than on
passive margins, because they may involve ocean-ocean or ocean-continent in character.
When a new subduction zone is initiated, the oceanic lithosphere may fracture
such that a remnant of this oceanic lithosphere is left attached to the adjacent
continental plate, landward of the new trench. This gradually subsides to form a
broad fore-arc basin, which slowly fills with sediment. The trench itself receives
sediment originally of volcanic origin, from submarine fans. The trench sediments

69

are largely scraped off as the plate descends to form slices of sediment sheets. With
continued subduction, these sheets, together with the sediments of the fore-arc basin,
build an even wider accretionary prism between the volcanically active portion of the arc
and the trench itself (Figure 1.61). If continental collision occurs, the sediments of the
accretionary prism, together with the remnant of the oceanic lithosphere, are forced
upwards, deformed further and eroded. The former marine fore-arc basin becomes a
continental plain receiving vast quantities of sediment eroded from the fold mountain
belt.

FIGURE 1.61 (a) Crosssection through an active


continental margin to show
the pattern of sedimentation.
Four separate zones are
recognisable: the fore-arc
basin, thrust wedges of
sediment scraped off as the
oceanic plate is subducted,
trench infill and deep-sea
(pelagic) sediments.
(b) Enlargement of the
outline section shown in (a)
to show details of wedges of
sediment bounded by low
angle faults (thrusts).
Note that the sections are
drawn with considerable
vertical exaggeration.

7.6 Basin formation and classification


7.6.1 Thermal processes

Imagine for the moment, that a uniformly thick slab of continental lithosphere (as defined
in Section 7.1) is driven by plate tectonics over a part of the mantle that is anomalously
hot - a hot spot. These are known to exist as relatively fixed reference points for plate
tectonics over long periods of time. Continents can drift over these spots, and there is
some evidence from intracontinental volcanism that this has occurred in the past.
The extent to which the crust would be heated depends on the size of the spot, the
magnitude of associated heat flow above the normal, and the time involved. The
temperature of partial melting in the mantle would occur at shallower depth, so the
lithosphere would thin. The density underlying the slab would fall and the surface would
inevitably rise to form a bulge.
Increased slope angles in the vicinity of the bulge, increase the rates of erosion and
transport of debris. If the bulge remains over the hot spot for long enough, it would be
eroded away eventually. Not only would the continental surface regain its previous
monotony, but also the crust itself would have been thinned by erosion.
70

Eventually, the continent passes over the hot spot completely, or the loss of thermal
energy by volcanism causes the anomaly to decay away. The lithosphere beneath the
now-thinned crust cools by conduction and radiation to reach its former temperature. As a
consequence the underlying density must increase, because part of the low density crust
has been transferred elsewhere by surface processes. Therefore the surface has to subside
below its original level to maintain balance and a basin forms. Local slopes increase, and
the basin can be filled with sediments, if their supply is sufficient. This model for thermal
generation of basins is shown in Figure 1.62.
FIGURE 1.62 Model
of basin formation by
purely thermal means.
(a) A hot spot begins to
heat up the lithosphere.
(b) Decrease in
lithosphere density and
rise of athenosphere
due to increased heat
flow causes doming at
the surface, which is
quickly eroded away.
(c) Cooling restores
the density of the
lithosphere, and the
site of the crustal
thinning becomes a
subsiding basin.

This process does not happen instantly, for all means of energy transfer involve time
factors. The process, which dominates in the lithosphere, is conduction. How long a
thermal anomaly takes to decay away depends on three factors: (i) the temperature
involved, (ii) the thermal conductivity of the rocks, and (iii) their thermal capacity (the
product of volume, density and specific heat). The theory is mathematically complex, but
demonstrates that the temperatures involved have little influence on the time. The major
factor is the volume of the material involved, which in this case is a function of the
thickness of the lithosphere. Calculations show that thermal anomalies in the continental
lithosphere take around 50 million years to decay away. So this is the sort of time over
which we might expect a thermally generated sedimentary basin to evolve, from erosion
of the original bulge through to its progressive subsidence and infilling.
Within the major continents, there are numerous large sedimentary basins ranging from
Precambrian to Recent age. Many of these are remote from zones of contemporary
tectonic activity and are founded on stable Precambrian crust. Good examples are found
in Africa (Figure 1.63). As well as having intracontinental sedimentary basins, the
African continent is characterized by many areas of regional doming associated with
volcanism, including those aligned along the East African Rift. One theory is that the
now nearly stationary African Plate has settled over a series of mantle hotspots, which
caused the domes. From this broad standpoint, the basins may reflect some former
positions of such sub-lithospheric anomalies, over which Africa has shifted before it
stopped drifting. More detailed analysis, however, does show that some basin margins are
faulted implying some tectonic control. An example is the oil-producing Sirte Basin of
Libya, which, although fault-bounded, may incorporate a high component of thermal
sagging.
71

FIGURE 1.63 The


distribution of Tertiary
to recent sedimentary
basins and uplifts
(swells) in Africa.

When the relations between sedimentary fill and the underlying crust in many
sedimentary basins are examined, it is found that there is little evidence for the erosion of
sufficient older crust to have allowed basin formation by purely thermal means. Some
other means of thinning the crust and generating density and gravity changes must be
sought.
7.6.2 Stretching of the lithosphere

Simple experiments with metal rods and toffee show that materials behave plastically
when they are slowly strained. During extensional strain, the material thins or necks
somewhere, and eventually breaks. This crude analogy has been applied to extension of
the lithosphere. The stresses involved are related to plate tectonic activity.
If the surface area of part of the lithosphere is extended by a factor , its volume does not
change and therefore it must decrease its thickness. The stretching factor (which can
also represent shortening if it is less than one) is applied to lengths of lithosphere or crust
in a two-dimensional cross-section.
Figure 1.64 shows, in sketch form, mechanisms for lithospheric extension.

72

FIGURE 1.64
Models for lithospheric
extension.

Five models for extension using detachments: (A) throughgoing detachment; (B)
ramps and flats on a detachment, leading to the development of marginal plateaux,
and ramp basins; (C) detachment plus pure shear stretching of the lithosphere at
the end of the detachment, with the zone of ductile stretching substantially offset
from the zone of brittle faulting, so rift basins do not expand to encounter ductilely
stretched rocks below the detachment; (D) detachment plus pure shear stretching
below the detachment but the zone of ductile stretching below the detachment is
73

You might also like