You are on page 1of 97

Lecture 1. Introduction: The context of course.

What is Geological bases of oil and gas

processing?

Petroleum geology comprises those geological disciplines which are of greatest significance for

the finding and recovery of oil and gas. Since most of the obvious and “easy to find” petroleum

already has been discovered it is necessary to use sophisticated methods in the exploration of

sedimentary basins. These include advanced geophysical techniques and basin modelling. There is

also much more emphasis now on enhanced recovery from the producing fields. Petroleum

technology has made great progress and many new tools and modelling programs have been

developed, both in exploration and production. It is however important to understand the geological

processes which determine the distribution of different sedimentary rocks and their physical

properties. This knowledge is fundamental to being able to successfully apply the methods now

available. It is difficult to know where to start when teaching petroleum geology because nearly all

the different disciplines build on each other. This introductory chapter will provide a short and

rather simple overview of some aspects of petroleum geology to introduce the subject and the

problems. Most of the other chapters will then expand on what is presented here to provide a better

background in relevant subjects. Since practically all petroleum occurs in sedimentary rocks,

sedimentary geology forms one of the main foundations of petroleum geology. Sedimentological

models are used to predict the location of different facies in the sedimentary basins, and from that

the likely presence of source rocks with a high content of organic matter, reservoir rocks and cap

rocks. The distribution and geometry of potential sandstones or carbonate reservoirs requires

detailed sedimentological models, and sequence stratigraphy has been a useful tool in such

reconstructions. The biostratigraphic correlation of strata encountered in exploration wells is

achieved by micropalaeontology (including palynology), a field developed very largely by the oil

industry. Due to the small size of the samples obtained during drilling operations one cannot rely on

macrofossils; even in core samples the chance of finding good macrofossils is poor. By contrast a

few grams of rock from the drill cuttings may contain several hundred microfossils or
palynomorphs. These also usually provide better stratigraphic resolution than macrofossils.

Reservoir rocks are mostly sandstones and carbonates which are sufficiently porous to hold

significant amounts of petroleum. The composition and properties of other rock types such as shales

and salt are also important. The sedimentary environments (sedimentary facies) determine the

distribution of reservoir rocks and their primary composition. Sediments do, however, alter their

properties with increasing overburden due to diagenesis during burial. Diagenetic processes

determine the porosity, permeability and other physical properties such as velocity, in both

sandstone and limestone reservoirs. Chemical processes controlling mineral reactions are important.

Organic geochemistry, which includes the study of organic matter in sediments and its

transformation into hydrocarbons, has become another vital part of petroleum geology. Tectonics

and structural geology provide an understanding of the subsidence, folding and uplift responsible

for the creation and dynamic history of a basin. The timing of the folding and faulting that forms

structural traps is very important in relation to the migration of hydrocarbons. Seismic methods

have become the main tool for mapping sedimentary facies, stratigraphy, sequence stratigraphy and

tectonic development. Marine seismics recorded from ships have become very efficient and seismic

lines are shot at only a few 100 m spacing or less. Because of the rapid improvement in the quality

of seismic data processing techniques, geological interpretation of seismic data has become an

entirely new and expanding field. Seismic and other geophysical data are often the only information

we have, particularly for offshore exploration where drilling is very costly. Shooting seismic lines

with a close spacing allows high resolution 3D seismic imagery to be produced for critical parts of

sedimentary basins. By repeating a 3D reservoir seismic survey during production, one can observe

how the gas/oil and oil/water contacts move as the reservoir is depleted. This is called 4D seismic

because time provides the fourth dimension. Geophysical measurements may include gravimetry

and magnetometry; electromagnetic methods that were used mostly in ore exploration have also

been applied to oil exploration. Electromagnetic methods have been used to detect sediments with

low resistivity due to the presence of oil instead of saline water. This method requires a few 100 m
of water and relatively shallow accumulation. Seismic surveys are more expensive on land than by

ship at sea because geophones have to be placed in a grid, often on uneven and difficult land

surfaces. Drilling on land, however, costs much less than from offshore rigs and a much denser well

spacing can be used during both exploration and production. Indirect methods of mapping rock

types employing geophysical aids are becoming increasingly important in petroleum geology, but it

is still necessary to take samples and examine the rocks themselves. A petroleum geologist should

have a broad geological training, preferably also from field work.

Geophysical well-logging methods have developed equally rapidly, from simple electric and

radioactive logs to highly advanced logging tools which provide detailed information about the

sequence penetrated by the well. Logs provide a continuity of information about the rock properties

which one can seldom obtain from exposures or core samples. This information makes it possible to

interpret not only the lithological composition of the rocks and the variation of porosity and

permeability, but also the depositional environment. Image logs make it possible also to detect

bedding and fractures inside the wells. Practical petroleum geology is not only based on many

different geological and geophysical disciplines. A good background in basic chemistry, physics,

mathematics and computing is also required, particularly for different types of basin modelling.

A Brief Petroleum History

There are many places where oil seeps out of the ground. Bitumen produced from such naturally

occurring crude oil has been collected and used since ancient times, both for lighting and medicine,

and by the Greeks even for warfare. In some places, for example Germany in the 1800s, small

mines were dug to get at the oil. Before 1859 oil was also recovered from coal for use in kerosene

lamps. It was not until Edwin Drake’s exploits in 1859 at Oil Creek near Titusville in West

Pennsylvania that oil was recovered in any quantity from boreholes. He drilled a well about 25 m

(70 ft) deep which produced 8–10 bbl/day, a huge production rate compared with anything earlier.

A few years later there were 74 wells round Oil Creek, and the USA’s annual production had risen
to half a million barrels. Outside the USA the calculated total production at that time was maximum

5,000 bbl. In 1870 production had increased tenfold, with 5 million bbl from the USA, and 538,000

bbl from other countries. In southern California oil production started early in 1864 (in Santa

Paula), but for many years oil was mined by driving shafts into the oil-bearing strata because it was

so heavy and biodegraded that it would not flow in a well. In its infancy, oil exploration consisted

largely of looking for oil seepage at the surface and drilling in the vicinity, which did not require

much geological knowledge. It was then realised that oil and gas occur where layers of sedimentary

rocks form domes or anticlinal structures since petroleum is less dense than water and a low

permeability (seal) layer is needed to prevent the oil and gas from rising and escaping. This led to

extensive geological mapping of anticlines and domes visible at the surface, particularly in the

USA. It was also found that oil fields had a tendency to lie along structural trends defined by

anticlines or faults and this “rule” was used in prospecting. This is also often the case with salt

domes, which became important prospecting targets. Oil production developed rapidly up to the end

of the nineteenth century, and more systematic geological principles for prospecting were gradually

developed. The geological information which one obtains at the surface is often not representative

of the structures deeper down. Structures which are not visible at the surface could be mapped by

correlation between wells using logs and cuttings from the drilling. One method was to measure the

depth of particularly characteristic strata through analysis of cuttings in different wells. Improved

electrical measurements (logs) from wells, developed during the 1920s and 1930s, made the whole

effort much simpler because they provided continuous vertical sections through the rocks. The first

logs were simple recordings of how well rocks conduct electrical currents (resistivity), and later

also gamma logs recorded the gamma-radiation emitted by the different sedimentary rocks. The

USA maintained its position as the major world producer of oil and gas well into the twentieth

century. Americans thus became leaders in the development of oil technology, which today is

strongly reflected in the industry’s terminology. USA also rapidly became the world’s greatest

consumer of oil and gas, and now has to import at least 60% of its oil consumption despite still
having a large home production (8 million bbl/day). The US consumption (21 million bbl/day) is a

very large fraction of the total world consumption (90 million bbl/day) It was first in the 1930–1940

period that the industry became aware of the vast oil resources of the Middle East, which now

account for about 60% of world reserves. Since then this region has dominated oil production. In

the 1950s and 1960s prospecting for oil and gas was extended onto the continental shelves, opening

up new reserves. Until the development of modern seismic it was not possible to effectively explore

deep below the seafloor in sedimentary basins offshore. As long as there was an abundance of oil to

be found in onshore basins there was little incentive to develop costly drilling rigs for offshore

exploration and platforms for production. Since the 1970s an increasingly large share of

international prospecting has taken place offshore, helped by improved seismic methods. Advanced

well log technology in particular made it possible to gain optimal information from each well.

Before the development of powerful computers, seismic recordings were based on analogue

methods which produced results very far removed from the standard of modern seismic data.

Recording of seismic data from ships is much less expensive than onshore where geophones have to

be laid out manually. Onshore, extensive drilling may be cheaper than expensive seismic mapping.

Rising oil prices and new technology have made exploration financially attractive in areas which

previously were of little interest, including in very deep waters. High oil prices can also pay for

more enhanced hydrocarbon recovery from reservoirs. There are now relatively few sedimentary

basins in the world that have not been explored and it is getting increasingly difficult to find new

giant fields (>500 million bbl). There is now increasing interest in heavy oil, tar sand and oil shale.

Oil shale is a source rock exposed near the surface. If the source rock (shale) is mature it will have a

characteristic smell of hydrocarbons, but it may not be mature so that hydrocarbons have not been

generated. If the oil shale is mature much of the oil has escaped by primary migration. Since the

hydrocarbons are thoroughly disseminated in the fine-grained sediment, oil cannot be produced in

the same way as from sandstone or carbonate reservoirs. The hydrocarbons can only be obtained by

breaking and crushing the shale and heating to distill off the interspersed hydrocarbons. Shales can
however contain gas which can be produced when there is a network of small fractures. Gas shale is

expected to be an important source of petroleum in the years to come, particularly in the US. Very

large amounts of fossil fuels are stored in organic-rich mudstones or shales that have not been

buried deeply enough for the organic matter to be converted to petroleum. In this case very little

hydrocarbon has excaped but these deposits must be mined and heated to 400–500◦C in ovens to

generate petroleum (pyrolysis). The Tertiary Green River Shale in Colorado, Utah and Wyoming

represents one of the largest petroleum reservoirs in the world. This is a lake deposit, and the

organic matter consisted mainly of algae. Although very large quantities of petroleum can be

produced from oil shale, production costs are at present too high compared to conventional oil.

There are also serious environmental problems involved in production from oil shale, and the

process requires very large quantities of water, a resource which is not always plentiful. The oil

reserves in such deposits exceed conventional oil reserves, but the expense and environmental

issues involved with production from these types of reservoirs clearly limit their exploitation. This

is particularly true of production from oil shale.

Accumulations of Organic Matter

It is well documented that oil accumulations are of organic origin and formed from organic

matter in sediments. Methane can be formed inorganically and is found in the atmosphere of several

other planets, but inorganic methane from the interior of the earth is likely to be well dispersed and

thus not form major gas accumulations in the earth’s crust. The organic matter from which

petroleum is derived originated through photosynthesis, i.e. storage of solar energy (Fig. 1.1).

Sunlight is continuously transformed into such energy on Earth but only a very small proportion of

the solar energy is preserved as organic matter and petroleum. The oil and gas which forms in

sedimentary basins each year is thus minute in comparison with the rate of exploitation (production)

and consumption. In practice petroleum must therefore be regarded as a non-renewable resource

even though some petroleum is being formed all the time. Most of the organic materials which

occur in source rocks for petroleum are algae, formed by photosynthesis. The zooplankton and
higher organisms that are also represented grazed the algae and were thus indirectly dependent on

photosynthesis too. The energy which we release when burning petroleum is therefore stored solar

energy. Since petroleum is derived from organic matter, it is important to understand how and

where sediments with a high content of organic matter are deposited. The total production of

organic material in the world’s oceans is now 5×1010 tonnes/year. Nutrients for this organic

production are supplied by erosion of rocks on land and transported into the ocean. The supply of

nutrients is therefore greatest in coastal areas, particularly where sediment-laden rivers discharge

into the sea. Plant debris is also supplied directly from the land in coastal areas. Biological

production is greatest in the uppermost 20–30 m of the ocean and most of the phytoplankton growth

takes place in this zone. In clear water, sunlight penetrates much deeper than in turbid water, but in

clear water there is usually little nutrient supply. At about 100–150 m depth, sunlight is too weak

for photosynthesis even in very clear water. Phytoplankton provides nutrition for all other marine

life in the oceans. Zooplankton feed on phytoplankton and therefore proliferate only where there is

vigorous phytoplankton production. Organisms sink after they have died, and may decay so that

nutrients are released and recycled at greater depths. Basins with restricted water circulation will

preserve more organic matter and produce good source rocks which may mature to generate oil and

gas (Fig. 1.2a, b). In polar regions, cold dense water sinks to great depths and flows along the

bottom of the deep oceans towards lower latitudes. This is the thermal conveyor belt transporting

heat to higher latitudes and it keeps the deep ocean water oxidizing. In areas near the equator where

the prevailing winds are from the east the surface water is driven away from the western coast of the

continents. This generates a strong upwelling of nutrient-rich water from the bottom of the sea

which sustains especially high levels of primary organic production (Fig. 1.3). The best examples of

this are the coast of Chile and off West Africa. Through photosynthesis, low energy carbon dioxide

and water are transformed into high energy carbohydrates (e.g. glucose):

CO2 + H2O → CH2O (organic matter) + O2


The production of organic matter is not limited by carbon dioxide or water, but by nutrient

availability.

Phosphorus (P) and nitrogen (N) are the most important nutrients, though the supply of iron can

also be limiting for alga production. It is this process of photosynthesis, which started 4 billion

years ago, that has built up an atmosphere rich in oxygen while accumulating reduced carbon in

sedimentary rocks as oil, gas and coal. Most of the carbon is nevertheless finely divided within

sedimentary rocks, for example shales and limestones, in concentrations too low to generate

significant oil and gas. Energy stored by photosynthesis can be used directly by organisms for

respiration. This is the opposite process, breaking carbohydrates down into carbon dioxide and

water again, so that the organisms gain energy. This occurs in organisms at night when there is no

light to drive photosynthesis. Also when we burn hydrocarbons, e.g. while driving a car, energy is

obtained by oxidation, again essentially reversing the photosynthesis equation quoted above.

Oxidation of 100 g glucose releases 375 kcal of energy. Carbohydrates that are produced but not

consumed by respiration can be stored as glucose, cellulose or starch in the cell walls.

Photosynthesis is also the biochemical source for the synthesis of lipids and proteins. Proteins are

large, complex molecules built up of condensed amino acids (e.g. glycine (H2NCH2−COOH)).

Dried phytoplankton contains 45–55% carbon, 4.5–9% nitrogen, 0.6–3.3% phosphorus and up to

25% of both silica and carbonate. Planktonic algae are the main contributors to the organic matter

which gives rise to petroleum. Among the most important are diatoms, which have amorphous silica

(opal A) shells.

Diatoms are most abundant in the higher latitudes and are also found in brackish and fresh

water. Blue– green algae (cyanobacteria) which live on the bottom in shallow areas, also contribute

to the organic material in sediments. In coastal swamps, and particularly on deltas, we have

extensive production of organic matter in the form of plants and trees which may avoid being

oxidised by sinking into mud or bog. The residues of these higher land plants may form peat, which

with deeper burial may be converted into lignite and bituminous coal. But such deposits are also a
potential source of gas and oil. Plant matter, including wood, also floats down rivers and is

deposited when it sinks to the bottom, usually in a nearshore deltaic environment. When the trees

rot they release CO2 and consume as much oxygen as the plant produced during the whole period

when it was growing. There is thus no net contribution of oxygen to the atmosphere. This also

applies to the bulk of the tropical rainforests. Where trees and plants sink into black mud,

preventing them from being oxidised, there is a net contribution of oxygen to the atmosphere and a

corresponding reduction of CO2 in the atmosphere. All animal plankton (zooplankton) live on plant

plankton, and in turn are eaten by higher organisms in the food chain. At each step in the food

chain, which we call a trophic level, the amount of organic matter (the biomass) is reduced to 10%.

Ninety percent of the production of organic matter is therefore from algae. This is why algae and to

some extent zooplankton account for the bulk of the organic material which can be transformed into

oil. Larger animals such as dinosaurs are totally irrelevant as sources of oil. The most important of

the zooplankton which provide organic matter for petroleum are:

1. Radiolaria – silica shells, wide distribution, particularly in tropical waters.

2. Foraminifera – shells of calcium carbonate.

3. Pteropods – pelagic gastropods (snails) with a foot which has been converted into wing-

shaped lobes; carbonate shells.

This is the second lowest level within the marine food chain. These zooplanktonic organisms

are eaten by crustaceans which themselves are eaten by fish. The total amount of organic matter that

can be produced in the ocean is dependent on the nutrient supply from rivers, but river water does

not only carry inorganic nutrients. It also contains significant amounts of organic matter, in

particular humic acid compounds, lignin and similar substances formed by the breakdown of plant

material which are weakly soluble in cold water. When the river water enters the sea, there is

precipitation due to the increased pH and lower surface temperature in the ocean. Other plant

materials, like waxes and resins, are more chemically resistant to breakdown and are insoluble in
water. Such organic particles tend to attach themselves to mineral grains and accompany sediment

out into the ocean. Most of the oil reservoirs which have been formed since the Palaeozoic have

been uplifted and eroded, and over time vast quantities of oil have flowed (seeped) out onto the land

or into the sea. In this sense, oil pollution is a natural process. Only a small proportion of the

petroleum that has been formed in source rocks has actually become trapped in a reservoir. One

might expect this seepage to have provided a source of recycled petroleum in younger sediments,

but petroleum breaks down extremely rapidly when subjected to weathering, oxidising to CO2, and

the nutrients (P, N) that were required to form the organic matter are released and may act like a

fertilizer. On land, evaporation will remove the lighter components while bacteria will degrade the

heavier components. Fossil asphalt lakes consist of heavy substances which neither evaporate nor

can be easily broken down by bacteria. In the ocean, the lighter components will dissolve quite

rapidly, while the heavier asphalt fraction will sink to the bottom and be degraded and recycled. In

uplifted sedimentary basins like the Ventura Basin and the Los Angeles Basin in Southern

California there are abundant natural oils seeps both onshore and offshore. On the beaches from

Santa Barbara towards Los Angeles there are many natural oil seeps.

Lecture 2. Formation of an oil accumulation

Breakdown of Organic Matter

Almost all (>99%) of the organic matter which is produced on land and in the oceans is broken

down through direct oxidation or by means of microbiological processes. If oxygen is present,

organic matter will be broken down in the following manner:

CH2O + O2 → CO2 + H2O

Where oxygen is available, organic matter is oxidised relatively rapidly both on land and in the

sea. As organisms die, organic material suspended in seawater sinks through the water column

consuming oxygen. If water circulation is restricted due to density stratification of the water
column, the oxygen supply will be exhausted. Instead, the bound oxygen in sulphates or nitrates is

used by sulphate-reducing and denitrifying bacteria which decompose organic material in an anoxic

environment. The first few centimetres below the seabed are usually oxidised, while reducing

conditions prevail 5–30 cm below the sea floor. Below this redox boundary where there is no free

oxygen, sulphate-reducing bacteria react with organic matter as indicated below:

2CH2O + 2H+ + SO– – 4 → H2S + 2CO2 + 2H2O NH3 + H+ + SO2 → NO3 + H2S + H2O

H2S is liberated, giving stagnant water and mud a strong smell. Through denitrification we get

5CH2O + 4H + + 4NO– 3 → 2N2 + 5CO2 + 7H2O

When the rate of accumulation of organic matter exceeds the rate of oxygen supply the redox

boundary will be in the water column, separating the oxidising surface water from the reducing

bottom water. This is typical of basins separated from the deep ocean by a shallow sill, like the

Black Sea and some of the deep Norwegian fjords. Fresh or brackish surface water floating on more

saline water also helps to maintain a stable water stratification with little vertical mixing. Lakes may

have good water stratification because warm surface water is less dense than the colder bottom

water. Black mud deposited at the bottom of lakes may produce good source rocks. In cold

climates, however, the water in the lakes overturns in the winter because the maximum water

density is at 4◦C, preventing the stable stratification required to form source rocks.

Formation of Source Rocks

All marine organic material is formed near the surface of the ocean, in the photic zone, through

photosynthesis. For the most part this is algae. Some phytoplankton are broken down chemically

and oxidised and some are eaten by zooplankton. Both types of plankton are eaten by higher

organisms which concentrate the indigestible part of the organic matter into fecal pellets which may

be incorporated into sediments. Plankton is made up of very small organisms which sink so slowly

that they are in most cases almost entirely degraded (oxidised) before they reach the bottom. Pellets,
on the other hand, are the size of sand grains and sink more rapidly, and this organic matter is more

likely to be preserved in the sediments. On the bottom, organic matter will be subjected to

breakdown by micro-organisms (bacteria). It will also be eaten by burrowing organisms which live

in the top portion of the sediments. The activity of these organisms contributes to reducing the

organic content of the sediments because most of the organic matter is digested when the sediment

is eaten. Bioturbation also stirs up the sediments, exposing them more to the oxygen-bearing bottom

water. However, if the bottom water is stagnant, the lack of oxygen and the toxicity of H2S will

exclude most life forms. The resultant lack of bioturbation will thus preserve more organic matter in

the sediment together with perfect, undisturbed, lamination. Stagnant, or anoxic, conditions are

defined by an oxygen content of the Atlantic experienced stagnant bottom conditions during such

periods. Limited water circulation in semi-enclosed marine basins due to restricted outflow over a

shallow threshold is a common cause of stagnant water bodies (Fig. 1.2a). The Black Sea is a good

example. In response to an abundant freshwater supply from rivers and a relatively low evaporation

rate, a low salinity surface layer leads to density stratification in the water column and a consequent

reduction in circulation. In basins with little precipitation and where there is net evaporation, the

surface water will have higher salinity and density than the water below it, and will sink down. This

circulation brings with it oxygen from the surface and can give oxidising bottom conditions with

little chance for organic matter to survive to form source rocks. Lakes or semi-enclosed marine

basins often have a temperature- or salinity-induced density stratification so that oxygenated surface

water does not mix with water in the deeper part of the basin. This leads to anoxic conditions and a

high degree of preservation of the organic matter produced in the surface waters. This aspect is

therefore of considerable interest in exploration for petroleum in freshwater basins, particularly in

Africa and China. The open oceans have normally had oxygenated water, but during the Cretaceous

most of the Atlantic Ocean is believed to have been stagnant during so-called “anoxic events”, and

substantial amounts of black shale were deposited in the deeper parts of the ocean during these

periods.
Lecture 3. Sediments to rocks to traps

Early Diagenesis of Organic Matter

Microbiological breakdown of organic matter in sediments is due to the activity of bacteria,

fungi, protozoa, etc. and under oxidising conditions these are extremely effective. However, the

porewater quickly becomes reducing if the oxygen is not replenished. In relatively coarse-grained

sediments (sand), oxygen may diffuse to depths of 5–20 cm below the seabed, while in clay and

fine-grained carbonate mud the boundary between oxidising and reducing water (redox boundary)

may be a few millimetres below the seafloor. The pores in the sediments here are so small that

water circulation and diffusion are insufficient to replace the original oxygen in the porewater as it

gets used up by oxidation of organic matter. Clay-rich sediments soon become a relatively closed

system, and the downward diffusion of oxygen from the seabed is very slow in fine-grained

sediments. Aerobic breakdown is therefore much more effective in coarse-grained sediments than in

fine-grained ones. In anaerobic transformation bacteria use organic matter, e.g. short carbohydrate

chains. Cellulose is broken down by fungi, and finally by bacteria. The end products are methane

(CH4) and carbon dioxide (CO2). Methane, however, is the only hydrocarbon produced in any

quantity at low temperatures by bacteria close to the surface of the sediment. Gas occurring at

shallow depths (shallow gas) therefore consists largely of methane (dry gas) unless there has been

addition from much deeper strata. Biogenic gas may form commercial accumulations, as in Western

Siberia and also in the shallow part of the North Sea basin. The presence of abundant shallow gas

may represent a hazard in the form of blowouts and fire during drilling. Gas occurring at shallow

depth may also have a deeper source generated from a gas-prone source rock (coaly sediments) or

by cracking of oil, but such gas has a very different isotopic signature than biogenic gas.

Kerogen

As organic material becomes buried by the accumulation of overlying sediments, water is

gradually expelled during compaction. Complex organic compounds like proteins are broken down
into amino acids, and carbohydrates into simpler sugar compounds. These are able to recombine to

make larger compounds, for example by amino acids reacting with carbohydrates (melanoid

reaction). As this type of polymerisation proceeds, the proportion of simpler soluble organic

compounds diminishes at depths of a few tens of metres down in the sediment. It is these newly-

formed complex organic structures which are called kerogen. Kerogen is a collective name for

organic material that is insoluble in organic solvents, water or oxidising acids. The portion of the

organic material soluble in organic solvents is called bitumen, which is essentially oil in a solid

state.

Kerogen consists of very large molecules and is a kind of polymer. When it has been exposed to

sufficient time and temperature these large molecules will crack into smaller molecules, mostly

petroleum. When the temperature is about 100◦C a long period of geological time is required. In

rapidly subsiding basins the exposure time is shorter and oil generation may only start at about 140–

150◦C. In the North Sea basin the “oil window” may typically be between 130 and140◦C.

Sedimentary Rocks

The sedimentary rocks form an outer skin on the Earth’s crust, covering three-quarters of the

continental areas and most of the sea floor. They vary in thickness up to 10 km. Nevertheless, they

only comprise about 5% of the crust. Most sedimentary rocks are of secondary origin, in that they

consist of detrital material derived by the breakdown of pre-existing rocks. Indeed, it has been

variously estimated that shales and sandstones, both of mechanical derivation, account for between

75 and 95% of all sedimentary rocks. However, certain sedimentary rocks are the products of

chemical or biochemical precipitation whereas others are of organic origin. Thus, the sedimentary

rocks can be divided into two principal groups, namely, the clastic (detrital) or exogenetic, and the

non-clastic or endogenetic types. Nevertheless, one factor that all sedimentary rocks have in

common is that they are deposited, and this gives rise to their most noteworthy characteristic, that

is, they are bedded or stratified. As noted above, most sedimentary rocks are formed from the
breakdown products of pre-existing rocks. Accordingly, the rate at which denudation takes place

acts as a control on the rate of sedimentation, which in turn affects the character of a sediment.

However, the rate of denudation is not only determined by the agents at work, that is, by

weathering, or by river, marine, wind or ice action, but also by the nature of the surface. In other

words, upland areas are more rapidly worn away than are lowlands. Indeed, denudation may be

regarded as a cyclic process, in that it begins with or is furthered by the elevation of a land surface,

and as this is gradually worn down, the rate of denudation slackens. Each cycle of erosion is

accompanied by a cycle of sedimentation. The particles of which most sedimentary rocks are

composed have undergone varying amounts of transportation. The amount of transport together

with the agent responsible, be it water, wind or ice, play an important role in determining the

character of a sediment. For instance, transport over short distances usually means that the sediment

is unsorted (the exception being beach sands), as does transportation by ice. With lengthier transport

by water or wind, not only does the material become better sorted but it is further reduced in size.

The character of a sedimentary rock is also influenced by the type of environment in which it

has been deposited, the presence of which is witnessed as ripple marks and cross bedding in sands

that accumulate in shallow water. The composition of a sedimentary rock depends partly on the

composition of the parent material and the stability of its component minerals, and partly on the

type of action to which the parent rock was subjected and the length of time it had to suffer such

action. The least stable minerals tend to be those that are developed in environments very different

from those experienced at the Earth’s surface. In fact, quartz, and, to a much lesser extent, mica, are

the only common detrital constituents of igneous and metamorphic rocks that are found in

abundance in sediments. Most of the other minerals are ultimately broken down chemically to give

rise to clay minerals. The more mature a sedimentary rock is, the more it approaches a stable end

product, and very mature sediments are likely to have experienced more than one cycle of

sedimentation. The type of climatic regime in which a deposit accumulates and the rate at which

this takes place also affect the stability and maturity of the resultant sedimentary product. For
example, chemical decay is inhibited in arid regions so that less stable minerals are more likely to

survive than in humid regions. However, even in humid regions, immature sediments may form

when basins are rapidly filled with detritus derived from neighbouring mountains, the rapid burial

affording protection against the attack of subaerial agents. In order to turn unconsolidated sediment

into solid rock, it must be lithified. Lithification involves two processes, consolidation and

cementation. The amount of consolidation that takes place within a sediment depends, first, on its

composition and texture and, second, on the pressures acting on it, notably that due to the weight of

overburden. Consolidation of sediments deposited in water also involves dewatering, that is, the

expulsion of connate water from the sediments. The porosity of a sediment is reduced as

consolidation takes place, and, as the individual particles become more closely packed, they may

even be deformed. Pressures developed during consolidation may lead to the differential solution of

minerals and the authigenic growth of new ones. Fine-grained sediments possess a higher porosity

than do coarser types and, therefore, undergo a greater amount of consolidation. For instance, muds

and clays may have original porosities ranging up to 80%, compared to 45 to 50% in sands and silts.

Hence, if muds and clays could be completely consolidated (they never are), they would occupy

only 20 to 45% of their original volume. The amount of consolidation that takes place in sands and

silts varies from 15 to 25%. Cementation involves the bonding together of sedimentary particles by

the precipitation of material in the pore spaces. This reduces the porosity. The cementing material

may be derived by partial intrastratal solution of grains or may be introduced into the pore spaces

from an extraneous source by circulating waters. Conversely, cement may be removed from a

sedimentary rock by leaching. The type of cement and, more importantly, the amount, affect the

strength of a sedimentary rock. The type also influences its colour. For example, sandstones with

siliceous or calcium carbonate cement are usually whitish grey, those with sideritic (iron carbonate)

cement are buff coloured, whereas a red colour is indicative of hematitic (iron oxide) cement and

brown of limonite (hydrated iron oxide). However, sedimentary rocks are frequently cemented by

more than one material. The matrix of a sedimentary rock refers to the fine material trapped within
the pore spaces between the particles. It helps to bind the latter together. The texture of a

sedimentary rock refers to the size, shape and arrangement of its constituent particles. Size is a

property that is not easy to assess accurately, for the grains and pebbles of which clastic sediments

are composed are irregular, three-dimensional objects. Direct measurement can only be applied to

large individual fragments where the length of the three principal axes can be recorded. But even

this rarely affords a true picture of size. Estimation of volume by displacement may provide a better

measure. Because of their smallness, the size of grains of sands and silts has to be measured

indirectly by sieving and sedimentation techniques, respectively. If individual particles of clay have

to be measured, this can be done with the aid of an electron microscope. If a rock is strongly

indurated, its disaggregation is impossible without fracturing many of the grains. In such a case, a

thin section of the rock is made and size analysis is carried out with the aid of a petrological

microscope, mechanical stage and micrometer. The results of a size analysis may be represented

graphically by a frequency curve or histogram. More frequently, however, they are used to draw a

cumulative curve. The latter may be drawn on semi-logarithmic paper (Fig. 1.15). Various

statistical parameters such as median and mean size, deviation, skewness and kurtosis can be

calculated from data derived from cumulative curves. The median or mean size permits the

determination of the grade of gravel, sand or silt, or their lithified equivalents. Deviation affords a

measure of sorting. However, the latter can be quickly and simply estimated by visual examination

of the curve in that the steeper it is, the more uniform the sorting of the sediment. The size of the

particles of a clastic sedimentary rock allows it to be placed in one of three groups that are termed

rudaceous or psephitic, arenaceous or psammitic and argillaceous or pelitic. Reference to size scales

is made in Chapter 5, where a description of mixed aggregates also is provided.

Shape is probably the most fundamental property of any particle, but, unfortunately it is one of

the most difficult to quantify. Shape is frequently assessed in terms of roundness and sphericity,

which may be estimated visually by comparison with standard images (Fig. 1.16). However,

because the latter is a subjective assessment, the values obtained suffer accordingly. A sedimentary
rock is an aggregate of particles, and some of its characteristics depend on the position of these

particles in space. The degree of grain orientation within a rock varies between perfect preferred

orientation, in which all the long axes run in the same direction, and perfect random orientation,

where the long axes point in all directions. The latter is found only infrequently as most aggregates

possess some degree of grain orientation. The arrangement of particles in a sedimentary rock

involves the concept of packing, which refers to the spatial density of the particles in an aggregate.

Packing has been defined as the mutual spatial relationship among the grains. It includes grain-to-

grain contacts and the shape of the contact. The latter involves the closeness or spread of particles,

that is, how much space in a given area is occupied by grains. Packing is an important property of

sedimentary rocks, for it is related to their degree of consolidation, density, porosity and strength.

Bedding and Sedimentary Structures

Sedimentary rocks are characterized by their stratification, and bedding planes are frequently

the dominant discontinuity in sedimentary rock masses (Fig. 1.17). As such, their spacing and

character (are they irregular, waved or straight, tight or open, rough or smooth?) are of particular

importance to the engineer. Several spacing classifications have been advanced (see Chapter 2). An

individual bed may be regarded as a thickness of sediment of the same composition that was

deposited under the same conditions. Lamination, on the other hand, refers to a bed of sedimentary

rock that exhibits thin layers or laminae, usually a few millimetres in thickness. The laminae may be

the result of minor fluctuations in the velocity of the transporting medium or the supply of material,

both of which produce alternating thin layers of slightly different grain size. Generally, however,

lamination is associated with the presence of thin layers of platy minerals, notably micas. These

have a marked preferred orientation, usually parallel to the bedding planes, and are responsible for

the fissility of the rock. The surfaces of these laminae are usually smooth and straight. Although

lamination is most characteristic of shales, it also may be present in siltstones and sandstones, and

occasionally in some limestones. Cross or current bedding is a depositional feature that occurs in

sediments of fluvial, littoral, marine and aeolian origin, and is found most notably in sandstones
(Fig. 1.18). In wind-blown sediments, it generally is referred to as dune bedding. Cross bedding is

confined within an individual sedimentation unit and consists of cross laminae inclined to the true

bedding planes. The original dip of these cross laminae is frequently between 20 and 30∞. The size

of the sedimentation unit in which they occur varies enormously. For example, in microcross-

bedding, it measures only a few millimetres, whereas in dune bedding, the unit may exceed 100 m.

Although graded bedding occurs in several different types of sedimentary rock, it is characteristic of

greywacke. As the name suggests, the sedimentation unit exhibits a grading from coarser grain size

at the bottom to finer at the top. Individual graded beds range in thickness from a few millimetres to

several metres. Usually, the thicker the bed, the coarser it is overall.

Sedimentary Rock Types

Gravel is an unconsolidated accumulation of rounded fragments, the lower size limit of which is

2 mm. The term rubble has been used to describe those deposits that contain angular fragments. The

composition of a gravel deposit reflects not only the source rocks of the area from which it was

derived but also is influenced by the agents responsible for its formation and the climatic regime in

which it was (or is being) deposited. The latter two factors have a varying tendency to reduce the

proportion of unstable material present. Relief also influences the nature of a gravel deposit. For

example, gravel production under low relief is small, and the pebbles tend to be inert residues such

as vein quartz, quartzite, chert and flint. Conversely, high relief and the accompanying rapid erosion

yield coarse, immature gravels. When gravel and larger-size material become indurated, they form

conglomerate; when rubble is indurated, it is termed a breccia (Fig. 1.19). Those conglomerates in

which the fragments are in contact and so make up a framework are referred to as

orthoconglomerates. By contrast, those deposits in which the larger fragments are separated by

matrix are referred to as paraconglomerates. Sands consist of a loose mixture of mineral grains and

rock fragments. Generally, they tend to be dominated by a few minerals, the chief of which is

frequently quartz. Usually, the grains show some degree of orientation, presumably related to the

direction of movement of the transporting medium. The process by which sand is turned into
sandstone is partly mechanical, involving grain fracturing, bending and deformation. However,

chemical activity is much more important. The latter includes decomposition and solution of grains,

precipitation of material from pore fluids and intergranular reactions. Silica (SiO2) is the

commonest cementing agent in sandstones, particularly older sandstones. Various carbonate

cements, especially calcite (CaCO3), are also common cementing materials. Ferruginous and

gypsiferous cements also are found in sandstones. Cement, notably the carbonate types, may be

removed in solution by percolating pore fluids. This brings about varying degrees of decementation.

Quartz, feldspar and rock fragments are the principal detrital components of which sandstones are

composed, and consequently they have been used to define the major classes of sandstone. Pettijohn

et al. (1972) also used the type of matrix in their classification.

In other words, those sandstones with more than 15% matrix were termed wackes. The chief

type of wacke is greywacke, which can be subdivided into lithic and feldspathic varieties. Those

sandstones with less than 15% matrix were divided into three families. The orthoquartzites or quartz

arenites contain 95% or more of quartz; 25% or more of the detrital material in arkoses consists of

feldspar; and in lithic sandstones, 25% or more of the detrital material consists of rock fragments

(Fig. 1.20). Silts are clastic sediments derived from pre-existing rocks, chiefly by mechanical

breakdown processes. They are composed mainly of fine quartz material. Silts may occur in

residual soils, but they are not important in such instances. However, silts are commonly found in

alluvial, lacustrine, fluvio-glacial and marine deposits. These silts tend to interdigitate with deposits

of sand and clay. Silts are also present with sands and clays in estuarine and deltaic sediments.

Lacustrine silts are often banded. Marine silts may also be banded. Wind-blown silts are generally

uniformly sorted. Siltstones may be massive or laminated, the individual laminae being picked out

by mica and/or carbonaceous material. Micro-cross-bedding is frequently developed and the

laminations may be convoluted in some siltstones. Siltstones have high quartz content with

predominantly siliceous cement. Frequently, siltstones are interbedded with shales or fine-grained

sandstones, the siltstones occurring as thin ribs. Loess is a wind-blown deposit that is mainly of silt
size and consists mostly of quartz particles, with lesser amounts of feldspar and clay minerals. It is

characterized by a lack of stratification and uniform sorting, and occurs as blanket deposits in

western Europe, the United States, Russia and China (Fig. 1.21). Deposits of loess are of

Pleistocene age and, because they show a close resemblance to fine-grained glacial debris, their

origin has customarily been assigned a glacial association. For instance, in the case of those regions

mentioned, it is presumed that winds blowing from the arid interiors of the northern continents

during glacial times picked up fine glacial outwash material and carried it for hundreds or thousands

of kilometres before deposition took place. Deposition is assumed to have occurred over steppe

lands, and the grasses left behind fossil root holes, which typify loess. These account for its crude

columnar structure. The lengthy transport explains the uniform sorting of loess. Deposits of clay are

composed principally of fine quartz and clay minerals. The latter represent the commonest

breakdown products of most of the chief rock-forming silicate minerals.

The clay minerals are all hydrated aluminium silicates and possess a flaky habit, that is, they are

phyllosilicates. The three major families of clay minerals are the kandites (kaolinite), illites (illite)

and smectites (montmorillonite). Kaolinite [Al4Si4O10(OH)8] is formed by the alteration of

feldspars, feldspathoids and other aluminium silicates due to hydrothermal action. Weathering

under acidic conditions is also responsible for kaolinization. Kaolinite is the chief clay mineral in

most residual and transported clays, is important in shales, and is found in variable amounts in

fireclays, laterites and soils. It is the most important clay mineral in china clays and ball clays.

Deposits of china clay (kaolin) are associated with acid igneous rocks such as granites,

granodiorites and tonalites, and with gneisses and granulites. Illite [K2-3Al8(Al2-3,Si13-

14)O40(OH)8] is of common occurrence in clays and shales, and is found in variable amounts in

tills and loess, but is less common in soils. It develops as an alteration product of feldspars, micas or

ferromagnesian silicates upon weathering or may form from other clay minerals during diagenesis.

Like kaolinite, illite also may be of hydrothermal origin. The development of illite, both under

weathering and by hydrothermal processes, is favoured by an alkaline environment.


Montmorillonite [(Mg,Al)4(Al,Si)8O20(OH)4.nH2O] develops when basic igneous rocks in

badly drained areas are subjected to weathering. The presence of magnesium is necessary for this

mineral to form, if the rocks were well drained, then the magnesium would be carried away and

kaolinite would develop. An alkaline environment favours the formation of montmorillonite.

Montmorillonite occurs in soils and argillaceous sediments such as shales derived from basic

igneous rocks. It is the principal constituent of bentonitic clays, which are formed by the weathering

of basic volcanic ash, and of fuller’s earth, which is also formed when basic igneous rocks are

weathered. In addition, when basic igneous rocks are subjected to hydrothermal action, this may

lead to the development of montmorillonite. Residual clay deposits develop in place and are the

products of weathering. In humid regions, residual clays tend to become enriched in hydroxides of

ferric iron and aluminium, and impoverished in lime, magnesia and alkalies. Even silica is removed

in hot humid regions, resulting in the formation of hydrated alumina or iron oxide, as in laterite. The

composition of transported clays varies because these materials consist mainly of abrasion products

(usually silty particles) and transported residual clay material. Shale is the commonest sedimentary

rock and is characterized by its lamination. Sedimentary rock of similar size range and composition,

but which is not laminated, is referred to as mudstone. In fact, there is no sharp distinction between

shale and mudstone, one grading into the other. An increasing content of siliceous or calcareous

material decreases the fissility of shale, whereas shales that have a high organic content are finely

laminated. Laminae range from 0.05 to 1.0 mm in thickness, with most in the range of 0.1 to 0.4

mm. Clay minerals and quartz are the principal constituents of mudstones and shales. Feldspars

often occur in the siltier shales. Shale may also contain appreciable quantities of carbonate,

particularly calcite, and gypsum (CaSO4.2H2O). Indeed, calcareous shale frequently grades into

shaly limestone. Carbonaceous black shales are rich in organic matter, contain a varying amount of

pyrite (FeS2), and are finely laminated. The term limestone is applied to those rocks in which the

carbonate fraction exceeds 50%, over half of which is calcite or aragonite (CaCO3). If the carbonate

material is made up chiefly of dolomite (CaCO3.MgCO3), the rock is named dolostone (this rock
generally is referred to as dolomite, but this term can be confused with that of the mineral of the

same name). Limestones and dolostones constitute about 20 to 25% of the sedimentary rocks,

according to Pettijohn (1975). This figure is much higher than some of the estimates provided by

previous authors. Limestones are polygenetic. Some are of mechanical origin, representing

carbonate detritus that has been transported and deposited. Others represent chemical or

biochemical precipitates that have formed in place. Allochthonous or transported limestone has a

fabric similar to that of sandstone and also may display current structures such as cross bedding and

ripple marks. By contrast, carbonate rocks that have formed in situ, that is, autochthonous types,

show no evidence of sorting or current action and at best possess a poorly developed stratification.

Exceptionally, some autochthonous limestones show growth bedding, the most striking of which is

stromatolitic bedding, as seen in algal limestones. Lithification of carbonate sediments often is

initiated as cementation at points of intergranular contact rather than as consolidation. In fact,

carbonate muds consolidate very little because of this early cementation. The rigidity of the weakest

carbonate rocks, such as chalk, may be attributed to the mechanical interlocking of grains with little

or no cement. Although cementation may take place more or less at the same time as deposition,

cemented and uncemented assemblages may be found within short horizontal distances. Indeed, a

recently cemented carbonate layer may overlie uncemented material. Because cementation occurs

concurrently with or soon after deposition, carbonate sediments can support high overburden

pressures before consolidation takes place. Hence, high values of porosity may be retained to

considerable depths of burial. Eventually, however, the porosity is reduced by postdepositional

changes that bring about recrystallization. Thus, a crystalline limestone is formed in this manner.

Folk (1973) distinguished two types of dolostone. First, he recognized an extremely finegrained

crystalline dolomicrite (less than 20 microns grain diameter), and secondly, a more coarsely grained

dolostone in which there was plentiful evidence of replacement. He regarded the first type as of

primary origin and the second as being formed as a result of diagenetic replacement of calcite by

dolomite in limestone. Primary dolostones tend to be thinly laminated and generally are
unfossiliferous. They are commonly associated with evaporates and may contain either nodules or

scattered crystals of gypsum or anhydrite (CaSO4). In those dolostones formed by dolomitization,

the original textures and structures may be obscured or may even have disappeared. Evaporitic

deposits are quantitatively unimportant as sediments. They are formed by precipitation from saline

waters, the high salt content being brought about by evaporation from inland seas or lakes in arid

areas. Salts can also be deposited from subsurface brines, brought to the surface of a playa or

sabkha flat by capillary action (Fig. 1.22). Seawater contains approximately 3.5%, by weight, of

dissolved salts, about 80% of which is sodium chloride. Experimental work has shown that when

the original volume of seawater is reduced by evaporation to about half, a little iron oxide and some

calcium carbonate are precipitated. Gypsum begins to form when the volume is reduced to about

one-fifth of the original, rock salt begins to precipitate when about one-tenth of the volume remains,

and, finally, when only 1.5% of the seawater is left, potash and magnesium salts start to crystallize.

This order agrees in a general way with the sequences found in some evaporitic deposits, however,

many exceptions are known. Many complex replacement sequences occur among evaporitic rocks,

for example, carbonate rocks may be replaced by anhydrite and sulphate rocks by halite (NaCl).

Organic residues that accumulate as sediments are of two major types, namely, peaty material that

when buried gives rise to coal, and sapropelic residues. Sapropel is silt rich in, or composed wholly

of, organic compounds that collect at the bottom of still bodies of water. Such deposits may give

rise to cannel or boghead coals. Sapropelic coals usually contain a significant amount of inorganic

matter as opposed to humic coals in which the inorganic content is low. The former are generally

not extensive and are not underlain by seat earths (i.e. fossil soils). Peat deposits accumulate in

poorly drained environments in which the formation of humic acid gives rise to deoxygenated

conditions. These inhibit the bacterial decay of organic matter. Peat accumulates wherever the

deposition of plant debris exceeds the rate of its decomposition. A massive deposit of peat is

required to produce a thick seam of coal; for example, a seam 1 m thick probably represents 15 m of

peat.
Chert and flint are the two most common siliceous sediments of chemical origin. Chert is a

dense rock composed of one or more forms of silica such as opal, chalcedony or microcrystalline

quartz. Sponge spicules and radiolarian remains may be found in some cherts, and carbonate

material may be scattered throughout impure varieties. Gradations occur from chert to sandstone

with chert cement, although sandy cherts are not common. Chert may suffer varying degrees of

devitrification. Chert may occur as thin beds or as nodules in carbonate host rocks. Both types are

of polygenetic origin. In other words, chert may be a replacement product, as in siliceous limestone,

for example, or it may represent a biochemical accumulate formed in a basin below the calcium

carbonate compensation depth. In yet other cases, chert may be the product of an ephemeral silica-

rich alkaline lake environment. Some sediments may have a high content of iron. The iron

carbonate, siderite (FeCO3), often occurs interbedded with chert or mixed in varying proportions

with clay, as in clay ironstones. Some iron-bearing formations are formed mainly of iron oxide,

hematite (Fe2O3) being the most common mineral. Hematite-rich beds are generally oolitic.

Limonite (2Fe2O3.3H2O) occurs in oolitic form in some ironstones. Bog iron ore is chiefly an

earthy mixture of ferric hydroxides. Siliceous iron ores include chamositic ironstones (chamosite,

Fe3Al2Si2O10.3H2O), which are also typically oolitic. Glauconitic [glauconite,

K(Fe3Al)2(Si,Al)4O10(OH)2] sandstones and limestones may contain 20% or more FeO and

Fe2O3. On rare occasions, bedded pyrite has been found in black shale.

Lecture 4. Generation, migration and trapping of hydrocarbons

Oil and gas traps

Oil and gas traps, sometimes referred to as petroleum traps are below ground traps where a
permeable reservoir rock is covered by some low permeability cap rock. This combination of rock
can take several forms, but they all prevent the upward migration of oil and natural gas up through
the reservoir rock. Once oil and natural gas are in the reservoir rock, they continue to migrate
upwards through the pore spaces of the rock until blocked by some sort of seal with a cap rock.
[1]
 The low permeability cap rocks are generally shale or low permeability sandstones and carbonate
rocks.[2]
There are two major categories of traps, and are classified based on how the petroleum accumulates.
The two main groups are structural traps and stratigraphic traps.[2] For a more in-depth
description of the overall structure of oil and gas deposits, see oil and gas reservoir.

Figure 1. Diagram showing the structure of several different types of oil and gas traps.[3]

Structural Traps

These traps are types that form as a result of some structural deformation - a bend or dip - of rock.
These traps take on several forms and shapes as a result of different types of deformation. This
deformation tends to take place over tens or hundreds of millions of years after sediment that
creates the seals and rocks themselves have been deposited. [1] In these traps, the pores of the
reservoir rock contain oil, gas, or water. Gas moves up in the trap as it is the lightest, with oil below
it and water at the bottom. The cap rock prevents upward migration of these fluids.
Anticline Traps

These types of traps are formed by a folding of rock. Specifically, a sandstone bed covered with low
permeability shale is folded into a trap that contains petroleum products. Hydrocarbons are trapped
in the peak of this fold.[2] An anticline trap is shown in the upper-left of Figure 1. Most anticline
traps are created as a result of sideways pressure, folding the layers of rock, but can also occur from
sediments being compacted. For more information on anticline traps, click here.[4]
Fault Traps

These types of traps are formed when reservoir rock is split along a fault line. Between the walls of
the split reservoir, clay traps oil and prevents it from leaving the trap. [4] Other times there exists a
pressure differential across the two sides of the fault that prevents the fluids from migrating. [2] A
fault trap is shown in the upper-right of Figure 1. Although faulting is common in many petroleum
fields, traps that result from faulting alone are not very common. Generally speaking, faulting
creates some other type of structure that creates the main trap.[2]
Salt Domes

Salt dome traps are formed as a result of below ground salt - which is less dense than the rock above
it - moving upwards slowly. This upward migration of the salt can deform and break up rock along
the way.[4] The process of this salt deforming rock is known as salt tectonics, and take place over
hundreds of millions of years.[2] Oil and gas that flows through the reservoir rock will come to rest
when it reaches the salt dome and is then trapped. A salt dome trap is shown on the lower left of
Figure 1.
Stratigraphic Traps

These traps are formed as a result of the deposition in sedimentary rocks. When the sediment that
creates the reservoir rock is deposited in a discontinuous layer, the seals are created beside and on
top of the reservoir. In some cases, these seals are made of impermeable or low permeability shale
deposited around the reservoir, blocking the oil and gas inside. The seals themselves may also be
source rocks.[1] A stratiographic trap is shown on the bottom right of Figure 1.
There are two main types of stratigraphic traps that are classified by when changes occur relative to
the sedimentation process. Primary stratigraphic traps result from changes that develop during the
sedimentation process. These are generally structural changes that arise as a result of discontinuous
deposition of sediment. Secondary stratigraphic traps result from changes that develop after
sedimentation has occurred. These changes can involve changes in porosity of the rock that lead to
formation of a cap-like rock.[1]
Migration of Petroleum

Petroleum migrates from low permeability source rocks into high permeability reservoir rocks

from which the petroleum can be produced (Fig. 1.2b). The main driving force for petroleum

migration is buoyancy because it is less dense than water. The forces acting against migration are

the capillary forces and the resistance to flow though rocks with low permeabilities Migration of oil

and gas will therefore nearly always have an upwards component. We distinguish between primary

migration, which is the flow of petroleum out of the source rock and secondary migration, which is

the continued flow from the source rock to the reservoir rock or up to the surface (Fig. 1.4). Oil and

gas may also migrate (leak) from the reservoir to a higher trap or to the surface. Hydrocarbons are
relatively insoluble in water and will therefore migrate as a separate phase. Solubility varies from as

little as 24 ppm for methane to 1,800 ppm for benzene. Other compounds, such as pentane, are even

less soluble (2–3 ppm). However, solubility increases markedly with pressure. Many hydrocarbons

have solubilities of less than 1 ppm in water It is difficult to envisage oil being dissolved in water

and transported in an aqueous solution, both because of the solubility and the low flow rates. It

would also be difficult to explain how the oil would come out of solution in the reservoirs (traps).

Gas, in particularly methane, has a fairly high solubility in water, especially under high pressure. If

methane-saturated water rises to lower pressures, large quantities of methane can bubble out of a

solution. It is therefore necessary to assume that oil is mostly transported as a separate phase. Oil is

lighter than water, and oil droplets would be able to move through the pores in the rocks but the

caplliary restance is high for separate oil drops in a water-wet rock (Fig. 1.5). In order to pass

through the narrow passage between pores (pore throat), the oil droplets must overcome the

capillary forces. When the pores are sufficiently small in a fine-grained sediment, these forces will

act as a barrier to further migration of oil. The small gas molecules, however, can diffuse through

extremely small pores and thus escape from shales which form tight seals for oil.

Fig. 1.4 Schematic illustration of primary migration (expul- sion) of petroleum from a source
rock and secondary migration into a reservoir (trap). This example is from the northern North
Sea where rifting in Upper Jurassic time produced good condi- tions for the formation of a source
rock and also traps on the uplifted fault blocks
Oil can therefore not migrate as small discrete droplets, but moves as a continuous string of oil
where most of the pores are filled with oil rather than water (highly oil-saturated). The pressure
in the oil phase at the top is then a function of the height of the oil- saturated column (string) and
the density difference between oil and water.
The rate of migration is a function of the rate of petroleum generation in the source rocks.
This is a function of the temperature integrated over time (Fig. 1.6).
The temperature history is a function of the burial depth and the geothermal gradients.
Deep burial over long time will cause all oil to be decomposed (cracked) into gas.
The degree of alteration of organic matter can be measured in different ways. Plant material
is altered from a dull material to a material which becomes more shiny with increasing
temperature. This can be quan- tified by measuring the amount of light reflected from a piece
of plant material (vitrinite) under the micro- scope. A vitrinite reflectivity of 1.2 indicates that
the source has generated much of the oil that can be gener- ated. We will say that the source
rock is in the middle of the “oil window” (Fig. 1.7). Values below 0.7–0.8. are found in
source rocks which have not been heated enough (immature source rocks).
Vitrinite reflectivities close to 2.0 and above indi- cate that the source rock has generated
all the oil and can generate only gas.

Hydrocarbon Traps

Traps consist of porous reservoir rocks overlain by tight (low permeability) rocks which do not

allow oil or gas to pass. These must form structures closed at the top such that they collect oil and

gas, which is lighter than water. We can think of an oil trap as a barrel or bucket upside down (Fig.

1.9) which can then be filled with petroleum which rises through the water until it is full. The point

where the petroleum can leak from this structure is called the spill point. The closure is the

maximum oil column that the structure can hold before leaking through the spill point (Fig. 1.10).

The cap rock may not be 100% effective in preventing the upward flow of hydrocarbons, but these

will still accumulate if the rate of leakage is less than the rate of supply up to the trap. Cap rocks are

usually not totally impermeable with respect to water, but may be impermeable to oil and gas due to

capillary resistance in the small pores. Traps can be classified according to the type of structure that

produces them. We distinguish between: (1) Structural traps that are formed by structural

deformation (folding, doming or faulting) of rocks. (2) Stratigraphic traps which are related to

primary features in the sedimentary sequences and do not require structural deformation like
faulting or folding. This may be sandstones pinching out in shales due to primary changes in facies

(Fig. 1.11).

Carbonate reefs tend to form primary structures which function as stratigraphic traps. There are

also combinations between stratigraphic and structural traps (Fig. 1.12). It is important to establish

when structural traps were formed in relation to the migration of the petroleum. Structures formed

after the main phase of source rock maturation and associated migration will not be effective traps.

In some cases traps formed late can collect gas which normally is generated and migrates later than

oil. Stratigraphic traps, by contrast, have been there all the time, and the timing of the migration is

not so important. They may however depend on slight tilting of the strata involved.

Structural Traps

(a) Anticlinal Domes

Domes formed by diapirism or other processes may form closures in all directions (four-way

closure). A simple anticline is not sufficient to trap oil. Anticlines with an axial culmination are

needed to provide four-way closure. This means that the fold axis must be dipping in both

directions (Fig. 1.9). Anticlinal traps can form in association with faulting. This is especially true in

connection with growth faults (roll-overs) (see below), but also with thrust zones.

(b) Salt Domes

Salt domes are formed because salt (specific gravity c.1.8–2.0) is lighter than the overlying

rock, and the salt therefore “floats” up due to buoyancy. The quantitatively most important salt

minerals are halite (NaCl – density 2.16 g/cm3), gypsum (CaSO4. 2H2O – density 2.32 g/cm3).

Anhydrite (CaSO4 – density 2.96 g/cm3) is too dense to contribute to the formation of diapers. In

order for the salt to move upwards and form a salt dome, a certain thickness of overburden is

required and the salt beds themselves must be at least 100–200 m thick. The upward movement of

salt through the overlying sequence, and the resultant deformation of the latter, is called
halokinetics or salt tectonics. The rate of salt movement is extremely slow and a dome may take

several million years to form. Movements of the earth’s surface may, however, also be recorded in

recent history as is the case onshore Denmark. Salt may break right through the overlying rocks and

rise to the surface, or form intrusions in younger sediments. If gypsum has been deposited this will

be altered into anhydrite at about 1 km burial depth, with a consequent 40% compaction and the

increase in density will remove the buoyancy relative to the surrounding sediments. A comparable

expansion occurs when rising anhydrite comes into contact with groundwater and reverts to

gypsum. Traps may be created

(1) in the layers above the salt dome,

(2) in the top of the salt dome (cap rock),

(3) in the beds which are faulted and turned up against the salt structure and

(4) through stratigraphic pinching out of beds round the salt dome.

Reservoirs may form by solution and brecciation at the top of salt domes. Salt tectonics is of

great importance in many oilbearing regions where there are thick salt deposits in the passive

margin sequences of the South Atlantic and the Gulf of Mexico. In the eastern USA we find

extensive tracts with Silurian salt, and in Texas and New Mexico we have Permian salt. Salt layers

are the ideal cap rock because of salt’s low permeability and ductile properties, which prevent

fracturing and leakage. Salt deposits are particularly common in the PermoTriassic around the

Atlantic. This is because prior to the opening of the Atlantic there were vast areas with fault-

controlled basins (rifts) in the middle of a supercontinent (America + Europe, Asia and Africa) with

little precipitation. We find similar conditions today around the Red Sea and the Dead Sea. The

Permian Zechstein salt in Germany and Denmark continues below the North Sea, and halotectonic

movements have formed dome structures in the Chalk, for example in the Ekofisk area.

(c) Growth Anticlines


These are dome-like structures formed when part of a basin subsides more slowly than its

surroundings, resulting in least sedimentation on the highest part. The sediment thickness decreases

towards the dome centre, which also compacts less than the adjacent thicker sediments and thus

contributes to the formation of an anticlinal structure. Growth anticlines form contemporaneously

with sediment accumulation, not through later folding. Growth anticlines can be formed above salt

domes, reefs or buried basement highs, through differential compaction.

(d) Fault Traps

In fault traps, the fault plane forms part of the structure trapping the oil and hindering its further

upward migration. The fault plane must therefore be sealing for vertical flow in order to function as

a barrier and a cap rock for the reservoir rocks. If the reservoir rock is juxtaposed against a

sandstone or other permeable rocks the fault must also be impermeable for flow across the fault

plane. Most frequently, however, the reservoir rock is faulted against a tight shale or mudrock and

the fault is then in most cases sealing (Fig. 1.9). When there is sandstone on both sides of the fault

plane, the permeability across it will amongst other things depend on how much clay has been

smeared along the junction. At greater depths (>3–4 km) there may be diagenetic changes such as

quartz cementation, which can make the fault plane tighter. There are many different types of faults:

Normal faults – often in connection with graben (rift) structures.

Strike-slip faults.

Reverse faults formed by tectonic stress.

Growth faults.

The displacement along faults can be both vertical (normal faults) and horizontal (strike slip

faults). Reverse faults are faults where the hangingwall is moved upwards relative to the footwall

below the fault plane. These are typical of areas with high horizontal stresses i.e. due to converging
plate movements. Growth faults are driven by gravity-sliding along curved (listric) fault planes and

are typical of sedimentary sequences such as deltas deposited with relatively rapid sedimentation.

The fault plane is often (though not always) sealing and can stop oil and gas from migrating further

upward. However, oil traps are equally often formed in anticlines on the upper side of the fault

plane. These are rollover anticlines. Because the faulting is active during sedimentation, the layers

on the downthrown side will be thickest. The name growth fault goes back to the early days of oil

exploration without seismic data. It was noticed that the layers had “grown” in thickness in the

wells on the downthrown side of the fault. The displacement of the beds decreases upwards along

the fault plane. Smaller, antithetic faults often develop in the opposite direction in the beds which

are turned inwards towards the main fault plane. Growth faults tend to have low permeability and

may contribute greatly to reduced porewater circulation in sedimentary basins, and we often find

undercompacted clay, which can turn into clay diapirs in association with growth faults.

Stratigraphic Traps

These are traps which are partially or wholly due to facies variation or unconformities, and not

primarily the result of tectonic deformation. Porous and permeable sands which pinch out up-dip in

less permeable rocks, e.g. shale (Figs. 1.11 and 1.12) are good examples. Barrier islands often form

stratigraphic traps because they may be separated from the coast by fine-grained lagoonal facies.

The main types are:

(a) Fluvial channel sandstones may be isolated and surrounded by impermeable clay-rich

sediments, or they may be folded so that we obtain a combination of stratigraphic and structural

traps (Figs. 1.9 and 1.10).

(b) Submarine channels and sandstone turbidites in strata rich in shale. Here we will often find

pinchout of permeable layers up-dip from the foot of the continental slope. This will result in

stratigraphic traps without any further folding being necessary.


(c) Reefs often form stratigraphic traps. A reef structure projects up from the sea bed and often

has shale sediments surrounding it, so that oil could migrate from the shale into the reef structure.

(d) Traps related to unconformities. Sandstones or other porous rocks may be overlain with an

angular unconformity by shales or other tight sediments, forming a trap underneath the

unconformity (Fig. 1.12). Topographic highs in the basement overlain with shales can also provide

good traps in fractured basement rocks. Remember that oil can migrate upwards into

stratigraphically lower rocks. In China there are numerous examples of this type of trap.

Much of the oil generated in sedimentary basins has not been trapped in reservoirs but reached

the surface on the seafloor or on land. There it is then broken down by bacteria and becomes heavy

oil, which is not very toxic. In California there are many examples of natural oil seeps which can be

observed along roads, on the beach (Fig. 1.13) and also offshore.

Other Types of Trap

More unusual kinds of trap can be encountered. If the porewater in a sedimentary basin has

sufficiently strong flow of meteoric water into the basin, the oil/water contact may diverge

markedly from a horizontal plane due to the hydrodynamic stresses. This has implications for

calculations of oil volumes within a structure, and in some instances oil can accumulate without

being sealed in, within a so-called hydrodynamic trap. The circulation of fresh (meteoric) water

down into oil-bearing rocks will, however, lead to biodegradation and the formation of asphalt.

Asphalt can then become a tight cap rock for the oil. At greater depths, beyond the reach of

meteoric water, water movement is limited and any deflection of the oil/water contact is more likely

to reflect pressure differences within the reservoir. Water will also flow then because of the pressure

gradient, but unless there are low permeability barriers the pressure will soon equalise. Tectonic

tilting will also tilt the oil/water contact. Reservoir Geology is not a well-defined discipline. It

includes many aspects of geology that are of special relevance to the production of petroleum. It is
also linked to engineering aspects of petroleum production. Reservoir geophysics has in recent

years become very important and is now well integrated with reservoir geology.

Lecture 5. Reservoir porosity and permeability

Any rock with sufficiently high porosity and permeability may serve as a reservoir rock

provided that there is a source of petroleum, a structure, and a tight cap rock. Sediments consist of

solid grains and of fluids which for the most part are water but may be oil and gas. Porosity (ϕ) is an

expression of the percentage (or fraction) of fluids by volume (Vf) compared to the total rock

volume with fluids (Vt), so that ϕ = Vf / Vt. Porosity is often expressed as a percentage, but in

many calculations it is easier to express it as a fraction, for example 0.3 instead of 30% porosity.

The void ratio (VR) is the ratio between pore volume (ϕ) and the volume of the grains (1–ϕ).

VR = ϕ/ (1 − ϕ)

Void ratio is often used in engineering and it has certain advantages in some mathematical

expressions. If we assume that we know the density of the mineral grains, the porosity can be found

by measuring the density of a known volume of the sediment. The density of the sediments (ρs) is

the sum of the density of the grains, which are mostly minerals ρm, and the density of the fluids

(ρf).

ρs = ϕρf + ρm (1 − ϕ)

Well sorted, rounded sand grains are almost spherical in shape. If we have grains of the same

size, which are all quite well rounded and with a high degree of sphericity, we will be able to pack

the grains so as to get minimum porosity. Rhombic is the densest packing, resulting in 26%

porosity, but this can not be obtained naturally. Cubic packing, where the grains are packed directly

one above another, results in about 48% porosity and this does not occur in nature either. Most well

sorted sandstones have a porosity which lies between these two values, typically around 40–42%.

Poorly sorted sand may have lower primary porosity and will also compact more at moderate burial
depths. Clay-rich sediments have a much greater porosity immediately after deposition, typically

60–80%. This means that immediately following deposition a sand bed is denser than a bed of clay

or silt. However, clay and silt lose their porosity more rapidly with burial. Porosity may be

classified into different types depending on its origin. Pore space between the primary sediment

grains is often referred to as primary porosity. Intergranular porosity simply means porosity

between the grains whereas intragranular porosity means porosity inside the sediment grains. The

latter may be cavities in fossils, e.g. foraminifera, gastropods, molluscs, but also partly dissolved

feldspar and rock fragments. Pore space formed by dissolution or fracturing of grains is called

secondary porosity. Cavities formed by selective solution of sediment grains or fossils are classified

as mouldic porosity. A typical example is when dissolution of aragonite fossils like gastropods

leaves open pore spaces (moulds). Particularly in carbonates we may also have porosity on a large

scale i.e. as caverns (karst) and in reefs. Pore space produced by fracturing is called fracture

porosity. Permeability is an expression of the ease with which fluids flow through a rock. It will

depend on the size of the pore spaces in the rocks, and in particular the connections between the

pore spaces. Even thin cracks will contribute greatly to increasing the permeability. Permeability

can be measured by letting a liquid or gas flow through a cylindrical rock sample under pressure.

The pressure difference P1−P2 between the two ends of a horizontal cylinder is P, the cylinder

length L, and the flow rate of water (or another fluid) through the cylinder, is Q (cm3/s). A is the

crosssection and μ the viscosity of the fluid Q = k· A·P L·μ where k is the permeability, which is

expressed in Darcy. The volume of water which flows through each surface unit in the cross-section

A is thus equal to the flux F = Q/A. F can be measured in cm3/cm2/s or in m3/m2/s. This is equal to

the Darcy velocity which is m/s. Well-sorted sandstones may have a permeabilities exceeding 1

Darcy and values between 100 and 1,000 mD are considered to be extremely good. Permeabilities

of 10–100 mD are also considered to be good values for reservoir rocks. Permeabilities of 1–10 mD

are typical of relatively dense sandstones and limestones, so-called tight reservoirs. There are also

examples of rocks with even lower permeabilities being exploited commercially for oil production,
for example in the Ekofisk Field where the generally low permeability of a chalk matrix is enhanced

by fractures which increase the overall permeability. In the great majority of rocks, the permeability

differs according to flow direction. In sedimentary rocks the permeability is much higher parallel to

the bedding compared with normal to the bedding. Channel sandstones can also have a marked

directional impact on the permeability. In well-cemented sandstones and limestones, and also in

certain shales, the matrix permeability is extremely low and the effective permeability may be

mostly controlled by fractures if they are present. Claystones and shales have very low permeability

and can be almost completely tight. In the laboratory shale permeabilities as low as 0.01 nanodarcy

have been measured. Samples from cores or outcrops can contain minute fissures formed in

response to unloading during retrieval to the surface and these must be closed to replicate the in situ

permeability prior to unloading. Most rocks are far from homogeneous. We may measure the

porosity and permeability of a hand specimen or core plug, but it is not certain that these are

representative of a larger volume. Fractures occur at varying intervals, and range in size from large,

open joints down to microscopic cracks which can barely be seen in a microscope. Rocks with low

porosity and permeability may fracture and sufficiently increase their porosity, and particularly

permeability, to form large oil reservoirs. This means that reservoirs may be good producers despite

relatively low porosity. Occasionally we find petroleum in fractured metamorphic and igneous

rocks but reservoirs normally consist of sedimentary rocks. Sandstones make up about 50–60% of

the reservoirs in the world and carbonate reservoirs may account for almost 40%. Many of the

reservoirs in the Middle East are carbonate rocks but in the rest of the world the percentage of

carbonate reservoirs is lower. The most important aspects of reservoir rocks include:

(1) The external geometry such as the thickness and extent of the reservoir rock in all

directions.

(2) The average porosity, pore size and pore geometry.


(3) The distribution of permeability in the reservoirs, particularly high permeability conduits

and low permeability barriers to fluid flow.

(4) Mineralogy and wettability of the pore network.

The properties of sandstones and carbonate reservoirs are primarily linked to the depositional

environment, the textural and mineralogical composition and the burial history. A good background

in general sedimentology, facies analysis and sequence stratigraphy is therefore important.

Nevertheless, many of the important properties of reservoir rocks linked to changes in facies and

smaller faults are below the vertical resolution of exploration seismic (15–30 m) and it is important

to establish relationships between facies models, diagenetic processes and reservoir properties. The

properties of faults are also very important factors determining oil flow during production.

Lecture 6. General reservoir properties

Petroleum reservoirs may contain any of the three fluid phases—water (brine), oil, or gas. The
initial distribution of phases depends on depth, temperature, pressure, composition, historical
migration, type of geological trap, and reservoir heterogeneity (that is, varying rock properties). The
forces that originally distribute the fluids are gravity, capillary, molecular diffusion, thermal
convection, and pressure gradients. It is generally assumed that reservoir fluids are in a static state
when discovered or, more correctly, that fluids are moving at a very slow rate relative to the time
required to extract the fluids (10 to 50 years). Clearly the fluids may still be in a dynamic state in
terms of geological time.
Because gravity is the dominant force in distributing fluids through geological time, hydrocarbons
migrate upward and are trapped against impermeable cap rock. Gas overlies oil, which overlies
water. However, because the reservoir pores are usually saturated completely by water before
[[hydrocarbon migration]] and because capillary forces acting to retain water in the smallest pores
exceed gravity forces, an initial (connate) [[water saturation]] will always be found in hydrocarbon-
bearing formations. The connate water saturation may vary from 5 to 50% with the hydrocarbons
still having sufficient mobility to produce at economical rates.
This article, along with the See also articles, reviews the physical and thermodynamic properties of
gas, oil, and reservoir brine. As commonly done, the phase and volumetric behavior of petroleum
reservoir fluids is referred to as PVT (pressure-volume-temperature). Two important general
references on PVT are Katz et al.[1] and Society of Petroleum Engineers.[2]

Property definitions

Some basic fluid property definitions are provided here:


Formation volume factor (FVF)

The ratio of a phase volume (water, oil, gas, or gas plus oil) at reservoir conditions, relative to the
volume of a surface phase (water, oil, or gas) at standard conditions resulting when the reservoir
material is brought to the surface. Denoted mathematically as Bw(bbl/STB),
B0 (bbl/STB), Bg (ft3/SCF), and Bt (bbl/STB).
Solution gas-oil ratio (GOR)

The amount of surface gas that can be dissolved in a stock tank oil when brought to a specific
pressure and temperature. Denoted mathematically as Rs (SCF/STB).
Solution oil-gas ratio (OGR)

The amount of surface condensate that can be vaporized in a surface gas at a specific pressure and
temperature; sometimes referred to as liquid content. Denoted mathematically as rs (STB/MMSCF).
Liquid specific gravity

The ratio of density of any liquid measured at standard conditions (usually 14.7 psia and 60 °T) to
the density of pure water at the same standard conditions. Denoted mathematically as γo (where
water = 1).
API specific gravity

Another common measure of oil specific gravity, defined by γAPI = (141.5/γO) –131.5, with units in
°API.
Gas specific gravity

The ratio of density of any gas at standard conditions (14.7 psia and 60 °F) to the density of air at
standard conditions; based on the ideal gas law (pV = nRT), gas gravity is also equal to the gas
molecular weight divided by air molecular weight (M air = 28.97). Denoted mathematically as
γg (where air = 1).
Bubblepoint pressure

At a given temperature, this condition occurs when an oil releases an infinitesimal bubble of gas
from solution when pressure drops below the bubblepoint.
Retrograde dewpoint pressure

At a given temperature, this condition occurs when a gas condenses an infinitesimal drop of oil
from solution when pressure drops below the dewpoint.
Saturation pressure

An oil at its bubblepoint pressure or a gas at its dewpoint pressure.


Critical point

The pressure and temperature of a reservoir fluid where the bubblepoint pressure curve meets the
retrograde dewpoint pressure curve (see Figures 1 and 2), representing a unique state where all
properties of the bubblepoint oil are identical to the dewpoint gas.
Composition or feed

Quantifies the amount of each component in a reservoir mixture, usually reported in mole fraction.
Typical components in petroleum reservoir mixtures include the nonhydrocarbons N 2, CO2, and H2S
and the hydrocarbons C1 C2, C3iC4nC4, iC5, nC5, C6, and C7+ (C7+, or “heptanes-plus,” includes many
hundreds of heavier compounds, such as paraffins, napthenes, and aromatics). Asphaltenesare also
found in reservoir oils.
Saturated condition

A condition where an oil and gas are in thermodynamic equilibrium, that is, the chemical force
exerted by each component in the oil phase is equal to the chemical force exerted by the same
component in the gas phase, thereby eliminating mass transfer of components from one phase to the
other.
Undersaturated condition

A condition when an oil or a gas is in a single phase but not at its saturation point (bubblepoint or
dewpoint), that is, the mixture is at a pressure greater than its saturation pressure.

Reservoir water

The water found in petroleum reservoirs is usually a brine consisting mostly of sodium chloride
(NaCl) in quantities from 10 to 350 ppt (‰); seawater has about 35 ppt. Other compounds
(electrolytes) found in reservoir brines include calcium (Ca), magnesium (Mg), sulfate (SO 4),
bicarbonate (HCO3), iodide (I), and bromide (Br). Brine specific gravity increases with salinity in
units of about 0.075 per 100 ppt.
At reservoir conditions, the brine that is sharing pore space with hydrocarbons always contains a
limited amount of solution gas (mainly methane), from about 10 SCF/STB at 1000 psia to about 35
SCF/STB at 10,000 psia for gas-water systems and slightly less for oil-water systems. Increasing
salinity decreases gas in solution. Water compressibility ranges from 2.5 to 5 × 10 –6 psi–1,
decreasing with increasing salinity. Water viscosity ranges from about 0.3 cP at high temperatures
(>250°F) to about 1 cP at ambient temperatures, increasing with increasing salinity. Finally,
reservoir brines exhibit only slight shrinkage (<5%) when produced to the surface.

Petroleum reservoir classifications

Petroleum reservoirs are usually classified into five fluid categories[3]:

 Dry gas
 Wet gas
 Gas condensate
 Volatile oil
 Black oil
The first three of these are gas reservoir fluid types, which are in a gaseous state at virgin reservoir
conditions, meaning that the critical temperature of the reservoir fluid is less than the reservoir
temperature. Dry gas and wet gas fluids consist mainly of light and intermediate hydrocarbons (N 2,
CO2, H2S, and C1 to C2), in which no liquids will condense in the reservoir rock during pressure
depletion. Wet gases produce high API condensate (distillate) at surface conditions in amounts
usually less than about 5 STB/MMSCF. The OGR should remain constant throughout the depletion
of a wet gas reservoir.
Gas condensates, in contrast, contain significant amounts of C5+ components, and they exhibit the
phenomenon of retrograde condensation at reservoir conditions, in other words, as pressure
decreases, increasing amounts of liquid condenses in the reservoir (down to about 2000 psia). This
results in a significant loss of in situ condensate reserves that may only be partially recovered by
revalorization at lower pressures. Gas condensate reservoirs exhibit producing gas-oil ratios from
2500 to 50,000 SCF/STB (400 to 10 STB/MMSCF). Gas cycling projects designed to avoid liquid
loss from retrograde condensation can usually be justified for fluids with liquid content higher than
about 50 to 100 STB/MMSCF. Offshore, the minimum liquid content to justify cycling is about 100
STB/MMSCF.
Reservoir oils are classified as either black oil or volatile oil, the former being more commonly
discovered in the first 50 years of the oil industry. Volatile oil reservoirs have become the norm in
the past 20 years, mainly because discoveries are at greater depths with higher initial pressures. A
clear demarcation between these two oil types is not easily made, although a gas-oil ratio of about
750 SCF/STB is probably a good indicator (black oils have lower GORs). Volatile oils may have
GORs up to 2500 SCF/STB and formation volume factors as large as three (meaning that the oil
shrinks by a factor of three when produced to the surface). Another characteristic of volatile oil
reservoirs is that the reservoir gas that evolves and flows into the wellbore will contain significant
quantities of liquids that may eventually contribute the majority of surface oil production at late
stages of depletion.
Table 1 gives some typical reservoir fluid compositions and properties. Figure 1 shows a pressure-
temperature diagram for a specific reservoir fluid composition. Depending on reservoir temperature,
this fluid would be defined as an oil or a gas. An oil exhibits a bubblepoint pressure at saturated
conditions, while a gas condensate exhibits a dewpoint pressure at saturated conditions.

Component and Dry Wet Volatile Black


Condensate
Property gas Gas oil oil

CO2 0.10 1.41 2.37 1.82 0.02

N2 2.07 0.25 0.31 0.24 0.34

C1 86.12 92.46 73.19 57.60 34.62

C2 5.91 3.18 7.80 7.35 4.11


C3 3.58 1.01 3.55 4.21 1.01

i C4 1.72 0.28 0.71 0.74 0.76

n C4 — 0.24 1.45 2.07 0.49

i C5 0.50 0.13 0.64 0.53 0.43

n C5 — 0.08 0.68 0.95 0.21

C6s — 0.14 1.09 1.92 1.16

C7+ — 0.82 8.21 22.57 56.40

GOR (SCF/STB) ∞ 69,000 5965 1465 320

OGR
0 15 165 680 3125
(STB/MMSCF)

γAPI — 65.0 48.5 36.7 23.6

M7+ — 132 184 240 274

γ7+ — 0.750 0.816 0.864 0.920


If a reservoir contains both a gas cap and an oil zone, then both fluids are normally at saturated
conditions initially. Initial pressure equals the dewpoint of the gas cap fluid, and it equals the
bubblepoint of the underlying oil (Figure 2). The repeat formation tester (RFT) has made the
determination of initial fluid contacts possible in reservoirs with reasonable permeability, that is, >1
md. A saturated gas cap in equilibrium with an underlying saturated oil, for example, will be seen as
a sharp discontinuity in RFT pressures at the gas-oil contact.
In the past 20 years, deeper petroleum reservoirs have been discovered and the traditional
interpretation of a reservoir containing both gas and oil has changed. An alternative interpretation in
some gas-oil reservoirs is that composition varies continuously with depth. Here the fluids at the
shallowest elevations are gas condensates, while the fluids at greater depths are oils. Sometimes the
initial reservoir pressure may be greater than the saturation pressure of all mixtures in the reservoir,
implying that the reservoir is entirely undersaturated even though a gas is at the top and an oil is at
the bottom of the reservoir. Reservoirs of this type would not show a sharp contrast in RFT
pressures at the depth where the fluid changes from a near-critical gas to a near-critical oil. Instead
they would show a continuously increasing pressure gradient (for example, from 0.2 to 0.3 psi/ft).

Fluid property correlations

Relatively accurate correlations are available for estimating the key fluid properties of reservoir
systems (Table 2). Standing [4] and McCain [5] give useful reviews of property correlations for oil
and gas, and other correlations are available. Note, however, that for specific producing provinces
(such as the Gulf Coast or the [[North Sea]]) more accurate correlations may exist.
Gas Oil and water

Pseudocritical properties Bubblepoint pressure

Z factor Solution GOR

Bubblepoint FVF

Density

Isothermal compressibility

Viscosity

K values

Interfacial tension

Diffusion coefficients

Equations of state (EOS) are now commonly used to calculate phase and volumetric behavior of
reservoir mixtures. In particular, EOS are useful for predicting phase behavior of miscible and
immiscible displacement processes resulting from the injection of gases such as carbon dioxide,
nitrogen, and lean or enriched natural gas in oil and gas condensate reservoirs. EOS do not usually
predict phase and volumetric behavior of reservoir mixtures accurately, thereby requiring
adjustment of component properties to match experimental PVT data[6].
Laboratory pressure-volume-temperature (PVT) experiments

Experimental PVT measurements are usually obtained for (1) large oil and gas fields, (2) volatile oil
and gas condensate reservoirs, and (3) reservoirs where gas injection is a potential EOR ([[enhanced
oil recovery]]) method. Two types of fluid samples can be taken during production, or when a well
is shut-in:

 Bottomhole samples, preferred for oils


 Separator samples, which must be recombined at the producing GOR during sampling
Recombined separator samples are standard for gas condensate fluids, but they may also be used for
oil reservoirs. Bottomhole sampling is preferred for oils if the reservoir is undersaturated (that is,
the initial pressure is higher than the bubblepoint pressure).
Standard PVT experiments include compositional gas chromatography (GC) analysis through
heptanes-plus (C7+), constant composition expansion, differential liberation expansion, constant
volume depletion, and multistage surface separation. Other PVT measurements include true boiling
point (TBP) distillation of the C7+ material and multicontact gas injection experiments. Table 3
summarizes these experiments, indicating when they are performed and on what types of reservoir
fluids.

Oil
Laboratory Analysis Condensates
s

Bottomhole sample * +

Recombined composition + *

C7+ TBP distillation + +

C7+ simulated distillation + +

Constant composition expansion * *

Multistage surface separation * +


Differential liberation * -

Constant volume depletion + *

Multicontact gas injection + +

Key: * standard, + can be done, – not done


Compositional analyses are used to describe the reservoir fluid makeup on a component basis,
including calculation of British thermal unit (BTU) (energy content) of gases, optimization of
separator conditions for liquid yield, and characterization of an EOS for compositional simulation.
Differential liberation and constant volume depletion experiments are designed to provide
quantitative information about the volumetric behavior of oil and gas condensate reservoirs during
pressure depletion. The multistage separator test is used together with differential liberation and
constant volume depletion data to calculate black oil properties Rs, Bo, Bg, and rs. Multicontact gas
injection experiments provide important volumetric and compositional data that can be used to
“tune” an equation of state (or alternative) model for simulation of gas injection processes.
Reservoir properties

This topic describes the key terminology for reservoir properties.

Drainage area

This is the areal extent of the reservoir. It is most often determined from geological
information about the formation / region, but it can also be estimated from performance studies
such as material balance, interference analysis, and simulation studies.

Expected ultimate recovery

The expected ultimate recovery (EUR) is the amount of oil or gas that can be economically
recovered from a reservoir under current operating and economic conditions.

Formation compressibility

Formation compressibility is the change in pore volume per unit of pore volume per unit
change in pressure. Even though the formation is a solid material, it is compressible. The
formation compressibility is used in the calculation of total compressibility. It is of the same
order of magnitude as the oil compressibility, or the water compressibility, approximately 10 -
6
 psi-1.

Thus, in the absence of gas saturation, the formation compressibility is relatively significant
and must not be ignored. However, when there is gas present in the pores, the gas
compressibility is often one or two orders of magnitude higher, and so the formation
compressibility is relatively small and may be ignored.

The formation compressibility may be measured in the laboratory, but it is usually derived
from correlations. It is obvious that the nature of the formation, its degree of consolidation, its
stress field, etc. all affect the formation compressibility, yet the correlations that exist do not
take all of these factors into account. They are very simplified and relate the formation
compressibility to the total porosity. The correlation used in Harmony Enterprise is derived
from laboratory measurements [Hall, 1953], and is applicable to sandstones and carbonates.
For overpressured reservoirs, the formation compressibility can be 10 times larger than
suggested by the Hall correlation. In practice, it has been found that formation compressibility
for overpressured reservoirs is approximately equal to 10-5 psi-1.

Gas saturation

Gas saturation is the fraction of the pore space occupied by gas. Most gas reservoirs also
contain some connate (non-movable) water. If the gas reservoir is part of a gas cap in a
saturated oil reservoir, the space may contain gas, oil, and water. Thus, the gas saturation is
rarely 100% but varies from 30% to 90%.

Gas saturation is used directly in the calculation of the reserves, and also in the calculation of
total compressibility.

Note that the gas saturation has a significant effect on the total compressibility because the gas
compressibility is much higher than the fluid compressibility of oil and water.

Initial pressure

Reservoir pressure is the pressure of fluids within the pores of a reservoir measured at a
specific point in time. The initial reservoir pressure is the average reservoir pressure measured
in a discovery well before any fluid is produced. This value is necessary for many reservoir
engineering calculations, such as reserve determination. Units are typically psi (field units) or
kilopascals (SI).

For well testing, the initial pressure is the average reservoir pressure before the start of a test.
For a new well, a well that has not been put on production, the initial pressure equals the virgin
or original pressure and corresponds to a drillstem test pressure. For a well that has been on
production for a long time, the initial pressure may or may not be equal to the original
reservoir pressure. The following example illustrates this concept.
 If Test 1 is being analyzed, the initial pressure is 2000.

 If Test 2 is being analyzed, the initial pressure is 1500 because the reservoir
pressure is fully built up before this test, and the historical pressure prior to
Test 2 does not apply. Thus, it behaves like a new well with an initial
pressure 1500.

 If Test 3 (which is a combination of Test 2 and some production prior to


Test 2) is being analyzed, the initial pressure is 2000, because it must reflect
the reservoir pressure before the production that is being analyzed during
the test.

Mid-point of perforations

Perforations are the communication tunnel created from the casing or liner into the reservoir
formation through which the oil or gas is produced. The mid-point of perforations is the
measured depth from the surface to the mid-point of the perforated interval of the formation.

Net pay

Net pay is the thickness of the formation that contributes to the flow of fluids, or the sum of the
productive intervals of a reservoir. It is determined from core or log analysis, and can be
different from the gross pay or the perforated interval. In the case of an inclined or deviated
wellbore in a declining formation, the net pay is measured perpendicular to the angle of
decline. Several examples of net pay are shown below.
Oil saturation

Oil saturation is the fraction of the pore space occupied by oil. Most oil reservoirs also contain
some connate (non-movable) water. Thus, the oil saturation is rarely 100% and usually ranges
from 10% to 90% (in the oil / water transition zones). The oil saturation affects the oil and gas
reserves, and is also used in the calculation of total compressibility.

Original fluid-in-place

Original fluid-in-place (OFIP) is the amount of fluid contained in a reservoir at standard


conditions, as calculated from the reservoir and fluid properties. It is a general term that refers
to the calculation of original gas-in-place (OGIP) and/or original oil-in-place (OOIP).
Porosity

The total porosity is the percentage volume occupied by the pore space, regardless of the type
of fluids contained in the pore space. It is obtained from core or log analysis. Total porosity is
used to calculate volumetric reserves and formation compressibility, and is used in the pressure
transient flow equations.

The total porosity has only a small effect during the transient flow period (it occurs within the
log term in the skin calculation equation for radial flow), but it has a significant effect during
pseudo-steady state (PSS) flow, as it directly affects the reserves contained in the reservoir.

Rock volume

The total volume of reservoir rock, often referred to as the bulk volume.

Pore volume

The total volume of pore space within the reservoir as determined by the porosity of the rock.

Hydrocarbon pore volume

The total pore volume containing hydrocarbons. This is generally the product of pore volume
and water saturation.

Recovery factor

The recovery factor (RF) is the ratio of the expected ultimate recovery and original gas/oil-in-
place. It indicates how much of the original gas/oil-in-place can be recovered under current
operating and economic conditions.

Reservoir temperature

The reservoir temperature represents the temperature of the formation.  It increases with
reservoir depth and differs widely depending on the reservoir location's geothermal gradient.
The geothermal gradient is the rate of increase in temperature per unit depth in the earth.
Although the gradient is variable depending on location, a typical average value is 15°F / 1000
ft.

In the analysis of gas wells, the reservoir temperature is used directly in the pressure transient
analysis calculations, as well as calculating all the gas fluid properties, such as the gas
formation volume factor, gas viscosity, and gas compressibility.
In the analysis of oil wells, the reservoir temperature does not enter into the pressure transient
analysis equations directly; rather it is used in calculating oil fluid properties such as the oil
formation volume factor, oil viscosity, oil compressibility, and solution gas-oil ratio.

Units are typically Rankine (field units) or Kelvin (SI).

Total compressibility

The compressibility of a substance is the change in volume per unit volume per unit change in
pressure. In a reservoir that consists of rock and pore space occupied by oil, water, and gas, the
total compressibility is defined as follows:

From this definition, the total compressibility takes into account the compressibility and
saturation of gas, oil, and water, as well as the (rock) formation compressibility. Note that the
gas saturation has a significant effect on the total compressibility because the gas
compressibility is much higher than the other fluid compressibility (oil and water). Thus, if
there is any gas present in the reservoir, the total compressibility is dominated by the gas
compressibility component.

In well testing and rate transient analysis, the analysis of pressure and production data is based
on the solution of the diffusivity equation. The analytical solution of this equation assumes that
the total compressibility is constant. This assumption is valid for single-phase liquid systems.
In gas systems, because compressibility of gas is highly dependent on pressure, the assumption
of a constant value of total compressibility can lead to serious errors, especially when the
reservoir has started to deplete (that is, during boundary-dominated flow). To account for this
changing gas compressibility, the concept of pseudo-time is invoked. It accounts for the
variation with pressure of compressibility, viscosity, and permeability.

The units of compressibility are 10-6 psi-1. (1 * 10-6 psi-1 (sometimes referred to as 1 Microsip.)

 co ≈ cw ≈ cf ≈ 10-6 psi-1

 c
g ≈ 10-3 psi-1 @ 500 psi

 c
g ≈ 10-3 psi-1 @ 1000 psi

 c
g ≈ 10-4 psi-1 @ 2500 psi

 c
g ≈ 10-4 psi-1 @ 5000 psi

 c
g ≈ 10-5 psi-1 @ 10,000 psi

 c
g ≈ 10-5 psi-1 @ 15,000 psi
Water saturation

Water saturation is the fraction of the pore space occupied by water. Most reservoirs are water
wet and contain connate water. Water saturation may range from 10% to 50% for an oil or gas
reservoir, and it is 100% for an aquifer. The water saturation obviously affects oil and gas
reserves, and it is also used in the calculation of total compressibility.

Wellbore radius

The wellbore radius is half of the measured inside diameter of the borehole at the point of
perforation. This is not to be confused with the effective wellbore radius, which accounts for
changing skin during production or after stimulation. The wellbore radius and effective
wellbore radius are equivalent when the skin is equal to zero.
Lecture 7. Anticlinal theory

WITHIN the last few months much interest has been excited by what is known as the it anticlinal

theory" of gas, a theory so simple, so plausible, and so easily understood by everyone, that its

adoption by the gas-producer was an almost foregone conclusion. The public has also accepted, but

not so generally, the assertion that the gas exists underground under any necessary pressure as a

liquid, thus accounting for the enormous quantities of gas obtained, apparently, from a small area.

Some, however, explain this by supposing that the manufacture of gas is constantly going on in the

formations underlying the gas-sand. It has also been asserted that the gas may exist (where the rock

is not porous) in crevices, and that these crevices or fissures will be found largely developed along

the anticlinal axes.

On the other hand, we have the opinion expressed by Professor Lesley that gas, water and oil

existing underground at great pressure will not separate, but that the gas will exist in solution,

"mixed" like carbonic acid gas in a soda-water fountain. This hypothesis, if correct, is fatal to the

anticlinal theory. - The anticlinal theory would also become less plausible if it were shown that the

gas exists as a liquid. The difference in specific gravity between it and water might not then seem

sufficient to force each through miles of pores until the hydrocarbon liquid had found its way to the

crest of the anticlinal, and the water to the lower portions of the rock, since these rocks, even in the

vicinity of the anticlinal axes, have a very gentle inclination, commonly less than one degree and
very rarely exceeding one degree and a half.t We may dismiss this consideration for the present, for

it will be seen further on that the gas cannot exist as liquid; but even if this were not so, time only

would be required to effect separation. The supposition of continuous manufacture in rocks

underlying the gas sands is entertained by few who are familiar with the phenomena presented by

oil- and gas-wells. The geologist is not able to deny the possibility of this state of affairs: but even if

such be the case, the limited life of many gas-wells as great producers is sufficient evidence that the

natural production underground can only be a fraction of the rate of consumption. However, this

supposition involves the question of the origin of gas and oil, whether in the rocks forming their

present reservoirs or in underlying bituminous or oleaginous shales or slates. The latter theory is

popular, but encounters one fatal objection which its adherents, with surprising complacency,

entirely ignore. Ascension from below implies the existence of open fissures,-these would permit

the escape of gas to the surface. As stated by Prof. Lesley: ••.. "This is the argument which weighs

most in my mind against one of the popular theories entertained by geologists, that all our

petroleum has ascended to the oil-rocks from far greater depths in the lower formations or in the

interior of the earth; for I confess myself unable to understand how the material could ascend 3000

feet or more from the Silurian or Lower Devonian rocks and there have been arrested, unable to

continue its ascent through the remaining one, two or three thousand feet to the surface. In addition

to this capital objection there is another, namely, that it has never been shown by anyone that

hydrocarbons in the gaseous form, ascending from great depths, could be condensed, or in any way

changed into liquid petroleum, and held in that form ever after by the porous sand-rocks." There is

no valid objection to the supposition that gas and oil have their origin in organic matter enclosed in

the porous sandstones in which they are now found. However, be the origin domestic or foreign to

the producing rock, experience has shown, in the oil-development at least, that the porous rock-

reservoir holds a definitely limited quantity, and that when this is once exhausted we may expect no

further accumulation; and the history of gas-wells in adjacent territory points to the same conclusion

as respects gas. Will gas and water existing underground at great pressure,-from 150 to 450 pounds
per square inch, for instance,-be found" mixed," or will they separate, the gas rising to the top? The

value of the anticlinal theory depends largely upon the solution of this question. As the propane

(C3H8) and ethane (C2H6) are present in small quantities and both probably exist as liquids, which

may, however, be held by the marsh-gas in suspension (just as water is evaporated and held in

suspension by air), they need not be here considered; and we may take the gas as a mixture of

marsh-gas, hydrogen and nitrogen, the former (marsh-gas CH4) in much larger percentage than the

latter combined. Marsh-gas is soluble in or absorbed by water at atmospheric pressure to only about

five per cent., * that is, it takes twenty volumes of water to absorb one volume of marsh-gas. The

solubility of marsh gas in salt water is not accurately known, but may be still less. Hence, if 20

(cubic feet of water absorb 1 cubic foot at atmospheric pressure, the same quantity of water would

by Henry's law (ibid., p. 236), absorb at 150 pounds or 10 atmospheres 10 cubic feet, or 50 per

cent.; and at 450 pounds or 30 atmospheres, 30 cubic feet or 150 per cent. In other words, one cubic

foot of water would absorb one and a half cubic feet of gas if compressed under 450 pounds

pressure,-any excess of gas present must separate and rise to the higher portions of the rock, the

water sinking to the lower portions. It is thus evident that the quantity of gas held by the water is

from A. practical stand-point small; and that a well piercing the rock where it i8 filled with salt

water, and producing a large flow of brine, could produce only one, two or three cubic feet (varying

with pressure and temperature) of gas for each cubic foot of water, if the amount of gas mixed with

the water were it"! only supply. Hence the argument, based on a supposed inseparable mixture of

gas, water and oil in depth, appears to me insufficient to overthrow the anticlinal theory; for it

seems certain that under all but exceptional conditions the great bulk of gas will naturally separate

from the water and rise through the interstices to the higher portions of the rock. t I deem it

unnecessary to discuss the possibility of a similar separation of oil and salt water, and of oil and

gas; but I may say in general that the separation of water and oil is probably not so easily

accomplished and will not occur so readily in territory where the dip is very slight, from the fact

that any porous rock is more permeable to gas than to a liquid.


One reason why the adoption of the anticlinal theory as a guide in developing gas-territory is not

likely to prove altogether satisfactory, is because its adherents are apparently losing sight of the fact

that gas-wells may also occur in sync1inals, or where there is no trace of either an anticlinal or

synclinal. We know that the porous portions of the productive oil- and gas rocks exist in belts and

pools isolated from each other by areas of rock too fine-grained and close to hold or yield quantities

commercially important. The quantity of gas that may exist in anyone of these patches or belts of

porous rock is necessarily limited by the quantity of water present. If the rock is entirely filled with

water we can expect no gas; if on the other hand the rock contains no water, we may expect gas

throughout the porous portion of the rock whether this be found in a synclinal or upon an anticlinal.

Between these two extremes we may have any and every variable quantity of water and gas. If the

water present is sufficient to fill only 25 per cent. of the voids in the rock, 75 per cent. of the area of

that particular porous patch wiII be productive gas territory, and if one end of this reaches an

anticlinal, and is punctured by wells located on the anticlinal theory, that portion equally good for

gas, extending down the slope of the anticlinal and reaching almost perhaps to the bottom of the

synclinal, will be passed over as worthless by the strict adherent to the anticlinal theory as at present

used. The" belt-line" theory of oil was found to be true within certain limits and with certain

qualifications; and any map of the productive oil-districts shows that the productive (porous)

streaks, belts and patches are more or less continuous along certain lines, which may be either

curved or straight. This theory proved useful, and would have given better results had not its

believers persistently adhered to it when and where it could be of no further use. This course was

followed by so many disheartening failures that some of the most enthusiastic" belt-line" operators

finally went to the other extreme and denounced it as entirely false. As we have good reason to

believe that the areas of porous gas rock will be found in patches, belts and streaks similar to the

distribution of porous oil-rock in Butler, Clarion and Venango counties, the belt-line theory should

appeal to the gas-producer as an aid in tracing out productive territory. We have, then, the anticlinal

theory and the belt-line theory, neither of which is complete or entirely sufficient in itself, but both
alike valuable and useful within certain limits; and by judiciously combining these theories, it

seems probable that, the prospect of being able to locate good territory in advance of developments

would be considerably increased. In other words, the natural-gas expert will probably double or

treble the chances of obtaining a good location for wens in both productive and undeveloped

territory if he relies, not upon either of these theories alone, but upon the two combined.

The geologist can trace out the anticlinals by surface observations, and locate them with sufficient

accuracy for all practical purposes; but in determining the trend of the porous streaks or belts he

must rely upon the records of wells already drilled, just as the oil-producer in the past has

determined his belt-lines" from data furnished by preceding operations. Thus by reasoning from the

known to the unknown, all of' the great Clarion-Butler oil belt was eventually developed. In the

same way the gas-producer must learn the trend or bearing of each productive (porous) belt or rock,

and trace it from point to point, the most promising locality being found where one of these belt

lines crosses an anticlinal axis. In the area known as the Pittsburgh gas-district (embracing all that

portion of Pennsylvania within some forty miles of Pittsburgh), the direction of these porous belts

has not yet been accurately defined. Some of the lines of' porous rock will doubtles.'3 be found to

vary considerably in bearing from others. The belt-lines of the Butler fields, for instance, will

doubtless have a trend not parallel to that of the Murraysville district. We must also not lose sight of

the fact that some of the productive areas may be merely circumscribed pools or patches; but even

in this case a knowledge of the trend of the porous patches will enable us the more readily to find

other such patches of productive rock. The oil-producer is already at work defining belt-lines in the

Washington field, and these lines may also be of use to the gas-producer; but no greater mistake

could be made than in attempting to apply them to gas-developments in other districts. Each district

must have its own lines of porous deposition worked out separately. To do this thoroughly will

require in each case some ti.me and labor, but this seems insignificant compared with the interests

at stake. In attempting to define the trend of the porous belt (belt-line) it will be necessary to

discriminate carefully between the different gas-producing rocks; for if in one locality the gas mmes
from one rock, and a well in the same or different locality is producing from another-a higher or

lower-gas-rock, it is evident that we have two entirely independent belts to define, which may have

very different trends. Thus, in the Butler oil district the" third sand" beltline was commonly taken at

about N. 22° E., or N. 22tO E., while the "fourth sand" belt-line curved from N. 45° E. at Modoc to

nearly due east and west near Brady's Bend; and the" second sand" belt at Martinsburg had a stilI

different bearing.

Anticline

Cross-sectional diagram of an anticline

Anticline exposed in road cut (small syncline visible at far right). Note the man standing in front

of the formation, for scale. New Jersey, U.S.A.

In structural geology, an anticline is a type of fold that is an arch-like shape and has its

oldest beds at its core, whereas a syncline is the inverse of a anticline. A typical anticline

is convex up in which the hinge or crest is the location where the curvature is greatest, and the

limbs are the sides of the fold that dip away from the hinge. Anticlines can be recognized and

differentiated from antiforms by a sequence of rock layers that become progressively older


toward the center of the fold. Therefore, if age relationships between various rock strata are

unknown, the term antiform should be used.

The progressing age of the rock strata towards the core and uplifted center, are the trademark

indications for evidence of anticlines on a geologic map. These formations occur because

anticlinal ridges typically develop above thrust faults during crustal deformations. The uplifted

core of the fold causes compression of strata that preferentially erodes to a deeper stratigraphic

level relative to the topographically lower flanks. Motion along the fault including both

shortening and extension of tectonic plates, usually also deforms strata near the fault. This can

result in an asymmetrical or overturned fold.[1]

Terminology of different folds [edit]

Diagram showing constituent parts of an anticline

Antiform vs anticline[edit]

An antiform can be used to describe any fold that is convex up. It is the relative ages of the rock

strata that distinguish anticlines from antiforms.[1]

Elements[edit]

The hinge of an anticline refers to the location where the curvature is greatest, also called

the crest.[1] The hinge is also the highest point on a stratum along the top of the fold.

The culmination also refers to the highest point along any geologic structure. The limbs are the
sides of the fold that display less curvature. The inflection point is the area on the limbs where

the curvature changes direction.[2]

The axial surface is an imaginary plane connecting the hinge of each layer of rock stratum

through the cross section of an anticline. If the axial surface is vertical and the angles on each

side of the fold are equivalent, then the anticline is symmetrical. If the axial plane is tilted or

offset, then the anticline is asymmetrical. An anticline that is cylindrical has a well-defined axial

surface, whereas non-cylindrical anticlines are too complex to have a single axial plane.

Types[edit]

An overturned anticline is an asymmetrical anticline with a limb that has been tilted

beyond perpendicular, so that the beds in that limb have basically flipped over and may dip in

the same direction on both sides of the axial plane.[3] If the angle between the limbs is large

(70–120 degrees), then the fold is an "open" fold, but if the angle between the limbs is small (30

degrees or less), then the fold is a "tight" fold.[4] If an anticline plunges (i.e., the anticline crest

is inclined to the Earth's surface), it will form Vs on a geologic map view that point in the

direction of plunge. A plunging anticline has a hinge that is not parallel to the earth's surface.

All anticlines and synclines have some degree of plunge. Periclinal folds are a type of anticlines

that have a well-defined, but curved hinge line and are doubly plunging and thus

elongate domes.[5]

Model of anticline. Oldest beds are in the center and youngest on the outside. The axial plane

intersects the center angle of bend. The hinge line follows the line of greatest bend, where the

axial plane intersects the outside of the fold.


Folds in which the limbs dip toward the hinge and display a more U-like shape are

called synclines. They usually flank the sides of anticlines and display opposite characteristics.

A syncline's oldest rock strata are in its outer limbs; the rocks become progressively younger

toward its hinge. A monocline is a bend in the strata resulting in a local steepening in only one

direction of dip.[2] Monoclines have the shape of a carpet draped over a stairstep.[4]

An anticline that has been more deeply eroded in the center is called a breached or scalped

anticline. Breached anticlines can become incised by stream erosion, forming an anticlinal

valley.

A structure that plunges in all directions to form a circular or elongate structure is a dome.

Domes may be created via diapirism from underlying magmatic intrusions or upwardly mobile,

mechanically ductile material such as rock salt (salt dome) and shale (shale diapir) that cause

deformations and uplift in the surface rock. The Richat Structure of the Sahara is considered a

dome that has been laid bare by erosion.

An anticline which plunges at both ends is termed a doubly plunging anticline, and may be

formed from multiple deformations, or superposition of two sets of folds. It may also be related

to the geometry of the underlying detachment fault and the varying amount of displacement

along the surface of that detachment fault.

An anticlinorium is a large anticline in which a series of minor anticlinal folds are

superimposed. Examples include the Late Jurassic to Early Cretaceous Purcell Anticlinorium in

British Columbia[1] and the Blue Ridge anticlinorium of northern Virginia and Maryland in the

Appalachians,[6] or the Nittany Valley in central Pennsylvania.

Formation processes[edit]
Anticline near Ehden, Lebanon

Anticlines are usually developed above thrust faults, so any small compression and motion

within the inner crust can have large effects on the upper rock stratum. Stresses developed

during mountain building or during other tectonic processes can similarly warp or bend bedding

and foliation (or other planar features). The more the underlying fault is tectonically uplifted,

the more the strata will be deformed and must adapt to new shapes. The shape formed will also

be very dependent on the properties and cohesion of the different types of rock within each

layer.

During the formation of flexural-slip folds, the different rock layers form parallel-slip folds to

accommodate for buckling. A good way to visualize how the multiple layers are manipulated, is

to bend a deck of cards and to imagine each card as a layer of rock stratum.[7] The amount of

slip on each side of the anticline increases from the hinge to the inflection point.[2]

Passive-flow folds form when the rock is so soft that it behaves like weak plastic and slowly

flows. In this process different parts of the rock body move at different rates causing shear stress

to gradually shift from layer to layer. There is no mechanical contrast between layers in this

type of fold. Passive-flow folds are extremely dependent on the rock composition of the stratum

and can typically occur in areas with high temperatures.[4]

Economic significance[edit]
Anticlines, structural domes, fault zones and stratigraphic traps are very favorable locations

for oil and natural gas drilling. About 80 percent of the world's petroleum has been found in

anticlinal traps.[8] The low density of petroleum causes oil to buoyantly migrate out of its

source rock and upward toward the surface until it is trapped and stored in reservoir rock such

as sandstone or porous limestone. The oil becomes trapped along with water and natural gas by

a caprock that is made up of impermeable barrier such as an impermeable stratum or fault zone.

[9] Examples of low-permeability seals that contain the hydrocarbons, oil and gas, in the ground

include shale, limestone, sandstone, and rock salt. The actual type of stratum does not matter as

long as it has low permeability.

Water, minerals and specific rock strata such as limestone found inside anticlines are also

extracted and commercialized. Lastly, ancient fossils are often found in anticlines and are used

for paleontological research or harvested into products to be sold.

1.1 Basic Trap-Fluid Nomenclature: Rocks above (cap rock) and alongside (wall rock) of a trap

are commonly impermeable not only for oil or gas but also for water under reservoir pressure

condition. Hence, water which originally present in a reservoir get displaced downward with

accumulation of oil and gas. This pool of displaced water is referred as 'bottom water'. The pool

of water that is bounded laterally is called edge water (Fig. 1). The structurally lowest point in a

trap that can retain hydrocarbon is referred to as spill point.

Fig. 1 Anticlinal structural trap with oil-water, oil-gas contact and bottom, edge water shown.
1. Classification of Traps

Traps are usually classified according to the mechanism that produces the
hydrocarbon accumulation. The two main groups of traps are those that are formed
by structural deformation of rocks (structural traps), and those that are related to
depositional or diagenetic features in the sedimentary sequence (stratigraphic
traps).

Many traps result from both of these factors (combination traps). A common
example is stratigraphic pinch-out (e.g., a sandstone lens wedging into mudstone)
that is combined with tectonic tilting (which allows hydrocarbons to pond in the
updip part of the sandstone wedge). Other traps result mainly from fracturing
(which creates the reservoir porosity) or hydrodynamic processes. There are many
classifications of hydrocarbon traps in use, but most have ~90% in common.

1.1 Structural Traps: Structural traps are primarily the result of


folding and (or) faulting, or both.

Fig. 2 Examples of structural trap: Reverse fault, Normal fault and faulted limb of
an anticline.

1.1.1 Anticlinal (fold) and dome traps: Necessary conditions: An


impervious cap rock and a porous reservoir rock; closure occurs in all directions to
prevent leakage (i.e. four-way closure necessary for a dome).

Simple fold traps (anticlinal) with axial culmination (fold axis dipping in two or more
directions). The simplest type of trap is formed when a sandstone bed that is
overlain by tight (i.e. low permeability) shale is folded into an anticline. A simple
anticline, however, may not necessarily be a trap. The crest of the anticline must
have an apical culmination (i.e. a peak) somewhere along the fold axis so that
hydrocarbons can be trapped. Closure is defined as the height to the crest of a
structure above the spill point of a trap. Anticlinal traps are commonly detected by
seismic reflection. In mature oilfields, most of these simple traps have probably
been found, but many anticlinal traps remain to be discovered offshore and in new
prospective areas.

 Salt domes: Strata around the salt dome curve upward creating traps against the
sealing salt layers (see below for more details).

 Growth domes: Domes or anticlines form during sedimentation when one area
subsides more slowly than the surrounding areas. Their formation is concurrent
with sedimentation (i.e. they form during deposition), and not due to later
(tectonic) folding. Growth anticlines may form due to differential compaction over
salt domes or other upward-projecting features in the substrate (topographic highs
on the buried landscape).

1.1.2 Fault traps: The fault plane must have a sealing effect so that it
functions as a fluid migration barrier for reservoir rocks. There are several common
types of fault trap:

 Normal faults- commonly associated with graben (rift) structures. b) Strike-slip


faults- these may not be sealed due to incremental movements, but basement-
controlled strike-slip faults commonly produce good anticlinal structures in
overlying softer sediments.

 Thrust faults- commonly associated with compressional tectonics (e.g., the Front
Ranges in Alberta).
Growth faults- Growth faults typically form in sediments that are deposited rapidly,
especially at deltas. Faulting occurs during sedimentation (i.e. syndepositionally),
such that the equivalent strata on the downthrow side will be thicker than on the
upthrow side. The throw between corresponding strata declines upwards along the
fault plane. Minor fault planes with an opposite throw (antithetic faults) may also
form in the strata that curve inward towards the main fault plane. The fault plane is
commonly sealed, preventing further upward migration of oil and gas. Fault traps
may also form when sandstone beds are offset against the fault plane. Some
petroleum traps, however, form in "roll-over" anticlines on the down-faulted block.
Growth faults may reduce porewater circulation in sedimentary basins;
consequently, undercompacted clays, which may develop into clay diapirs, are
often associated with growth faults.

Some general points about fault traps…

The geometry and timing determine whether faults will be effective in forming fault
traps:

 Dead faults that predate basinal sediments only affect the underlying basement –
they play no direct role in hydrocarbon trapping in the younger sedimentary pile.
 Continuously developing faults (growth faults) — these are active during
sedimentation and are major petroleum traps (e.g., Niger Delta).
 Young (late) faults —these form late during sedimentation; depending on their
initiation and growth, they may or may not be effective as traps. 4. Late
regenerated faults —these are new movements on old faults — they are more likely
to destroy than form traps, but may be effective.

Many petroleum fields are closely linked to faulting, but traps that result from
faulting alone are less common.

There are three common fault – petroleum pool associations:

 The fault itself makes the trap without an ancillary trapping mechanism such as a
fold — normal faults are the most common examples.
 The fault creates another structure (e.g., a fold or horst) that in turn forms the main
trap.
 The fault may be a consequence of another structure that forms the main trap- e.g.,
the extensional crestal faults that form above some anticlines.
Important point: Faults are highly ambiguous features. They may leak, acting as
permeable conduits for fluid flow (including oil and gas migration), but more
commonly act as seals unless they are rejuvenated after petroleum has pooled.
They form seals because of fault gouges or differential pressures either side of the
fault plane (Fig. 3):
Fig. 3 Model of a fault zone with formation of fault gouze. A fault may create lateral
seal for a trap.

SALT DOME RESERVOIRS

Salt domes form when salt is less dense than the overlying rock, and the salt moves
slowly upwards due to its buoyancy. For this to happen, there must be a minimum
overburden and the thickness of the salt deposits must be more than ~100 m.
Upward movement of salt through the sedimentary strata, and associated
deformation is termed halokinetics or salt tectonics. Movements may continue for
several hundred million years.

Traps may form (1) in the strata overlying the salt dome, (2) in the top of the salt
domes (the cap rock - caused by brecciation and dissolution), (3) in the strata that
curve upward against the salt intrusion, (4) due to stratigraphic pinch-out of strata
around the salt dome (Fig. 4). Salt dome reservoirs produce major oilfields where
basinal sediments contain thick salt deposits. Salt deposits are common in Permian-
Jurassic sediments around the Atlantic Ocean. Examples include the Gulf of
Mexico, where there is Permian and Jurassic salt, the Permian Zechstein salt in NW
Europe and the North Sea.
Fig. 4 Ideal salt-dome traps based on examples from U.S. Gulf Coast. Note
development of trap at the top, curved strata alongside and because of pinch out.
1.2 Stratigraphic Traps: Stratigraphic traps are created by any variation
in the stratigraphy that is independent of structural deformation, although many
stratigraphic traps involve a tectonic component such as tilting of strata.

Two main groups can be recognized —

Primary stratigraphic traps result from variations in facies that developed during
sedimentation. These include features such as lenses, pinch-outs, and appropriate
facies changes.

Examples include (Fig. 5):


 Primary pinch out of strata, e.g., strata that pinch out updip in less permeable rocks
such as shale;
 Fluvial channels of sandstone that are isolated and surrounded by impermeable
clay-rich sediments;
 Submarine channels and sandstone turbidites in strata rich in shale;
 Porous reefs that are surrounded by shale, etc.

Fig. 5 Examples of stratigraphic traps: Pinch out, Unconformity, platform reefal


carbonate.
Secondary stratigraphic traps result from variations that developed after
sedimentation, mainly because of diagenesis. These include variations due to
porosity enhancement by dissolution or loss by cementation. Paleogeomorphic traps
are controlled by buried landscape. Some are associated with prominences (hills);
others with depressions (valleys). Many are also partly controlled by
unconformities so are also termed unconformity traps.

Combination Traps: Many oil fields around the world are not solely the result of
structure or stratigraphy. They are result of combined effects of structural and
stratigraphical processes. Pinchout, onlap, truncation traps all require some closure,
which is often caused by structural deformation along the strike. Similarly,
folded/faulted beds may be sealed by unconformity planes to form trap (Fig. 6).
Fig. 6 Trap formation by truncation of folded strata by unconformity plane. Traps
can form both below or above the unconformity plane.
1.3 Other Types of Traps

1.3.1 Fractured Reservoirs: Fractured basement rock that projects upward


locally into overlying shales can also provide good traps. Fractured, well-cemented
rocks (limestones and cherts) along faults may also form good traps of limited
lateral extent (Fig. 7).

Fig. 7 Trap formation in the basement because of fracture and suitable source rock
above in form of shale. Map view on the right.

1.3.2 Hydrodynamic Traps: If pore water flow in a sedimentary basin is


strong enough, the oil-water contact may deviate from the horizontal because of the
hydrodynamic shear stress that is set up (Fig. 8). In some cases, oil may
accumulate without closure. Flow of fresh (meteoric) water down through oil-
bearing rocks commonly results in biodegradation of the oil and formation of
asphalt, which may then form a cap rock for oil.

Fig. 8 Hydrodynamic Trap. Oil-water contact deviated from horizontal because of

water movement.
Summary
Hydrocarbon traps are formed when impermeable/ less permeable strata caps porous
and permeable reservoir strata. This typical geological condition allows entry of
hydrocarbon within the reservoir but prevents its further migration upward.
Displaced water from a reservoir with emplacement of oil/gas forms ‘Bottom
water’ and ‘Edge water’ in a trap. The lowest structural point up to which a trap
can retain hydrocarbon is referred to as spill point for a trap. Broadly, Traps are
classified under stratigraphic, structural and combination categories. Both folding
and faulting may result in formation of structural trap. Uprise of low-density salt
through sedimentary strata and deformations resulted thereby may give way to the
formation of varied structural traps. Stratigraphic traps are independent of
structural deformation and can be both primary and secondary in nature. Primary
stratigraphic traps are result of facies variation in course of deposition whereas
secondary stratigraphic traps are those, which are formed in course of diagenetic
history. Some other trapping mechanisms include Fracture trap and Hydrodynamic
trap.

Lecture 8 .Introduction of rotary drilling

Drilling for oil is a costly process, especially offshore. The object of a well is to prove the
presence of, or produce, oil or gas. Sometimes wells are also drilled to inject water, chemicals or
steam into the reservoir during production. Even a well which fails to find hydrocarbons (a dry
well) is still of great value, because of the information it provides about the rocks in the area. This
information forms part of the basis for the geological maps and profiles which are used in further
exploration for oil and can be sold or exchanged for data from other companies. This is the reason
why oil companies wish to keep the geological results of oil drilling confidential for some years
after a well has been completed. The first well in a new area is called a wildcat well, while
appraisal wells are drilled to estimate the extension of an oil field. They may also become
production wells. Stratigraphic wells are drilled mainly to obtain stratigraphical information from
the basin. Oil drilling used to be carried out largely on land, but now offshore drilling takes place
not only on our continental shelves, but also in deep water (1–3 km). This type of drilling is many
times more costly than drilling in shallow water or on land. This has led to increased efforts to gain
maximum information from wells. The cost of analysing samples and logs is small in relation to
the cost of drilling the well. We shall not go very deeply into the technical aspects of drilling for
oil here, but merely look briefly at some of the most important principles. When drilling
commences at the surface, the diameter of the well may be 20"–30" (50–75 cm), but decreases
downwards to 3"–6" (7–15 cm) at great depths. Normally a roller bit is used, which crushes the
rock into small pieces (about 2–5 mm) called cuttings. Core samples are only taken when drilling
through especially important rock strata (usually reservoir rocks) where large intact samples are
needed for detailed examination. A circular diamond core drill bit must then be used. This takes
time and costs a lot more per running metre, as the entire drill string has to be recovered to get
each core section to the surface. Only the most critical sections are therefore cored. Drilling mud is
pumped down through the drill string into the well during drilling. This mud has several functions.
When one drills several hundred or a 1,000 m down into rock, one encounters water, gas or oil
which may be under high pressure. The drilling mud acts as a counterweight to prevent the
uncontrolled gush of water or petroleum into the well and up to the surface in a blow-out. The
pressure exerted by the drilling mud must exceed the pressure of oil and water in the surrounding
formation. Heavy minerals such as barytes are frequently added to increase density; the main
components of drilling mud are montmorillonite (smectite) containing clays, with a large number
of different additives. The drilling mud also serves to cool the drill bit, and cuttings are brought
back to the surface suspended in the circulating mud. The cuttings are then washed out from the
drilling mud onto a sieve (shale shaker) and the mud can be used again. The cuttings are
continuously analysed on the drilling platform by a geologist who logs the composition of the
cuttings, making a preliminary description of the rocks which are being drilled, and their
mineralogy. Samples of cuttings are usually taken every 10–30 ft drilled. More detailed analyses
are carried out in the laboratory. The cuttings are often poorly washed and need extra cleaning to
get rid of the drilling mud; in weakly indurated mudstones it may be difficult to separate the
drilling mud from the soft cuttings. Organic additions to drilling mud may cause problems when
analysing cuttings, and sometimes oil-based rather than water-based drilling mud is used, which
confuses analyses for oil. The fossil content of the rock fragments, largely microfossils, is used to
determine the age of the strata (see Biostratigraphy). Not all the cuttings which come up with the
drilling mud have necessarily come from precisely the strata being drilled through at that time.
There may also be contamination due to the caving in of overlying strata into the rising drilling
mud. This means we can find material from younger rocks with a different composition mixed in
with the formation being drilled, together with younger fossils. This demands considerable care
when making stratigraphic interpretations based on microfossils identified in cuttings. The safest
way is to register the first occurrences of a species when proceeding downwards from the top in
the well. The last occurrence of a fossil may be the result of cave-ins from younger strata. Since
the pressure of the drilling mud is being monitored and adjusted to prevent oil and gas from
penetrating into the well, significant oil and gas occurrences may be drilled through without being
registered. This should be detected on the logs but it may be advisable to carry out special tests in
the most promising strata to find out if there is petroleum present, and in what quantities. As
drilling proceeds, the well is lined with steel casing to prevent rock and loose sediment falling into
it, but prior to casing, each section of the well has to be logged with different logging tools which
require physical contact with the wall of the well. Radioactive logs, however, can also be run after
the casing has been installed. It is useful to note when the different casings are installed, because
that limits the strata which could have “caved in” and contaminated the cuttings. If the well is
going to produce oil, a production pipe is used and installed running through the petroleumbearing
strata. It is then perforated by shooting holes in the steel casing (in the oil column) so that
petroleum can flow into the well.

For an oil field to be capable of full production, several wells are normally required. Up to
1990–1995 most of the exploration wells and production wells were nearly vertical. Horizontal
barriers due to changes in facies or faults could then be critical barriers during production. It was
difficult to produce oil from thin sandstones or carbonate beds because of vertical flow of water
from below or gas from above. Onshore this could be compensated for by having a dense well
spacing but offshore that would be too expensive. Horizontal drilling has revolutionised oil
production. It is now possible to follow thin oil columns laterally and to make complex wells to
drain different compartments in the reservoir. Large oil fields with thin oil columns like the Troll
field in the North Sea would have been difficult to produce without horizontal drilling. Horizontal
and deviated wells may extend 8–9 km from the drilling platform, enabling production from
relatively small reservoir compartments away from the platform. Earlier, geophysics was used
mainly to define structures but the quality of the seismic data was not good enough to provide much
detail. 3D seismic based on 50 m line spacing allows the construction of a threedimensional cube of
geophysicial data, which provides much more detailed information. By repeating seismic surveys
during production at 2–5 year intervals, the effect of changes in the fluid composition on the
seismic data can be seen, thus adding the time dimension to 3D seismics. 4D seismics has made it
possible to follow the depletion of an oil or gas field by monitoring the GWC and often also the
OWC during production. By this means, parts of a reservoir that have not been drained by the
production wells can be detected. If the isolated reservoir compartments are large it may be
economic to drill additional wells to drain them.
Rotary drilling
Rotary drilling is used to form a deep observation borehole or for obtaining representative samples
of rock. The drilling method involves a powered rotary cutting head on the end of a shaft, driven
into the ground as it rotates. The system requires lubrication (air, water or drilling mud) to keep the
cutting head cool.

There are two types of rotary boring, open-hole and core drilling. Material recovered from open-
hole drilling is mixed with the drilling lubricant. It is unsuitable for effective sampling, and it is
often difficult to observe and record the strata. Open-hole boring is only suitable for rapid drilling to
enable core recovery at a greater depth or for the installation of monitoring wells. Core drilling is
carried out using wire-line, double or triple-tube core barrels with diamond or tungsten-tipped core
bits.

Wire-line core barrels are rotated from the surface by rods normally of the same diameter as the
outer core barrel. The core is brought to the surface within the inner barrel using a wire rope or
attached line to a recovery tool. This system is particularly suitable for superficial or weak deposits,
as any vibration from the drilling action is minimised due to close-fitting rods used within the hole.

The conventional double-tube core consists of two barrels; the outer barrel is rotated by the drill
rods and carries the coring bit. The inner barrel does not rotate, and the core passes up into this
inner barrel, enabling the sample to be recovered and brought to the surface.

With triple-core barrels, the non-rotating inner barrel contains a removable tube or liner. At the end
of each core run, this liner with the core it contains is extracted and stored in a core box. This
method does not increase core recovery but is more likely to preserve the core in original condition.

Lecture 9. Petroleum system elements


The Petroleum System consists of a mature source rock, migration pathway,reservoir rock, trap and
seal. Appropriate relative timing of formation of these elements and the processes of
generation, migration and accumulation are necessary for hydrocarbons to accumulate and be
preserved.
The components and critical timing relationships of a petroleum system can be displayed in a chart
that shows geologic time along the horizontal axis and the petroleum system elements along the
vertical axis.

Exploration plays and prospects are typically developed in basins or regions in which a complete
petroleum system has some likelihood of existing.

Source Rock Hydrocarbon Generation

The formation of hydrocarbon liquids from an organic rich source rock with kerogen and bitumen
to accumulates as oil or gas.
Generation depends on three main factors:
 the presence of organic matter rich enough to yield hydrocarbons,
  adequate temperature,
 and sufficient time to bring the source rock to maturity.
 Pressure and the presence of bacteria and catalysts also affect generation.
 Generation is a critical phase in the development of a petroleum system.
Migration

The movement of hydrocarbons from their source into reservoir rocks.


 The movement of newly generated hydrocarbons out of their source rock is primary
migration, also called expulsion.
 The further movement of the hydrocarbons into reservoir rock in a hydrocarbon trap or other
area of accumulation is secondary migration.
 Migration typically occurs from a structurally low area to a higher area in the subsurface
because of the relative buoyancy of hydrocarbons in comparison to the surrounding rock.
 Migration can be local or can occur along distances of hundreds of kilometres in large
sedimentary basins, and is
 critical to the formation of a viable petroleum system.
Accumulation

The phase in the development of a petroleum system during which hydrocarbons migrate into and
remain trapped in a reservoir.

Reservoir

A subsurface body of rock having sufficient porosity and permeability to store and transmit fluids.

 Sedimentary rocks are the most common reservoir rocks because they have more porosity
than most igneous and metamorphic rocks and
 they form under temperature conditions at which hydrocarbons can be preserved.
 A reservoir is a critical component of a complete petroleum system.
Seal (cap rock)

An impermeable rock that acts as a barrier to further migration of hydrocarbon liquids.


Rocks that forms a barrier or cap above and around reservoir rock  forming a trap such that fluids
cannot migrate beyond the reservoir. The permeability of a seal capable of retaining fluids through
geologic time is   ~  10-6 to 10-8 darcies.  commonly
 shale, mudstone
 anhydrite 
 salt, 
 A seal is a critical component of a complete petroleum system.
Trap

A configuration of rocks suitable for containing hydrocarbons and sealed by a relatively


impermeable formation through which hydrocarbons will not migrate.
Traps are described as
 structural traps
 Hydrocarbon traps that form in geologic structures such as folds and faults
 stratigraphic traps
 Hydrocarbon traps that result from changes in rock type or pinch-outs,
unconformities, or other sedimentary features such as reefs or buildups
 A trap is an essential component of a petroleum system.

Petroleum System Analysis

The establishment of a petroleum system is essential to form oil and gas fields. The petroleum system is

a unifying concept that encompasses all of the disparate geological or geochemical elements and processes of

petroleum geology. Oil and gas fields will not be formed if any of the elements or processes of the petroleum

system are missing.


Elements and processes

The essential elements and processes of the petroleum system include the following:

 Source: rock that is capable of generating or that has generated movable quantities of hydrocarbons.
 Reservoir: rock that has the ability to store hydrocarbons inside its pores.
 Seal: rock that impedes the escape of hydrocarbons from the reservoir rock.
 Trap: A trap consists of a geometric arrangement of reservoir and seal rocks that allow hydrocarbons to
accumulate.
 Migration: After generation and expulsion from a thermally-mature source rock, hydrocarbons must migrate
through carrier beds, faults, and fractures into the reservoir rock. If an adequate trap and seal are present,
hydrocarbon migration will cease and the reservoir will be filled.

Elements and processes of a petroleum system


Improving the discovery rate of oil and gas fields

 These essential elements and processes must be correctly distributed in time and space. To improve

the discovery rate of oil and gas fields, comprehensive evaluation of each element and process of the

petroleum system is important. JOGMEC is conducting basin modeling which is a computer simulation

technique that models generation, migration, and accumulation of oil and gas in a sedimentary basin.

Results of 3D basin modeling in a Middle-East marine oil field.

Lecture 10. Form of natural reservoirs

Oil and gas reservoir


An oil and gas reservoir is a formation of rock in which oil and natural gas has accumulated. The
oil and gas collected in small, connected pore spaces of rock and are trapped within the reservoir by
adjacent and overlying, impermeable layers of rock.[1] Typical reservoirs are not "pools" or "lakes"
of oil beneath the surface, as there are no vast open cavities that contain oil. Oil and gas reservoirs
can also be referred to as "hydrocarbon reservoirs".
Figure 1. An anticline oil and gas reservoir.[2]

Conventional hydrocarbon reservoirs consist of three main parts: the source rock, the reservoir rock,
and the cap rock (Figure 1). The source rock is the rock that contains the kerogen that the oil and
gas forms from. The reservoir rock is the porous, permeable rock layer or layers that hold the oil
and gas. The cap rock seals the top and sides so that the hydrocarbons are trapped in the reservoir,
while water often seals the bottom.[3]
For a reservoir to exist, oil and gas from the source rock must migrate into the reservoir rock, which
takes millions of years. This migration occurs because oil and gas are less dense than water. This
difference in density causes the oil and gas to rise towards the surface so that they are above
groundwater with the gas settling above the oil because of its lighter densities. Migration pathways -
a set of well connected fractures - must exist for this rising to occur.[4]
Reservoir rocks need to be both porous and permeable. This means that there are small pockets of
space within the rock where oil or gas can settle and small channels connecting these pockets to
allow the oil or gas to flow out of this rock easily when it is drilled. [4] These spaces between grains
can develop as the formation of rock occurs or afterwards, usually as a result
of groundwater passing through the rock and dissolving some of the cement between sediment
grains.
The rock formation must be formed or deformed in such a way to create a trap for the oil and
gas. Anticlines are the most common formation shape for this to occur. Anticlines for a rough "A"
shape, with the cap rock making the sides of the "A". The fossil fuels accumulate in the peak of the
"A" and the bottom is sealed (usually with water), preventing the oil and gas from seeping out and
escaping. [4]
The Purpose of Reservoir Engineering

It is the core responsibility of the reservoir engineering discipline to learn as much as possible about
reservoir fluids and their flow capacities through the rock matrix, in which they are trapped. Reservoir
engineers conduct studies, tests, sampling runs, modeling exercises, and other activities in an effort to
provide a clear understanding of the properties of the reservoir fluids. These properties change over
time, and reservoir engineers must understand to what degree they change. This includes as the fluids
move from within the reservoir, into the bottom of a producing well, into the tubing string and up the
well, through the tree and choke, into gathering lines, and finally to the entrance of the first-stage
separator. In addition, the reservoir’s unique and operative drive mechanism is changing as reservoir
pressure decreases over time, at constant reservoir temperature. Reservoir engineers spend
considerable time planning and gathering data early in the producing life of a hydrocarbon
accumulation, with the express purpose of defining exactly what is occurring in the reservoir. Having
such information in hand makes the medium- and longer-term planning of a well completion’s life more
readily manageable, technically accurate, and doable. Especially when later decisions concerning
workover and artificial lift planning are made.
Reservoir engineering has been defined as “the art of developing and producing oil and gas fluids in
such a manner to maximize reservoir economics.” The definition has a broad and varied meaning,
but the goal of ultimate hydrocarbon recovery and maximum revenue generation are closely linked.
Before establishing a plan to produce a reservoir, the development engineers must understand key
parameters about the subject reservoir:

Only with a good understanding of these variables can the engineering team plan and execute an
efficient job of drilling, completing, and producing wells, to efficiently recover the maximum
volume of hydrocarbons from a reservoir.
Petroleum Formation and Occurrence

There are three requirements which must be met before a petroleum reservoir can exist. These are:
 

Source Rock: The primary theory accepted for the generation of hydrocarbon is that it is formed by
organic evolution. This is a result of the decomposition of vegetable and animal organisms that
lived during previous geologic ages. The accepted scenario is that organic material was deposited at
the bottom of seas and lakes, and was later subject to the deposition of fine sediments over a period
of time. At depths where anaerobic bacteria can prosper and thrive in the absence of oxygen,
deposits are rich in carbon and hydrogen. Many oil and gas fields in various parts of the world are
found at the depths of current the continental shelves of ancient seas and lake beds.

Fig. 1 –  Marine shale,  an example of a shale source rock.

Reservoir Rock: The term reservoir implies storage. Reservoir rock is therefore "that rock in which
the hydrocarbon can be stored and from which it can be produced." This reservoir rock may or may
not be the rock in which the hydrocarbon was generated (its source rock). For a rock to be a
potential reservoir rock, two properties, porosity and permeability, must exist with sufficient
magnitudes to justify economic development of the hydrocarbon reservoir. These beds are primarily
sand, sandstone, limestone, and dolomite. The fluids present are found in the intergranular spaces
between grains called pores.

Fig 2.  – Example of a reservoir rock including petroleum migration and entrapment.

Hydrocarbon Trap: Traps must also exist to trap the hydrocarbon in place forming a hydrocarbon
reservoir. The fluids of the subsurface migrate according to density. As previously discussed, the
dominant fluids present or potentially present are hydrocarbon gas, hydrocarbon liquid, and salt
water. Since the hydrocarbons are less dense than the salt water, they will tend to migrate upward to
the surface, displacing the heavier water down elevation. These fluids will continue to migrate until
they encounter impermeable rock, which will serve as a reservoir seal or trap. If impermeable rock
does not exist, the hydrocarbons will migrate to the surface and be dissipated into the environment
in geologic time. See Figure 2 for an example of a trap. 
A reservoir is that portion of a trap that contains oil and/or gas as a single hydraulically connected
system. And, many hydrocarbon reservoirs are hydraulically connected to various volumes of
water-bearing rock, called aquifers. A typical reservoir rock is porous and permeable, which means
that the pores interconnect so that fluids can migrate through the rock. When “reservoir fluids” is
used, the meaning of the term “fluids” is defined as reservoir oil, gas, condensate, and water.
Distribution of Hydrocarbon Fluids in Traps

Porous rocks contain oil, gas, and water within the connected permeable pore space. Gravitational
and capillary forces largely control the distribution of these fluids in petroleum accumulations. As
oil and gas are lighter than water, the upward movement of petroleum must be restricted in order for
hydrocarbon to accumulate at depth. That is, a natural barrier or trap must exist for petroleum
accumulation to form. The reservoir rock must be overlain by impermeable beds so that oil and gas
accumulation cannot seep from the trap and migrate to higher beds in the stratigraphic sequence.
Gravitational forces cause the lighter fluids to rise to higher positions in the trap. Capillary forces,
on the other hand, cause a wetting fluid to rise into pore space containing a non-wetting fluid. Water
is a wetting fluid with respect to oil and gas, and oil is a wetting fluid with respect to gas. Even
though capillary forces counteract the force of gravity in segregating the fluids, an equilibrium
exists between these two forces before any disturbance occurs, such as a drill bit penetrating a trap.
A typical fluid distribution in a trap is illustrated in Figure 3.
 
Fig. 3 - A reservoir trap.

The following are the three possible modes for gas occurrence in reservoirs:
1. In some reservoirs, only oil and water are present at the reservoir’s pressure and temperature
conditions. Between a reservoir’s oil and water zones, there is a section of the vertical column
referred to as an oil-water transitional zone. Natural formation water, or connate water, is present in
the pore spaces of the rock in the oil zone with increasing depth. In the water zone, these pore
spaces are fully saturated with water. In this type of reservoir, natural gas occurs as solution gas,
and is evident or present only when reservoir pressure sufficiently decreases, and gas breaks out of
the oil.  
Fig. 4 – Example of Solution Gas
 
2. In other reservoirs, the associated free gas is found at the structural top of the trap, forming a gas
cap. Beneath the gas cap is the oil accumulation and beneath the oil is water (again, refer to Figure
2-1). Such a reservoir could produce both the associated free gas and solution gas dissolved in the
oil with well completions established in each zone.   
Fig. 5 –  Example of a Gas Cap

3. In still other natural gas reservoirs, the gas zone structurally is above a gas-water transition zone.
The gas zone often contains connate water saturation, which increases with depth, to eventually
become a completely water saturated zone. The gas in this type of hydrocarbon accumulation is
referred to as non-associated gas, as there is no oil present. 
Fig. 6  – Example of Non-Associated Gas

Rock Types

The most common sedimentary rocks associated with petroleum reservoirs are:

Figure 7 shows the lithologic relationship of these common rock materials, indicating the transition
from one type of rock to another, and the composition and nomenclature of multi-compositional
reservoir rocks.
Fig. 7 - Sedimentary rock constituents – three component diagram.  (Classification of the common
sedimentary rocks by mineral composition)
Reservoir Pressure and Temperature

The pressure at depth is usually expressed as:

Gradients within the range of 0.433 to 0.5 psi/ft are considered normal. Abnormal pressures can
occur when some part of the overburden load is transmitted to the formation fluids during major
tectonic events. Abnormal pressures caused by closure of rock structures may lead to very high
gradients (0.8 to 0.9 psi/ft), and may even approach the geostatic gradient of 1.0 psi/ft. 
Reservoir temperatures conform to the regional geothermal gradient. A normal value is often
approximately 1.6°F/100 ft of depth. The geothermal gradient varies from one location to another,
and also varies with depth, according to the rock type. The flowing temperature gradient varies
from the static geothermal gradient and can be measured in a well while the well is flowing under
stabilized conditions.
Because of the large thermal capacity of the rock matrix, conditions within the reservoir at any
single depth across the reservoir may be considered isothermal in most cases.
Three Types of Oil Reservoirs
There are different types of hydrocarbon fluids that can be found in a reservoir. Understanding the
characteristics of these different fluid types will help in determining how best to deplete the
reservoir. But that is a topic for another time; for now we will examine the different types of
reservoirs.

Two Broad Types: Oil and Gas

First, petroleum reservoirs are broadly divided into oil reservoirs and gas reservoirs. Oil reservoirs
around the world are not the same; the fluid composition, the prevailing temperature and the
pressure (called the reservoir pressure) all vary. We can further break down oil reservoirs into
different types based on the interaction between this reservoir pressure and the hydrocarbon fluids.

Classifying with Help from the Phase Diagram

To understand this classification, we’ll be looking at a very important chart; a pressure-temperature


envelope or phase diagram. True, there are different types of crude oil, such as black oil, volatile
oil, etc. For all oil reservoirs, the basic concept of the phase envelope is the same for each type of
crude oil even though there are slight variations. Therefore, even though fluid types vary, we can
study the three kinds of oil reservoirs by looking at a single phase envelope.

This is the phase envelope for an oil reservoir that contains black oil. The bubble point line on the
envelope is our area of concentration for oil reservoirs. We already know that reservoir temperature
and pressure for different oil reservoirs around the world varies. In this particular phase envelope,
the numbers 1, 2 and 3 trace the different possible points the reservoir pressure can be. One
reservoir in the North Sea can have the reservoir pressure at point 1 while another can have
reservoir pressure at point 3. These two reservoirs may be oil reservoirs in the same region, but
since they are found at different reservoir pressures, this pressure difference will affect the kind of
oil reservoir and our depletion plan. Thus, our strategies to produce from a reservoir with a reservoir
pressure at point 1 will be different from that of the reservoir whose reservoir pressure is at point 3
even though they are both oil reservoirs.
Undersaturated Oil Reservoirs

Take a look at the phase envelope. If the reservoir pressure is at point 1 then that reservoir is an
undersaturated oil reservoir. What this means is that the reservoir pressure (point 1) is higher than
the bubble point pressure.
The bubble point pressure (on the bubble point line) is a very important concept for every oil
reservoir. This pressure is obtained at the laboratory from tests conducted on samples of the
reservoir fluid. To conduct this test, a sample of the reservoir oil is placed in a chamber (container)
at the reservoir pressure. We slowly lower the pressure exerted on this oil. It’s not difficult to
imagine, just like your soft drink or beer, there is always some gas bubbled into the liquid under
pressure. This pressure is sealed by the bottle cap. When you open the bottle, you release the
pressure and expose the liquid to the atmosphere thereby lowering the pressure, so the gases bubble
and come out of the liquid. Let’s assume that the bottling company did not force enough gases into
this drink and the soft drink or beer could still have absorbed more gases if they were bubbled at
that pressure. This assumption leads us to understanding undersaturated oil reservoirs.
The word “undersaturated” also means that this kind of oil reservoir is not saturated with gas
bubbles. The saturation here is referring to gas bubbles, and so the pressure has kept all the lighter
gaseous hydrocarbons (bubbles) inside the crude oil. At the same time, this undersaturated state tells
us that assuming there were more gas bubbles, the crude oil in this reservoir would have
comfortably absorbed them. An understanding of this will guide us in coming up with the best
strategies to produce the hydrocarbons in this reservoir.
Saturated Oil Reservoirs

Any reservoir having a pressure that falls at point 2 on the phase envelope is a saturated oil
reservoir. A saturated reservoir is fully saturated with lighter hydrocarbon gases at that reservoir
pressure. Saturation means that the crude oil here is fully occupied with dissolved gases and will not
take any extra gas bubbles at that pressure (point 2) unless we increase the pressure. Remember, it
is the position of the reservoir pressure on the phase envelope that helps us distinguish what type of
oil reservoir we are dealing with.
Back to our analogy with a bottle of beer or soft drink, in this case the bottling company kept
forcing gas bubbles into the drink in the bottle until it could not take any extra gas. This state is the
saturation state, and oil reservoirs found at this state are termed saturated oil reservoirs. It is not
enough to know that we just drilled into an oil reservoir — knowing the kind of oil reservoir it is
will help us choose the right methods to efficiently produce the hydrocarbons in the reservoir.
(Related reading: The Different Ways of Perforating a Well.)
Gas-Cap Reservoirs

Point 3 on the phase envelope is called the 2-phase region; remember that at points 1 and 2 we only
have crude oil in the reservoir with some form of gas saturation at that pressure.
In the beer and soft drink analogy, what happens the moment we open the bottle cap? Some gases
come out of solution because we just disturbed the pressure equilibrium of the drink in the bottle.
Gas-cap reservoirs are operating under similar conditions — some of the gases have come out of
solution and are now occupying the top of the reservoir, whereas in the case of the soft drink or
beer, the gases simply escape to occupy space in the room.
For an oil reservoir, because gas is less dense than oil, the gas will occupy the upper portions of the
reservoir rock. If the reservoir pressure was higher (like at points 2 or 1) then the gases would have
been forced back into solution. In other words, gas-cap oil reservoirs are supersaturated with gas.
Gas-cap reservoirs have gases at the upper portions (cap) of the reservoir trap. Taking the time to
know what kind of oil reservoir we are dealing with is never a waste of time, and this understanding
could be the difference that allows us to maximize production from the reservoir.
Conclusion
It is not enough to know that we just drilled into an oil reservoir; an understanding of the kind of oil
reservoir we are dealing with ensures efficient production from the reservoir. (For more
information, see the article Understanding Reservoir Drive Mechanisms.) For example, if we know
that the reservoir is a gas-cap oil reservoir, then we can engineer our flow rate in such a way that we
do not produce too fast from the reservoir.
Producing too fast will make us produce the gases at the top sooner than we want to and we don’t
want that. This is called a “blowdown”. The gas at the cap helps to push the oil down through the
perforations into the drilled well. The longer we can keep the gas in the cap while producing
economically the better. So an understanding of the kind of oil reservoir we are dealing with helps
us make the best decisions to efficiently produce the hydrocarbons in the reservoir.
Types of Reservoir Fluids

The reservoir can produce different types of fluids. Each different fluid type has characteristics that
petroleum engineers need to be aware of in order to produce the reservoir efficiently. Generally the
reservoir fluid types are classified into 5 separate categories from a phase diagram. They are:

1. Black Oil
2. Volatile Oil
3.  Gas Condensate
4. Wet Gas
5. Dry Gas
The fluid type can be identified by the shape of it’s phase diagram and the position of it’s critical
point relative to the reservoir temperature. In the following figures, we assume that the reservoir
pressure declines at a constant temperature (isothermal condition).

Black Oil

The figure below shows the pressure decline of a black oil reservoir on a phase diagram. The
vertical line from point 1 to point 3 represents the reservoir pressure decline path at a constant
temperature. The pressure and temperature of a separator at the surface is also shown in the
diagram.
For a black oil reservoir, the initial pressure at point 1 is much larger than the bubble point
pressure. When reservoir pressure is above the bubble point (between point 1 and point 2), we call
this the undersaturated reservoir state. Undersaturated implies that the oil can absorb more gas if
more gas became present in the reservoir. When the pressure hits the bubble point line at point 2,
the first bubble of gas forms. Below point 2, gas starts to evolve out of the oil into the reservoir. We
call this the saturated reservoir state. Saturated implies that the oil cannot absorb more gas if more
gas became present in the reservoir. As the pressure declines further to point 3, more gas evolves
out of the oil into the reservoir. At point 3, the reservoir pressure can no longer push the fluids to
the surface under it’s current state. At the surface separator condition, more gas evolves out of the
oil. For black oil reservoirs, a large percentage of the oil phase is recovered at the surface .  
Volatile Oil

The figure below show the pressure decline of a volatile oil reservoir on a phase diagram.
A volatile oil reservoir starts out like a black oil reservoir with the initial reservoir pressure lying
above the bubble point line. However for a volatile oil, the initial pressure is much closer to the
bubble point pressure. This is because a volatile oil reservoir contains more lighter hydrocarbon
components than the black oil reservoir, thus the phase diagram shifts up and to the left. Further, the
quality lines are closer together. This means that a small reduction of pressure below the bubble
point line will release a large quantity of gas. For a volatile oil reservoir, more gas is produced at
the surface because more solution gas evolves out of the oil. This lead to less liquid recovery as
indicated in the phase diagram above. The dividing line between volatile oils and black oils are
somewhat arbitrary. But it should be clear that volatile oils produce a higher quantity of gas than
black oils.

Gas Condensate

The figure below shows the pressure decline of a gas condensate oil reservoir on a phase diagram:
A gas condensate reservoir is identified by its temperature. The reservoir temperature falls between
the critical point temperature and the cricondentherm. As pressure declines below the dew point
line, the heavier hydrocarbons fall out in the reservoir as liquid. This liquid more or less is
immobile and blocks the pore space near the wellbore. Thus less gas is produced because the liquid
partially blocks the gas flow. Liquid continues to form in the reservoir as long as the pressure is
within the retrograde region. As pressure falls below the retrograde region, the liquid re-vaporizes
into gas. This type of reservoir is pretty bizarre, but it can be explained by a phase diagram.

Wet Gas

The figure below shows the pressure decline of a wet gas reservoir on a phase diagram:
A wet gas reservoir has a reservoir temperature above the cricondentherm. As the pressure declines,
the reservoir never enters into the two-phase region. This means liquid will never drop out in the
reservoir and it will only contain gas over its producing life. However, liquid will form at surface
separator conditions as shown in the figure above. That is why they call it “wet” gas because
condensate drops out of the gas at the surface.

Dry Gas

The figure below shows the pressure decline of a dry gas reservoir on a phase diagram:
A dry gas is primarily pure methane and the phase diagram is short and condensed. As the reservoir
pressure declines, the pressure depletion path never enters the two-phase region. Further, the
pressure and temperature at surface conditions falls outside of the two-phase region. This means
that the fluid stays in the gas phase at reservoir and surface conditions. A dry gas reservoir is simply
a gas reservoir and is the easiest reservoir to model analytically. 

In summary, we defined the five types of reservoir fluids utilizing phase diagrams. Understanding
the type of fluid in a reservoir greatly effects the depletion plan of a well and how petroleum
engineers perform certain calculations. 

Lecture 11. The future of Petroleum Geoscience

Petroleum geoscience is geology and geophysics applied to petroleum exploration and


production. In this book we will try to show the wide range of disciplines that are relevant and
useful for this purpose. Many of the disciplines in the geosciences are highly specialised and
there is often too little commu- nication between the different fields. Most researchers naturally
Dissolution of
CO
focus on a very small area because of the requirements with respect to methods and analytical
2
carbonate rocks
CO
techniques, and the demands of following the litera- ture. Petroleum geoscience requires a broad
Photosynthesis

overview of substantial parts ofmatter


CO Oxidation of
2
organic geology and geophysics and provides good training in the
integration of very differentTransport of types of data Delta and models. These skills are also applicable in many
organic matter Addition of organic matter
types of environmental research and when solving
by rivers
Organic
practical
+ nutrients to environment
the ocean problems.
The petroleum industry employs a large percentage plant
matter in of the
Photosynthesis
of biogenicworld’s
CO geologists
2 and geophysicists
Breakdown
carbonate
delta
and funds much of the research in this field. sediment Oxidation of
organic matter
CO 2

Most of the obvious petroleum-bearing structures of carbonate have already been found in the explored
Deposition

CO Ocean floor
sedimentary basins and there are now rather Oxidation few areas that have yet to be explored seismically
2
and
biological processingDeposition of of organic
and by drilling. The large, easy-to-find structures did not
materialorganic usually require very advanced methods
material
C
and geological skills. Kerogen in sediments

More and more sophisticated methods are therefore CO 2


used in modern exploration. We are on a
global basis not finding enough new oil fields oil to replace the pro- duced oil. Global reserves
Carbon in sedimentary rocks
have however not changed very much because of higher estimates of recovery from existing fields
and because unconventional oil like tar sand is now included in the reserves.
There is a major challenge for geoscientists to develop ever better exploration methods and to
opti- mise production.
Even if the global production of conventional oil may be reduced there will be significant
production for many decades. This will mostly come from the tail production of giant fields and
from small reservoirs, but it is rather labour intensive. This is also the case with unconventional
oil (tar sand and oil shale) and also tight gas reserves and shale gas.
Until enough alternative sources of energy are developed it is necessary to extract fossils fuels
from these resources. This should be done with as little environmental damage as possible and this
requires a new generation of highly skilled geoscientists.
Dissolution
CO2 of carbonate
rocks
CO
Photosynthesis
CO2 Oxidation
of organic
matter
Transport Delt
of organic a Addition of organic matter
matter by + nutrients to the ocean
rivers Organic
plant Photosynthesis
matter in of biogenic CO2
Breakdown
delta carbonate
sediment Oxidation
of organic
CO2 matter
Depositi
of carbonate
on
CO2Ocean floor
Oxidation and
biological processingDeposition of of
organic materialorganic material
C
Kerogen in sediments

CO2

Carbon in sedimentary rocks

Fig. 1.14 Illustration of the carbon cycle. Carbon from organic matter and carbonate rocks are the
major sinks for carbon (CO2). The rate of precipitation of carbonate in the ocean by organisms is
limited by the supply of Ca++ and Mg++ from weathering of silicate rocks brought in by rivers.

You might also like