You are on page 1of 86

Module Code and Name: DTME216: BASICS OF OIL AND GAS SYSTEMS

Module Level: YEAR II SEMESTER I


Module Credit: 3 CU
Module Overview
Introduces the student to Oil and Gas systems in a reservoir
Module Learning Outcome
By the end of this module the learner should be able to:
1. Describe the composition of petroleum and its different characteristics of the reservoir rock.
2. Explain gas and oil gas systems.
3. Describe the distribution of oil, gas and water in the reservoir.
Competences
1. Describe and make a sketch reservoir where the petroleum is stored.
2. Explain the composition of petroleum and its different characteristics of the reservoir rock.
3. Explain gas and oil gas systems.
4. Describe the distribution of oil, gas and water in the reservoir.
Preparatory Assignment
Case studies: in the oil and gas producing states of the world
Detailed Module Description Duration

Sub-module 1: Development of Oil and Gas


Upstream, Mid-Stream and Down Stream processes
6 hours
See Basics of Oil and Gas
Sub-module 2: Oil and gas source rocks
Oil and gas reservoir rocks, oil and gas trap, coalbed methane, shale gas
6 hours

Exploration for oil and gas is today strongly dependent on the recognition and understanding of
some basic geological facts and principles. For convenience, they are listed here in their simplest
form.
Hydrocarbons
Oil and gas are derived from organic-rich source rocks comprising mainly the remains of marine
algae and bacteria, and plant matter of continental origin. Oil and gas occur underground in the
pore spaces of sedimentary rocks and are trapped there if prevented from migrating further.

Rocks
Rocks are divided into three main groups: igneous rocks, which include granites and volcanic
rocks consolidated from hot, liquid material; sedimentary rocks, which are either fragments of
other rocks deposited on land or under the sea by wind and water, chemically precipitated from
evaporating waters, or of organic origin; metamorphic rocks, which comprise rocks originally of
igneous or sedimentary origin whose composition and structure have been profoundly changed
by heat and pressure.
During the excavation, geologists carry out drilling in the swamps or earth crust different types
of layers of soil and rocks can be encountered in the excavation process. The circular drill, drills
out cylindrical rock cores shown is the photo below.
Most hydrocarbon accumulations are limited to sedimentary rocks although some significant oil
and gas accumulations are contained in fractured igneous and metamorphic rocks. The
occurrence of many metals, on the other hand, is largely confined to igneous and metamorphic
rocks, with the exception of some iron, and sulphide ores such as those of copper, zinc an d lead,
and "placer" deposits like those of gold, tin and uranium.
Photos taken by Tukahirwa Gilbert. Rock cores that depict geological rock structure of a swamp
at Bukasa In-Land port.

Types of Rocks Encountered in Wells


Rocks are classified as clastic or nonclastic.
Clastic rocks are those made up of particles derived from some preexisting rock. Nonclastic
sedimentary rocks such as limestone or dolomite are formed by chemical precipitation. Important
clastic rocks include sand and sandstones, silt and siltstones, conglomerates, arkoses, oolites, and
shales.
The nonclastic sediments are chiefly limestone, dolomite, and chalk. Limestone in its pure form
is calcium carbonate, or calcite. Dolomite is the double carbonate of calcium and magnesium.
Limestones and dolomites are likely to have low porosities and permeabilities within the matrix,
and they may have been subjected to solution processes which developed channels, cavities,
vugs, and even caverns, which provide the more prolific reservoirs for petroleum. Chalk is a
form of limestone that is soft, low in permeability, and high in porosity. Salt, anhydrite, and
gypsum are common nonclastic rocks.
Sedimentary Basins
Hydrocarbons are found in sedimentary basins. It is important, therefore, to understand
something of the origin of sediments and of the basins in which they accumulate.
Erosion, Sediment Transport and Deposition
Wherever rocks are elevated and exposed to the elements, they become subject to weathering
and erosion. Assisted by the force of gravity, the products of erosion are carried away by water,
ice and wind and are deposited as sediment in the valleys and plains and in the seas beyond (Fig.
2.1).

Figure 2.1: Erosion, transport and deposition of sediment


Figure 2.2: Vertical and lateral changes in composition of sedimentary rocks. Schematic
section through the edge of the southern Sumatra basin. Note: Tuff is another name for
volcanic ash.

Under arid climatic conditions, salt and gypsum deposits, referred to generally as evaporites,
may form by the evaporation of sea water in shallow lagoons, for instance, or in desert basins of
in land drainage. The shallow waters of warm, clear tropical seas favour the growth of corals and
algae, which are important contributors to the formation of carbonate rocks such as limestone
and its related alteration product, dolomite. Where vegetable matter accumulates, peat is formed
which, after burial, converts into lignite and eventually coal. The beds of sedimentary rocks
deposited in this way are seldom uniform in thickness, composition or texture. The variations
may be small or large depending on many factors, and a section through a series of sedimentary
rocks usually shows lateral and vertical changes in both lithology and thickness (Fig. 2.2).

Basin Development
Sediments are deposited preferentially in topographic depressions, which may then be referred to
as sedimentary basins. Sedimentary basins are of several types. Rift basins (e.g. Fig. 2.3) are
relatively long and narrow because they occupy the depressions formed when the crust fractures
and pulls apart under tension.

Figure 2.3 Typical locations of sedimentary basins within continental areas and overlying
the transition from oceanic to continental crust The sedimentary fill of these basins
becomes increasingly metamorphosed with depth.

Diagenesis of Sediments
After burial, sediments react to increasing temperature and pressure, and commonly also to the
corrosive action of water saturated with other chemicals. The result is a change in the
composition of the rock (diagenesis) which, in general, has the effect of reducing its porosity.
Limestones, however, may alter to dolomite with the creation of extra porosity. If subjected to
very high temperatures and pressures, the buried sedimentary rock may be completely
recrystallised; it is then said to have been metamorphosed from, say, sandstone to quartzite, or
limestone to marble.

Folds and Faults


Basins develop under essentially tensional conditions, and the unsupported sides of the fracture
tend to subside along "normal" fault planes so as to fill the gap (Fig. 2.4A). Most tensional
situations in the crust are relieved by faults of this type. Depending on other factors, a group of
faults either have the same sense of movement and result in overall subsidence by the creation of
a series of "half grabens" (Fig. 2.4D), or the sense of movement alternates to form a series of
"horsts" (highs) and "grabens" (lows) and Fig. 2.4C).
Horizontal stresses sometimes result in major linear "wrench "faults (Fig.2.4E). If accompanied
by local lateral tension, then fault-bounded subsidence of the crust may occur (Fig. 2.4G), over
which a basin will develop. If, on the other hand, shearing is accompanied by local lateral
compression, then linear slivers of the more rigid rocks will be squeezed upwards as narrow
horsts bounded by outward-curving faults ("flower structures", Fig. 2.4F) and the overlying
sediments will be uplifted ("inverted") into a series of fault-bounded anticlines. Provided other
criteria for the accumulation of hydrocarbons are met, these anticlines can form traps for oil and
gas.
With mild horizontal compression, sediments deform into a series of anticlines and synclines
(Fig. 2.41). If compression continues, a plane of detachment may develop between the relatively
plastic sedimentary sequence and the underlying more rigid crystalline basement. Under these
circumstances, folds may go through the stages of being asymmetric (Fig. 2.4J) and recumbent
(Fig. 2.4K), until finally the upper limb of a recumbent fold shears at the point of maximum
curvature and a thrust develops (Fig. 2.4L). Repeatedly folded and thrust sequences of this type
develop over the subduction zones of active continental margins.
Figure 2.5 FAULT AND FOLD TYPES

Source Rocks, Oil Generation and Migration


Hydrocarbons are formed by the thermal conversion of organic matter trapped in sedimentary
rocks (source rocks). With increasing burial, the temperature of the source rock increases and at
a given temperature the organic matter (kerogen) transforms into oil and gas. After expulsion
from the source rock, oil and gas migrate to the reservoir formations (Fig. 2.6).
In aquatic environments, abundant life is seen only in the upper water layers (down to 100
metres depth), where algae and bacteria create organic matter by photosynthesis from carbon
dioxide and water. Under normal circumstances, the organic matter is recycled into carbon
dioxide and water by oxidation and by aerobic bacteria.
Only under certain conditions, for example when photosynthetic activity is abnormally high, will
organic matter accumulate on the sea floor in large quantities. This is related to an unusually
large supply of nutrients, supplied, for instance, by up-welling water. In such cases bacteria and
algae consume so much oxygen that it causes a drastic reduction of the aerobic zone of the water.

Figure 2.6: The burtal/temperature-related maturation of a source rock and the migration
of the generated hydrocarbons into both structural and stratigraphic traps

The algae and aerobic bacteria die and are deposited in large amounts on the bottom. Here,
anaerobic bacteria convert the degradable parts into carbon dioxide and water, simultaneously
synthesizing their own bodies. In this way, the original organic matter is reworked and converted
into a (sapropelic) biomass. This biomass consists of bacterial bodies (the "reworkers") plus the
microbially resistant part of the planktonic bodies (lipids) together with the resistant organic
matter originally present or supplied from elsewhere, such as pollen, spores, plant waxes and
plant resins. As the anaerobic degradation is relatively slow and the biodegradation of bacteria
and algae in these circumstances very incomplete, organic-rich sediments (3 to 20 per cent
carbon) can be deposited. The type of organic material contained in such rocks, which are termed
"source rocks", determines whether they are capable of generating predominantly oil or gas.
Thus hydrogen-rich, amorphous, sapropelic organic material is an excellent source for both crude
oil and gas, while hydrogen-poor, coaly organic matter is mainly a source for gas. The quality
and composition of crude oil is also t o a large extent determined by the type of organic material.
Structureless organic matter of marine origin will result in an oil different from that formed from
organic matter of lacustrine (i.e. from lakes) or terrestrial origin containing plant waxes and
resins.
Rich oil-source rock, generally known as oil shale, is exposed at the surface in a number of
places in the world. If heated in a retort, it will give off oil. Oil shales formed the basis of local
synthetic oil production in several areas in the past and, in a few areas, production continues
today largely on a pilot-plant scale.

Figure 2.7: Trinidad Pitch Lake.

Coal is also an important source of hydrocarbons, especially gas. It originates largely by burial of
forest and swamp types of vegetation. As the result of progressive depth/temperature-related
carbonisation of the original plant material, the volatile gaseous and liquid hydrocarbons are
driven off to leave a sequence that ranges from peat, through lignite and the bituminous coals, to
anthracite, which is almost pure carbon.
Oil and/or gas is said to migrate when it leaves the source rock in which it was generated and
moves either to a reservoir rock where it is trapped, or to the earth' s surface where it escapes as a
seepage (Fig. 2.7). Migration is not fully understood, but is thought to take place along faults and
minor fractures within the rock sequence. As hydrocarbons are lighter than water, they generally
migrate in an upwards or sideways direction from areas of higher to lower pressure.

Accumulation of Oil and Gas


Migrating oil and gas can accumulate if the following essential geological conditions are
satisfied:
• The presence of reservoir rock, i.e. formations containing interconnected pores (e.g. sands and
sandstones) or cracks and voids (e.g. some limestones).
• The presence, at the top of the reservoir rock, of a formation that is impervious to the passage
of hydrocarbons (e.g. anhydrite, salt or shale). When this lies directly over the reservoir rock, as
in most oil and gas accumulations, it is called a "caprock", "roof rock" or "seal".
• The presence of a trap, i.e. a geometrical configuration of the reservoir rocks and seal that
prevents the lateral escape of fluids, such as when the cap rock is concave when viewed from
below. Geometrical shapes of this type are depicted on maps as being enclosed by a series of
depth contours: thus these potential hydrocarbon traps are commonly referred to as "closures".
These conditions define a potential trap in which oil and gas migrating from the source
rock may accumulate. Because of a difference in density, oil will displace downwards the water
previously filling the reservoir, and free gas (gas not dissolved in oil as a result of high pressure),
if present, will collect in the highest part of the reservoir to form a "gas cap" with the oil below
it. Below the oil, the pores in the reservoir rock will remain full of formation water, usually
saline.

Oil and Gas Traps


Traps. Geologists commonly describe traps on the basis of their origin. Structural traps are
closures formed by structural movements within the Earth, and stratigraphic traps are closures
formed by sedimentation and diagenesis, without the need for structural movements.
Structural/stratigraphic traps are closures formed by patterns of reservoir rock that impinge upon
a structure.
Structural traps are formed most commonly by structural uplift and differential compaction.
Typical structural traps are structural domes and doubly plunging anticlines (see Fig. 2.8). These
traps have a structural high and quaquaversal dips (the seal dips away from a structural high in
all directions). The bulk of the world’s oil is found in these four-way-closure traps which were
the first type to be exploited by surface mapping. Many major oil fields in the world were
discovered by using surface mapping to locate domal structures.

Fig. 2.8—Diagram description of hydrocarbon traps.


A more complex method of forming a structural trap is by faulting and structural uplift (see Fig.
2.8). Faulted structures can vary from a simple faulted anticline to complex faulting around
piercement structures and domal uplifts. Faulted structures are very common and form some of
the most complex reservoirs known. Types of faults include normal, listric, reverse, and thrust,
which are related to the stress fields generated during structural movement. Piercement traps
(diapirs) are formed typically by salt moving up through a stack of sediment driven by the
density difference between salt and quartz or carbonate. Closure is achieved by the uplift of
sediments juxtaposed to the piercement dome, by the top seal being an overlying impervious bed
and the lateral seals being formed by structural dip, by sealing faults, or by the piercement salt.
Faulted reservoirs commonly have a bottom seal formed by the lower contact of sand with shale.
The bottom seal, along with the oil/water contact within the sand body, forms the base of the
reservoir.
Structural/stratigraphic traps are formed by a combination of structure, deposition, and
diagenesis. The most common form, the updip pinchout of reservoir lithology into a sealing
lithology (see Fig. 2.8), is found in the flanks of structures. The top and updip seal is normally an
impervious rock type, and the lateral seals are formed by either structural dip or the lateral
pinchout of reservoir rock into seal material. The base of the reservoir is defined by a bottom seal
composed of impervious rock and by an oil/water contact. During relative sea-level fall, streams
may erode deep valleys, thus forming lateral seals for fluvial sediments. Onlap of sand onto a
paleotopographic high during relative sea-level rise can produce an updip seal for a sand body.
Unconformity traps are formed by the truncation of dipping strata by overlying bedded sealing
lithology. The reservoir rock may be found in the form of buried hills formed by erosion during
the time of the unconformity. The oil/water contact forms the base of these reservoirs. A
stratigraphic trap may be partly related to diagenetic processes; for example, the updip seal for
the supergiant Coalinga field, California, is tar- and asphalt-filled sandstone and conglomerate.
Many traps in the Permian reservoirs of west Texas, are formed by lateral changes related to
stratigraphy from porous to dense dolomite in an updip direction.
Stratigraphic traps are formed by depositional processes that produce paleotopographic highs
encased in impermeable material, such as evaporite or shale (Fig. 2.8). Closure occurs when
there is contact between seal material and underlying sediment. The most common type is a
carbonate buildup, usually erroneously called a ―reef.‖ Piles of sand deposited on the seafloor by
density currents often form broad topographic highs that, in turn, form stratigraphic traps.
Structure may also play a part in the geometry of stratigraphic traps, although the defining
characteristic is that structure is not required to form the trap.

Seals A seal is a low-permeable to impermeable rock or immobile fluid, such as tar, with a
capillary entry pressure large enough to dam up or trap hydrocarbons. Typical seals include top,
bottom, lateral, and fault, as shown in Fig. 2.9. Faults may be sealing or nonsealing, depending
on whether the sand offsets another sand (nonsealing) or shale (sealing). Any lithology can be a
seal or flow barrier. The requirement is that the minimum capillary displacement pressure of the
seal or flow-barrier material be greater than the buoyancy pressure of the hydrocarbons in the
accumulation. The continuous, small, pore-throat sizes create a barrier to moving hydrocarbons,
causing them to dam up or become trapped. Therefore, the size of the continuous pore throats
and the density of the hydrocarbons and water are critical elements in evaluating a seal or flow
barrier.

Fig. 2.9 —Some typical seals; HCH is the hydrocarbon-column height that the weakest seal
will hold. Figure courtesy of Sneider Exploration Inc.

Unconventional sources of oil and gas


The reservoirs described earlier are called conventional sources of oil and gas. As demand
increases, prices soar and new conventional resources become economically viable. At the same
time, production of oil and gas from unconventional sources becomes more attractive.
These unconventional sources include: very heavy crudes, oil sands, oil shale, gas and
synthetic crude from coal, coal bed methane, methane hydrates and biofuels.
Extra heavy crude
Very heavy crude are hydrocarbons with an API grade of about 15 or below.
If the reservoir temperature is high enough, the crude will flow from the reservoir. In other areas,
such as Canada, the reservoir temperature is lower and steam injection must be used to stimulate
flow from the formation. When reaching the surface, the crude must be mixed with diluents
(often LPGs) to allow it to flow in pipelines. The crude must be upgraded in a processing plant
to make lighter Syn-Crude with a higher yield of high value fuels. Typical SynCrude has an API
of 26-30. The diluents are recycled by separating them out and piping them back to the wellhead
site. The crude undergoes several stages of hydrocracking and coking to form lighter
hydrocarbons and remove coke. It is often rich in sulfur (sour crude), which must be removed.

Tar sand

Typical tar sand contains sand grains with a water envelope, covered by a bitumen film that may
contain 70% oil. Various fine particles can be suspended in the water and bitumen.
This type of tar sand can be processed with water extraction. Hot water is added to the sand, and
the resulting slurry is piped to the extraction plant where it is agitated and the oil skimmed from
the top.

Oil shale
Most oil shales are fine-grained sedimentary rocks containing relatively large amounts of organic
matter, from which significant amounts of shale oil and combustible gas can be extracted by
destructive distillation.
Oil shale differs from coal in that organic matter in shales has a higher atomic hydrogen to
carbon ratio. Coal also has an organic to inorganic matter ratio of more than 4, i.e., 75 to 5, while
oil shales have a higher content of sedimentary rock.
Oil shales are thought to form when algae and sediment deposit in lakes, lagoons and swamps
where an anaerobic (oxygen-free) environment prevents the breakdown of organic matter, thus
allowing it to accumulate in thick layers.

Shale gas and coal bed methane


Oil shales are also becoming an important source of shale gas, and some analysts expect that this
source of natural gas can supply half of the gas consumption.
This form of production is different from oil shale gas, which is produced by pyrolysis (heating
and hydrocarbon decomposition) of mined oil shale.
Coal deposits also contain large amounts of methane, referred to as coal bed methane. The
methane is absorbed in the coal matrix and requires extraction techniques similar to shale gas.
Often the coal bed is flooded, so after well completion and fracturing, the coal seam (layer of
coal) must be dewatered. A common solution is to extract water through the well tubing.
Generally, the water needs to be pumped out and therefore control is needed to prevent the gas
from entering the water in the tubing (the well becomes gassy). This reduces the pressure and
allows methane to desorb from the matrix and be produced through the casing.

Coal, gas to liquids and synthetic fuel


Coal is similar in origin to oil shales, but typically formed from the anaerobic decay of peat
swamps and relatively free from non-organic sediment deposits, reformed by heat and pressure.
Sub-module 3: Main types of hydrocarbons
Petroleum, Diesel, paraffins, naphthenes, aromatics and impurities
12 hours

HYDROCARBONS
A hydrocarbon is an organic compound composed of two elements, hydrogen and carbon. A
large part of the composition of petroleum is made up of hydrocarbons of varying lengths. The
smallest hydrocarbon, methane, is composed of a single carbon atom and four hydrogen atoms.
However, hydrocarbons can literally consist of hundreds or thousands of individual atoms that
are linked together in any number of ways, including chains, circles, and other complex shapes.
Hydrocarbons may be gaseous, liquid or solid at normal temperature and pressure, depending on
the number and arrangement of the carbon atoms in their molecules. Those with up to four
carbon atoms are gaseous; those with twenty or more are solid; those in between are liquid.
Liquid mixtures, such as most crude oils, may contain either gaseous or solid compounds or both
in solution.
Hydrocarbons are carbon containing organic compounds that provide a source of energy and raw
materials. Real world relation: If you are driven in a car, bus, jet or truck, you have used
hydrocarbons. The gasoline and diesel fuel that we used in cars, jets and buses, are
hydrocarbons.

Organic compounds.
These are carbon containing compounds with the primary exception of carbon oxides, carbide,
and carbonated, which are considered inorganic. In organic compounds, carbon atoms are
bonded to hydrogen atoms or atoms of other elements that are near carbon in the periodic table,
especially nitrogen, oxygen, sulphur, phosphorous and the halogens.
The simplest organic compounds are hydrocarbons. Which contain only elements of carbon and
hydrogen. Thousands types of hydrocarbons are known each containing only the elements
carbon and hydrogen. The simplest hydrocarbon molecule, CH4, consists of a carbon atom
bonded to four hydrogen atoms. This substance called methane. It’s is an excellent fuel and it’s
the main component of natural gas.
The larger hydrocarbon molecules have two or more carbon atoms joined to one another as well
as to hydrogen atoms. The carbon atoms may link together in a straight chain, a branched chain
or a ring. Examples are:
From these three basic configurations a considerable number of hydrocarbons can be built up,
especially since more complicated compounds may be formed by combinations of chains and rings,
for example:

The number of hydrogen atoms associated with a given skeleton of carbon atoms may vary.
When a chain or ring carries the full complement of hydrogen atoms, the hydrocarbon is said to
be saturated, and such hydrocarbons are known as paraffins, paraffinic hydrocarbons or
alkanes/cycloalkanes. Straight-chain structures are normal paraffins, branched-chain structures
are iso-paraffins, and ring-type structures are cycloparaffins or naphthenes. Thus for three
hydrocarbons with five carbons atoms, all pentanes, we have, amongst others, the following structures:

Thus normal pentane, normally abbreviated to n - pentane, is a straight-chain, i.e. unbranched,


paraffin, isopentane is branched and cyclopentane a ring compound.
When less than the full complement of hydrogen atoms is present in a hydrocarbon chain or ring,
the hydrocarbon is said to be unsaturated. Unsaturated hydrocarbons are characterized by having
two adjacent carbon atoms linked by two or three bonds instead of only one. These links are
known as double bonds and triple bonds, respectively; they are not stronger than the single bond,
but on the contrary surprisingly vulnerable, with the result that the unsaturated compounds are
chemically more reactive than the saturates.
Straight- or branched-chain hydrocarbons with one double bond are called mono-olefins or
alkenes, hydrocarbons with a double b o n d in a ring are cycloolefins, or cycloalkenes, and those
with two double bonds in the structure diolefins or dienes. Hydrocarbons with a triple bond are
called acetylenes or alkynes.
The simplest members of the olefin and acetylene series are ethylene and acetylene, and
butadiene is the simplest diolefin:

Neither olefins nor acetylenes occur in crude oil or natural gas, but are produced by conversion
processes in the refinery and are important raw materials for chemical syntheses. Ring
compounds containing one or more six-membered rings with three alternate double bonds form
an important group known as aromatics because most of them have a characteristic smell. The
simplest member is benzene, C6H6, in which each carbon atom carries only one hydrogen atom:

More complex molecules of the aromatic series are obtained by replacing one or more hydrogen
atoms by hydrocarbon groups or by "condensing" one or more rings:

From these few examples it will be obvious that there is no end to the number and complexity of
hydrocarbon structures. By introducing other elements, in particular oxygen, nitrogen and
sulphur, the number of possibilities based on a carbon skeleton (and thus the number of possible
organic chemicals), increases tremendously.

Petroleum is a naturally occurring flammable liquid consisting of a complex mixture of


hydrocarbons of various molecular weights and other liquid organic compounds that are found in
geologic formations beneath the Earth's surface.

Hydrocarbons in Fuel
Hydrocarbons containing between six and 10 carbon molecules are the top components of most
fuels, regardless of whether they are alkanes, alkenes, or cyclic. In general, these molecules are
burned to produce energy.
Burning hydrocarbons requires oxygen. The hydrocarbon and oxygen combine, in a process
called combustion, to produce water, carbon dioxide, and energy. Of course, these molecules are
not the only 3 products of the combustion of hydrocarbon. Hydrocarbons that are contaminated
with atoms such as sulfur and nitrogen will also produce nitrogen dioxide and sulfur dioxide.
Because hydrocarbons are composed purely of carbon and hydrogen, their combustion with
oxygen can only produce water as a result of the combination between hydrogen and oxygen and
carbon dioxide as a result of the combination of carbon and oxygen. The energy produced by
burning a hydrocarbon comes from breaking both carbon-hydrogen and carbon-carbon bonds
and recombining them into carbon-oxygen and hydrogen-oxygen bonds.
Because an unsaturated hydrocarbon has fewer hydrogen carbon bonds, it has less hydrogen per
molecule than a similar unsaturated hydrocarbon and will produce more carbon dioxide. This
also means unsaturated hydrocarbons produce less energy when burned than do saturated
hydrocarbons. In order to gain the same amount of energy, a greater quantity of unsaturated
hydrocarbon must be burned and as a result more carbon dioxide is created in the process. Thus,
unsaturated hydrocarbons are less environmentally friendly than saturated hydrocarbons.
Beyond the release of carbon dioxide, burning hydrocarbons also releases other contaminants
into the atmosphere. Because refining hydrocarbons is not perfect process, all fuels will contain
some level of contaminants. During combustion, sulfur combines with oxygen to produce sulfur
dioxide. Sulfur dioxide later combines with hydrogen in the atmosphere to produce the weak
sulfurous acid as well as the strong sulfuric acid. Both of these contribute to acid rain. In addition
to sulfur, nitrogen is also a common contaminant in hydrocarbons. Nitrogen dioxide can react
with hydrogen in the atmosphere to produce nitric acid, which also contributes to acid rain.

How Hydrocarbons burn

Ideal Reaction
Under ideal settings, where only hydrocarbon and oxygen are present, the chemical reaction
commonly called combustion or burning produces only water, carbon dioxide, and energy as the
following basic equation shows. In the above ideal reaction, the energy gained from the reaction
is greater than the energy put into the reaction. It is common knowledge that a spark is needed to
make a hydrocarbon burn. The spark represents the energy need to break the carbon-carbon and
carbon-hydrogen bonds of the hydrocarbon molecule as well as the oxygen-oxygen bond of the
oxygen molecule. The typical C-C bond requires 350 kJ/mol to break, the typical C-H bond
requires 413 kJ/mol, and the O-O bond requires about 498 kJ/mol.
We know, however, the energy is released from these reactions and it is released when new
bonds are formed. The H-O bonds of water release about 464 kJ/mol of energy when formed and
the C=O bonds of CO2 release about 800 kJ/mol when formed. The net outcome is the release of
energy in the form of heat.

Oxygen as Limiting Factor

The atmosphere is not pure oxygen. It contains a number of other gases including nitrogen
(78%), argon, hydrogen, iodine, and other trace compounds. The effect of these other compounds
is two-fold. First, they act as contaminants in the reaction, which is discussed in this section.
They also act to effectively limit the concentration of oxygen. Rather than 100% oxygen,
reactions that occur under atmospheric conditions are subject to only 21% oxygen.
When oxygen is a limiting factor, it is not possible to pair every carbon atom with two oxygen
atoms during a combustion reaction. Thus, some carbon atoms end up with only one oxygen
atom. This produces carbon monoxide. When a combustion reaction produces CO, it is referred
to as incomplete combustion.
Atmospheric combustion would then be more closely modeled by the equation that follows (note
this is not a balanced equation).

Atmospheric Contaminants

Contamination can come from the petroleum refining process or from the atmosphere itself. The
most common atmospheric contaminant (for purposes of combustion) is nitrogen in the form of
N2. When nitrogen is burned at high temperatures in the presence of oxygen, it produces
nitrogen oxide and nitrogen dioxide. These two compounds represent about 1% of the output of a
common hydrocarbon combustion reaction under atmospheric conditions. Thus, the equation can
be rewritten to reflect this new addition.

Both nitrogen compounds can participate in the formation of nitric acid (HNO3), which is a
component of acid rain. The reactions are as follows:

Common Hydrocarbons and Their Uses


Name Number of Carbon Atoms Uses
Methane 1 Fuel in electrical
generation. Produces least
about of carbon dioxide.
Ethane 2 Used in the production of
ethylene, which is utilized
in various chemical
applications.
Propane 3 Generally used for heating
and cooking
Butane 4 Generally used in lighters
and in aerosol cans
Pentane 5 Can be used as solvents in
the laboratory and in the
production of polystyrene.
Hexane 6 Used to produce in glue for
shoes, leather products, and
in roofing
Heptane 7 The major component of
gasoline
Octane 8 An additive to gasoline that
reduces knock, particularly
in its branched forms
Nonane 9 The component of fuel,
particularly diesel
Decane 10 A component of gasoline,
but generally more
important in jet fuel and
diesel

Hydrocarbons longer than 10 carbon atoms in length are generally broken down through the
process known as ―cracking‖ to yield molecules with lengths of 10 atoms or less.
COMPOSITION OF PETROLEUM
The hydrocarbons in petroleum are mostly alkanes, cycloalkanes and various aromatic
hydrocarbons while the other organic compounds contain nitrogen, oxygen and sulfur, and trace
amounts of metals such as iron, nickel, copper and vanadium. The exact molecular composition
varies widely from formation to formation. Other compounds such as Sulphur, nitrogen and
oxygen combined with carbon and hydrogen are also found.
Metals in the forms of inorganic salts or organometallic compounds are present in the crude
mixture in trace amounts.
The ratios of the different constituents in crude oils, however, vary appreciably from one
reservoir to another. Normally, crude oils are not used directly as fuels or as feedstocks for the
production of chemicals. This is due to the complex nature of the crude oil mixture and the
presence of some impurities that are corrosive or poisonous to processing catalysts.
The exact molecular composition varies widely from formation to formation but the proportions
of chemical elements vary over fairly narrow limits as follows:

Composition by weight

Element Percent range


Carbon 83 to 87%
Hydrogen 10 to 14%
Nitrogen 0.1 to 2%
Oxygen 0.05 to 1.5%
Sulphur 0.05 to 6.0%
Metals < 0.1%

Four different types of hydrocarbon molecules appear in petroleum.


The relative percentage of each varies from oil to oil, determining the properties of each oil.

Composition by weight

Hydrocarbon Average Range


Paraffins 30% 15 to 60%
Naphthenes 49% 30 to 60%
Aromatics 15% 3 to 30%
Asphaltic 6% remainder

Petroleum is used mostly, by volume, for producing fuel oil and petrol, both important "primary
energy" sources. 84 vol. % of the hydrocarbons present in petroleum is converted into energy-
rich fuels (petroleum-based fuels), including petrol, diesel, jet, heating, and other fuel oils, and
liquefied petroleum gas.

Properties of petroleum

1) API gravity
API (American Petroleum Institute) gravity of petroleum fractions is a measure of density of the
stream. Usually measured at 60 oF, the API gravity is expressed as API = 141.5/specific gravity –
131.5 where specific gravity is measured at 60 oF.
According to the above expression, 10 o API gravity indicates a specific gravity of 1 (equivalent
to water specific gravity). In other words, higher values of API gravity indicate lower specific
gravity and therefore lighter crude oils or refinery products and vice-versa. As far as crude oil is
concerned, lighter API gravity value is desired as more amount of gas fraction, naphtha and gas
oils can be produced from the lighter crude oil than with the heavier crude oil. Therefore, crude
oil with high values of API gravity are expensive to procure due to their quality.
2) Watson characterization factor.
The K factor or characterization factor is a systematic way of classifying a crude oil according to
its paraffinic, naphthenic, intermediate or aromatic nature. 12.5 Or higher indicate a crude oil of
predominantly paraffinic constituents, while 10 or lower indicate crude of more aromatic nature.
The K factor is also referred to as the UOP K factor or just UOPK.
3) Sulphur content
Since crude oil is obtained from petroleum reservoirs, sulphur is present in the crude oil. Usually,
crude oil has both organic and inorganic sulphur in which the inorganic sulphur dominates the
composition. Typically, crude oils with high sulphur content are termed as sour crude. On the
other hand, crude oils with low sulphur content are termed as sweet crude. Typically, crude oil
sulphur content consists of 0.5 – 5 wt % of sulphur. Crudes with sulphur content lower than 0.5
wt % are termed as sweet crudes. It is estimated that about 80% of world crude oil reserves are
sour.
4) TBP/ASTM distillation curves
The most important characterization properties of the Crude/intermediate/product streams are the
TBP/ASTM distillation curves. Both these distillation curves are measured at 1 atm pressure. In
both these cases, the boiling points of various volume fractions are being measured. However,
the basic difference between TBP curve and ASTM distillation curve is that while TBP curve is
measured using batch distillation apparatus consisting of no less than 100 trays and very high
reflux ratio, the ASTM distillation is measured in a single stage apparatus without any reflux.
Therefore, the ASTM does not indicate a good separation of various components and indicates
the operation of the laboratory setup far away from the equilibrium.
5) Viscosity
Viscosity is a measure of the flow properties of the refinery stream. Typically in the refining
industry, viscosity is measured in terms of centistokes (termed as cst) or saybolt seconds or
redwood seconds. Usually, the viscosity measurements are carried out at 100oF and 210oF.
Viscosity is a very important property for the heavy products obtained from the crude oil. The
viscosity acts as an important characterization property in the blending units associated to heavy
products such as bunker fuel. Typically, viscosity of these products is specified to be within a
specified range and this is achieved by adjusting the viscosities of the streams entering the
blending unit.
6) Flash and fire point
Flash and fire point are important properties that are relevant to the safety and transmission of
refinery products. Flash point is the temperature above which the product flashes forming a
mixture capable of inducing ignition with air. Fire point is the temperature well above the flash
point where the product could catch fire. These two important properties are always taken care in
the day to day operation of a refinery.
7) Pour point
When a petroleum product is cooled, first a cloudy appearance of the product occurs at a certain
temperature. This temperature is termed as the cloud point. Upon further cooling, the product
will ceases to flow at a temperature. This temperature is termed as the pour point. Both pour and
cloud points are important properties of the product streams as far as heavier products are
concerned. For heavier products, they are specified in a desired range and this is achieved by
blending appropriate amounts of lighter intermediate products.
8) Octane number
Though irrelevant to the crude oil stream, the octane number is an important property for many
intermediate streams that undergo blending later on to produce automotive gasoline, diesel etc.
Typically gasoline tends to knock the engines. The knocking tendency of the gasoline is defined
in terms of the maximum compression ratio of the engine at which the knock occurs. Therefore,
high quality gasoline will tend to knock at higher compression ratios and vice versa.
However, for comparative purpose, still one needs to have a pure component whose compression
ratio is known for knocking. Iso-octane is eventually considered as the barometer for octane
number comparison. While iso-octane was given an octane number of 100, n heptane is given a
scale of 0. Therefore, the octane number of a fuel is equivalent to a mixture of an iso-octane and
n-heptane that provides the same compression ratio in a fuel engine. Thus an octane number of
80 indicates that the fuel is equivalent to the performance characteristics in a fuel engine fed with
80 vol % of isooctane and 20 % of n-heptane.
Octane numbers are very relevant in the reforming, isomerization and alkylation processes of the
refining industry. These processes enable the successful reactive transformations to yield long
side chain paraffins and aromatics that possess higher octane numbers than the feed constituents
which do not consist of higher quantities of constituents possessing straight chain paraffins and
non-aromatics (naphthenes).

Dry gas is a natural gas stream that consists almost entirely of methane, with possibly a few
percent ethane and propane.
Wet gas is a natural gas stream that consists of a high percentage of methane (80-90%), plus
natural gas liquids, ethane, propane, butane, and natural gasoline, and small amounts of other
constituents.
Condensate is a very light crude oil-type hydrocarbon that comes from a well producing
predominantly natural gas (a gas well). Condensate generally has some natural gas liquids in it.
Light crude oil probably has a few percent natural gas liquids. Heavy crude has little or no
natural gas liquids and a high percentage of the heavy hydrocarbons. Sometimes the heaviest of
the heavy crudes have to be heated to make them fluid enough to pump out of the ground and
through a pipeline.

Bitumen is a composition of heavy hydrocarbons, including the very complex asphaltenes.


Together, they form solids at ambient temperatures. Bitumen is a hard substance to handle. In the
reservoir, bitnmen is a thick, sluggish fluid like road asphalt and has to be extracted from the
ground with much heat and effort and hauled away to a facility that can upgrade it to a useful
product.
Diesel
Diesel fuel is any liquid fuel used in diesel engines, originally obtained from crude-oil
distillation (petro diesel), but alternatives are increasingly being developed for partial or total
substitution of petro diesel, such as biodiesel (from vegetal oils), and synthetic diesel (usually
from a gas fuel coming from coal reforming or biomass, also named gas to liquid fuels, GTL). In
all cases, diesel nowadays must be free of sulphur.
Diesel types
Type A for road vehicles, B for industries (agriculture, fishing; same properties as type A, but
red-colored for different taxation), C for heating (not for engines; blue-colored).
Properties
i. Density = 830 kg/m3 (780-860 kg/m3 at 40 ºC). Thermal expansion coefficient=800⋅10-6 K-1.
880 kg/m3 for biodiesel (860-900 kg/m3 at 40 ºC).
ii. Boiling and freezing points. Not well defined because they are mixtures. In general, these
fuels remain liquid down to −30 ºC (some antifreeze additives may be added to guarantee that).

iii. Viscosity=3⋅10-6 m2/s (2.0⋅10-6-4.0⋅10-6 m2/s at 40 ºC) for diesel; 4.0⋅10-6-6.0⋅10-6 m2/s for
biodiesel.

iv. Vapor pressure=1-10kPa at 38 ºC for diesel and JP-4, 0.5-5kPa at 38 ºC for kerosene.

v. Cetane number =45(between 40-55); 60-65 for biodiesel. This is a measure of a fuel's ignition
delay; the time period between the start of injection and start of combustion (ignition) of the fuel,
with larger cetane numbers having lower ignition delays. This is only of interest in compression-
ignition engines, and only valid for light distillate fuels (because of the test engine; for heavy
fuel oil, a different burning-quality index is used, calculated from the fuel density and viscosity).

vi. Flash-point=50 ºC typical (40 ºC minimum). In the range (310-340)K (370-430K for
biodiesel).

vii. Heating value. HHV=47 MJ/kg, LHV=43 MJ/kg (HHV=40 MJ/kg for biodiesel).

PARAFFINS
Paraffin hydrocarbon, also called alkane, any of the saturated hydrocarbons having the general
formula CnH2n+2, C being a carbon atom, H a hydrogen atom, and n an integer. Paraffin wax,
colorless or white, somewhat translucent, hard wax consisting of a mixture of solid straight-chain
hydrocarbons ranging in melting point from about 48° to 66° C (120° to 150° F). Paraffin wax is
obtained from petroleum by dewaxing light lubricating oil stocks. It is used in candles, wax
paper, polishes, cosmetics, and electrical insulators. It assists in extracting perfumes from
flowers, forms a base for medical ointments, and supplies a waterproof coating for wood. In
wood and paper matches, it helps to ignite the matchstick by supplying an easily vaporized
hydrocarbon fuel.
The paraffins are major constituents of natural gas and petroleum. All paraffins are colorless.
Alkanes such as methane, ethane, propane, n and iso butane, n and iso pentane. These
compounds are primarily obtained as a gas fraction from the crude distillation unit. The simplest
alkane, methane (CH4), is the principal constituent of natural gas. Alkanes (paraffins), are
saturated hydrocarbons with straight (normal) or branched (iso) chains which contain only
carbon and hydrogen and have the general formula CnH2n+2.

They generally have from 5 to 40 carbon atoms per molecule, although trace amounts of shorter
or longer molecules may be present in the mixture.
- the alkanes from pentane (C5H12) to octane (C8H18) are refined into petrol, the ones from
nonane (C9H20) to hexadecane (C16H34) into diesel fuel, kerosene and jet fuel.
- alkanes with more than 16 carbon atoms can be refined into fuel oil and lubricating oil.
- at the heavier end of the range, paraffin wax is an alkane with approximately 25 carbon atoms,
while asphalt has 35 and up, although these are usually cracked by modern refineries into more
valuable products.
- the shortest molecules, those with four or fewer carbon atoms, are in a gaseous state at room
temperature. They are the petroleum gases.
- Depending on demand and the cost of recovery, these gases are flared off, sold as liquefied
petroleum gas under pressure, or used to power the refinery's own burners.
- During the winter, butane (C4H10) is blended into the petrol pool at high rates, because its high
vapor pressure assists with cold starts. Heavier petroleum fractions such as kerosene and gas oil
may contain two or more cyclohexane rings fused through two vicinal carbons.
Methane, ethane, propane, and butane are gaseous hydrocarbons at ambient temperatures and
atmospheric pressure. They are usually found associated with crude oils in a dissolved state.
Normal alkanes (n-alkanes, n-paraffins) are straight-chain hydrocarbons having no branches.
Branched alkanes are saturated hydrocarbons with an alkyl substituent or a side branch from the
main chain.
A branched alkane with the same number of carbons and hydrogens as an n-alkane is called an
isomer.
For example, butane (C4H10) has two isomers, n-butane and 2-methyl propane (isobutane). As
the molecular weight of the hydrocarbon increases, the number of isomers also increases.
Pentane (C5C12) has three isomers; hexane (C6H14) has five. The following shows the isomers of
hexane:
C5 - C17 / liquids
C16 – C78 / solids
Crude oils contain many short, medium, and long-chain normal and branched paraffins.
A naphtha fraction (obtained as a light liquid stream from crude fractionation) with a narrow
boiling range may contain a limited but still large number of isomers.
Properties of Paraffins
i. Substitution reactions
One or more hydrogen atoms of alkanes can be replaced by halogens, nitro group and sulphonic
acid group. Halogenation takes place either at higher temperature (573-773 K) or in the presence
of diffused sunlight or ultraviolet light. Lower alkanes do not undergo nitration and sulphonation
reactions. These reactions in which hydrogen atoms of alkanes are substituted are known as
substitution reactions.
ii. Combustion
Alkanes on heating in the presence of air or dioxygen are completely oxidized to carbon dioxide
and water with the evolution of large amount of heat.
iii. Controlled oxidation
Alkanes on heating with a regulated supply of dioxygen or air at high pressure and in the
presence of suitable catalysts give a variety of oxidation products.
iv. Isomerization ; n-Alkanes on heating in the presence of anhydrous aluminium chloride and
hydrogen chloride gas isomerize to branched chain alkanes. Major products are given below.
Some minor products are also possible which you can think over. Minor products are generally
not reported in organic reactions.

v. Pyrolysis
Higher alkanes on heating to higher temperature decompose into lower alkanes, alkenes etc.
Such a decomposition reaction into smaller fragments by the application of heat is called
pyrolysis or cracking.
vi. Aromatization; n-Alkanes having six or more carbon atoms on heating to 773K at 10-20
atmospheric pressure in the presence of oxides of vanadium, molybdenum or chromium
supported over alumina get dehydrogenated and cyclised to benzene and its homologues. This
reaction is known as aromatization or reforming.

vii. Reaction with steam


Methane reacts with steam at 1273 K in the presence of nickel catalyst to form carbon monoxide
and dihydrogen. This method is used for industrial preparation of dihydrogen gas.

CYCLOPARAFFINS (NAPHTHENES)
The cycloalkanes (naphthenes), are saturated hydrocarbons which have one or more carbon
rings to which hydrogen atoms are attached according to the formula CnH2n. Naphthenes or
cycloalkanes such as cyclopropane, methyl cyclohexane are also present in the crude oil. These
compounds are not aromatic and hence do not contribute much to the octane number. Therefore,
in the reforming reaction, these compounds are targeted to generate aromatics which have higher
octane numbers than the naphthenes.
Cycloalkanes have two fewer hydrogen atoms than alkanes, because another carbon-carbon
bond is needed to form the ring. Cycloalkanes are drawn as simple polygons in which the sides
represent the carbon-carbon bonds. It is understood that each corner of the polygon is a carbon
atom bonded to two hydrogen atoms.
Cycloalkanes have similar properties to alkanes but have higher boiling points.
Saturated cyclic hydrocarbons, normally known as naphthenes, are also part of the hydrocarbon
constituents of crude oils. Their ratio, however, depends on the crude type. The lower members
of naphthenes are cyclopentane, cyclohexane, and their mono-substituted compounds.
They are normally present in the light and the heavy naphtha fractions.
Cyclohexanes, substituted cyclopentanes, and substituted cyclohexanes are important precursors
for aromatic hydrocarbons.
The examples shown here are for three naphthenes of special importance.
If a naphtha fraction contains these compounds, the first two can be converted to benzene, and
the last compound can dehydrogenate to toluene during processing. Dimethylcyclohexanes are
also important precursors.
AROMATICS
Aromatics, so called because of their distinctive perfumed smell, are substances derived from
crude oil and, in small quantities, from coal. Aromatics are hydrocarbons, organic compounds
that consist exclusively of the elements carbon and hydrogen – without which life would not be
possible on Earth.
The aromatic hydrocarbons are unsaturated hydrocarbons which have one or more planar six-
carbon rings called benzene rings, to which hydrogen atoms are attached with the formula CnHn.
Aromatic hydrocarbons, also called arenes, are a unique class of carbon molecules in which
carbon atoms are connected by alternating double and single bonds.
They tend to burn with a sooty flame, and many have a sweet aroma. Some are carcinogenic.
Aromatics such as benzene, toluene o/m/p-xylene are also available in the crude oil. These
contribute towards higher octane number products and the target is to maximize their quantity in
a refinery process. Lower members of aromatic compounds are present in small amounts in
crude oils and light petroleum fractions.
The simplest mononuclear aromatic compound is benzene (C6H6). Toluene (C7H8) and xylene
(C8H10) are also mononuclear aromatic compounds found in variable amounts in crude oils.
Benzene, toluene, and xylenes (BTX) are important petrochemical intermediates as well as
valuable gasoline components.

How are aromatics produced?


Benzene, toluene, and xylenes can be made by various processes. However, most BTX
production is based on the recovery of aromatics derived from the catalytic reforming of naphtha
in a petroleum refinery. ... The BTX aromatics can be extracted from catalytic reformate or from
pyrolysis gasoline by many different methods
Separating BTX aromatics from crude oil distillates is not feasible because they are present in
low concentrations. Enriching a naphtha fraction with these aromatics is possible through a
catalytic reforming process.
Binuclear aromatic hydrocarbons are found in heavier fractions than naphtha. Trinuclear and
polynuclear aromatic hydrocarbons, in combination with heterocyclic compounds, are major
constituents of heavy crudes and crude residues. Asphaltenes are a complex mixture of aromatic
and heterocyclic compounds.

Benzene (C6H6)
The structure of benzene can be drawn in two ways. In the first, the double bond character is
explicitly drawn. In the shorthand version, a circle is simply drawn inside the ring to demonstrate
the structure. Each carbon atom in benzene has single hydrogen attached to it. In the diagram, the
hydrogen has been omitted for clarity.

Basic Properties
Like other hydrocarbons, benzene is a natural component of petroleum. It is a colorless,
flammable, sweet-smelling liquid at room temperature and is a component of most gasoline
mixes as it has a high octane number. Benzene is also highly carcinogenic and is well-known to
cause bone marrow failure and bone cancer. Of course, its carcinogenicity was not well known
when it was being used as an additive in after shave and other cosmetics due to its ―pleasant
aroma.‖
Cancer
What makes benzene carcinogenic is its interaction with DNA. It breaks the bonds between
subunits of DNA in the body, which in turn causes cells to either die or reproduce without
normal controls (cancer). It is general consensus that no amount of benzene is safe, but this has
not stopped its use in many industrial and laboratory applications where is properties have yet to
be supplanted by another substance. Toluene is used where possible as a less carcinogenic and
less toxic alternative to benzene.
Currently, the largest exposures to benzene occur through smoking as it is a byproduct of
burning the tar found in cigarettes. The second largest exposure is in gasoline service stations,
which, when combined with industrial emissions and the emissions from car exhaust, accounts
for 20% of all atmospheric benzene.
Uses
The largest use of benzene (50%) is in the product of styrene and polystyren plastics. It is also
converted to a molecule known as cyclohexane, which is important in the production of Nylon.
About 15% of benzene is used to produce cyclohexane. Smaller amounts are used in everything
from pesticides to rubber to pharmaceuticals.
Impurities
All sorts of cats and dogs can turn up in a hydrocarbon reservoir, whether it contains oil or gas or
both. Along with natural gas and its natural gas liquid constituents, other gases can present
themselves (table 3-1). Carbon dioxide (C02), oxygen (02), nitrogen (N2), and hydrogen sulfide
(H2S) are the most common. Not only is hydrogen sulfide a lethal gas in very small
concentrations, but burning has it created a pollutant. That gives two good reasons why
environmental laws and safety considerations require it be removed. Excess carbon dioxide has
to be eliminated because it can cause corrosion of transportation and processing equipment.
A gas stream can contain minor amounts of the rare gases-argon, helium, neon, and xenon.
Sometimes they are in sufficient quantities to make separation and recovery commercially
attractive.
Crude oils can have the same contaminant problems. Most serious for crude oil is the presence of
sulfur and metals. Sulfur can reside in crude oil in the form of dissolved hydrogen sulfide, or it
can be sulfur atoms chemically attached to the hydrocarbon molecules like thiophene (CsHsS).
Depending on the country, environmental regulations require almost all the sulfur be removed
from refined products before they can be sold. Special units in refineries are required in order to
remove it, and as a result, the more sulfur there is in the crude, the lower will be its value to
refiners. Metals such as vanadium, nickel, and copper, can damage the catalysts used to process a
crude oil in refineries and can also cause debits to the values of crude oils so contaminated.
Petroleum Contaminants
Though sulfur is not a major component of the atmosphere, it is often found in petroleum. In
fact, sour versus sweet petroleum is determined by the sulfur content where sour petroleum
contains more than 0.5% sulfur. Sulfur, when burned during hydrocarbon combustion produces
sulfur dioxide, which acts as a precursor to sulfuric acid. Like nitric acid, this contributes to acid
rain. We can once again update our hydrocarbon combustion reaction to reflect this new
contaminant (reaction not balanced and H2S is not the only sulfur contaminant. Others include
COS, CS2, SO2 and more).

The reaction that produces sulfuric acid from sulfur dioxide is more complicated than that which
governs nitric oxide production. In general, the reaction proceeds as follows (note that a dot
indicates an extra electron on a compound, making it a radical in chemistry lingo).

Soot and Smoke


Soot and smoke both refer to particulate matter that gets trapped in gases during combustion.
The visible, dark black component of smoke is carbon that has incompletely burned and, rather
than forming CO2, has formed solid carbon compounds known as amorphous carbon. These
carbon compounds are collectively referred to as soot. Diesel exhaust accounts for 25% of all
smoke and soot in the atmosphere. This is generally the result of the lower quality of fuel (less
refined) that constitutes diesel fuel. Soot, in particular, contains large amounts of polycyclic
aromatic hydrocarbons, which are well-known mutagens and carcinogens. Diesel exhaust is
considered the most carcinogenic of fossil fuels used in transportation.

NON-HYDROCARBONS
A brief reference has already been made to the non-hydrocarbons that may occur in crude oils
and oil products. Although small in quantity, some of them have a considerable influence on
product quality. In many cases they have noxious or harmful effects and must be removed, or
converted to less harmful compounds, by refining processes. In a few cases their presence is
beneficial and they should not be removed or converted.
The most important elements occurring in non-hydrocarbons are sulphur (S), nitrogen ( N ) or
oxygen (O); in some crude oils there are small amounts of metal compounds, of vanadium (V),
nickel (Ni), sodium (Na) or potassium (K) for example.
Sulphur compounds
Many types of sulphur compounds occur in crude oils in widely varying amounts from less than
0.2 per cent by weight in some Pennsylvanian, Algerian and Russian crudes to over 6 per cent by
weight in some Mexican a n d Middle East crudes.
A distinction is often made between corrosive and non-corrosive sulphur compounds. The
corrosive ones are free sulphur, hydrogen sulphide and thiols (mercaptans) of low molecular
weight. Moreover, they have an obnoxious smell.
Hydrogen sulphide, H2S, has the structure H-S-H. If one of the hydrogen atoms is replaced by a
hydrocarbon group, the compound is called a mercaptan or thiol, for example:

The compounds are formed during the distillation of crude oils; they may cause severe corrosion
of the processing units, and addition of chemicals, proper temperature control and the application
of special alloys in plant equipment are required to control them.
The non-corrosive sulphur compounds are sulphides (thioethers), disulphides and thiophenes. If
both of the two hydrogen atoms in hydrogen sulphide are replaced by hydrocarbon groups, the
compound is called a sulphide or thioether, for example:

The disulphides are formed either from mercaptans by oxidation or from sulphides and sulphur:

Thiophenes are sulphur compounds with a ring structure containing five atoms:
The non-corrosive sulphur compounds, although not directly corrosive, may cause corrosion on
decomposition at higher temperatures and therefore also require careful temperature control in
processing units.
Apart from their unpleasant smell, b o t h corrosive a n d non-corrosive sulphur compounds are
undesirable in most products. In fuels, the sulphur burns to sulphur dioxide and sulphur trioxide;
these oxides combine with the water formed by combustion to give sulphurous and sulphuric acids,
which may cause serious corrosion in the colder parts of engines or furnaces. Furthermore, some
sulphur compounds reduce the effect of anti-knock additives (tetraethyllead and tetramethyllead)
on the octane rating of gasolines. Sulphur compounds in illuminating kerosine promote charring
of the wick and cause a bluish white deposit on the lamp glass. In dry-cleaning solvents they may
give a bad odour to cleaned goods and in paint thinners may affect the colour of the dried film.
Some natural gases have a high content of hydrogen sulphide; that from Lacqin France contains
15 per cent by volume, and in Canada there are wells producing natural gas with even 32 per
cent by volume.
The lower thiols are insoluble in water, but soluble in hydrocarbons, and have an intolerable
odour. They react with sodium and copper to form sodium and copper mercaptides and with
oxygen to form disulphides.
Thioethers or sulphides are also insoluble in water, but soluble in hydrocarbons, and have an
offensive odour. However, because of their relatively unreactive nature, drastic treatment is
necessary for their removal. Disulphides are more reactive than thioethers, on account of the S-S
linkage, and can readily be oxidised to compounds soluble in water. Thiophenes have a pleasant
odour, comparable with that of benzene, and are relatively stable; they may even be beneficial.

Nitrogen compounds
Most crude oils contain less than 0.1 per cent by weight of nitrogen. The nitrogen compounds in
the crude are complex and for the most part unidentified, but on distillation they give/rise to
nitrogen bases (compounds of pyridine, a six-membered nitrogen-containing ring) in the derived
products.
Nitrogen bases often cause discoloration of heavy gasolines and kerosines, particularly when
associated with phenols. In gasolines they may also cause engine fouling and in lubricating oils
engine "lacquer". In heavy gas oil feedstocks for catalytic cracking they may reduce the activity
of the catalyst by increasing coke deposits. Nitrogen bases can be removed by acid treatment and
recovered by neutralisation of the acid extract.

Oxygen compounds
Some crude oils contain oxygen compounds. Their structure has not yet been established, but on
distillation of the crudes the oxygen compounds decompose to form ring compounds with a
carboxylic acid group (COOH), in the side chain, for

These compounds are known as "naphthenic acids", large quantities having been originally found
in distillation products of Russian naphthenic crudes. The carboxylic acid group(s) may,
however, be attached to hydrocarbon groups other than naphthenes, and "petroleum acids" would
be a more accurate term; however, "naphthenic acids" is generally accepted. Some of these acids
are highly corrosive a n d special alloys have t o b e used in processing equipment.
Naphthenic acids are extracted from distillates by alkali treatment, either during distillation or
afterwards, and are recovered by acidifying the extract. They are valuable by-products used in
the manufacture of paint-driers, emulsifiers and cheap soaps.
Phenolic compounds occur in some crude and are formed during cracking. They are oxygen
compounds containing one or more OH groups, derived from aromatic hydrocarbons. The
simplest members are phenol, the cresols and the xylenols, which are recovered during refining:
Other compounds
Several other elements occur in crude oils, either as inorganic or organic compounds, and remain
in the ash on burning. They vary from crude to crude, but many types of crude contain vanadium
and nickel. Sodium and potassium are usually present, derived from saline water produced
together with oil. Copper, zinc and iron are also found. These elements are generally of little
account, but sometimes they are important e.g. vanadium is recovered as vanadium ashes from
deposits on furnace walls, or from flue gases, when high vanadium fuels are burnt in refinery
furnaces. Vanadium metal is an important component for the manufacture of special steels.
Vanadium, iron and nickel in the feedstocks for catalytic cracking may spoil catalyst activity,
and so the feedstocks have to be carefully distilled or redistilled to leave the metal compounds in the
residue.

HYDROCARBON REACTIONS
Of the four main groups of hydrocarbons (paraffins, olefins, naphthenes and aromatics), the
olefins are the most reactive a n d the paraffins the least. In the refining of crude oil and in the
manufacture of petrochemicals, certain basic reactions play an important role. Some of them are
also of interest in connection with the performance properties of oil products, e.g. in the
deterioration of gasoline and lubricating oils through oxidation and polymerization.
The following are t h e most i m p o r t a n t of these reactions:

Dehydrogenation — the elimination of hydrogen atoms from a molecule. A saturated


hydrocarbon becomes unsaturated, and a chemical substance changes its type:
Hydrogenation — the reverse process to dehydrogenation; the filling up of the "free" places or
double bonds in unsaturated structures by hydrogen atoms (addition):

Cracking — disruption of the carbon-carbon bonds in large hydrocarbon molecules by heat, so


that smaller molecules (both saturated and unsaturated) are obtained:

Pyrolysis — a severe form of thermal cracking; the disruption reaction is usually accompanied
by a rearrangement of the fragments:
Isomerization — the rearrangement of the carbon skeleton of a molecule, conversion of a
straight chain into a branched chain and the reverse:

Cyclisation — conversion of a chain into a ring molecule, hydrogen being lost:

Alkylation — the introduction of a straight- or branched-chain hydrocarbon group, into an


aromatic or branched-chain hydrocarbon:
Polymerization and copolymerization — the combination of a number of unsaturated
molecules of the same or different compounds to form a single large molecule, called a polymer
or homopolymer when it is built up from a number of identical monomers, and a copolymer
when it is a combination of two or more different types:

Polymers are often solids (such as plastics a n d synthetic fibres), the properties of which depend
largely on their molecular size.

Oxidation — the reactions of oxygen with a molecule that may or may not already contain
oxygen. Oxidation may be partial, resulting in the incorporation of oxygen into the molecule or
in the elimination of hydrogen from it, or it may be complete, forming c a r b o n dioxide a n d water
(combustion):
Chlorination — in the reaction of a saturated hydrocarbon with chlorine one or more of the
hydrogen atoms may be replaced by chlorine atoms with the formation of hydrochloric acid. The
replacement of hydrogen by another atom in this way is called substitution:

In the reaction of an unsaturated hydrocarbon with chlorine, two chlorine atoms are directly
attached to the double bond. This is known as an addition reaction:

Hydration — the addition of water to a double bond without breakdown of the molecular
structure:

Dehydration — the reverse process in the chemical field:


However, in oil manufacturing the term is also used for simple drying of a product (elimination
of dissolved or emulsified water).

Esterification — the reaction of an alcohol with an organic or mineral acid with elimination of
water to form an ester:

Hydrolysis — the decomposition of a molecular structure by the action of water. The hydrolysis
of an ester results in the formation of an alcohol and an acid, and is the reverse of esterification:

Condensation — the coupling of organic molecules accompanied by the separation of water or


some other simple substance, e.g. alcohol. A catalyst is usually required to promote the reaction:

Sulphonation — the action of concentrated sulphuric acid on an aromatic hydrocarbon, e.g.


benzene, to form a sulphonic acid. The hydrocarbon group in a sulphonic acid is directly linked to
the sulphur atom:
Sulphation — the reaction of an olefin with sulphuric acid. An ester is produced by addition of
the sulphuric acid to the double bond and the hydrocarbon group is linked to the sulphur atom
through a n oxygen atom:

Hydrodesulphurisation — the elimination of sulphur from sulphur-containing chain molecules


in crudes or distillates by the action of hydrogen under pressure over a catalyst:

Catalysis — the alteration of the rate of a chemical reaction by the presence of a "foreign"
substance (catalyst) that remains unchanged at the end of the reaction, for instance hydrogenation
using metallic platinum or nickel, and the cracking of a hydrocarbon using a silicate.

TYPES OF CRUDE OIL


Crude oils vary widely in appearance and consistency from country to country and from field to
field. They range from yellowish brown, mobile liquids to black, viscous semi-solids. However,
all crude oils consist essentially of hydrocarbons.
Their differences are due to the different proportions of the various molecular types and sizes of
hydrocarbons previously described. One crude oil may contain mostly paraffins, another mostly
naphthenes. Whether paraffinic or naphthenic, one may contain a large quantity of lower
hydrocarbons and be mobile or contain a lot of dissolved gas; another may consist mainly of
higher hydrocarbons and be highly viscous, with little or nodissolved gas.
The nature of the crude governs to a certain extent the nature of the products that can be
manufactured from it and their suitability for special applications. A naphthenic crude will be
more suitable for the production of asphaltic bitumen, a paraffinic crude for wax. A naphthenic
crude, and even more so an aromatic one, will yield lubricating oils whose viscosities are rather
sensitive to temperature. However, modern refining methods permit greater flexibility in their
use of crudes to produce any desired type of product.
Crudes are usually classified into three groups, according to the nature of the hydrocarbons they
contain.

Paraffin-Base Crude Oils


These contain paraffin wax (higher molecular weight paraffins which are solid at room
temperature), but little or no asphaltic (bituminous) matter. They consist mainly of paraffinic
hydrocarbons and usually give good yields of paraffin wax and high-grade lubricating oils.
Asphaltic-Base Crude Oils
These contain little or no paraffin wax, but asphaltic matter is usually present in large
proportions. They consist mainly of naphthenes and yield lubricating oils whose viscosities are
more sensitive to temperature than those from paraffin-base crudes, but which can be made
equivalent to the latter by special refining methods. These crudes are now often referred to as
naphthene-base crude oils.
Mixed-Base Crude Oils
These contain substantial proportions of both paraffin wax and asphaltic matter. Both paraffins
and naphthenes are present, together with a certain proportion of aromatic hydrocarbons. This
classification is a rough-and-ready division into types and should not be used too strictly. Most
crudes exhibit considerable overlapping of the types described and by far the majority is of the
mixed base type.
DISTILLATION
The first step in the manufacture of petroleum products is the separation of crude oil into the
main fractions by distillation. This is the most important process in the refinery, because, in
addition to its use for separation, it plays an important part in refining the products to marketing
specifications.
A main distinguishing feature of the various petroleum products is their volatility, or ability to
vaporise. This is associated with the size of the molecule; in compounds of a similar type, the
larger the molecule, the lower the volatility. At ambient temperatures and pressure, gasoline is a
liquid that vaporises readily, while kerosine and fuel oils are liquids requiring higher
temperatures to vaporize them. Products such as paraffin wax, solid under normal conditions,
require heating to a relatively high temperature before they liquefy and to still higher
temperatures before they vaporise.
Volatility is related to the boiling point; a liquid with a low boiling point is more volatile than
one with a higher boiling point. When a liquid, say water, is heated, the energy of its molecules
increases and more molecules are able to pass through the surface of the liquid into the space
above, i.e. more molecules pass into the vapour state. The pressure in the space above the
surface, normally atmospheric pressure, tends to restrict the formation of vapour, but the
temperature of the liquid determines the number of molecules leaving the surface of the liquid,
and this in turn determines the vapour pressure of the liquid at that temperature.
When the vapour pressure is equal to or slightly higher than atmospheric pressure, vapour forms
freely throughout the whole liquid, as is shown by the disturbance of the liquid surface and the
formation of vapour bubbles in the liquid; the liquid is said to boil. The temperature at which a
pure liquid boils is its boiling point and remains constant until all the liquid has evaporated, an
important characteristic of a pure substance. The boiling point varies with pressure. At normal
atmospheric pressure pure water boils at 100°C (212°F), ethyl alcohol at 78°C (172°F).
Similarly, each of the individual hydrocarbons present in crude oil has its own characteristic
boiling point. The boiling point is lowered by reducing the pressure in the space above the liquid
(by creating a vacuum) and raised by increasing the pressure.
The heat transferred to the liquid in the process of boiling is retained in the vapour (latent heat of
evaporation), and if this heat is removed, the vapour condenses back into the liquid state, giving
off the heat of condensation. This is seen when steam (water vapour) from a kettle of boiling
water strikes a cold surface.

Simple Distillation
The series of operations comprising boiling and condensation is known as distillation. A simple
laboratory distillation apparatus is shown in Figure 5.1. The liquid is boiled in a flask or "still",
the vapour is condensed in a tube or "condenser" surrounded by cold running water, and the
distillate collected in a receiver.
In a mixture of several liquids of different boiling points, each component has its own
characteristic vapour pressure, and the total vapour pressure above the liquid is the sum of the
partial vapour pressures of the components. The mixture boils when the total vapour pressure is
equal to the (external) pressure above the liquid.
When such a mixture is distilled, molecules of each component will vaporise, and the
composition of the vapour phase will depend on the vapour pressures and the concentrations of
the components in the liquid phase. Since the lower-boilingpoint components have the higher
vapour pressures, the distillate will at first be richer in these than in the higher-boiling-point
components, whereas the liquid in the still will have a higher concentration of high-boiling-point
components. As distillation proceeds, the composition of both distillate and residue will change
progressively until all the liquid has been distilled into the receiver.
Boiling starts at a temperature that lies somewhere in the range of the boiling points of the
components and depends on their ratio in the mixture. The initial boiling point (IBP) is defined
as the temperature at which the first drop distils over. The temperature gradually increases during
distillation, and the more volatile components distil over. The liquid becomes richer in higher-
boiling-point components until the last drop of liquid evaporates at the highest temperature, the
final boiling point (FBP).
Fractional Distillation
Using a simple distilling apparatus as described above, it is not possible to effect sharp
separation between the components of a mixture in one distillation. By redistilling the first
portion, a distillate richer in the more volatile components will be obtained, but the yield will be
low, since part of the components always remains in the still. To effect a good separation it is
necessary to modify the apparatus for continuous condensation and redistillation by inserting a
still-head or "fractionating" column between still and condenser, as shown in Figure 5.2.
Some of the vapour from the boiling liquid condenses as a liquid fraction in each bulb of the
column. The condensation of further vapour from the still supplies heat, which re-evaporates the
lighter or lower-boiling-point components from the liquid in the bulbs. These components
condense in the next higher bulb, and so on up the column. As it becomes richer in the heavier,
less volatile and higherboiling-point components, the liquid in the bulbs flows back to the still.
Thus there is a countercurrent flow of vapour and liquid, the vapour ascending the column and
becoming lighter as the heavier components condense, and the liquid descending and becoming
heavier as the lighter components re-evaporate. The vapour passing over the top into the
condenser consists at first of the low-boiling components, and as these are removed the
temperature of the liquid in the still increases steadily and higher-boiling components distil over.
By changing the receiver at intervals, several different fractions are obtained. A fraction
separated in this manner may consist of a relatively pure component from a simple mixture or a
number of components from a complex mixture, depending on the composition of the mixture
distilled and the type of apparatus. This process is called "fractional distillation".

Distillation of Crude Oil


The products obtained by distillation of crude oil do not consist of single hydrocarbons, except in
the case of simple gases such as ethane and propane. Each product fraction contains many
hydrocarbon compounds boiling within a certain range and these can be broadly classified in
order of decreasing volatility into gases, light distillates, middle distillates a n d residue. The
gases consist chiefly of methane, ethane, propane and butane. The first two are utilised as fuel or
petrochemical feedstocks. Propane and butane may also be liquefied by compression and
marketed as liquefied petroleum gas (LPG).
Butane may to some extent be added to motor gasoline. The light distillates comprise fractions
which may be used directly in the blending of motor and aviation gasolines or as catalytic
reforming and petrochemical feedstocks; these fractions are sometimes referred to as tops or
naphtha.
Sub- module 4: Two. Phase behaviour
Phase diagram, dew point and bubble point lines, critical point line. Dew point pressure and
bubble point pressure
8 hours

The process of vaporizing liquids by gas pressure is a phenomenon common in the petroleum
industry but uncommon elsewhere. An understanding of the phase of hydrocarbon mixtures is
necessary for the treating of nature gas–liquid systems. The goal is to predict, when the
composition of the mixture is known, the quantities and compositions of the phases that occur at
equilibrium for any pressure and temperature. An understanding of the behavior and the
nomenclature for pure substances is helpful in introducing the properties of the single-phase
fluids. Knowledge of the nature of simple mixtures provides a background for studying the
behavior of naturally occurring natural gas systems.

PURE SUBSTANCES –VAPOR PRESSURE


Pure substances may occur in vapor, liquid, and solid phases depending upon the temperature
and pressure. When a substance is in a single phase, its temperature T and pressure P define the
volume V of the component. Figure 4-1 in three dimensions is a plot of these three variables (P,
V, T) for a pure substance. The phase behavior is represented by the surface exposed in the
cutaway view. Since three-dimensional figures are more difficult to use than two, pressure-
temperature projections of the intersections of the surfaces (figure 4-2) and pressure-volume
sections at constant temperature (fig. 4-3) are more commonly employed. Diagrams such as
these are used to portray the phase behavior of natural gas constituents.
Figure 4-1 Pressure-volume-temperature diagram for a pure substance

The conditions at which vapor and liquid may coexist are represented in figure 4-1 as the area
HbCdI. The surface is ―ruled‖; that is, the lines generating the surface are all parallel to the base
of the figure. The projection of CbH or CdI on a pressure-temperature plane becomes a single
curve (HC of Fig. 4-2) known as the vapor-pressure curve. The two-phase surface BDHG for
solid and liquid is ruled likewise, and projects as curve in Fig. 4-2. The solid-vapor surface as
HF. Consider methane in the solid state at m, and increase its temperature at constant pressure
P1. At n, liquid begins to form and, upon the addition of heat at constant temperature and
pressure; the methane will become all liquid at o. Further temperature rise accompanied by the
formation of a less dense fluid phase at b, called vapor. Vaporization takes place at constant
temperature until at d it is completed. Temperature rise beyond d to q simply causes the volume
to increase without any phase change.
At the triple point H, all three phases coexist; only at this single temperature and pressure can
one have three phases together at equilibrium for a pure compound.

Figure 4-2: Pressure-temperature diagram.

The specific-volume sections at constant temperature (such as abde of Fig. 4-1) may be plotted
as isotherms (Fig. 4-3). At the critical point C, the specific volumes and densities of vapor and
liquid merge. At this critical temperature and pressure, all other properties of the phase merge.

Phase Behaviour of Hydrocarbon Systems (PV diagram)


Figure 4.3 shows the pressure versus volume per mole weight (specific volume) characteristics
of a typical pure hydrocarbon (e.g. propane). Imagine in the following discussion that all changes
occur isothermally (with no heat flowing either into or out of the fluid) and at the same
temperature. Initially the component is in the liquid phase at 1000 psia, and has a volume of
about 2 ft3/lb.mol. (point A). Expansion of the system (A-B) results in large drops in pressure
with small increases in specific volume, due to the small compressibility of liquids (liquid
hydrocarbons as well as liquid formation waters have small compressibilities that are almost
independent of pressure for the range of pressures encountered in hydrocarbon reservoirs). On
further expansion, a pressure will be attained where the first tiny bubble of gas appears (point B).
This is the bubble point or saturation pressure for a given temperature. Further expansion (B-C)
now occurs at constant pressure with more and more of the liquid turning into the gas phase until
no more fluid remains. The constant pressure at which this occurs is called the vapour pressure
of the fluid at a given temperature. Point C represents the situation where the last tiny drop of
liquid turns into gas, and is called the dew point. Further expansion now takes place in the
vapour phase (C-D). The pistons in Figure 4.3 demonstrate the changes in fluid phase
schematically. It is worth noting that the process A-B-C-D described above during expansion
(reducing the pressure on the piston) is perfectly reversible. If a system is in state D, then
application of pressure to the fluid by applying pressure to the pistons will result in changes
following the curve D-C-B-A.

We can examine the curve in Figure 4.3 for a range of fluid temperatures. If this is done, the
pressure-volume relationships obtained can be plotted on a pressure-volume diagram with the
bubble point and dew point locus also included (Figure 4.4). Note that the bubble point and dew
point curves join together at a point (shown by a dot in Figure 4.4). This is the critical point. The
region under the bubble point/dew point envelope is the region where the vapour phase and
liquid phase can coexist, and hence have an interface (the surface of a liquid drop or of a vapour
bubble). The region above this envelope represents the region where the vapour phase and liquid
phase do not coexist. Thus at any given constant low fluid pressure, reduction of fluid volume
will involve the vapour condensing to a liquid via the two phase region, where both liquid and
vapour coexist. But at a given constant high fluid pressure (higher than the critical point), a
reduction of fluid volume will involve the vapour phase turning into a liquid phase without any
fluid interface being generated (i.e. the vapour becomes denser and denser until it can be
considered as a light liquid). Thus the critical point can also be viewed as the point at which the
properties of the liquid and the gas become indistinguishable (i.e. the gas is so dense that it looks
like a low density liquid and vice versa).
Suppose that we find the bubble points and dew points for a range of different temperatures, and
plot the data on a graph of pressure against temperature. Figure 4.2 shows such a plot.
Note that the dew point and bubble points are always the same for a pure component, so they
plot as a single line until the peak of Figure 4.4 is reached, which is the critical point.

Figure 4.3 PV phase diagram of a pure component


Figure 4.4 PV Phase Behavior of a Pure Component
Figure 4.5 PV Phase Behavior of a Multi-Component system

The behaviour of a hydrocarbon fluid made up of many different hydrocarbon components


shows slightly different behaviour (Figure 4.5). The initial expansion of the liquid is similar to
that for the single component case. Once the bubble point is reached, further expansion does not
occur at constant pressure but is accompanied by a decrease in pressure (vapour pressure) due to
changes in the relative fractional amounts of liquid to gas for each hydrocarbon in the vaporising
mixture. In this case the bubble points and dew points differ, and the resulting pressure-
temperature plot is no longer a straight line but a phase envelope composed of the bubble point
and dew point curves, which now meet at the critical point (Figure 4.6). There are also two other
points on this diagram that are of interest. The cricondenbar, which defines the pressure above
which the two phases cannot exist together whatever the temperature, and the cricondentherm,
which defines the temperature above which the two phases cannot exist together whatever the
pressure. A fluid that exists above the bubble point curve is classified as undersaturated as it
contains no free gas, while a fluid at the bubble point curve or below it is classified as saturated,
and contains free gas.
Figure 4.6 PT phase diagram for a multicomponent system

MIXTURES –RETROGRADE PHENOMENA


Unusual phase changes occur for mixtures which do not occur for pure substances, these
abnormal phenomena are called ―retrograde‖. The behavior of a hydrocarbon mixture is
illustrated by figure 4.7.
The P-T diagram of Fig. 4.7 will be used to illustrate the language used in describing the
behavior of a mixture. Inside the curve ABCDIE, the mixture is in two phases, vapor and liquid.
Outside the border curve, only one phase occurs. The mixture at K is in one phase, and it remains
in a single phase with a pressure reduction at constant temperature until point B is reached. Here,
a bubble of vapor appears at equilibrium; point B is called a bubble point. Further reduction in
pressure causes more vapor to form and, at F, 20% is in the vapor state. Points along the curve
ABC are bubble points for a mixture.
Figure 4.7: phase diagram for a mixture

The mixture at J is in a single phase and it continues as such with pressure increase until point E
is reached. Here, a droplet of liquid or dew appears at equilibrium; the name ―dew point‖ is
given to this condition. As further pressure increase occurs from point E, more liquid is
condensed as indicated. Likewise, a pressure reduction from L to D will cause dew to form.
A pressure reduction on the fluid from M to C causes the single-phase fluid to change abruptly to
approximately 50 per cent vapor and 50 per cent liquid at C. it so happens that the properties of
the vapor along the dew-point curve and those of the liquid along the bubble-point curve merge
at C, and this is the critical point of the mixture. Physically, the critical point is very difficult to
identify, and the point is more readily determined by plotting the bubble point, the dew point,
and per cent liquid lines, and then finding the common point C.
Compressing pure substances in the vapor state to the vapor-pressure curve will cause dew to
form; therefore, the behavior from J to E is to be expected. However, anyone unfamiliar with the
type of phase behavior from L to D to G –the formation of liquid by pressure reduction on a gas
–would be surprised, for this behavior is unlike that of pure substances. The phenomenon of
liquid formation by isothermal expansion of a single-phase fluid is ―retrograde condensation‖.
The expansion from D toward G yields a maximum percentage of liquid at G, and the liquid
vaporizes in a normal manner as the pressure is reduced still further from G to E. since liquid is
vaporized when the mixture at G is compressed to D at constant temperature, this process is
called ―retrograde vaporization‖. The term applies to vaporization of a liquid by the compression
of a vapor in contact with the liquid. The terms ―retrograde condensation‖ and ―retrograde
vaporization‖ may also be applied to isobaric phenomena when the phase change occurring is
opposite to that which would occur if the same change in conditions were applied to a pure
substance. The area represented by CDIGC in figure 4.7 is the only part of the diagram where
retrograde phenomena occur.
Sub- module 5: Oil and Gas systems
Main types of gas systems: wet, dry and gas condensate; oil system
8 hours

Hydrocarbon reservoirs are usually classified into the following five main types
· Dry gas
· Wet gas
· Gas condensate
· Volatile oil
· Black oil
Each of these reservoirs can be understood in terms of its phase envelope. The schematic
diagram of their PT phase envelopes is shown in Figure 5.1.

Figure 5.1 PT curves for different types of hydrocarbon reservoir

Dry Gas Reservoirs


A typical dry gas reservoir is shown in Figure 5.2. The reservoir temperature is well above the
cricondentherm. During production the fluids are reduced in temperature and pressure. The
temperature-pressure path followed during production does not penetrate the phase envelope,
resulting in the production of gas at the surface with no associated liquid phase. Clearly, it would
be possible to produce some liquids if the pressure is maintained at a higher level. In practice, the
stock tank pressures are usually high enough for some liquids to be produced (Figure 5.3).

Wet Gas Reservoirs


A typical wet gas reservoir is shown in Figure 5.3. The reservoir temperature is just above the
cricondentherm. During production the fluids are reduced in temperature and pressure. The
temperature-pressure path followed during production just penetrates the phase envelope,
resulting in the production of gas at the surface with a small associated liquid phase. The GOR
(gas-oil ratio) has fallen as some liquid is being produced.

Fig 5.2 PT phase diagram for dry gas Fig 5.3 PT phase diagram for wet gas

Gas Condensate Reservoirs


A typical gas condensate reservoir is shown in Figure 5.4. The reservoir temperature is such that
it falls between the temperature of the critical point and the cricondentherm. The production path
then has a complex history. Initially, the fluids are in an indeterminate vapour phase, and the
vapour expands as the pressure and temperature drop. This occurs until the dewpoint line is
reached, whereupon increasing amounts of liquids are condensed from the vapour phase. If the
pressures and temperatures reduce further, the condensed liquid may reevaporate, although
sufficiently low pressures and temperatures may not be available for this to happen. If this
occurs, the process is called isothermal retrograde condensation. Isobaric retrograde
condensation also exists as a scientific phenomenon, but does not occur in the predominantly
isothermal conditions of hydrocarbon reservoirs. Thus, in gas condensate reservoirs, the oil
produced at the surface results from a vapour existing in the reservoir. The GOR has decreased
significantly, the OGR has increased, and more liquid is present in this mixture.

Figure 5.4 PT phase diagram for a gas condensate

Volatile Oil Reservoirs


A typical volatile oil reservoir is shown in Figure 5.5. The reservoir PT conditions place it inside
the phase envelope, with a liquid oil phase existing in equilibrium with a vapour phase having
gas condensate compositions. The production path results in small amounts of further
condensation, and re-evaporation can occur again, but should be avoided as much as possible by
keeping the stock tank pressure as high as possible. The fraction of gases is reduced, and the
fraction of denser liquid hydrocarbon liquids is increased, compared with the previously
discussed reservoir types. Changes in the GOR, OGR and specific gravities are in agreement
with the general trend.

Black Oil Reservoirs


A typical gas condensate reservoir is shown in Figure 5.6. The reservoir temperature is much
lower than the temperature of the critical point of the system, and at pressures above the
cricondenbar. Thus, the hydrocarbon in the reservoir exists as a liquid at depth. The production
path first involves a reduction in pressure with only small amounts of expansion in the liquid
phase. Once the bubble point line is reached, gas begins to come out of solution and continues to
do so until the stock tank is reached. The composition of this gas changes very little along the
production path, is relatively lean, and is not usually of economic importance when produced.
The produced hydrocarbon fluid that is now dominated by heavy hydrocarbon liquids, with most
of the produced gas present as methane. The GOR, OGR and specific gravities mirror the fluid
composition.

Fig 5.5 PT phase diagram for a volatile oil Fig 5.6 PT phase diagram for black oil
Sub-module 6: Properties of oil and gas
Viscosity, formation volume factor and gas/ oil ration: rock properties: porosity permeability
saturation
12 hours

Rock Properties: Porosity Permeability Saturation


The nature of the reservoir rock bearing petroleum determines the fluid content and the ability of
the fluids to move through the rock. Oil and gas are contained in the pores and fractures of
sandstones, dolomite, limestones, chert, chalk and other porous media. To be more inclusive, and
consider rocks such as shales, evaporates, and diatomites because these provide the seals,
bounding materials, or source rocks to our reservoirs. It is important to note that shales and
claystones make up the most abundant rock type in the typical sedimentary column.
Some rock properties include; porosity, permeability, retaining capacity for water under capillary
equilibrium and the tortuous nature of the interstices. Water is believed to have been present
when the petroleum accumulated in most reservoirs. In the accumulation, the oil and gas did not
completely replace the water adjacent to the solid surface. This interstitial water found with
petroleum in the rock is called connate water.
Cores and cuttings taken during the drilling of wells are analyzed in the laboratory to determine
porosity and permeability, and sometimes oil content, water content, and other properties. The
connate-water content may be estimated by determining the water content as a function of
capillary pressure. Electrical logs of wells reveal the properties of the surrounding rock.

Density and Porosity


Density; Density is defined as the mass per volume of a substance.

Typically, with units of g/cm3 or kg/m3. Other units that might be encountered are lbm/gallon or
lbm/ft3
For rocks, porosity (Φ) is defined as the nonsolid or pore-volume fraction.
Porosity is a volume ratio and thus dimensionless, and is usually reported as a fraction or
percent.

Several volume definitions are required to describe porosity:


Total volume of rock = VT or Vrx
Volume of mineral phase = Vg or Vm
Volume of pores or openings = Vpor
Volume of interconnected pores = Vp − con
Volume of isolated pores = Vp − iso
Volume of cracks or fractures = Vf x orVc x
Volume of fluid Phase 1, 2 = Vf 1, Vf 2, etc.

From these we can define the various kinds of porosity encountered:

Total porosity Φ = Vpor /VT

Effective porosity Φ p − e = Vp − con /VT

Ineffective porosity Φ p − iso = Vp − iso /VT

Crack or fracture porosity Φ f x = Vf x /VT

Similarly, the definitions of the standard densities associated with rocks then follows:
Grain density ρg = Ms /Vg
Dry density ρd = Md /VT
Saturated density ρsat = Msat /VT
Buoyant density ρb = Mb /VT
Fluid density ρ f l = Ms /VT
Where, Ms, Md, Msat, Mb, and Mfl are the mass of the solid, dry rock, saturated rock, buoyant rock,
and fluid, respectively.
Effective porosity is defined as the ratio (in per cent) of connected pore volume to total volume.
Absolute porosity is defined as the ratio of total void volume to bulk volume.
Determination of porosity
The per cent porosity of an extracted dry rock or core specimen is 100 times the void divided by
the bulk volume. The problem is to determine the volume of the rock solids and of the bulk
volume. The volume of rock solids may be determined by
(1) Weighing the dry core plug;
(2) Placing the plug in the vessel to evacuate the air;
(3) Admitting a liquid of known density (tetrachlorethane) to submerge the plug and
returning atmospheric pressure to the vessel; and
(4) Weighing the saturated plug, taking care to remove liquid draining from the outside the
plug when removing it from the liquid.
The bulk volume may be determined by finding the liquid displaced in a pycnometer with
mercury, water, or a heavy organic liquid such as tetrachlorethane.

The porosity is calculated as follows:

Example Determination
Weight of dry core =19.810 grams
Weight of core saturated with tetrachlorethane=22.413 grams
Weight of liquid in pores=2.603 grams
Density of tetrachlorethane=1.600 grams/cu cm
Volume of core specimen=9.05 cu cm

Alternatively definition and formula of porosity

ROCK POROSITY
Rock porosity is a measure of the pore space available for the storage of fluids in rock.
In general form:

Where, porosity is expressed in fraction


Vb = (VP + Vm)
Vb = Bulk volume of the reservoir
VP = Pore volume
Vm = Matrix volume

Classification of rocks porosity


a) Primary (original) Porosity: These are developed at time of deposition of the
particles/debris.
b) Secondary Porosity: They are developed as a result of geologic process occurring after
deposition.

c)

d)

Factors affecting porosity


Factors include:
 Particle shape: Porosity increases as particle uniformity decreases.
 Packing Arrangement: Porosity decreases as compaction increases.
 Particle Size Distribution: Porosity decreases as the range of particle size increases.
 Interstitial and Cementing Material: Porosity decreases as the amount of interstitial and
cementing material increases. Clean sand-little interstitial material, while Shaly sand –has
more interstitial material.
 Vugs, Fractures: Contribute substantially to the volume of pore spaces and highly
variable in size and distribution. There could be two or more systems of pore openings-
extremely complex.

Calculations of porosity
Table of matrix densities
Lithology (g/cm3)
Quartz 2.65
Limestone 2.71
Dolomite 2.87

Laboratory measurement of porosity


Conventional core analysis: measure any two;
 Bulk volume, Vb,
 Matrix volume, Vm and
 Pore volume, Vp

Measurement of bulk volume: can be determined by:


1) Calculating from dimensions or
2) Displacement method

(a) Volumetric (measure volume) the volume of the sample is measured.


 Drop the sample into liquid and observe volume charge of liquid
 Must prevent test liquid from entering pores space of sample by coat with paraffin or pre-
saturate sample with test liquid (use mercury as test liquid).

(b) Gravimetric (measure mass).


 Change in weight of immersed sample and prevent test liquid from entering pore space.
 Change in weight of container and test fluid when sample is introduced.

Boyle's Law: P1V1 = P2V2


The volume and pressure of a gas P1 and V2 is fixed in a chamber 1, the sample (core) is put into
the second chamber and the valve is opened.

V2 = Volumetric of first chamber & volume of second chamber-matrix volume or core


(calculated)
VT = Volume of first chamber + volume second chamber (known)
Vm = VT-V2

Determining Pore volume (Vp)


(1) Gravimetric

V2 = Volume of first chamber + pore volume of core (calculated)


Vp = V2-V1

The dry sample coated with paraffin weighed 20.9gm. The paraffin coated sample displaced
10.9cc of liquid. Assume the density of solid paraffin is 0.9gm/cc. What is the bulk volume of
the sample?
Solution:
Weight of paraffin coating = 20.9gm -20.0gm = 0.9gm
Volume of paraffin coating =1.0cc

Bulk volume of sample = 10.9cc-1.0cc = 9.9cc


The core sample of problem above was stripped of the paraffin coat, crushed to grain size, and
immersed in a Russell tube. The volume of the grains was 7.7cc. What was the porosity of the
sample? Is this effective or total porosity?
Solution:
Bulk Volume = 9.9cc
Matrix Volume = 7.7cc

It is total porosity.

Calculate the porosity of a core sample when the following information is available:
 Dry weight of sample = 427.3gm
 Weight of sample when saturated with water =448.6gm
 Density of water = 1.0gm/cm3
 Weight of water saturated sample immersed in water = 269.6gm

What is the lithology of the sample?

The lithology is limestone.


Is the porosity effective or total? Why?
Effective, because fluid was forced into the pore space

A carbonate whole core (3 inches by 6 inches, 695cc) is placed in cell two of a Boyle’s Law
device. Each of the cells has a volume of 1,000cc. Cell one is pressured to 50.0psig. Cell two is
evacuated. The cells are connected and the resulting pressure is 28.1psig. Calculate the porosity
of the core.

Solution

Permeability
Permeability is the ability of a rock to serve as a conduit for the flow of fluids. The capacity to
flow fluids is one of the most important properties of reservoir rocks. Permeability (k) is a rock
property relating the flow per unit area to the hydraulic gradient by Darcy’s law,

⁄ ⁄
Where, p is pressure, ρ is fluid density, g is gravitational acceleration, z is elevation, and μ is the
dynamic viscosity. The ratio q/A has the units of velocity and is sometimes referred to as the
―Darcy velocity‖ to distinguish it from the localized velocity of flow within pore channels.
Reservoir engineers use relative permeability and capillary pressure relationships for estimating
the amount of oil and gas in a reservoir and for predicting the capacity for flow of oil, water, and
gas throughout the life of the reservoir. Relative permeabilities and capillary pressure are
complex functions of the structure and chemistry of the fluids and solids in a producing
reservoir. As a result, they can vary from place to place in a reservoir.
Before defining relative permeability and capillary pressure, let us briefly review the definition
of permeability. Permeability represents the capacity for flow through porous material. It is
defined by Darcy’s law (without gravitational effects) as

Darcy’s law relates the flow rate q to the permeability k, cross-sectional area A, viscosity μ,
pressure drop ΔP, and length L of the material. High permeability corresponds to increased
capacity for flow. The dimensions of permeability are length squared, often expressed as darcies
(1 darcy = 0.987×10–8 cm2), millidarcies, or micrometers squared. Some writers use ―absolute
permeability‖ or ―intrinsic permeability‖ in place of permeability.
Capillary pressure relationships are dimensional functions that range from large negative to large
positive values.
Capillary pressure is often defined as the pressure of the less-dense phase minus the pressure of
the more-dense phase.
Saturation is the fraction of pore space that is occupied by a phase.
Viscosity
Knowledge of viscosity of hydrocarbon fluids is essential for a study of dynamic or flow
behavior of these fluids through pipes, porous media, or, more generally, wherever transport of
momentum occurs in fluid motion. The influence of fluid viscosity on flow is especially
important in petroleum reservoirs, since the flow is predominantly in the laminar-flow region
where the pressure drop is proportional to the viscosity. In the laminar or viscous region,
viscosity is defined by the relationship

Where, μ = coefficient of viscosity, or absolute viscosity, grams mass/ (cm) (sec)


ω = shear stress per unit area in the shear plane parallel to the direction of flow, grams
force/ sq cm

⁄ Velocity gradient perpendicular to the plane of shear, cm/ (sec) (cm)

mass/ gram force) (cm/sec2)

The unit of absolute viscosity is the gram mass / (cm) (sec), or the poise. Other units are the
centipoise, or 0.01 poise; the millipoise, or 0.001 poise; and the micropoise, or 0.000001 poise.
Water at 68.4oF has a viscosity of 1.0 centipoise.
Viscosity measurements on liquids flowing through capillaries under the driving force of the
head of liquid yield kinematic viscosities.
The rolling-ball viscosimeter is used for measuring viscosity, with the ball forcing the fluid
through the crescent between the ball and the tube wall.

Fig.6.4 Principle of rolling ball viscosimeter

Viscosity of liquids
Liquids decrease in viscosity with increased temperature and increase in viscosity with increased
pressure.

Gas Density and Formation Volume Factor


The formation volume factor of gas is defined as the ratio of the volume of gas at the reservoir
temperature and pressure to the volume at the standard or surface temperature and pressure (ps
and Ts). It is given the symbol Bg and is often expressed in either cubic feet of reservoir volume
per standard cubic foot of gas or barrels of reservoir volume per standard cubic foot of gas. The
gas-deviation factor is unity at standard conditions; hence, the equation for the gas formation
volume factor can be calculated using the real gas equation:

The density of a reservoir gas is defined as the mass of the gas divided by its reservoir volume,
so it can also be derived and calculated from the real-gas law:

Isothermal Compressibility of Gases


The isothermal gas compressibility, cg, is a useful concept that is used extensively in determining
the compressible properties of the reservoir. The isothermal compressibility is also the reciprocal
of the bulk modulus of elasticity. Gas usually is the most compressible medium in the reservoir;
however, care should be taken so that it is not confused with the gas-deviation factor, z, which is
sometimes called the compressibility factor.
The isothermal gas compressibility is defined as:
Sub-module 7: Initial fluid distribution: free water, free oil level oil water contact, transition
zone, and residual oil
8 hours

Introduction
Reservoir fluids fall into three broad categories; (i) aqueous solutions with dissolved salts, (ii)
liquid hydrocarbons, and (iii) gases (hydrocarbon and non-hydrocarbon). In all cases their
compositions depend upon their source, history, and present thermodynamic conditions. Their
distribution within a given reservoir depends upon the thermodynamic conditions of the reservoir
as well as the petrophysical properties of the rocks and the physical and chemical properties of
the fluids themselves. This chapter briefly examines these reservoir fluid properties.

Fluid Distribution
The distribution of a particular set of reservoir fluids depends not only on the characteristics of
the rock-fluid system now, but also the history of the fluids, and ultimately their source. A list of
factors affecting fluid distribution would be manifold. However, the most important are:
Depth The difference in the density of the fluids results in their separation over time due to
gravity (differential buoyancy).
Fluid Composition The composition of the reservoir fluid has an extremely important control on
its pressure-volume-temperature properties, which define the relative volumes of each fluid in a
reservoir. This subject is a major theme of this chapter. It also affects distribution through the
wettability of the reservoir rocks.
Reservoir Temperature Exerts a major control on the relative volumes of each fluid in a
reservoir.
Fluid Pressure Exerts a major control on the relative volumes of each fluid in a reservoir.
Fluid Migration Different fluids migrate in different ways depending on their density, viscosity,
and the wettability of the rock. The mode of migration helps define the distribution of the fluids
in the reservoir.
Trap-Type Clearly, the effectiveness of the hydrocarbon trap also has a control on fluid
distribution (e.g., cap rocks may be permeable to gas but not to oil).
Rock structure The microstructure of the rock can preferentially accept some fluids and not
others through the operation of wettability contrasts and capillary pressure. In addition, the
common heterogeneity of rock properties results in preferential fluid distributions throughout the
reservoir in all three spatial dimensions.
The fundamental forces that drive, stabilise, or limit fluid movement are:
· Gravity (e.g. causing separation of gas, oil and water in the reservoir column)
· Capillary (e.g. responsible for the retention of water in micro-porosity)
· Molecular diffusion (e.g. small scale flow acting to homogenise fluid compositions within a
given phase)
· Thermal convection (convective movement of all mobile fluids, especially gases)
· Fluid pressure gradients (the major force operating during primary production)
Although each of these forces and factors vary from reservoir to reservoir, and between
lithologies within a reservoir, certain forces are of seminal importance. For example, it is gravity
that ensures, that when all three basic fluids types are present in an uncompartmentalised
reservoir, the order of fluids with increasing depth is GAS:OIL:WATER, in exact analogy to a
bottle of french dressing that has been left to settle.
Microscopic Efficiency of Immiscible Displacement
The conceptual aspects of the displacement of oil by water in reservoir rocks. Fig. 7.1 is a
schematic diagram of the water/oil displacement process. At the pore level (i.e., where the water
and oil phases interact immiscibly when moving from one set of pores to the next), wettability
and pore geometry are the two key considerations.

Fig. 7.1—Saturation profile during a waterflood


Wettability
Wettability is defined in terms of the interaction of two immiscible phases, such as oil and water,
and a solid surface, such as that of the pores of a reservoir rock. For understanding wettability
concepts and for simple laboratory determinations, the solid surface is taken as a smooth flat
surface. Fig. 7.2 illustrates two styles of wettability: water-wet and oilwet.

Fig. 7.2—Wettability of oil/water/solid system

Eq. 7.1 describes the force relationship that is in balance for the drop of water that is on the solid
surface and is surrounded by oil. The interfacial tension (IFT) between the oil and water phases
varies depending on the compositions of the phases but generally is relatively high, in the 10- to
30-dyne/cm range. The contact angle θ is used to define which fluid phase is more wetting—for
low contact angles, the water phase is more wetting, whereas for high contact angles, the oil
phase is more wetting.
Where, σos = the IFT between the oil and solid phases,
σws = the IFT between the water and solid phases, and
σow = the IFT between the oil and water phases.
The particular contact angle depends on many variables, including the composition of the crude
oil and the amount of gas in solution; the salinity and pH of the connate brine; the mineralogy of
the rock surfaces; and the salinity and pH of the injected water that is used for waterflooding.
The concentration of surface-active components (e.g., asphaltenes) that are in the crude oil and
that can adsorb on the rock surfaces affects wettability.
Reservoir rocks typically are described as being water-wet, oil-wet, or intermediate-wet. A
water-wet rock surface is one that has a strong preference to be coated, or ―wetted,‖ by the water
phase, so that there will be a continuous water phase on the rock surfaces. Oil-wet rocks prefer to
be coated with oil instead of water. Strongly oil-wet rocks have been created for laboratory
studies but, as discussed below, are unlikely to exist in real reservoirs. Intermediate-wet reservoir
rocks have been found in several oil reservoirs. The term ―dalmatian wetting‖ describes reservoir
rocks that have both oil-wet and water-wet surfaces. Fig.7.3 illustrates two styles of
intermediate-wetting.

Fig. 7.3—Relationship of mineralogy to wetting conditions: (a) dalmatian wetting and (b)
mixed wetting
Pore Geometry. The pore geometry for any reservoir rock is the result of its depositional and
diagenetic history. The depositional environment determines a rock’s grain size and sorting.
Post-depositional diagenetic changes caused by various types of cementation, leaching, and clay
alteration will impact a rock’s pore characteristics whether the rock is primarily silica or
carbonate.

Mode of Delivery
The module should be taught using lectures, Industrial Visits, experiment and practical work
Assessment
The module is assessed through assignments, tests, practical reports and module examination.
Their relative contribution to the final grade is shown below:
Requirements Contribution
Assignments 5%
Tests 10%
Practical work 25%
Final module examination 60%
Total 100%

Reading List
1. Hunt, J, M, 1995, Petroleum geochemistry and geology, 2nd edition, W.H, Freeman and
Company, New York.
2. Broadhead, R, 2002, The origin of Oil and Gas in New Mexico’s energy : present and future
production, economics, and the environment, BristerB.L. New Mexico Bureau Geol Mineral
Resources.
3. Mike May 2011, investing in oil and gas 5th Edition University of Texas at Austin, Texas.
4. John Low 2011, Oil and Gas in a Nutshell, 5th edition, Nutshell series
5. Debby Denby 2011, Fundamentals of Petroleum, 5th edition, University of Texas at Austin.
6. J.S Archer &C.G Wall 2001, Petroleum Engineering, Principles and Practice-

You might also like