You are on page 1of 156

THE PETROLEUM INDUSTRY

BY

Dr. Mohammad A. Al-Marhoun


Professor Emeritus of Petroleum Engineering

And

Dr. Sidqi A. Abu-Khamsin


Professor of Petroleum Engineering

Department of Petroleum Engineering


King Fahd University of Petroleum and Minerals
Dhahran, Saudi Arabia

August 2015
Contents

CHAPTER 1 - PETROLEUM GEOLOGY 1


1.1 Introduction 1
1.2 The Earth 1
1.3 Types of Rocks 7
1.4 Definition of Petroleum 13
1.5 Origin of Petroleum 15
1.6 Oil and Gas Traps 17
1.7 Geological Structures 20
1.8 Structural Petroleum Traps 23

CHAPTER 2 - PETROLEUM EXPLORATION 25


2.1 Introduction 25
2.2 Geological Methods 25
2.3 Geophysical Methods 28
2.4 Geochemical Methods 35

CHAPTER 3 - DRILLING 39
3.1 Introduction 39
3.2 Rotary Drilling 42
3.3 The Drilling Rig 42
3.4 Drilling Fluids 52
3.5 Casing 59
3.6 Cementing 61
3.7 Well Completion 62
3.8 Rig Types 64
3.9 Rig Personnel 66
3.10 Directional Drilling 67

CHAPTER 4 - RESERVOIR EVALUATION 70


4.1 Introduction 70
4.2 Driller’s Log 70
4.3 Well Logging 70
4.4 Rock and Fluid Sampling 75
4.5 Production Tests 75
4.6 Rock Properties 76
4.7 Fluid Properties 90
4.8 Reservoir Water 93

CHAPTER 5 - PRODUCTION ENGINEERING 97


5.1 Introduction 97
5.2 Downhole Equipment 97
5.3 Wellhead Assembly 99

i
5.4 Natural Flow 101
5.5 Artificial Lifting Methods 101
5.6 Surface Production Facilities 105
5.7 Gas-Oil Separation 105
5.8 Oil and Gas Treatment 108
5.9 Oil and Gas Metering 111
5.10 Crude Oil Gathering and Storage 112
5.11 Transportation 112
5.12 Well Stimulation 113

CHAPTER 6 - PETROLEUM RESERVOIRS 116


6.1 Introduction 116
6.2 Petroleum Reserves 116
6.3 Types of Petroleum Reservoirs 118
6.4 Reservoir Driving Mechanisms 118
6.5 Material Balance 126
6.6 Well Performance 128
6.7 Enhanced Oil Recovery 131

CHAPTER 7 - PETROLEUM REFINING 138


7.1 Introduction 138
7.2 Crude Oil Distillation 139
7.3 Thermal Cracking 141
7.4 Catalytic Cracking 141
7.5 Hydrocracking 143
7.6 Catalytic Reforming 143

CHAPTER 8 - PETROCHEMICALS 146


8.1 Introduction 146
8.2 Basic Petrochemicals 146
8.3 Solvents 148
8.4 Detergents and Surfactants 149
8.5 Polymers 150

NOMENCLATURE 152

ii
Chapter 1

PETROLEUM GEOLOGY

1.1 Introduction
It is often very difficult to know where and how much petroleum could exist under the

surface. This information requires techniques based on different branches of science and

engineering. One such branch is petroleum geology, which is the utilization of geology in the

exploration for and exploitation of petroleum resources. Since geology is the science of the

earth, facts about the earth and the natural petroleum environment need to be reviewed first.

1.2 The Earth


The earth is an ellipsoidal planet whose average radius is about 6370 km. It consists

of the core (3420 km in radius), surrounded by a mantle (2900 km thick), and at the surface of

the earth there remains a crust of the lightest materials (7 to 70 km thick) - Fig. 1.1. These

layers and their subdivisions are described in Table 1.1. The age of the earth is approximated

at about five billion ( 5x109 ) years.

Table 1.1: Layers of the Earth

Layer Contains Depth Interval (Km)

Crust Ocean crust surface to 7-10

Continental crust surface to 20-70

Lithosphere Crust and Upper portion of mantle surface to 100

Mantle Asthenosphere 100 - 350

Remaining upper mantle 350 - 670

Lower mantle 670 - 2900

Core Outer core (liquid) 2900 - 5150

Inner core (solid) 5150 - center of earth

1
Fig. 1.1: The inferred structure of the earth

2
Atmosphere
The earth is surrounded by an atmosphere made up of a mixture of gases whose

composition is given in Table 1.2. The shape and composition of the surface of the earth is

greatly influenced by atmospheric events such as rain, snow, winds, etc. Also, the continuous

existence of living organisms on earth depends on the availability of the necessary gases in

the atmosphere.

Table 1.2: Average Composition of the Gas Mixture in the Atmosphere

Gas Mole %

Nitrogen 78.000

Oxygen 21.000

Carbon Dioxide 0.093

Argon 0.023

Others 0.884

Total 100.000

Hydrosphere
The hydrosphere is the surface of the globe that is covered by water. It encompasses

the oceans, lakes, rivers, and all other water bodies. About 77 percent of the salts contained in

the hydrosphere is Sodium Chloride (NaCl).

Crust of the Earth


Of all the earth's layers, petroleum geology is concerned only with the crust of the

earth since petroleum is never found outside the crust. The crust is so thin that if the earth is

shrunk down to the size of an onion, its crust will be thinner than the skin of the onion. Yet,

all the mountains, oceans, soils and living organisms are in the crust of the earth. Figure 1.2

summarizes the various topographic features of the earth's crust with respect to their

elevations referenced to sea level.

3
The major elements in the rocks of the continental crust vary mainly by location and

depth. However, they consist primarily of oxygen and silicon ( Table 1.3).

Table 1.3: Major Elements of the Earth's Crust

Element Weight % Volume %

Oxygen (O) 46.60 93.77

Silicon (Si) 27.72 0.86

Aluminum (Al) 8.13 0.47

Iron (Fe) 5.00 0.43


Calcium (Ca) 3.63 1.03

Sodium (Na) 2.83 1.32

Potassium (K) 2.59 1.83

Magnesium (Mg) 2.09 0.29

Total 98.59 100.00

Approximate Calculations
Volume of the Earth:

Assuming the earth to be a perfect sphere instead of an ellipsoid:


Ve = 4/3  re3 (1.1)

where
Ve : volume of the earth, m3

re : average radius of the earth  6,370,000 m.

Therefore,
Ve = 4/3  6,370,0003

= 1.08 x 1021 m3 or 1.08 x 1012 km3.

4
Fig. 1.2: Elevation of the earth's crust

5
Volume of the Crust:
Vce = 4/3  [ re3 - ( re - h ce )3 ] (1.2)

where
Vce : volume of the crust of the earth, m3

h ce : average thickness of the crust of the earth  20,000 m.

Therefore,
Vce = 4/3  (6,370,0003 - 6,350,0003)

= 1.02 x 1019 m3 or 1.02 x 1010 km3

which is approximately 1% of the earth's volume.

Surface Area of the Earth


A = 4  re 2 (1.3)

= 4  x 6,370,0002

= 5.10 x 1014 m2

= 510 million km2.

Mass of the Earth


From the universal law of gravitation, the force of attraction between the earth and a

body on the earth's surface is given by:


F = G m me/r2 (1.4)

where
F : force of attraction, N

m : mass of the body, kg


me : mass of the earth, kg

G : universal gravity constant = 6.67 x 10-11 Nm2/kg2

r : distance between the two masses, m.

Since the body is on the surface of the earth, then r is essentially equal to the radius of the
earth, re. Also, the gravitational force is essentially equal to the weight of the body. Since:

F = mg

6
Then,
G = m me/re2 (1.5)

where

g : gravitational acceleration of the earth  9.81 m/s2.

Hence,
me = g re2 / G (1.6)
9.81 m / s2 6,370,000 m 2
2

 
=
6.67 x1011 N m 2 / kg 2
m / s2 . kg 2
= 5.97 x 1024
N
By definition,
1N= 1 kg . m/s2

Therefore,
me = 5.97 x 1024 kg.

Average Density of the Earth


e = me / Ve (1.7)
24
597
. x10 kg
=
. x1021 m 3
108

= 5.53 x 103 kg/m3

= 5.53 g/cm3.

1.3 Types of Rocks


Rocks are classified into three main groups: igneous, sedimentary and metamorphic

rocks. These types are distributed throughout the crust according to the proportions shown in

Fig. 1.3.

Igneous Rocks
These are the magmatic type of rocks. Igneous rocks are produced by the

crystallization of magma, which is the molten fluid generated within the earth. Sometimes,

7
magma erupts to the surface as lava flow and volcanic ash; however, it doesn't always reach

the surface of the earth. It cools under subsurface conditions and forms rocks such as granite

and basalt. During the cooling period, these rocks develop crystalline, shiny and solid

characteristics. Volcanoes are prime sources of surface igneous rocks. Figure 1.4 shows a

volcano with its dikes and sills.

Igneous rocks are not important for hydrocarbon generation or accumulation because

of the very high temperatures of their origin - under such temperatures hydrocarbons cannot

form or exist - and because they usually do not have pore spaces.

Sedimentary Rocks
Sedimentary rocks are formed mainly by the accumulation of sediments, organic

materials or salts. The surface of the earth changes continuously because of erosion caused by

atmospheric events such as rain, snow, tornadoes, wind, waves, and day and night

temperature differences. Rivers and winds transport erosion products - as solutions or as small

rock particles - to various places such as river deltas or sea beds where they settle. These

particles are called sediments and the settling process is called sedimentation (Fig. 1.5). As

this sedimentation process continues, more and more layers accumulate and the older layers

become buried deeper. Due to the heavy loads on top of them, the bottom layers will be

compressed and dehydrated. This process causes the loose sediments to become hard,

compact, and consolidated, forming sedimentary rocks. Sandstone, for example, forms when

sand grains are compressed together over a long period of time.


Examples of rock types which originate from chemical solutions are salt, gypsum and

anhydrite. These minerals precipitate when the water of the solution evaporates. Another type

of sedimentary rock is limestone, the origin of which can be either organic - from the shells

and skeletons of mainly marine organisms - or inorganic - by the following reactions that can

occur in sea water:

CO2  H2O  H2CO3


H2CO3  Ca( OH )2  CaCO3  2H2O

8
Fig. 1.3: Proportions of rock types in the earth’s crust

Fig. 1.4: A volcano system

9
Fig. 1.5: Sedimentation and subsidence processes.

10
Sedimentary rocks are very important to petroleum geologists because within them

hydrocarbon generation and accumulation takes place. They contain almost all the

hydrocarbon deposits discovered so far. Sedimentary rocks are generally classified as clastic,

carbonate, evaporite, organic, or others ( see Table 1.4). However, the most common types

of sedimentary rocks are shale, sandstone and carbonates (limestone and dolomite) ( see Table

1.5 and Fig. 1.6).

Table 1.4: Sedimentary Rocks

Clastic Chemical Organic Others


Carbonate Evaporite

Conglomerate Limestone Gypsum Peat Chert


Sandstone Dolomite Anhydrite Coal

Siltstone Salt Diatomite

Shale Potash Limestone

Table 1.5: Most Common Types of Sedimentary Rocks

Type % Volume of % Volume of

Sedimentary Rocks Earth's Crust

Shale 53 4.2

Sandstone 22 1.7

Limestone and Dolomite 25 2.0

Grains of sedimentary rock are made up of mineral crystals and/or non-crystalline

materials. A mineral is a naturally-occurring crystalline inorganic material with a fixed

chemical composition. Its physical properties, however, may vary within a small range.

Silicates (most abundant rock forming minerals in the crust), carbonates, clays, evaporates,

and oxidates are common minerals that form sedimentary rocks. Detailed information on

these minerals is given in Table 1.7.

11
Fig. 1.6: Proportions of sedimentary rocks.

12
Table 1.7: Sedimentary Rock Forming Minerals

Mineral Chemical Formula Density (g/cm3)


1. Silicates
Quartz (sand) SiO2 2.65

2. Carbonates
Calcite (limestone) CaCO3 2.71
Dolomite CaCO3 MgCO3 2.87

3. Clays
Illite K, Al4 (Si7 Al) O20 (OH)4 2.53
Montmorillonite (Ca, Na)7 (Al, Mg, Fe) 4
(Si, Al)8 O20 (OH)4 (H2O)n 2.12
4. Evaporites
Anhydrite CaSO4 2.96
Gypsum CaSO4(H2O)2 2.32
Halite NaCl 2.17
Sylvite KCl 1.98

5. Oxidates
Hematite Fe2O3 5.27
Magnetite Fe3O4 5.18

Metamorphic Rocks
Metamorphic rocks are generated by the metamorphism of sedimentary and igneous

rocks. High pressure, heat and chemically active fluids change the composition of a rock and

result in a recrystalline structure. This process is called metamorphism, an example of which

is the change of limestone to marble as a result of heat.

The Rock Cycle


Because of continuous geologic processes, rocks can change their form from one type

to another as can be seen in Figs. 1.7 and 1.8.

1.4 Definition of Petroleum


Petroleum is a Greek word derived from petro (rock) and oleum (oil), meaning "rock

oil". It may occur as natural gas, crude oil or asphalt (bitumen) which is in the solid state. It

13
Fig. 1.7: Formation of rocks.

Fig. 1.8: The rock cycle.

14
is a mixture of hundreds of thousands of hydrocarbons and, therefore, has no unique formula.

It could also contain impurities like oxygen, nitrogen, sulfur, and even traces of metals.

1.5 Origin of Petroleum


Two groups of theories exist for the origin of petroleum: the abiogenic and biogenic

groups. Abiogenic theories propose that petroleum was formed from non-biological sources

such as the reaction between water and metallic carbides or between alkali metals, water and

carbon dioxide. Such theories did not gain acceptance because metallic carbides and alkali

metals are not known to exist in the earth's crust where all the petroleum is found.

The biogenic theories attribute petroleum to plant and animal origins. These theories

have become generally accepted due to many forms of chemical, geologic and field evidence.

It is believed that in some marine sedimentation areas, such as ancient river deltas and

shallow seas, accumulating sediments contained a large amount of organic material. The

main source of this material was the remains of dead microscopic plants and animals which

settled to the bottom and got mixed with the sediments. Over millions of years, the organic

material disintegrated under anaerobic conditions (absence of oxygen) into simpler

hydrocarbons as a result of temperature, pressure, bacterial action and radioactivity. The

nature of the produced hydrocarbons was dictated by the temperature, pressure and duration

of burial of the sediments. If the temperature was between 50 and 150 oC ( corresponding to

2 to 3 km depth) oil was usually generated; above 150 oC gas was usually generated. Figure

1.9 shows the distribution of discovered oil and gas deposits according to the depth of burial.
The sedimentary layer (usually shale) within which petroleum is generated is called

the source rock. However, generated petroleum seldom stays within the source rock; rather it

is expelled out to an adjacent porous layer as soon as it becomes fluid enough to move.

Several forces cause this primary migration of petroleum, the strongest of which is buoyancy.

Since the source rock also contains water - which is heavier than oil - the buoyancy force

causes the water to push the oil upwards. Another cause of expulsion is the capillary force

which sucks the water into the oil-occupied pores of the source rock which causes oil.

15
Fig. 1.9: Depths of burial of oil and gas reserves.

16
displacement. Compaction of the source rock under the heavy loads above it also causes oil

expulsion.

Expelled oil and gas usually percolate through the porous strata above the source rock

until a barrier stops them. Sometimes, they travel along a porous layer for hundreds of

kilometers before they are stopped. This second stage of movement is called the secondary

migration of petroleum. If the barrier is large enough, huge quantities of petroleum can

accumulate behind it in what is called a trap. Because of gravity, gas segregates from the oil

and accumulates on top of it; below the oil, the trap is filled with water.

The majority of petroleum deposits in the world are found in rocks belonging to the

Mesozoic era (220 to 65 million years ago), especially the Cretaceous period when climatic

and geographic conditions were very favorable for the generation of petroleum. In eastern

Saudi Arabia, the most important oil-bearing strata are the Arab (Jurassic) and Wasia

(Cretaceous) formations. Table 1.8 shows the positions of these formations relative to the

other sedimentary layers of Saudi Arabia.

1.6 Oil and Gas Traps


A trap is defined as any geological condition which presents an obstacle to migrating

oil or gas. Traps should also have porous and permeable rock for storing commercial

quantities of hydrocarbons. Rocks can store liquids just like a sponge containing water, but

with much smaller pores. The formation of traps is closely related to the movements of the

earth's crust and the formation of geological structures.

Plate Movements
The lithosphere is made up of several large plates of solid rock that cover the surface

of the earth. A plate may contain parts of many continents and oceans as shown in Fig. 1.10.

All lithospheric plates are known to move in a sliding fashion over the semiplastic and ductile

rock of the asthenosphere. This movement is caused by forces produced by convection

currents (currents of molten rock rising from the hot interior to the cold surface of the earth)

within the mantle and the rotation of the earth. In some places, plates may be colliding; in

17
Table 1.8: Sedimentary Sequence of Saudi Arabia

Age Formation Typical Thickness (m)


Surficial deposits
Kharj 28
Hofuf 95
CENOZOIC Dam 91
Hadrukh 84
Dammam 33
Rus 56
Umm er-Rdhuma 243
Aruma 142
Wasia 42
Biyadh 425
Buwaib 18
Yamama 46
Sulay 170
Hith 90
Arab 124
MESOZOIC Jubaila 118
Hanifa 113
Tuwaiq Mountain 203
Dhurma 375
Marrat 103
Minjur 315
Jilh 326
Sudair 116
Khuff 171
Wajid 950
PALEOZOIC Jauf 299
Tabuk 1072
Saq 600

other places, they may be moving away from, sliding by or sinking below each other. Great

geological phenomena occur along the boundaries of plates. These include volcanoes,

earthquakes, faults, folds, and mountain ranges. A prominent example is the Himalayas, the

largest and highest mountain range in the world, which were produced as a result of the

movement of the Australia plate northwards. Surprisingly, the Himalayas are still rising

because the Australia plate continues to press against the Eurasia plate at a rate of about five

centimeters a year. This example shows that the forces responsible for such movements must

18
Fig. 1.10: Crustal plates.

19
be very great. Such geological phenomena are studied under a branch of geology called

"Plate Tectonics".

Formation of Geological Structures


Plate movements cause three types of stresses: (1) compressive stress: which tends to

squeeze the rock and shorten the distance parallel to the direction of compression, (2) tensile

stress: which pulls the rock apart and is opposite of compressive stress, and (3) shear stress:

which acts in parallel but in opposite directions causing one part of a rock mass to slide past

another. These stresses, alone or together, cause the formation of many important geological

features and structures which make suitable petroleum traps.

1.7 Geological Structures


Depending on the crustal layers - whether brittle or flexible - the following structures

are commonly generated as a result of geoforces acting on the rocks.

Folds
If flexible layers are subjected to horizontal compression forces, upward or downward

bending may occur. If the bending is upward, the structure is called an "anticline", and if it is

downward, the structure is called a "syncline" (Fig. 1.11).

Faults
If either tension or compression forces act on brittle layers, fracturing may occur
which leads to separation of the layer into two blocks. Faults are of several types depending

on the movement of the blocks relative to each other. Two of these types are shown in Fig.

1.12.

Domes
A dome is generated when a subsurface body pierces into upper layers. Domes are

convex in shape when observed form the top and the dip (inclination) of the layers is uniform

in all directions. The common causes of piercement of the subsurface layers are magma

20
Fig. 1.11: Compression forces forming anticlines and syncline.

Fig.1.12: Tension forces forming faults.

21
Fig. 1.13: A salt dome formed by a salt diapire.

22
Intrusions and salt diapires. Under high pressure and temperature, a solid salt layer located

underground behaves like a liquid and flows towards the surface - in the form of a salt diapire

- by piercing through the upper layers (Fig. 1.13).

1.8 Structural Petroleum Traps


Several types of petroleum traps have been known, and they are classified into three

categories: structural, stratigraphic and combination traps. Structural traps are the most

important since more than 80% of all petroleum reserves have been found in this type.

Structural traps are associated with one of the geological structures explained in the previous

section. For example, if one layer in a sequence of layers which have been deformed into an

anticline is impermeable - does not allow flow of fluids - migrating oil or gas may collect

below it as shown in Fig. 1.14. Such a trap is called a fold and it is found in most Saudi

Arabian fields.

If the same sequence of layers had been faulted and the two sides slid in opposite

directions such that the imperable layer became facing a permeable layer, oil migrating along

the permeable layer may be trapped at the fault. Fault traps (Fig. 1.15) are common in the

UAE. The salt diapire in a salt dome can act as a seal to all layers pierced by it. Those layers

would then become excellent petroleum traps as shown in Fig. 1.16. A few Omani fields are

typical salt domes.

23
Fig. 1.14: An anticlinal petroleum trap.

Fig. 1.15: A fault petroleum trap.

24
Chapter 2

PETROLEUM EXPLORATION

2.1 Introduction
It is quite difficult to determine where and at what depths under the surface

hydrocarbons are located, and in what quantity. Exploration is the organized and systematic

search for petroleum. Nowadays, techniques based on geology, geophysics and geochemistry

are utilized.

2.2 Geological Methods


Geological Methods basically rely on maps in which surface and subsurface layers are

studied and correlated. Some of the techniques used in mapping are as follows:

Image Processing: Photographs taken from airplanes or satellites are used to make surface

and subsurface maps of the earth.

Contour Maps: A contour map consists of a number of contours. Contours are lines on

which every point has the same property. A structural contour map shows lines of equal

elevation measured from sea level, which is a reference datum for an underground layer as

depicted in Fig. 2.1. There are various types of contour maps among which the following are

the most utilized in petroleum engineering:

Isopacheous map : Lines of the same thickness of a subsurface layer

Isobaric map : Lines of the same pressure

Isothermal maps : Lines of the same temperature

Isoporosity map : Lines of the same porosity

Isopermeability map : Lines of the same permeability.

Cross-sections: A cross section is a profile portraying an interpretation of a vertical section of

the earth (Fig. 2.2).

25
Fig. 2.1: Contour map showing equal elevation above sea level.

26
Fig. 2.2: A cross section inferred from a contour map.

27
2.3 Geophysical Methods
These methods involve the application of the principles of physics and rock mechanics

to the earth. Four major geophysical methods are employed today: seismic, magnetic,

gravitational, and electrical. None of these methods directly detects the presence of

hydrocarbons within the subsurface layers. Instead, they outline prospective sedimentary

layers and traps beneath the surface. These methods can be applied for other purposes as

well, such as locating underground ore bodies, coal deposits, and underground water sources.

However, they are mostly used for petroleum exploration.

Seismic Method
This is the most widely used geophysical method for measuring the thickness and

mapping the structure of subsurface layers. The principle of this method is to generate sound

waves at the surface and record the time it takes these waves to travel down to a subsurface

layer, get reflected by it and arrive back at the surface as shown in Fig. 2.3. Once the sound

velocity in the layer above is known, the distance traveled can be calculated using the

measured travel time according to:

L = vt

where

L = travel distance, m

v = sound velocity, m/s

t = travel time, s.
The distance L can then be used to estimate the depth of that layer.

More than often, many layers lie under the surface, each with its own properties and

sound velocity. Moreover, sound waves arriving at the boundary between two layers both

reflect and refract through the lower layer. The refracted waves will then travel deeper and do

the same at the next boundary. Since all reflected waves ultimately reach the surface - though

at different times - many wave arrivals are detected after each sound generation. Adding the

fact that many sound detectors are usually used at a time, one can imagine the large amount of

data collected. Therefore, computers are always used to collect and analyze the data to

28
Fig. 2.3: Reflection of sound waves generated at the surface off a subsurface layer.

29
generate profiles of subsurface structures such as the one shown in Fig. 2.4. Prospective

underground hydrocarbon traps can then be located from these profiles.

On-shore, sound waves were traditionally generated by detonating explosives

(dynamite) or by dropping a heavy object to the ground. The sound generation point on the

earth's surface is called a shot point. There is usually a 50 to 100 meter interval between shot

points during seismic explorations. A recent development in sound wave generation is the use

of "vibroseis". This technique uses a heavy metal plate - weighing over one ton - which is

linked to a hydraulic vibrator and is carried by a truck. At the desired location, the plate is

lowered to the ground and is vibrated at a pre-programmed frequency (7 to 70 Hertz) for a

period of 5 to 10 seconds. The amount of data received is great, and its interpretation requires

advanced software and computers.

The reflected (striking back) or refracted (passing through) sound waves from the

subsurface layers are detected by geophones on the surface. Geophone spacing is about half

the shot point spacing and has usually 25-50 meter intervals.

Off-shore, sound waves are generated with an air-gun mounted on a ship and are

received by hydrophones which are placed in streamers dragged behind the ship. Therefore,

sound waves are both generated and received at the surface of the water.

Using Seismic Records


A seismic program for an area is the set of all shot point and geophone locations along

with the times, durations and frequencies of the sound waves generated. A seismic record is
the entire data sequences produced by a seismic program. The latest development in seismic

technology is the inference of detailed subsurface stratigraphy from seismic records. Much of

this is done by the geologist, who proposes an approximate model for the subsurface

stratigraphy of an area, reduces this to a geometric model involving layers with specific

density and velocity characteristics, and enters this model into a digital computer. The

geologist then inputs the seismic program used for this area and asks the computer to generate

a hypothetical seismic record for the model. Comparing this record with the real one, the

geologist modifies his model to minimize the differences. This procedure is repeated until

30
Fig. 2.4: Seismic reflections on a record section. (Courtesy of Western Geophysical)

31
the agreement between the two records is good, at which point the geologist adopts his model

as an accurate picture of the area.

Example 2.1 - Thickness Calculation - One Layer

Problem: Calculate the depth of a reflection point if the distance between a shot point and a

geophone is 100 m and the first reflected wave is received 0.2 s after the shot. Assume the

subsurface rock velocity is 5,500 m/s.

Solution: The distance from the shot point to the reflection point is:

L = vt/2

= (5,500 m/s) (0.2 s) /2

= 550 m

The depth of the reflection point is:

h2 = 5502 - 502

h = 547.7 m.

Example 2.2 - Thickness Calculation - Two Layers

Problem: Calculate the thickness of the bottom layer for the same conditions as Example 2.1

if the second reflecting wave is received 0.6 s after the shot.

Solution: Assuming the changes in velocity and refraction angle in the bottom layer are

negligible, then:

L = (5,500 m/s) (0.6 s) /2


= 1,650 m

The depth of the second reflection point is :

h2 = 16502 - 502
h22 = 1649.2

h = h2 - h1

= 1649.2 - 547.7

= 1,101.5 m.

32
Magnetic Method
The basic principle of this method is measurement of the intensity of the earth's

magnetic field. The magnetic field generated at the earth's surface depends on the distribution

and magnetic properties of the rocks located beneath the surface.

Since magnetic minerals in magma tend to line up along the earth's magnetic field as

the molten material crystallizes into solid igneous rock, large masses of such rocks generate a

significant local magnetic field. Metamorphic rocks may exhibit similar and fairly strongly

oriented magnetic fields because metamorphism usually involves considerable

recrystallization.

Sedimentary rocks, on the other hand, exhibit much weaker magnetic characteristics.

Magnetite, the only common strongly magnetic mineral, is commonly found in many igneous

and metamorphic rocks but is quite rare in most sedimentary rocks. Carbonate rocks very

seldom contain iron and have virtually no magnetic properties.

Therefore, in general, localities of strong magnetic flux have igneous and/or

metamorphic rocks near the surface while localities with a weaker magnetic flux have thick

sedimentary sections (Fig. 2.5). By mapping the earth's magnetic field at the earth's surface,

substructures containing sedimentary rocks beneath the surface can be determined. The shape

and the thickness of the sedimentary rocks may indicate prospective hydrocarbon traps.

Magnetic measurements are commonly done with geomagnetic gradiometers which

are suspended from low-flying airplanes. Gradiometers have two magnetometers separated

from each other by a known vertical distance. This allows simultaneous and continuous
measurement of the earth's vertical magnetic field gradient and the total magnetic field

intensities along a predetermined pattern. The data obtained from airborne magnetic surveys

are inexpensive compared to data from other sources.

Gravitational Method
In this method, variations in the gravitational acceleration of the earth (g) are

measured. The force of gravitational attraction between two objects is proportional to the

product of the masses of the two objects and is inversely proportional to the square of the

33
Fig. 2.5: Magnetic anomaly due to subsurface magnetic rocks.

34
distance between them. Local changes in the earth's gravitational force reflect variation in the

density of the subsurface rocks (Fig. 2.6). The surface gravitational attraction of overlying

layers of sand and shale with an overall relative density of 2.60 is measurably greater than

that of an overlying mass of salt with a relative density of 2.16. Such a distinction has been

used to locate deeply buried salt domes under flat and featureless coastal plains.

Modern gravity meters (gravitometers) have very sensitive balances which can easily

detect any small difference in gravitation. Like magnetic surveys, gravity surveys can be

recorded more or less continuously from low flying planes and from moving survey ships.

Often, magnetic and gravity surveys are recorded simultaneously. The principal use of gravity

surveys and magnetic data today is in initial reconnaissance mapping of large undrilled areas.

This data is often used to decide which one of several areas looks the most attractive and is

worth conducting seismic surveys.

Electrical Method (Resistivity Survey)


This is a geophysical method which is used for simple structures of up to 1 km in

depth. The principle of this method is to send electrical current into the ground and measure

the potential drop due to the resistance of the subsurface layers. Because oil is a poor

conductor, subsurface layers containing only water - especially salt water - would show a

much lower electrical resistivity than layers containing oil and water. Four electrodes are

usually used: two for sending current into the ground and the other two for measurement of

the potential drop. Investigation of the electrical anomalies can lead to the location of oil
deposits beneath the surface of the earth.

2.4 Geochemical Method


Geochemical surveys involve analyses of air and soil samples and surface and ground

waters. Soil analysis, for example, may indicate the presence of hydrocarbons beneath the

surface. Similarly, analysis of surface water from streams and rivers may indicate oil seeps

along the river banks or bed. Figure 2.7 illustrates various situations where hydrocarbons

escaping from an underground trap can be detected at the surface.

35
Fig. 2.6: Gravitational anomaly due to subsurface dense rocks.

36
Fig. 2.7: Origins of surface petroleum seeps.

37
Exercises
2.1 Calculate the thickness of a layer if the distance from shot point to geophone is 50 m

and the first sound wave arrival time is 0.4 s. Assume a subsurface sound velocity of

5,800 m/s. Ans: 1159.73 m.

2.2 Calculate the thickness of the bottom layer for the same conditions as above if the

second wave arrival time is 0.6 s. Ans: 580.09 m.

2.3 Name the various petroleum exploration methods and explain one of them with

sketches.

2.4 Which geophysical method is used most commonly ? Why?

2.5 What is the basic difference between the magnetic and the gravitational methods?

2.6 What are the main characteristics of the the geochemical method that the geophysical

methods do not have?

38
Chapter 3

DRILLING

3.1 Introduction
Once exploration work has identified a subsurface structure which looks like a

potential petroleum trap, the only way of finding that out is to drill a well down to that

structure. If petroleum is indeed found, the well becomes the simplest and least expensive

means of delivering the oil and gas to the surface.

For centuries, man had exploited surface and near-surface deposits of petroleum by

digging pits and primitive wells. These early wells were dug by hand according to various

techniques. The Baku and Burma methods were widely used.

A Baku-type well was started as a large-diameter hole and was then deepened through

a succession of smaller-diameter steps as shown in Fig. 3.1. Once completed, men could

walk up and down the hole carrying oil to the surface.

In the Burma type, one or two men dig at the bottom of the hole and load the rubble

into a basket. Other men at the surface pull the basket up as shown in Fig. 3.2. The diameter

of the hole in the Burma type remained the same along the full depth of the well.

For the purpose of salt mining, the Chinese later developed a better technique to drill

wells. A metal pipe with a sharp chisel-like end was raised and dropped repeatedly into the

hole. After several strikes, the pipe was lifted to the surface to clean it out of the cuttings

accumulated inside it. Men used to jump over a board to generate an up-and-down motion

which was conveyed to the pipe through a rope and pulley mechanism.

During the 1860's, when the modern oil industry was just beginning, the need to drill

deeper wells at faster rates was great. Early drillers quickly improved the Chinese method by

replacing the pipe with a solid bit and adding an engine as a power source to raise and drop

the bit. The rock cuttings are removed periodically by a bailer. The technique became popular

and later came to be called the cable-tool method, and it is still being used today to drill some

oil wells. Fig. 3.3 shows a modern cable-tool rig.

39
Fig. 3.1: Baku type well.

Fig. 3.2: Burma type well.

40
Fig. 3.3: A cable-tool rig.

41
Nowadays, the majority of oil and gas wells are drilled by the rotary method which is

described in the following section. This method was first applied to an oil well in 1901 and

has seen many developments and improvements since then.

3.2 Rotary Drilling


In rotary drilling, a bit is attached to the bottom end of a string of pipe joints (the drill

string), which is rotated at the surface. The first joint of the drill string (the kelly) is square in

shape to enable the rotating equipment at the surface to hold onto it.

Due to continuous rotation and the weight of the drill string, the bit crushes, scraps

and cuts the rock into small pieces. These cuttings are carried to the surface by means of

circulating a drilling mud. The drilling mud is pumped down the drill sting where it exits

through nozzles in the bit. As the mud flows back to the surface through the annulus, it

flushes out the cuttings. Once the cuttings are removed from the mud, it is pumped back into

the well through a device called the swivel which is attached to the top of the kelly. The

swivel along with the entire drill string is raised and lowered by a giant hook which is

suspended from the rig by cables.

3.3 The Drilling Rig


A drilling rig, shown in Fig. 3.4, is made up of two main sections. The bottom section

is called the substructure and the section at the top is called the "derrick". The derrick is a

strong steel structure capable of carrying huge loads and is installed on top of the
substructure. Derricks come in different sizes, but those less than 40 meters high are called

"short rigs". The substructure is a rigid platform which supports the derrick. Its height

ranges from 5 to 10 meters according to the clearance needed above the well head.

Components of a Rotary Rig


A rotary drilling rig is a complex piece of machinery with many systems and

components. Four of its major systems are discussed below.

42
Fig. 3.4: A rotary drilling rig.

43
1. Power System
Generally, rigs use diesel engines to generate electricity on location where each

generator can provide up to 500 kW in electric power. All the rig's systems are driven by

electric motors; however, most of the power is consumed by the mud circulation pumps

(85%) and the hoisting draw works (15%).

2. Hoisting System
The hoisting system consists of the drawworks, crownblock, traveling block, hook,

elevator, swivel, drilling line and dead line anchor (Fig. 3.5).

The drawworks drum is rotated in either direction to pull in (spooling) or release (feed

off) the cable, which raises the traveling block or lowers it accordingly. Since the entire drill

string is suspended from the traveling block, a very heavy load is exerted on the draw works;

therefore, a block-and-tackle system is used.

The crown block, the traveling block, and the drilling line make a block-and-tackle

system which helps reduce the tension in the cable. Due to the smaller force applied, the

power requirement from the engines is also reduced. However, the work expended will

remain the same, because the length of the cable is increased. In the block-and-tackle system,

the tension in any of the lines is given by:

T = W/n

where

T : tension in any line, N


W : total load suspended from the traveling block, N

n : number of lines.

The speed of the traveling block is related to the speed of the fast line by:
vf = n vh

where
vh : speed of the traveling block, m/s

vf : speed of the fast line, m/s.

44
Fig. 3.5: Components of the hoisting system.

45
Example 3.1 - Hoisting Calculation

Problem: A 10,000-kg drill pipe has to be hoisted at a speed of 1 m/s by a traveling block

strung with eight line. Calculate the speed of the fast line and its tension.

Solution:

W = mg

= (10,000 kg) (9.81 m/s2)

= 98,100 N

T = W/n

= 98,100/8

= 12,262.5 N
Vf = n vh

= 8 (1 m/s)

= 8 m/s.

3. Rotating System
The rotating system, shown in Fig. 3.6, consists of the rotary table, master bushing,

kelly bushing, kelly, drill pipe, and the bit. Rotation starts with the swivel which is made up

of two parts. The top part of the swivel is suspended from the hook and is, therefore, fixed.

The bottom part is free to rotate and is connected to a square - or hexagonal - pipe called the

kelly. The kelly passes through the rotary table which is installed at the rig floor.

Rotation is imparted to the whole system by the motor-driven rotary table. Through
the master and kelly bushings, the rotary table rotates the kelly while allowing it to slide

through the opening in the table. The drill pipe with the bit at its end is connected to the kelly

and, therefore, rotates with it.

As the hole gets deeper during drilling, the drillstring is lowered gradually to keep the

bit at the bottom of the hole. This is achieved by lowering the traveling block which causes

the kelly to slide through the rotary table while they both are rotating.

46
Fig. 3.6: Components of the rotating system.

47
Drilling Bits
The bit is a device that does the actual drilling. Its diameter - and, therefore, that of

the hole - is larger than the drill pipe diameter. This is to allow the mud to flow back through

the annulus. To facilitate drilling, the bit must press against the hole's bottom while rotating.

The weight of the drillstring is utilized for this purpose. However, if the drill string is allowed

to rest completely on the hole's bottom, it cannot withstand its own weight and might bend,

buckle or break. Therefore, the drillstring is always kept partly suspended from the hook to

provide some "weight on the bit". The amount of this weight depends on many factors namely

the rock hardness and type of bit. Several types of bits are available, and the choice depends

on the particular use. Some common types are shown in Fig. 3.7 and are described below.

Blade Drag Bits: Blade drag bits are used for very soft formations such as sticky shales.

They drill mainly by a scooping action like shoveling. Two nozzles direct drilling fluid to the

bottom and near the edge of the blades to remove the cuttings. Sometimes, the rock is so soft

that drilling is advanced just by jetting the drilling fluid to the ground. The common types of

drag bits are called fish tail bits.

Conical Drag Bits: A conical drag bit has a cone-shape body with side blades. As the bit

rotates, it exerts an abrading and scraping action on the rock with its teeth.

Rolling Cutter Jet Bits: These are used for most of the drilling. They usually have three
teethed cones which roll freely when the bit is rotated. Long teeth are used for soft

formations, while short teeth are used for hard formations. These bits have a scraping and

scooping action; but because of the rolling, they also exert a crushing and grinding action on

the rock. There are three nozzles for the drilling fluid both to clean the teeth and to flush the

rocks for softening and loosening purposes.

48
Fig. 3.7: Common types of drilling bits.

49
Diamond Bits: This type of bit has teeth made of industrial diamond and is used for

very hard formations. This bit is similar to the conical bit because it does not roll during

rotation and it drills by an abrading and scraping action.

Core Bits: Sometimes, a large and uniform sample of the subsurface rock is desired. For

such samples - called core samples - a special bit with a round and hollow interior is used.

The core bit grinds the rock on the circumference of the hole leaving a cylindrical center to

enter the bit's body where it passes on into a core holder. The core holder is closed on the

other end and is used to store the core sample and protect it from contamination by the

drilling fluid. Depending on the type of core bit assembly, core samples up to 20 feet long can

be cut.

4. Circulating System
The function of this system is to circulate a fluid (usually mud) through the well to

bring the cuttings to the surface (Fig. 3.8). Pumps draw mud from the suction pit - a tank

where mud is prepared and stored - and deliver it to the stand-pipe at high pressure. From the

stand-pipe, mud flows through the rotary hose (kelly hose) and enters the swivel via a

smoothly bent steel pipe called the "goose neck". From the swivel, the mud flows through the

drill string which leads directly to the bit where it exits through its nozzles. The mud then

rises in the hole - through the annulus - carrying with it all the rock cuttings to the surface.

Before returning back to the mud pit via the mud return line, the mud has to be cleaned. This
is done by first flowing the mud over a vibrating screen known as the "shale shaker", which

screens out the cuttings that are larger in diameter than the mesh size of the screen. The

screen's vibration causes the cuttings to roll off the screen where they are discarded into the

reserve pit. The mud with only smaller particles - such as sand and silt - falls through the

screen into an intermediate tank called the settling pit. Here, the small particles are removed

and chemicals are added to the mud to recondition it. The cleaned and treated mud flows over

to the suction pit where it is ready for recirculation.

50
Fig. 3.8: Components of the circulating system.

51
3.4 Drilling Fluid
The drilling mud is an extremely important component of the drilling operations as

will be shown shortly. Therefore, it is prepared carefully according to exact specifications

and its properties are continuously monitored. Chemicals are frequently added to it or

removed form it according to the needs and conditions of drilling.

In the early days, plain water was used at the start of drilling, and while circulating

through the well, the water gradually changed into mud as it contacted the subsurface layers.

As deeper wells were required, necessitating muds of specific properties, drillers began

mixing the mud at the surface. Eversince, the variety of mud ingredients grew larger every

year.

Besides water and clay, a drilling mud nowadays could contain oil, salt, coal powder,

saw dust, mica flakes and an array of chemicals. Foam - and even air - has also been used in

drilling. For this reason, a more appropriate name for the drilling mud today is the "drilling

fluid".

The drilling fluid performs the following vital functions:

1. It cleans the hole and transports the cuttings to the surface. These rock fragments can

provide information on the subsurface strata and the location of petroleum deposits.

2. It cools the bit and lubricates the outside of the drill string. Without the drilling fluid,

the bit temperature could increase beyond the melting point of steel. Friction with the

wall of the hole could also cause the drillstring to get stuck.

3. It controls the well. Due to its hydrostatic pressure, the drilling fluid balances the
pressure of any oil or gas layer that may be penetrated. This prevents such fluids from

entering the wellbore and flow up to the surface causing a "blow out".

4. It supports the wellbore. If a porous subsurface stratum is encountered, the water of

the drilling fluid could filter into it causing the fluid to thicken. However, clay

particles in the mud cannot enter the stratum and accumulate on the wall of the hole in

the form of a layer. This thin layer of clay is called the "mud cake" and it acts as a

plaster preventing the well bore from caving in. It also stops further water filtration.

52
5. When mud circulation is stopped for any reason, the thick mud prevents the cuttings it

is carrying from settling to the bottom of the well. Without such action, resuming

rotation of the drillstring could become impossible.

6. It partially supports the weight of the drill string due to buoyancy forces.

7. It enhances drilling by its jetting action through the bit nozzles which softens the rock.

Some soft formation can be drilled through this way without even bit rotation.

8. It provides a medium for well logging operations which will be explained in a later

chapter.

Example 3.2 - Buoyancy calculation

Problem: If the density of steel is 7800 kg/m3; the density of the drilling fluid is 1200

kg/m3; the outside diameter of the drillpipe is 10 cm, the internal diameter is 8.5 cm; and the

depth of the hole is 2000 m, calculate the buoyancy force of the drilling fluid.

Solution: The volume of the displaced mud is equal to the volume of the drillpipe wall:

Vd =  D 2  Di 2  L
4 o

= (0.102 - 0.0852) 2000
4

= 4.359 m3

The buoyancy force is


Fb = Vd m g

= 4.359 x 1200 x 9.81

= 51,314 N

Note that the weight of the drillpipe in air is:


Wd = Vd s g

= 4.359 x 7800 x 9.81

= 333,542 N

which means that the drilling fluid supports nearly 15.4% of the weight of the drill pipe.

53
To ensure that the drilling fluid performs its functions properly, its properties are

periodically measured on site. Techniques used to measure some of its most important

properties are described below.

Density: Density is normally measured with a mud balance (Fig. 3.9) which is a very simple

device. One just fills the balance cell with mud, closes the lid and pushes the rider to balance.

When it is balanced, the mud density can be read directly from the scale. One can calibrate

the mud balance simply by filling the cell with fresh water and set the rider at 1 g/cm3. Some

weight could then be added to or subtracted from the counterbalance weight source to

balance it.

Mud Viscosity: A funnel (Fig. 3.10) having a 1500 cm3 volume is filled (from the

screened side) with mud while the bottom of the funnel is closed with a finger. Mud is taken

from the settling pit just beneath the shale shaker. The funnel is then drained over a cup

having a 946 cm3 volume and the time taken to fill the cup is measured. This time is called

the mud funnel viscosity; for water it is about 26 seconds.

Filtration: As mentioned earlier, water filtration into a porous stratum ceases once a good

mud cake forms. Sometimes, however, a mud cake fails to form leading to severe "water loss"

which is a serious problem. Therefore, the filtration characteristics of a mud must be tested to

check if they are adequate. This is done according to a standard procedure in a special device
called the "Filter Press" (Fig. 3.11). The press has a cup which is filled with mud. When air

pressure is applied to the cup, the mud filters through the cup's base which is merely a screen

covered with a filter paper. The amount of water collected in a given interval is used to

calculate the mud's water loss tendency in standard units.

Example 3.4 - Typical Mud Calculations

Problem: a) Suppose the hole is 20 cm in diameter and 2500 m deep, how much mud is

needed to fill it?

54
Fig. 3.9: The mud balance.

55
Fig. 3.10: The mud viscosity funnel.

56
Fig. 3.11: The filter press.

57
Solution: The volume of the hole = cross-secitonal area x depth

=  (d/2)2 h

= (0.20/2)2 2500

= 78.5 m3

Generally, the volume of the hole is multiplied by 2 to obtain the mud volume needed for the

whole circulation system. Therefore, the mud needed is 157 m3.

b) If the pressure of the oil in the layer that the well has just reached is estimated at

28 MPa, what should the mud density be in order to balance this pressure?

Solution: The mud must create a hydrostatic pressure at the bottom of the hole equal to 28

MPa.
Since P = m g h

where

P : hydrostatic pressure, Pa
m  mud density, kg/m3

g : acceleration due to gravity, m/s2

h : depth of the hole, m


P
Then m =
gh
28 x106
= = 1142 kg/m3.
9.81 x 2500

Actually, the mud should be mixed to a density slightly larger than 1142 kg/m3.

c) How much water and clay need to be mixed to prepare this volume of mud with

this density?

Solution: The volume of mud is the sum of the volumes of its ingredients.
Vm = Vw + Vc

where
Vm : volume of mud, m3

Vw : volume of water, m3

Vc : volume of clay, m3

The density of mud is calculated according to :

58
w Vw  c Vc
m =
Vw  Vc

where
w : density of water (1000 kg/m3 )

c : density of clay (2670 kg/m3 )

substituting for Vc by (Vm - Vw) and rearranging:

Vm ( c  m )
Vw =
w  c
157 ( 2670  1142 )
=
2670  1000

= 143.7 m3
Therefore,
Vc = 157 - 143.7 = 13.3 m3.

3.5 Casing
Normally, drilling is stopped at certain depths - decided by the geologist - to set the

casing. The casing is a heavy steel pipe placed in the borehole to support the hole against

collapsing and to seal all subsurface layers permanently. The seal is needed to prevent the oil

or gas - once production starts - from contaminating or entering into other strata, or any

ground water from entering the borehole. Like the drill pipe, the casing comes in pieces called

joints but its diameter is much larger and close to that of the hole. Four strings of casing are

normally installed in the hole, each of a different size and to a different depth. They are the

conductor, the surface casing, the intermediate casing and the production casing (Fig. 3.12).

When a depth is reached where a casing string should be installed, the drillstring and

bit are first taken out of the hole. Then a round and smooth object called the guide shoe is

attached to the first joint of casing to help it slide easily into the mud-filled hole. Using the

rig's hoisting system, the casing is then run into the hole adding one joint at a time until the

guide shoe reaches the bottom. At this point, the casing string is ready to be cemented in

place.

59
Fig. 3.12: The types of casing.

60
3.6 Cementing
Cementing involves filling the annulus between the casing and the borehole with a

high strength cement. Once the cement sets, the casing string becomes fixed permanently into

the hole. The cement slurry is prepared on location using special truck mixers and is pumped

into the casing. Before pumping cement, a special rubber plug is inserted into the casing

ahead of the cement to separate it from any liquid - usually mud - that might be present inside

the casing. Sometimes, a batch of about 3 m3 of a cleaning liquid called the "preflush" is

placed after the plug. The preflush is intended to clean the inside of the casing and -

ultimately - the annulus from mud and mud cake. The preflush is then followed by cement.

Once the required volume of cement is pumped, another rubber plug is inserted behind it

followed by mud. Pumping mud into the casing continues until the bottom plug stops at the

float valve, which is a check valve placed one or two casing joints above the guide shoe. At

this point, the cementing operator notices a sudden increase in the pumping pressure.

Continued pumping of mud causes the bottom plug to rupture allowing the cement to

fill the bottom joints of casing and to rise in the annulus. When the top plug reaches the

bottom - as signaled by another pressure surge at the pump - the cementing operator knows

that most of the cement is now in the annulus. Pumping is then stopped and the casing is left

undisturbed until the cement sets. During this period, the cement cannot flow back into the

casing because of the float valve.

Once the waiting period is over, drilling is resumed with a smaller bit to fit inside the

casing string just installed. The new bit will first drill through the rubber plugs, float valve,
cement inside the bottom joints of the casing and the guide shoe before it reaches the bottom

of the hole.

The last casing string - the production casing - is run only after a petroleum reservoir

has been encountered. A reservoir is a trap proven to contain commercial quantities of

petroleum; and such proof is decided based on many tests performed on the formation. These

tests will be discussed in detail in the next chapter. Running the production casing is part of

the well completion procedure.

61
3.7 Well Completion
This is the last phase in drilling an oil well; and it involves preparing the formation

for oil or gas production. Depending on the properties of the fluids and rock of the formation

and its conditions, one of several completion methods is chosen. The most common ones are

discussed below and are illustrated in Fig. 3.13.

Open Hole Completion


If the formation rock is sufficiently hard and if no water or gas is produced with the

oil, the well is completed as an open hole. In these wells, the end of the production casing is

set just above the formation. This way, oil may enter the borehole at every point on the

formation face.

Perforated Casing Completion


In this type of completion, the production casing extends to a little below the reservoir,

sealing the entire formation face. For the oil to enter the borehole, holes (called perforations)

must be made in the production casing and the sheet of cement behind it. This is done with a

special device called the “perforator”. The perforator is a long and slim device which carries

bullets that can be fired electrically from the surface while the device is inside the hole. Once

the perforator is wired for the desired number and location of perforations, it is lowered to the

required depth and the bullets fired. The bullets will penetrate the casing and the cement

sheet, and will travel for a distance inside the formation leaving an open path for the oil to
flow through. The perforated casing completion allows oil production from certain intervals

only. This is useful when the reservoir has many water layers within it which need to be

sealed.

Liner Completion
This method is halfway between open hole and perforated casing completion. The

production casing is still set above the formation, but the hole is not left open. A special pipe -

called the liner and is made of two or three joints of a smaller size casing - is suspended from

62
Fig. 3.13: Common types of well completion.

63
the end of the production casing and is allowed to hang in the well opposite the formation.

Oil can enter the bore hole through holes in the liner; sometimes the liner is slotted and

covered with a screen. Such a completion is suitable for weak formations which cannot be left

open because of sand production. Sand can accumulate behind the liner and, therefore, stop

more sand production; gravel can also be packed in that space to act as a filter.

3.8 Rig Types


Drilling rigs are commonly classified into two major categories: (a) Onshore (land)

rigs and (b) Offshore (marine) rigs. Some common types are shown in Fig. 3.14.

Most onshore rigs today are mobile, which can be moved from one location to

another. Smaller rigs are moved as one piece while large ones are moved in sections that are

assembled on site. Offshore rigs come in a bigger variety because of the wide range of sea

conditions, especially water depth and weather. Some of the most common offshore rigs are

discussed below.

Barge: This type of drilling unit is used primarily in swampy areas, like in river deltas

or in the coastal areas of shallow lakes. It can be towed to the location and then ballasted to

rest on the bottom.

Jack-up: The jack-up is a mobile drilling rig, generally used in shallow waters of depth

less than 107 m (350 ft). These rigs are very stable drilling platforms because they rest on the
seabed and are not affected by sea currents. Most jack-up rigs have three, four or five legs, but

a few of the earlier models have eight or ten, and one has fourteen legs. These are cheaper rigs

as compared to other mobile offshore rigs.

Fixed Platforms: These are very common drilling and production units, and hundreds of

them are used offshore in many parts of the world. Fixed platforms are very stable under the

worst sea conditions, but they are virtually immobile. They are preferred over all others in

shallow waters, but they can also be used in very deep waters.

64
Fig. 3.14: Common types of drilling rigs.

65
Semi-Submersible: These are floating drilling units consisting of hulls or caissons which

carry a number of vertical columns to support a deck fitted with the derrick and other drilling

equipment. Semi-submersible drilling rigs are very stable for deep water drilling and they are

available with various hull configuration and number of stabilizing columns. Both hull and

columns are used for ballasting as well as storing supplies.

Drillship: Drillships, also known as "floaters", are specially built or modified for deep-

water drilling. Although these are not stable during drilling, they provide greater mobility

than other mobile drilling units. Their main advantage is their ability to drill at almost any

depth of water. Drillships are anchor-moored, but recent ships are equipped with dynamic

positioning components which enable them to keep on-station above the borehole. These

drilling vessels have greater storage capacity than other types of offshore rigs and operate

independent of service and supply ships.

3.9 Rig Personnel


As a standard procedure, a rig is operated on 3 shifts in order to drill continuously 24

hours a day. Usually, six persons work on each 8-hour working shift. The driller operates the

draw works and is responsible for the crew and the drilling operations. The derrickman works

at the monkey board during tripping ( pulling the drillstring out of or inserting it back into the

hole ) and at other times he is responsible for mud conditioning. Three floormen handle

drillpipe joints during drilling or tripping. The sixth person is the motorman who is
responsible for maintenance of the engines. Responsibility for the entire rig and drilling site

lies on one person called the "toolpusher". He is usually an experienced ex-driller who stays

at the location 24 hours a day until the drilling is finished. If the well is a wildcat (the first

well drilled in an exploration area), a geologist is present at the site to examine the cuttings

and a mud engineer is in charge of the drilling fluid.

66
3.10 Directional Drilling
An oil or gas well is seldom drilled perfectly vertical; a small offset of few meters

between the top and bottom is always expected. However, an offset hundreds of meters wide

can be created intentionally under certain circumstances. Such "deviated" wells are drilled

whenever the topography prevents drilling at the desired location, e.g., in cities and marshes.

Sometimes, only a limited number of drilling sites in an area can be afforded, e.g., an offshore

filed. In such cases, many deviated wells are drilled from one offshore platform, each

targeted to a different corner of the reservoir. Deviated wells are started as vertical wells

down to a certain depth, then an elbow-shaped device of a few degrees bend is fitted above a

special bit which is rotated by its own motor rather than from the surface. Drilling is then

resumed for a few hundred more meters before another elbow is fitted. This procedure is

repeated until the desired angle of deviation is achieved after which drilling is continued

straight until the target depth is reached. A typical deviated well is shown in Fig. 3.15, where

vertical drilling is not possible due to the populated area above.

The ultimate development in drilling deviated wells is horizontal drilling. This new

technology is capable of drilling a well whose last section extends horizontally for thousands

of meters, thus increasing well productivity many fold. Reservoirs as thin as 3 meters can now

be exploited economically via one horizontal well drilled right in the middle of the stratum. If

the need arises, the last section of the well can even be configured in any desired shape. It can

run horizontally for a distance, curve upwards, then horizontal again, turn right, then

downward .... and so on. Obviously, such technology requires highly advanced steering
components and direction sensors. Computer-controlled, subsurface devices can, for instance,

steer the bit, determine its exact location and bearing, test the rock and its fluid content, and

evaluate the bit performance all while drilling.

67
Fig. 3.15: A deviated well drilled to a reservoir under a populated area.

68
Exercises
3.1 What is the basic difference between Rotary Drilling and Cable Tool Drilling?

3.2 What are the functions of the B.O.P?

3.3 Why is the block-and tackle system used in the rig?

3.4 What are the differences between plain air and plain water considering the functions

of drilling fluids?

3.5 Why do you need casing and cementing operations?

3.6 How do you select a Well Completion Technique?

3.7 What are the basic differences between onshore and offshore drilling operations?

3.8 How do you give the desired vertical angle and the direction to a well to be deviated at

a certain depth?

3.9 Calculate the speed of the fast line and the tension in it if a 8,500 kg drill string has to

be hoisted at a speed of 1 m/s. Ans: 8 m/s; 10,423.0 N.

3.10 Calculate the weight of the casing in air, the buoyancy force of the drilling fluid, and

the weight of the casing in drilling fluid, when the density of steel is 7,800 kg/m3, the

density of the drilling fluid is 1,100 kg/m3, the outside diameter of the casing is 20

cm, the inside diameter is 18 cm, and the length of the casing is 1,500 m. Ans: 685 kN,

96.5 kN, 588.5 kN.

3.11 Calculate the amount of mud needed for a 60 cm diameter hole at a depth of 500 m.

Ans: 283.0 m3.

3.12 Calculate the hydrostatic pressure of 1,100 kg/m3 of mud at a depth of 6,000 m. Ans:
65.0 MPa.

3.13 Calculate the amount of cement slurry needed to cement a well when the average hole

diameter is 20 cm, the casing outer diameter is 17 cm, and the depth of the borehole is

2,200 m. Ans: 23.0 m3.

69
Chapter 4

RESERVOIR EVALUATION

4.1 Introduction
Well completion involves a lot of expenses in the form of production casing and other

equipment. Therefore, once an oil-bearing layer is encountered, the decision whether to

complete the well or not must be made very carefully and quickly. This involves many types

of tests which fall under a branch of petroleum engineering called Formation Evaluation.

Since the only direct access to the reservoir from the surface is through the borehole, all these

tests center around the well. In this chapter, tests performed on the well are discussed first

while tests performed on samples obtained from the well are discussed later in the chapter.

4.2 Driller's Log


The driller's log is a record showing the nature of the rock being drilled, the drilling

fluid being used and the progress of drilling. The cuttings which are transported to the

surface by the drilling fluid come over to the shale shaker where the larger pieces can be

picked off the screen. These are examined in a laboratory to find out the type of rock. They

are also checked for traces of oil by putting them under ultraviolet light where oil glows a

blue color. To detect gas, a gas sensor is placed just above the outlet of the mud return line.

4.3 Well Logging


One of the most important pieces of information needed on an oil-bearing layer is its

content of oil. This content depends upon: a) the fraction of rock volume that is pore space -

this is defined as the porosity - and b) the fraction of the pore space that is filled with oil - this

is defined as the oil saturation. It is important to remember that the pore space could be shared

by oil, water and gas, and the sum of all fluid saturations is simply equal to 1.

The oil content can be easily measured if a core sample is obtained at the surface. But

since coring is an expensive operation and is performed on proven reservoirs only, good

estimates on the oil content can be derived from well logging.

70
Well logging involves measurement of electrical, acoustic or radioactive properties of

the subsurface layers. When a well log is desired, the drill string is taken out of the hole and a

special device is lowered into the borehole by a cable. This device has two main parts, a

transmitter and a receiver. Depending on the type of well log, the transmitter sends electrical,

acoustic or radioactive signals to the formation and the receiver detects the responses and

transfers them to the surface via the cable. Since the length of the cable in the borehole is

known, the formation responses and their corresponding depths can be recorded

simultaneously. These recordings are called "well logs" and the operation is called "well

logging". The following are examples of typical well logging operations aimed at estimating

the formation's porosity and oil saturation.

Electrical Resistivity Log


Electrical resistivity of a substance is a measure of its opposition to the flow of

electrical current. When electric current is passed to a formation, the formation exhibits

electrical resistance to this current. In this type of log, a tool measures the resistivity of the oil

bearing formation along its thickness. The magnitude of this resistivity depends obviously on

the formation lithology and its content of hydrocarbons. Hydrocarbons offer much higher

electric resistance than does salt water and, therefore, the oil-bearing zone of the formation

will correspond to a higher resistivity than the water zone. In fact, the resistivity log could

show the boundary between these two zones. It is also possible to use the resistivity log to

compute the water saturation in the oil zone using Archie's law:
Sw2 = Ro / R t (4.1)

where
Rt : Resistivity of the oil zone of the formation, m

Sw : water saturation in the oil zone of the formation, fraction

Ro : Resistivity of the water zone of the formation, m.

71
Example 4.1 - Water Saturation in the Oil Zone

Problem: Calculate the water saturation in the oil zone of an oil-bearing formation if the oil

zone resistivity is 10 m and the water zone resistivity is 0.9 m (Fig. 4.1).

Solution:
Ro = 0.9 m

Rt = 10.0 m

then
Sw2 = Ro / R t

= 0.9 / 10.0 = 0.09


Sw = 0.3

And if there is no gas in the formation, the oil saturation is:


So = 1 - 0.3

= 0.7

Note that the log indicates that the oil zone extends from 2512 to 2590 meters of depth, i.e., it

is 78 meters thick.

Acoustic Log
Acoustic Logging is based on measuring the velocity of sound in rock formations.

Sound travels faster in a rock with no pores than in the same rock with pores containing fluid.

This is because the velocity of sound is higher in a solid medium than in a fluid medium.

Consequently, if the rock lithology is known, the porosity of a particular rock can be
estimated from its average sound velocity. The relationship between porosity and sound

travel time per foot - which is the inverse of the average sound velocity - is:
 tlog = tf + (1 - )tma (4.2)

where
tlog : sound travel time in the formation as shown by the acoustic log, s

 tma : sound travel time in the grain material of the formation, s

 tf : sound travel time in the fluids of the formation, s

  porosity of the formation, fraction

72
Fig. 4.1: A resistivity log across a reservoir.

73
Rearrangement of Equ. (4.2) yields porosity as:
   (tlog - tma) / (tf - tma)

Example 4.2 - Porosity from the Acoustic Log

Problem: Calculate the porosity of a limestone formation containing salt water of 1.05 g/cm3

density if the acoustic log reading was 70 s, where


tma = 45 s (known unit sound travel time for limestone grains)

 tf = 189 s (known for 1.05 g/cm3 salt water)

Solution:
  (70 - 45) / (189 - 45)

= 0.17

Radioactive Density Log


Radioactive density logging involves sending gamma rays into the formation and

detecting the rate of gamma rays returning to the logging tool. This can then be used to

compute the density of the formation. Since the rock contains pores which are filled with

liquids, this density is called the "bulk density". A simple equation can then be written for

the bulk density as:


b = f + (1 - )ma (4.3)

where
b : bulk density of the formation as shown by the density log, g/cm3

f : formation fluid density, g/cm3

ma : formation grain density, g/cm3

 : porosity of the formation , fraction.

Example 4.3 - Porosity from the Density Log

Problem: Calculate the porosity of a sandstone formation containing fresh water if the

density log reading is 2.32 g/cm3 and


ma = 2.65 g/cm3 (known density of sandstone grains)

74
f = 1.00 g/cm3 (density of fresh water).

Solution:
  (ma - b ) / (ma - f)

= (2.65 - 2.32) / (2.65 - 1.00)

= 0.20.

4.4 Rock and Fluid Sampling


When a general idea of the formation being drilled is desired, rock cuttings can

provide suitable information. For detailed information, however, the cuttings are inadequate

because they are insufficient, too small or contaminated with drilling fluid. Moreover, the

depth from which a rock cutting has come is not known accurately. Core samples provide a

wealth of knowledge on the reservoir and its productivity.

If the approximate depth of the reservoir is known beforehand, a large diameter core

can be planned and obtained with a core bit. On the other hand, if the formation has already

been drilled through, core samples can still be obtained from the side of the hole. A special

device is lowered with cable to the desired depth and small core samplers (tied by wires) are

fired by explosive charges. When the empty core samplers hit the formation face, they cut

small cylindrical samples called "sidewall cores" which can then be brought to the surface.

Oil samples can be obtained at the surface by producing the well temporarily as will

be discussed shortly. But the composition of the oil at the surface may be quite different from

its composition in the reservoir. Therefore, oil samples are sometimes obtained directly from
the formation using a bottom-hole sample bomb. This device is lowered into the hole by a

cable and is positioned opposite the formation. The bomb is then filled with oil by opening

and closing - from the surface - its main valve; after which the bomb is pulled back to the

surface.

4.5 Production Tests


A good oil content is only one factor that qualifies an oil-bearing formation to be

considered an oil reservoir; another factor is the rate at which the formation can produce oil.

Still another factor is how large - area wise - the trap is. These questions can be answered

75
quickly by a simple test performed using drilling rig equipment in addition to a few simple

tools. The drill stem test is performed usually after the logs show a good oil content.

Drillstem Test
In this test, the drillstem (drillpipe) is used to put the formation in a production mode

temporarily. The drillstem is lowered back into the hole with a packer (an inflatable rubber

object) installed above its bottom end. When the packer is inflated above the formation top

(Fig. 4.2), the borehole annulus is sealed and the formation and the bottom hole become

completely isolated from the rest of the well. Openning a valve installed at the top of the

drillstem allows oil to flow to the surface where it is collected and metered. The productivity

of the well can then be judged from the flowrate and types of fluids produced. After

monitoring the well for a few hours, the valve is closed and the well is left undisturbed for a

day or two. Throughout the test, a pressure gauge installed at the bottom of the drillpipe

records the bottomhole (formation face) pressure continuously. The difference between the

formation pressure recorded before the production phase and after the shut-in phase is

utilized to assess the size of the accumulation. The smaller this difference is, the larger is the

trap. Fig. 4.3 shows a typical pressure trace during a drillstem test.

Once the drillstem test confirms that the oil accumulation is indeed a reservoir and that

the well is a productive one, the next step would be to complete the well - as explained in the

previous chapter - and tie it to the production system - as will be described in the next chapter.

But before moving to that final stage, let us look at some routine laboratory tests that are
conducted on rock and fluid samples. The decision to complete the well may not depend on

the results of these tests since these could take time and cannot be carried out at the well site.

However, they are important for the characterization of the reservoir, the way it is produced

and the way it is managed in the future.

4.6 Rock Properties


Petroleum reservoir rocks are generally sedimentary rocks such as sandstone (SiO2),

limestone (CaCO3), and dolomite [CaMg(CO3)2]. A petroleum engineer is interested in

76
Fig. 4.2: Schematic of a drill-stem test.

77
Fig. 4.3: A pressure record during a drill-stem test.

78
mainly five basic rock properties; these are the porosity, compressibility, permeability,

resistivity and fluid saturation.

1. Porosity
Porosity is a very important property of reservoir rocks. The rock must have pore

spaces (voids) of such a size and character as to permit the storage of fluids. The absolute

porosity of a rock is the ratio of the pore volume within the rock to the bulk (total) volume of

the rock:
   Vp / Vb (4.4)

where

   absolute porosity, fraction


Vp : pore volume, m3

Vb : bulk volume, m3

Note that the difference between Vb and Vp is the volume of the rock matrix (grains), Vma.

Porosity can be classified according to its origin as follows:

Primary Porosity: Primary porosity is developed during the deposition of sediments.

Examples of primary porosity include the intergranular porosity of sandstone particles in a

clastic sediment, Fig. 4.4, and the oolitic porosity of limestone spheres.

Secondary Porosity: Secondary Porosity is developed after the sedimentation process as a

result of compaction, fracturing, chemical dissolution, cementation, recrystalization and

dolomitization (Fig. 4.5).

Not all pores within a rock sample are connected; some are completely isolated (Fig.

4.6). Since only the interconnected pores can supply oil, their volume is used to define the

effective porosity:

interconnec ted porevolume


 e =
bulk volume

Therefore, the effective porosity is less than or equal to the absolute porosity; and the effective

porosity is obviously the one used in the estimation of oil reserves.

79
Fig. 4.4: Intergranular porosity of sandstone.

80
Fig. 4.5: Types of secondary porosity.

81
Fig. 4.6: Interconnected and isolated pores within a rock matrix.

82
Measurements and Calculation of Porosity
There are various methods of determining the porosity of a core sample. All of these

are designed to calculate two out of the three required parameters: bulk volume, pore volume

and matrix volume. Bulk volume can be calculated easily from the dimensions of the core

sample if it has a regular shape. Pore volume can be measured by saturating the core sample

with a suitable liquid and noting the increase in its mass.

Example 4.4 - Porosity Calculation

Problem: Calculate the porosity of a regular cylindrical core sample if the length of the core

is 3.84 cm, its diameter is 2.53 cm, and the mass of the dry core sample is 37.75 g. When the

core sample is vacuumed and saturated with distilled water of 0.997 g/cm3 density, the mass

of the core increased to 41.60 g.

Solution:

Bulk Volume
L = 3.84 cm

d = 2.53 cm
Vb = cross-sectional area x length

   (d/2)2 L

   (2.53/2)2 x 3.84

= 19.30 cm3

Pore Volume
Mass of dry core sample = 37.75 g

Mass of saturated core sample = 41.60 g


Mass of water (mw) = 41.60 - 37.35

= 4.25 g
Volume of of water (Vw) = mw/w

= 4.25/0.997

= 4.26 cm3

83
This volume should be equal to the pore volume of the core sample since the water filled its

pores completely. Therefore,


Vp = Vw = 4.26

Porosity of the core sample () = Vp/Vb

= 4.26/19.30

= 0.22

= 22%

It should be noted that this porosity is obviously the effective porosity.

Reservoir quality is sometimes classified according to rock porosity in the following

manner:

* good porosity :  > 15%

* fair porosity : 15% > > 10%

* poor porosity : 10% > 

2. Rock Compressibility
Generally, the porosity of sedimentary rock varies with the depth of the rock.

Sediments which are deeply buried exhibit lower porosity values than sediments which are

buried at a shallower depth. This change in porosity is due to the compressibility and

compaction of the rocks. Besides the external stress of overlying rocks, reservoir rocks are

also subjected to the internal stress exerted by the reservoir fluids. During production, the

drop in reservoir pressure results in a change in the internal stress of the rock which introduces

changes in the pore and bulk volumes of the rock. Based on these changes, the bulk rock

compressibility is defined as the change in bulk volume per unit bulk volume per unit change

in pressure.
1 Vb
cb = - (4.5)
Vb P

where
cb : rock compressibility, kPa -1
Vb : average bulk volume, m3

84
 Vb : change in bulk volume, m3

 P : change in pressure, kPa.

Rock compressibilities range usually between 35 x 10-6 to 70 x 10-6 kPa-1; and they

should be taken into consideration when measuring the porosity of a core sample in the

laboratory or estimating the remaining reserves in a reservoir.

3. Permeability
Permeability is the ability of a rock to allow fluids to flow through its pores. A rock is

termed impermeable if the rate of a fluid passing through it is negligible under practical

conditions; and unless an oil bearing formation possesses good permeability, its commercial

value may not be attractive.

Permeability is defined according to Darcy's law which is the basic equation describing

the flow of fluids through porous media:

q k P
= (4.6)
A  L

where

q : volumetric flow rate of the fluid, m3/s

A : cross-sectional area of the porous medium, m2

 L : length of the porous medium, m

P : pressure drop across L, Pa

k : permeability of the porous medium, m2

  viscosity of the fluid, Pa.s

A permeability of one m2 means that a one meter long rock having a cross-sectional

area of one m2 will transmit a fluid of one Pa.s viscosity at a flow rate of one m3 per second

under a pressure difference of one Pa. This is clearly shown in the following calculation:

qL ( m3 / s )( Pa. s )( m )
k = = = 1.0 m2
AP ( m 2 )( Pa )

85
The unit of m2 is very large, however, for normal reservoir permeability. Therefore,

permeability is expressed in terms of m2 :

 m2 = (10-6 m)2 = 10-12 m2

The permeability of a reservoir rock is generally determined in the laboratory using various

types of permeameters. A simple permeameter is shown in Fig. 4.7.

Example 4.5 - Permeability Calculation

Problem: Calculate the permeability of a core sample from the following data:
Volume of water passed through the core sample, Vw = 10.00 cm3

Flow time (t) = 30.00 s

Pressure drop during flow ( P) = 200 kPa

Diameter of core sample (d) = 1.90 cm

Length of core sample (L) = 1.90 cm

Viscosity of water at 25oC (  0.895 mPa.s

Solution:
Cross-sectional area (A) = (d/2)2

   (1.90/2)2

= 2.84 cm2
Flow rate (q) = Vw/t

= 10.00 / 30.00

= 0.33 cm3/s
k P
Permeability (k ) =
 L
( 0.33 x106 m3 / s)( 0.895 x103 Pa.s)(190
. x102 m )
=
( 2.84 x104 m 2 )( 200 x103 Pa )

= 9.88 x 10-14 m2

 0.1 m2

86
Fig. 4.7: Schematic of a permeameter.

87
4. Resistivity
The resistivity of a core sample can be computed by applying a potential drop across

the core sample and measuring the resultant current passing through it.

R = V A/(I L) (4.7)

where

R : resistivity of the core sample, . m

A : cross-sectional area of the core sample, m2

L : length of the core, m

V : potential drop, Volt

I : current, Amp.
When the core sample contains both oil and water, its resistivity is denoted Rt. If an

identical core sample obtained from the same formation is fully saturated with water, its
resistivity – denoted Ro – can also be measured using the same technique. The water

saturation of the first core sample can then be computed using Archie’s law (Equ. 4.1):
Sw = Ro / R t

The laboratory determined water and oil saturations of a reservoir can then be compared with

the values obtained with the electrical resistivity log.

5. Fluid Saturation
In addition to the resistivity log and the core resistivity measurement, the fluid

saturation of a reservoir can be estimated directly from a core sample. The reservoir core

sample is first heated inside a retort distillation apparatus (Fig. 4.8). The oil and water will

evaporate and condense into the graduated cylinder where their volumes are measured. With

the pore volume of the core sample already determined, the oil and water saturation can then

be computed. Remember that:


Sw = Vw / Vp (4.8)

and
Sw + So + Sg = 1.0 (4.9)

88
Fig. 4.8: A retort distillation apparatus.

89
4.7 Fluid Properties
Detailed knowledge of the reservoir fluid properties is of vital importance for the

efficient exploitation of an oil reservoir. Such knowledge influences the way wells are

completed and produced and the various strategies of managing the reservoir.

Relative Density
Since no two oils are identical, a very wide range of physical and chemical

characteristics have been observed across the fields of the world. To keep matters simple, one

criterion has been adopted to characterize the quality of an oil; this is the oil's relative density
(  o ) defined as:

o = o / w

where
o = density of oil, kg/m3 w = density of water, kg/m3

However, the relative density is seldom used in the petroleum industry; rather it is

incorporated into another unit defined by the American Petroleum Institute and known as the

API gravity:
1415
.
API gravity = - 131.5
o

Since water has a relative density of 1, its API gravity is 10. Therefore, oils heavier than

water will have API gravities less than 10, and oils lighter than water will have API gravities

greater than 10. Usually, the higher the API gravity of an oil is, the lighter - and, hence, more

valuable - it becomes. Saudi Arabian crude oils, for example, are classified into the standard

grades listed in Table 4.1. Other important oil properties are the solution gas/oil ratio, the

formation volume factor and the viscosity.

Solution Gas/Oil Ratio


Reservoir oil always contains some quantity of natural gas dissolved in it. The gas is

kept in solution as a result of the pressure in the reservoir; and it will remain so as long as the

90
pressure is above a certain value called the bubble point pressure (Pbp). If the pressure drops

below Pbp, gas will evolve (bubble out or separate) from the oil.

Table 4.1: Grade classification of Saudi Arabian crude oils

Grade API Gravity Relative Density

Arabian Super Light > 40 < 0.82

Arabian Extra Light 37 - 40 0.82 - 0.84

Arabian Light 34 – 37 0.84 - 0.86

Arabian Medium 30 – 33 0.86 - 0.88

Arabian Heavy < 30 > 0.88

The amount of gas that is released from the oil depends on how far the pressure is
dropped below Pbp; the lower the pressure, the more gas released. If the oil is brought from

reservoir pressure directly to atmospheric pressure, all the dissolved gas will be released. At

an intermediate pressure, some gas will be released and some will remain in solution.

Therefore, the amount of gas remained in solution - expressed as a ratio of gas volume to oil
volume - should be constant above Pbp and should decrease below Pbp down to zero at

atmospheric pressure (zero gauge pressure). For the sake of uniformity, the volume of gas

dissolved or released is expressed in standard (at 100 kPa pressure and 15 oC temperature)

cubic meters, and the volume of oil is in stock tank cubic meters. Stock tank conditions are
identical to the standard conditions. The solution gas/oil ratio (Rs) is, therefore, defined as:
volume of dissolved gas, std. m3
Rs =
volume of stock tan k oil, ST m3

Fig. 4.9 shows a typical variation of Rs with pressure.

Formation Volume Factor


As oil travels from the reservoir, up the well and finally arrives at the stock tank its

pressure is decreasing continuously. With this decrease in pressure, the oil releases dissolved

91
Fig. 4.9: Variation of solution gas-oil ratio with pressure.

92
gas and its volume shrinks. This is similar to what happens to a soft drink after it loses its gas.
Naturally, the degree of shrinkage increases the further the pressure is taken below Pbp

reaching its maximum at the stock tank where the pressure is atmospheric. Therefore, a m3 of

oil leaving the reservoir will end up, for example, as 0.6 or 0.7 m3 of oil in the stock tank.

Conversely, a m3 of oil reaching the stock tank may have been 1.4 or 1.6 m3 of oil when it left

the reservoir rock. The ratio of the volume of oil at the reservoir conditions to its volume at the
stock tank conditions is called the formation volume factor (Bo) and is expressed as:
Bo = Vor / Vo

where
Vor : volume of oil at reservoir conditions, m3

Vo : volume of oil at stock tank conditions, m3

Bo is, therefore, always greater than or equal to 1 as shown in Fig. 4.10. The utility of Bois

that it allows us to calculate the amount of processed (stock tank) oil we will obtain from a

given volume of reservoir oil.

Viscosity
Viscosity is a measure of the resistance of a fluid to flow. The more viscous a fluid is,

the more difficult it gets for it to flow through the reservoir rock, the well or the surface pipes.

In qualitative terms, the larger the viscosity is, the greater the force (pressure) needed to push

a given flowrate through those media. The viscosity of all fluids vary with temperature and

pressure; but for crude oil, it is especially influenced by pressure because of dissolved gas.

Normally, the oil becomes less viscous as more gas is dissolved in it. Therefore, the oil
viscosity increases if the pressure is reduced below Pbp because the oil starts losing gas. The

maximum oil viscosity is obviously encountered in the stock tank as shown in Fig. 4.11.

4.8 Reservoir Water


Water is always present in the pore spaces of petroleum reservoirs. Knowledge of the

properties of this water is very important to many petroleum engineering calculations. The

composition of reservoir water varies from one reservoir to another depending on the amount

93
Fig. 4.10: Variation of oil formation volume factor with pressure.

94
Fig. 4.11: Variation of oil viscosity with pressure.

95
and type of dissolved solids. All reservoir waters contain dissolved solids, primarily sodium

chloride. Reservoir waters have been reported with total dissolved solids up to approximately

300,000 parts per million (ppm) of solution (30% of the total mass), while seawater contains

approximately 35,000 ppm total solids. Table 4.2 lists typical compositions of waters from a

variety of sources.

Table 4.2: Typical Water Analyses (ppm)

Ion Fresh Water Ocean Water Arabian Gulf Abqaiq Arab-C


Water Reservoir Water
Ca  50 420 641 27257
Mg  20 1,300 1860 3730
Na  300 10,710 16840 44760
K 30 - 624 -
SO4 50 2,700 4100 734
Cl  400 19,410 28800 126810
Br  50 - - -
I 50 - - -
Total 1,000 34,540 52,030 204,391

96
Chapter 5

PRODUCTION ENGINEERING

5.1 Introduction
Production engineering deals with the various operations required to maintain the well

in a good producing condition, and to process the oil and/or gas before delivery to the end

user. In other words, the path beginning at the point where the oil approaches the wellbore and

ending at the point where it enters a pipeline lies within the domain of production engineering.

Operations which deal with the well - or the reservoir portion that is close to the borehole -

are normally classified as Workover Operations; those involving processing the oil after it

leaves the well are classified as Surface Operations. But before these operations are explained,
some basic equipment that is installed in the well to facilitate production are discussed first.

5.2 Downhole Equipment


As explained in the preceding chapter, all that remains after the well has been

completed is to tie it to the production network of the field. This procedure is not as simple as

it may sound, but rather requires special equipment that has to be installed at various depths

inside the well. One such equipment is the tubing, which is a small-diameter pipe that runs

along the entire depth of the well and is used as the production string, i.e., to conduct the

produced fluids to the surface.

Theoretically, it is possible to use the production casing as a production string, but

such option is rarely used because of the high risk of damage to the casing. Casing damage

results primarily from corrosion caused by water and/or acidic gases accompanying oil

production. Such corrosion could develop holes and cracks in the casing through which oil

may leak out or ground water may leak into the well. While small holes and cracks in the

casing can be easily repaired, extensive casing damage could necessitate abandoning the well

since replacing a cemented casing is impossible. For this reason the tubing is used instead.

Figure 5.1 shows a basic tubing production string. The casing-tubing annulus is

isolated with a production packer which prevents flow and pressure communication between

97
Fig. 5.1: Tubing production string with packer.

98
the formation and the annulus. The production packer ( also called the tubing packer ) is a

device that seals the casing-tubing annulus or the tubing-borehole annulus while allowing

fluids to flow through the packer. The seal is made by inflating a rubber element in the packer

using gas pressure or other mechanical means. The packer is usually installed between two

tubing joints close to the producing formation. The tubing can be easily retrieved for repair or

replacement since it is light and has no cement bonding.

At some depth below the surface another special device is usually installed in the

tubing. This is the sub-surface safety valve (SSV) and its purpose is to stop flow through the

tubing in the event that the wellhead assembly ( explained below ) is damaged. The SSV can

be the direct-control type which is set to close automatically if the flowrate exceeds a certain

value. It can also be the remote-control type which can be operated from the surface. Should

the wellhead assembly become destroyed by accident, wear and tear or even an act of

sabotage, the SSV could prevent the disastrous consequence of a well blowout or a well fire.

5.3 Wellhead Assembly


The wellhead assembly is a group of valves, gauges and special connections which are

attached to the top of the well at the surface (Fig. 5.2). The basic function of this equipment is

to allow full control over the flow of produced fluids, which are under high pressure and could

contain toxic gases such as carbon dioxide and hydrogen sulfide. It also allows access to the

tubing and the casing-tubing annulus for whatever later need. Through special connectors in

the wellhead assembly, logging tools, well testing tools and other well maintenance devices

can be lowered into the well.

The Produced fluids (oil, water and gas) are transported from the well through a

flowline connected to the wellhead assembly. A special device called the choke is installed at

the inlet to this flowline as shown in Fig. 5.2. The function of the choke is to set the well’s rate

of production at a specific level. If a higher or lower production rate is desired, the choke size

is selected accordingly. With the installation of the wellhead assembly and tying the well in,

the well is now ready for production.

99
Fig. 5.2: Wellhead assembly.

100
5.4 Natural Flow
Fluids flow from one point to another when a driving force is provided that is

sufficient to overcome the resistance to flow that exists in between. In petroleum wells, the

driving force is the reservoir pressure; and the main resistance to flow between the reservoir

and the surface is the hydrostatic pressure created in the well by the weight of the oil column

in the tubing.

Some reservoirs posses sufficient pressure initially to overcome the hydrostatic

pressure and push the oil to the surface and beyond, providing good production rates at good

surface pressures. Wells producing in this manner are called natural flow wells, and such

desirable situation may last for many years. Some wells in the Middle East, for example, have

been producing by natural flow for more than forty years. The ability of a well to produce by

natural flow depends obviously on the reservoir pressure, the oil properties, the required

production rate, the pressure required at the surface to push the oil through the surface

facilities, and many other factors.

Other reservoirs, on the other hand, do not have sufficient pressure to produce by

natural flow. After a well in such a reservoir is completed and tied in and the drilling fluid is

pumped out or displaced by gas, the well is put on production. Oil will start flowing from the
reservoir to the borehole due to the difference between the reservoir pressure ( Pr ) and the

borehole pressure ( Pw ), which is nearly atmospheric. As the tubing gradually fills up with oil,

Pw increases becauses of the hydrostatic pressure created at the well’s bottom. However, the
oil level soon reaches a height when Pw equals Pr ; at which point, oil ceases to flow from the

formation. In such a reservoir, oil can be brought to the surface only if supplemental energy is

provided from the surface. This is achieved by various techniques called artificial lifting

methods. These methods are also resorted to if a well exhausts the reservoir pressure and

ceases to flow naturally.

5.5 Artificial Lifting Methods


These methods fall under basically two categories according to the approach in which

the problem is handled. In some methods, the pressure at a certain point in the well is reduced

101
by decreasing the density of the oil above that point. In other methods, the pressure near the

top of the oil column - which is essentially atmospheric - is boosted to push the oil to the

surface. Two primary examples of artificial lifting are given below.

Gas Lift: In this method, the density of the oil column in the wellbore is reduced by mixing

the oil with gas. At a specific depth, a special valve called the gas lift valve is installed in the

tubing. Compressed natural gas is then injected into the casing-tubing annulus with a

sufficient pressure to enable it to enter the tubing through the valve. As the gas bubbles

upward through the oil, the oil in the tubing above the gas lift valve is lightened by mixing

with the gas causing an overall reduction in the hydrostatic pressure in the well. The
consequent reduction in Pw enables the low reservoir pressure to raise the column of liquid

and gas mixture to the surface aided by the force of the expanding gas (Fig. 5.3).

The pressure at the gas lift valve can be calculated by the following equation:
Pv = mix g h + Pt (5.1)

where
Pv : pressure at the gas lift valve, Pa

mix : required density of the oil-gas mixture, kg/ m 3

h : depth of the gas lift valve, m


Pt : desired pressure at the surface, Pa

The density of an oil-gas mixture can be estimated from:


mix =   g + (1-  )  o (5.2)

where
g : gas density, kg/ m 3

o : oil density, kg/ m 3

 : volume fraction of gas in gas-oil mixture

Example 5.1 - Average Density Calculation


Problem: Given that the gas lift valve is installed at 300 m below the top of the oil column in

the tubing, and that this top is 100 meters below the surface, calculate (a) the pressure at the

102
Fig. 5.3: Mechanics of gas lift.

103
gas lift valve, (b) the required density of the oil-gas mixture, and (c) the amount of gas to be

injected for every m 3 of oil produced. The density of the crude oil is 0.85 g/cm3, density of

gas is 0.0012 g/cm3, and the required pressure at the surface is 980 kPa.

Solution:
a. Pressure at gas lift valve:

Before gas injection, the pressure at the valve must be equal to the hydrostatic pressure of the

oil column above it:


Ph = o g h
106
= 0.85 x x 9.807 x 300
103
= 2500785 N/ m2

 2500 kPa

Therefore,
Pv = Ph = 2500 kPa

b. Density of oil-gas mixture

After gas injection, the pressure at the valve must be equal to the hydrostatic pressure of the

oil-gas column above it ( which must extend to the surface) plus the required surface pressure.

Therefore, from Equ. 5.1


Pv  Pt
mix =
gh
2500000  980000
=
9.807 x 400

= 387.5 kg/m3

= 0.387 g/cm3

Which means that the oil-gas mixture must have this density to enable it to reach the surface at

a pressure of 980 kPa.

c. From Equ. 5.2, the gas volume fraction in the mixture is:
 o   mix 0.85  0.387
 = = = 0.556
o   g 0.85  0.0012

which means that about 1.257 m3 of gas must be injected for every 1 m3 of oil to be produced.

104
Downhole Pump: In this method, a pump is installed into the tubing somewhere below the

top of the oil column and is used to raise the oil pressure to enable it to reach the surface.

Several pump designs and operating methods are available. For high production rates, the

electric submersible pump is employed. This is a slim pump of the impeller type which is

supplied with power by a cable running from the surface. For low rates, a sucker-rod pump is

used. This is a plunger type pump whose rod is moved up and down by an electric motor at the

surface.

5.6 Surface Production Facilities


An oil well seldom produces pure oil; there is nearly always some quantity of gas,

water and sediment mixed with the oil. The gas could have come from the gas cap, a region in

the reservoir which contains free gas, or could have been originally dissolved in the oil.

Dissolved gas usually starts evolving from the oil during its flow through the tubing where the

pressure drops below the bubble-point. Reservoir water may be produced as vapor in the gas

or as liquid water which is either free or sometimes emulsified with the oil. Fluid flow through

the reservoir may cause grains of sand, silt, and clay to break loose from the reservoir rock and

become entrained with the fluids as sediment. The oil may also contain other contaminants
such as CO2, H2S, dissolved salts and heavy metals.

The main purpose of the various surface production facilities is to separate the

produced mixture into free oil, gas, and water streams and to process the hydrocarbons (oil

and gas) to an acceptable state of purity.

5.7 Gas-Oil Separation


An oil containing a quantity of dissolved gas is extremely volatile and, therefore,

highly inflammable and dangerous. Thus, the gas (free or dissolved) accompanying the oil

must be completely removed first. This is achieved in the Gas-oil Separation Plant (GOSP)

whose main component is the gas-oil separator. The separator is a vessel large enough to

cause the oil pressure to drop to near atmospheric and, thus, bring all the dissolved gas out of

105
solution. Due to gravity, the oil would then collect at the bottom of the separator while the gas

rises to the top; and both streams are removed for further processing.

Separators come in different designs and are classified according to their shape as

horizontal, vertical or spherical. Of these, horizontal separators are the most common whereas

spherical separators are very rarely used.

Horizontal Separator : This type is shown in Fig. 5.4. The oil and gas enter the vessel at one

end where they pass through deflectors. The separator should be long enough to allow

complete separation at the given flowrate and gas-oil ratio.

Vertical Separator : The oil and gas enter the vertical separator (Fig. 5.5) at a point near the

top. Inside, they are forced into a circular motion where centrifugal and gravitational forces

cause oil and gas separation. Because they occupy a relatively small area, vertical separators

are most common in offshore platforms where space is limited.

Spherical Separator : The working principle of this type is similar to that of the other two;

but, because of its shape, it is used mainly for high-pressure separation. Nevertheless, its use is

limited due to its high fabrication and installation costs.

Multi-Stage Separation
For an oil with a large solution gas-oil ratio, complete gas-oil separation may not be

achieved in one stage, or no ordinary separator can handle the large quantity of gas. In this

case, separation is conducted over several stages in series where the oil leaving a separator is

flowed into the next. The stock tank is always considered as the final stage. Therefore, two-

stage separation involves one separator and the stock tank; three-stage separation involves two

separators and the stock tank; and so on. The most commonly used arrangement is three-stage

separation.

One purpose of multi-stage separation is to reduce the oil pressure one small step at a

time and, hence, evolve the gas in portions that can be handled by each separator. Another

purpose is to minimize the loss of intermediates into the gas stream. The intermediates are

106
Fig. 5.4: Horizontal gas-oil separator.

Fig. 5.5: Vertical gas-oil separator.

107
valuable constituents of the oil which include the butanes and pentanes. The larger and more

abrupt the pressure drop is ( as when few stages are used), the more intermediates evaporate

into the gas which would then necessitate additional processing of the gas to recover them. On

the other hand, smaller pressure drops ( as when several stages are used) preserve most of the

intermediates in the oil stream.

5.8 Oil and Gas Treatment


After the separation process, the oil and gas streams invariably require further

treatment for purification purposes. Some of the most common purification processes are

explained below.

Water Removal
In many fields, formation water is produced with the oil in various proportions. The

amount of water is expressed as the water cut, which is defined as the volume fraction of free

water in the total produced liquids. Because it is in the liquid state, this water remains with the

oil phase as it passes through the gas-oil separator. Since formation water is usually salty, it is

very corrosive and must be removed from the crude oil before the oil is pumped to the refinery

or storage tanks. Also, oil sales contracts usually specify a maximum allowable water content

of 2% by volume before the oil is considered as pipeline-quality oil.

Separation of free water takes place normally in a simple free-water knockout drum

where the oil/water mixture is fed continuously. Due to their higher density, free water

droplets settle out within a reasonably short time and collect at the bottom of the drum where

the water is removed.

Separation becomes more difficult, however, when oil and water are produced as an

emulsion. An emulsion is defined as a uniform mixture of two immiscible liquid phases in

which one phase is distributed as globules in the other. The mixture is maintained in a stable

state by the presence of an emulsifyer, which is a compound that keeps the water - or oil -

droplets in a dispersed state. Emulsions are formed when the oil contains a natural emulsifyer.

Most oil field emulsions are the water-in-oil type where individual water droplets are

108
dispersed in the oil. An inverted, or oil-in-water, emulsion can also occur, especially when the

water cut is very high.

Emulsion Treatment
In order to "break" the emulsion, three types of processes are generally utilized:

thermal, chemical or electrostatic treatment. The basic principle of all these methods is to

weaken or destroy the film of emulsifyer molecules which surrounds a water droplet and

prevents it from coalescing with other droplets. Once this is achieved, the water droplets could

combine together to form larger globules that would be heavy enough to settle out of the oil.

Thermal treatment involves weakening the emulsifyer film by heating the emulsion;

chemical treatment involves adding a de-emulsifyer, a compound that neutralizes the

emulsifyer; and electrostatic treatment employs a high electric voltage to break it. The method

adopted for any particular treatment depends on the severity of the emulsion problem. Usually

one process is sufficient; but sometimes a combination of the processes is used for stubborn

emulsions. For instance, the thermal as well as the chemical methods could be used. In this

case, the emulsifyer-breaking chemical is usually added to the produced oil before it reaches

the GOSP. This ensures thorough mixing with the emulsion and, consequently, weakening and

breaking the emulsifyer film. When the weakened emulsion is heated after it leaves the gas-oil

separator, the water globules separate readily from the crude oil.

Basic Sediment and Water


Most produced oils contain variable amounts of sand, silt, and other formation solids

that become mixed with some water and oil to form sludge. This sludge, called Basic Sediment

and Water (BS&W), usually settles at the bottom of the separator, emulsion treaters and stock

tanks where it is periodically removed.

Sour Crude
Sour crude is oil which contains a relatively high amount of sulfur. Sulfur can be in the
form of dissolved hydrogen sulfide (H2S) or as sulfur-containing compounds like the

109
mercaptans and alkyl sulfides. In whatever form, sulfur compounds are corrosive and could be
poisonous; furthermore, H2S has a very unpleasant smell. Refineries do not accept crude oil if

it contains more sulfur than a specified limit since sulfur tends to damage catalysts used in

some refinery processes.

Hydrogen sulfide is usually removed form oil by means of stripping. A stream of sweet
- low H2S content - natural gas is mixed with the oil in a stripping tower. This allows the gas

to strip the H2S from the oil which is then separated as sweetened oil. The H2S-rich gas is

then processed further to remove the H2S from it.

Other forms of sulfur are removed from the oil by means of reaction with an alkaline
base like NaOH or Ca(OH)2 solution. The base solution is thoroughly mixed with the crude in

a vertical mixing tower leading to sulfur exraction in the form of Na2S or CaS. The mixture is

then fed to a settling tank where the Na2S solution is removed from the bottom and the

sweetened crude is removed from the top. The processed crude oil is constantly checked for its

sulfur content; and if it is more than the accepted level, the crude is recycled back into the

mixing tower.

Sour Gas Treatment


H2S is removed from sour natural gas by treatment with mono ethanol amine (MEA)

solution which dissolves H2S gas. The MEA solution is purified of the H2S by a separate

process and is recycled back for reuse. The H2S is reacted with alumina at a high temperature

to convert it into sulfur.

Condensate (liquid) Separation


Some gas fields produce gas which is rich in intermediates. When the gas reaches the

surface where the temperature and pressure are low, the intermediates condense into liquid

which is called the condensate. This condensate, which flows with the gas stream like mist, is

extremely valuable and must be recovered. There are various types of separators that remove

condensate from the gas stream; one of which is called the scrubber. This is a separator

designed to handle streams with a large gas-liquid ratio. The basic operation is to pass the

110
stream through very fine electrically-charged screens which collect the liquid droplets. The

stream is usually passed first through another screen or oil bath to remove any dust, scale or

rust before the actual removal of condensate takes place.

5.9 Oil and Gas Metering


Oil and gas production rates from a well, a group of wells or a field are determined on

a regular basis. This data is collected for a variety of reasons such as reserve calculations,

reservoir and well monitoring, sales contract fulfillment and future planning. The rates could

be measured and recorded continuously by devices installed in the field or at scheduled times

by field crews according to certain procedures.

Oil Metering : Metering oil production can be by either manual or automatic means. Manual

measurement merely involves measuring the oil level in the stock tank before and after any

charge or discharge of oil from the tank. Automatic measurement relies on a liquid flowmeter

installed on the pipeline carrying the crude to the stock tank. The oil rate from an individual

well is determined by diverting the well’s production to a portable separator and then

measuring the separator’s rate.

Gas Metering : The most widely used device for metering natural gas production is the

orifice meter, which is a steel plate with a small hole in its center. The orifice meter is inserted

into the gas pipeline and the pressure drop across it is measured. The gas flowrate is then

computed by a standard formula utilizing this pressure drop. Another method utilizes the

pressure drop along a known length of pipeline which is then applied to a different formula.

In modern installations, electronic pressure and temperature gauges feed electrical

signals to a computer that processes them into flowrate, cumulative volume, temperature,

pressure, date and time data which is recorded continuously.

111
5.10 Crude Oil Gathering and Storage
Each oil or gas well is served by a flowline which carries the well’s production to the

area GOSP. At the GOSP, all flowlines converge on a single large pipe called the header

which leads to the main separator. The flow lines are also connected, via side branches, to

another header which leads to a small separator called the Test Trap. This vessel is used to test

the production rate, gas/oil ratio and water cut of a single well by diverting its production to

the test trap header. This network of pipes, headers and diversion valves is called the

Manifold.

Depending on the size of the field, several GOSPs could be required to process the

field’s production. Once the gas is separated from the oil, the gas flows through a separate

pipeline to the gas plant where natural gas liquids (NGL) are recovered. The separated oil is

processed further to remove free water, emulsion and other contaminants before it is pumped

to the stock tanks. From there, it could be pumped to an export terminal, refinery or directly to

a consumer (power station, petrochemical plant, etc.).

5.11 Transportation
From the stock tank to the consumer, crude oil could travel thousands of kilometers

across continents and oceans. The modern petroleum industry employs a variety of means

such as the pipeline, ocean tanker, railroad car and tanker truck to move oil from field to

market. However, the first two means account for the largest share in the global transportation

network.

The pipeline is the cheapest method of moving oil over land. Although the

construction cost is very large, the low operating cost and long sevice life make it more

economic than the railroad or truck especially for long distances. Pipelines vary in size

according to the required capacity and pressure; forty-eight-inch ( 120 cm ) and sixty-inch

( 150 cm ) diameter pipelines are common. Over long distances, pressure losses along the

length of the pipeline are too great for the flow to continue. For this reason, pump stations are

installed at selected intervals to boost the pressure and maintain the flow.

112
Pipelines can carry crude oil or oil products. It is possible to move shipments of two or

more different oil products in one pipeline by pumping one shipment at a time, separating it

from the previous shipment by a plug-type device called the Peg.

Although pipelines have been built across lakes and shallow seas, they are impossible

to lay across oceans. For trans-oceanic transportation, the oil tanker is the most efficient and

economic means. The oil tanker is a ship whose hull is converted into a tank of many

compartments. When the tanker is not carying oil, its tanks are filled up with water to maintain

ballast. Before reaching the loading port, the tanker gets rid of its ballast water. Loading a

tanker with oil can take up to 3 days, and is conducted at an onshore or offshore terminal.

The first oil tanker appeared late in the ninteenth century, but it did not become a

prominent link in the oil transportation business untill the second world war. Then, the Middle

East was becoming a major oil producing region supplying Europe and the Far East. This

necessitated construction of more and more tankers with ever increasing capacities. Between

1945 and 1975, the average tanker capacity increased from 34,000 tons ( approx. 260,000

barrels ) to 540,000 tons ( approx. 4 million barrels ).

5.12 Well Stimulation


Oil and gas wells require regular maintenance to keep them in optimum producing

conditions. Well maintenance procedures are given the special name of Workover Operations.

An especially important group of workover operations are well stimulation techniques which

are aimed at improving the productivity of a well by decreasing the resistance to flow near the

wellbore. Well stimulation can be performed right after the well is completed or periodically

during its life. In this section, some of the most common stimulation techniques are discussed.

Acid Washing
In this operation, a small quantity of acid - usually hydrochlric acid - is pumped and

withdrawn back and forth across the formation face for a period of time. It is usually

performed immediately after the well is completed to clean the formation face of any mud

cake and remove acid-soluble particles that may be present in the wellbore.

113
Matrix Acidizing
In some formations, mud filtration may cause a drastic reduction in permeability

around the wellbore because of plugging of rock pores by clay particles. In such cases, a

quantity of acid is injected into the formation at a low pressure followed by a shut-in period.

The acid dissolves the particles plugging the pores as well as some of the rock itself causing

enlargement of the formation pores. When the well is put back on production, a significant

improvement in productivity is noticed.

In general, 15% concentration hydrochloric acid (HCl) is used for limestone and

dolomite formations and a mixture of 12% HCl plus 3% hydroflouric acid (HF) is used for

sandstone formations. Typical acidizing reactions are:


2 HCL + CaCO3  CaCl2 + H2O + CO2 (limestone)

4 HCL + CaMg(CO3 )2  CaCl2 + MgCl2 + 2H2O + 2CO2 (dolomite)

6 HF + SiO2  H2SiF6 + 2H2O (sandstone)

Example 5.2 - Minimum Acid Volume Calculation


Problem: A damaged formation 36 m thick needs to be acidized 60 cm around the well bore.

What is the minimum volume of acid needed for this well whose diameter is 16 cm? The

formation porosity is 20%.

Solution:
 [ rac  rw ]h
2 2
The pore volume of acidized portion =

= . [(68)2  (8)2 ] 3600 x 0.20


3142

= 10,315,814 cm 3
 rw h
2
The volume of the wellbore =

= . (8)2 3600
3142 = 723,917 cm 3

Acid volume required = 10,315,814 + 723,917 = 11,039,731 cm 3

 11.0 m 3

Hydraulic Fracturing
Some formations have such low permeabilities that even matrix acidizing does not

improve productivity significantly. For such cases, fractures in the formation are created

114
hydraulically - by means of liquid pressure. These fractures, which are normally parallel to the

vertical plane, radiate from the well bore and do not extend deep into the formation. The

hydraulic fracturing procedure involves the injection under very high pressures of a fracturing

fluid into the formation. The injection rates and pressures are so high that the formation cannot

withstand the imposed stresses causing it to fracture, at which point injection is stopped.

Naturally, when the pressure is reduced, the fractures will close up again. Therefore, the

fracturing fluid is usually mixed with small solid particles ( the propping agent ) which are

carried into the formation and embedded inside the fractures to keep them open. However,

sometimes the propping agent is stuck into soft formations or crushed inside hard formations,

thus, failing to keep the fractures open.

Acid Fracturing
In this technique, acid is used as the fracturing fluid. The acid reacts with the walls of

the fracture creating open channels that do not close up after the high injection pressure is

released.

115
Chapter 6

PETROLEUM RESERVOIRS

6.1 Introduction
A petroleum reservoir is that portion of a trap which contains oil and/or gas. It is

sometimes referred to as a field, although technically this term refers to the area at the surface

that encompasses all wells and production facilities. In one field, there may exist under the

surface several reservoirs of different types and depths, and some of them may overlay each

other.

The responsibility of a petroleum engineer does not end when a reservoir is developed,

that is when several wells have been drilled and linked to a surface facility where the
reservoir’s production is processed, stored or shipped. Actually, this is only the beginning of a

long association between the engineer and the reservoir. In this association, which could span

a lifetime, the petroleum engineer has but one goal in mind: to produce as much oil from the

reservoir as economically possible. For this goal, the engineer monitors, tests and simulates

the performance of the reservoir in order to understand its mechanics and devise a

development plan to maximize its production. The branch of petroleum engineering that

applies modern technology to facilitate this goal is called Reservoir Engineering. In this

chapter, some basic definitions, concepts and techniques of reservoir engineering are

presented.

6.2 Petroleum Reserves


The amount of oil, or gas, stored within the pore space of a reservoir can be easily

estimated using a simple equation. This equation involves nothing but a computation of the

volume of the reservoir. If the area of a reservoir, A, its average thickness, h, its average

porosity,Ф , and its average water saturation,Sw , are all known, then
The bulk volume of the reservoir, Vb (m3) = Ah
The pore volume of the reservoir, V p (m3) = A h

The total volume of oil, Vo (m3) = A h  (1-Sw)

116
Since the total volume of oil is always estimated at stock tank conditions, Vo is divided

by Boi , the oil formation volume factor at the initial pressure of the reservoir ( Pi )

The total volume of oil, N (ST m 3 ) = A h  (1- S w ) / Boi (6.1)

Equation 6.1 is called the volumetric equation and the volume of oil estimated by it is called

the initial oil in place (IOIP). The IOIP, given the symbol N with units of stock tank m 3 ,

represents the total amount of oil that exists in the reservoir when it is first discovered. The

accuracy of our estimation of N depends naturally on the accuracy of the variables in the

volumetric equation, but these can be estimated fairly accurately after a few wells have been

drilled, cored, logged and tested.

The peculiar problem with petroleum reservoirs is that not all the oil they contain

initially can be ultimately produced. Many forces within the reservoir rock act to prevent some

of the oil to reach the wells. This means that the fraction of IOIP that is produced by the time

the reservoir is abandoned is never equal to 1. This fraction is called the oil recovery factor
with the symbol Er . The recovery factor is a parameter which is very difficult to estimate early

in the life of a reservoir. At best, a reservoir engineer can come up with a rough estimate that

is based on his experience and his understanding of the forces acting within the reservoir. His

estimate is also influenced by the petroleum recovery technology available at the time of

discovery.
With the above definition of Er , the oil reserves in a given field or reservoir are

estimated by multiplying IOIP into Er . Depending on how certain we are of our estimates of

these two parameters, two kinds of reserves are defined:

The Proven Reserves: This is the amount of oil which is certainly going to be recovered from

the reservoir. It is estimated from quantities of oil that are known to exist with great
confidence and assuming a conservative Er based on today’s technology.

The Probable Reserves: This is the amount of oil that is expected to be recovered from the

reservoir. It is estimated from quantities of oil that are believed to exist based on geologic and

117
seismic estimates of the size and properties of the reservoir. It also reflects speculated
improvements in Er as the technology improves.

The proven reserves is not a constant number; rather it changes with time as a result of

three factors. They decrease continuously as, obviously, more oil is produced from the

reservoir. However, as more oil is produced, more data become available that could give the

reservoir engineer more insight into the nature of the reservoir. Such data, which could also
come from new wells drilled, allows the engineer to revise his estimates of both IOIP and Er .

Lastly, technology developments in the form of new materials, processes, and practices
continuously improve Er .

6.3 Types of Petroleum Reservoirs


Petroleum reservoirs are classified into three basic types: oil reservoirs, gas reservoirs

and gas condensate reservoirs. Oil reservoirs produce mainly oil with some gas at a relatively

small gas/oil ratio. This gas is called associated gas because it is mostly liberated from the oil

as the oil is processed at the surface. Gas reservoirs contain and produce mainly gas, and the

producing gas/oil ratio is relatively large. Gas condensate reservoirs contain a mixture which

is rich in intermediate and heavier hydrocarbons. Under reservoir conditions, the mixture is

gaseous; but when the fluid reaches the surface, these hydrocarbons condense into liquid

resulting in medium gas/oil ratios. Another fluid that always exists in petroleum reservoirs is

water, and it is usually produced with the hydrocarbon fluids in both liquid and vapor forms.

6.4 Reservoir Driving Mechanisms


The basic force that pushes oil to flow through reservoir rock towards a well is

pressure. As long as the prevailing reservoir pressure is larger than the pressure at the bottom

of the well, oil flow shall occur. And the larger this pressure differential, the larger will the oil

production rate and, consequently, the oil recovery be. We have to remember, though, that the

pressure at the bottom of the well should be high enough to push the oil to the surface;

otherwise natural flow will not be realized.

118
Yet, with continued oil production, the reservoir pressure seldom remains constant.

This is dictated by the pressure-volume relationship of fluids, and can be compared with the

decrease in the pressure inside a balloon as air is released from it. Therefore, unless something

is done to the reservoir pressure, oil production rate should fall steadily with time.

The rate of decline of reservoir pressure with oil production varies from one oil

reservoir to another depending on the composition and satuartion of the various phases within

the reservoir pore space and the structure of the reservoir and its vicinity. Interaction between

these reservoir conditions creates a natural mechanism by which reservoir fluids are pushed

towards the wells. Such natural mechanism is called the Drive, and several forms have been

identified in oil and gas reservoirs. The following are three of the basic drives.

Solution-Gas Drive
This drive mechanism is also known as the depletion drive. It is characteristic of a

reservoir whose initial pressure is above its oil’s bubble-point pressure like the Arab-D

reservoir in Ghawwar field.

Soon after production commences, and depending on the size of the reservoir, the

pressure in such a reservoir begins to drop fast because of the relatively small compressibility

of oil. When the bubble-point pressure is reached gas starts evolving from the oil. More oil

production causes further pressure drop and more gas liberation. Initially, the gas saturation is

too small to enable it to flow. However, the free gas, having a larger compressibility than oil,

readily expands and becomes an extra driving force to push the oil and delay the pressure

decline.

As oil production continues, further pressure decline liberates more free gas (Fig. 6.1)

which, at some point, beings to flow towards the wells. From this point on, the production gas-

oil ratio increases rapidly with decline in reservoir pressure because gas flows easier than oil.

The gas-oil ratio reaches a maximum shortly before the reservoir pressure drops to a level

when fluid production ceases.

119
Fig. 6.1: Solution gas drive reservoir.

Fig. 6.2: Production trends - solution gas drive.

120
Dissolved gas drive is the least efficient of all drive mechanisms; it seldom yields more

than 10% in ultimate recovery. Figure 6.2 summarizes the characteristic trends of reservoir

behavior during the life of a typical dissolved-gas drive reservoir.

Gas-Cap Drive
A reservoir whose initial pressure is below its oil’s bubble-point pressure would have a

free gas phase, and this gas would naturally accumulate above the oil forming what is called a

Gas Cap - Fig. 6.3. Since oil wells in such a reservoir are normally completed in the oil zone,

no free gas production is seen.

Oil production causes a pressure drop in the oil zone which, in turn, is communicated

to the gas cap. In response, the gas in the gas cap expands arresting the pressure decline

because of its high compressibility. The gas also displaces the oil downwards as it expands,

thus help drain the oil towards the wells. However, too high a production rate may draw the

gas prematurely into the well causing the gas-oil ratio to increase sharply.

The recovery from gas-cap drive reservoirs depends on the size of the gas cap; but, this

recovery is normally greater than that realized from solution-gas drive reservoirs and could

reach 40%. The Khafji reservoir in the offshore Marjan field is gas-cap drive reservoir, and

Fig. 6.4 demonstrates typical performance trends of such a reservoir.

Water Drive
Most oil reservoirs are underlain and/or surrounded by a water-saturated zone called

the Aquifer - Fig. 6.5. Aquifers vary in size and, depending on the continuity of the formation,

could cover thousands of square kilometers. Given sufficient time, a pressure drop in the

reservoir would be ultimately felt by the aquifer too, and aquifer water will expand just like

oil.

Although water generally has a smaller compressibility than oil, in the case of an

extensive aquifer water flow, or influx, into the reservoir can be large enough to replenish the

reservoir pressure. In such water-drive reservoirs, the pressure experiences a small drop

initially and then stabilizes at a certain value as the aquifer water begins to expand and

121
Fig. 6.3: Gas cap drive reservoir.

Fig. 6.4: Production trends - gas cap drive.

122
Fig. 6.5: Water drive reservoir.

Fig. 6.6: Production trends - water drive.

123
invades the oil zone - Fig. 6.6. The stabilization time is dictated by several factors, chief

among which are the sizes of the reservoir and the aquifer and the rock and fluid properties.

Water drive is one of the most efficient of the natural drive mechanisms yielding

recoveries as high as 75% when a large, responsive aquifer is hydraulically connected to the

reservoir. Figure 6.6 summarizes the characteristic trends occurring during the production life

of a typical water-drive reservoir like the Khafji reservoir in Saffaniya field..

It is worth mentioning here that oil reservoirs whose aquifers are too small or have too

low a permeability to contribute any significant water influx are called volumetric reservoirs.

These reservoirs are monitored and managed more carefully; they also require more effort in

order to achieve a satisfactory recovery.

Combination Drive
Very often, two or more of the basic drives act simultaneously in a reservoir - Fig. 6.7.

For instance, an under-saturated oil (above the bubble point) reservoir could be receiving good

pressure support from an aquifer. Similarly, a gas-cap reservoir could be assisted by an aquifer

too. For this combination-drive reservoirs, the recovery will be the resultant of the actions of

all drives involved.

It should be pointed out that in all oil reservoirs added cushioning to the decline in

reservoir pressure is also provided by the reservoir water and rock. These two ingredients do

expand as the pressure drops; however, their effect becomes insignificant should a very

compressive fluid, such as gas, share the reservoir’s pore space.

Besides the collection and analysis of reservoir data for the purpose of estimating the

IOIP, a prime responsibility of the reservoir engineer is to determine, and as early as possible

in the life of the reservoir, what type of drive exists in the reservoir. Not only is this

information necessary to arrive at a reasonable recovery factor for proven reserves estimation,

it would also help the engineer devise an optimum field development plan for maximizing

these reserves. Such plan could involve a careful drilling and completion program and

implementation of one or more of the recovery improvement techniques discussed at the end

of this chapter.

124
Fig. 6.7: Combination drive reservoir.

125
6.5 Material Balance
One of the powerful tools of reservoir engineering is the Material Balance Equation

which is an account of the fluids that enter and exit the reservoir. This equation allows the

reservoir engineer to determine whether the reservoir is volumetric or under water drive; and

if so, how strong this drive is. It can also provide an estimate of the IOIP as will be shown

shortly. This feature is especially useful when the variables in the volumetric equation are

known with reasonable certainty, usually early in the life of the reservoir.
The volume occupied by the oil in a volumetric solution-drive reservoir, Vo , is

mathematically equal to:


Vo = A h  (1- S w ) m3
Ignoring rock and water expansion, Vo should remain constant with time. Since the IOIP is

computed by:
N = Vo / Boi ST m 3

then
Vo = N Boi
If N p is the total quantity of oil (in ST m 3 ) produced by the time the reservoir pressure has

declined from Pi to P ( P > Pbp ), then the amount of oil left in the reservoir at P is N - N p .

Yet, this amount would still occupy a volume Vo in the reservoir due to oil expansion though

at a smaller density or, equivalently, larger formation volume factor. This concept is depicted

in Fig. 6.8 and is described by the following equation:


N Boi = (N - N p ) Bo

This equation is a simple form of the material balance equation which is rearranged to give
N p Bo
N =
Bo  Boi

Since the variation of Bo with pressure is known early on and with good accuracy, the

reservoir engineer can obtain a firm estimate of N from early pressure and production data
only without even a knowledge of Vo .

126
Fig. 6.8: Material balance as applied to volumetric reservoirs.

127
For volumetric gas reservoirs, the material balance approach can be utilized in a

similar fashion to obtain the initial gas in-place, G. The material balance equation applicable

in this case is:


G p Bg
G =
Bg  Bgi

Example 6.1 - Oil in Place and Fractional Recovery


Problem: When a volumetric oil reservoir was discovered, its pressure was 21,300 kPa.

After eighteen months of operation, 53 million ST m 3 of oil have been produced and the
pressure has declined to 18,500 kPa. If Bo at the initial pressure is 1.604 m 3 /ST m 3 and at the
current pressure is 1.753 m 3 /ST m 3 , and if the Pbp is 17,250 kPa, estimate the IOIP and the

current recovery.
Solution: Since the reservoir pressure is still above the Pbp , the simple material balance

equation can be applied


N p Bo
N =
Bo  Boi
53x10 6 x1.753
= = 623,550,336 ST m 3
1.753 -1.604
 623.6 million ST m 3

The recovery is
53,000,000
r = = 0.084997
623,550,336

or 8.5% of the IOIP.

6.6 Well Performance


In a reservoir of homogeneous properties, the flow around the well is radial as depicted

in Fig. 6.9. At steady state, Darcy’s law in radial coordinates is used to compute the flowrate

of oil into a well:


0.542867 h k ( Pe  Pw )
q =
 o ln( re / rw )

Where

128
Fig. 6.9: Radial flow of oil into a well.

129
q : reservoir flowrate, m3/d

h : average reservoir thickness, m


rw : wellbore radius, m

re : reservoir (external) radius, m

Pw : bottomhole (wellbore) pressure, kPa

Pe : pressure at reservoir boundary, kPa

o : oil viscosity, mPa.s

k : 2

Note that q is the flowrate of oil in the reservoir at reservoir conditions. The surface

(production) flowrate in STm3 per day is computed by simply introducing the oil formation
volume factor, Bo , as follows:

qsc = q / Bo

where
qsc : surface production rate, ST m3/d

Bo : oil formation volume factor, reservoir m3 / ST m3

Example 6.2 - Flow Rate Calculation


Problem: Calculate the maximum possible production rate from an oil well in a reservoir

where the pressure at the reservoir’s boundary is constant at 15,000 kPa. The pressure at the

wellbore is kept at 10,000 kPa. The following data is also known:


rw = 0.15 m re = 1,000 m

o = 3.0 mPa.s k = 0.4 2

h = 10 m Bo = 1.7 res. m3 / ST m3

Solution:
0.542867 h k ( Pe  Pw )
qsc =
Bo  o ln( re / rw )
0.542867 (10)( 0.4)(15000  10000)
=
17
. (30. ) ln(1000 / 015
. )

= 241.8 ST m3/d

130
Darcy’s law can be written in terms of any combination of pressures and radii. For

example, if the pressure at a distance r from the well is P, the steady state flowrate is still
0.542867 h k ( P  Pw )
qsc =
Bo  o ln( r / rw )

This equation can then be used to compute the pressure profile in the reservoir as is shown in

Fig. 6.10 for the well of Example 6.2. Note that most of the pressure drop occurs within few

meters from the wellbore. For this reason, well stimulation techniques are very effective even

though the zone of improved permeability is only few meters around the well bore.

Productivity Index
As Darcy’s equation demonstrates, oil flow into the well is driven by a difference in

pressure between the reservoir and the wellbore. This difference is called the drawdown

pressure. However, the production rate is also dictated by the other parameters in the equation.

For this reason, the productivity index is a useful indicator of how prolific the well is. It is

defined as the cubic meters of oil per day the well can produce per unit drawdown. In equation

form, the productivity index is:


J = qsc / ( Pe - Pw )

Where

J : productivity index, ST m3/d / kPa

Example 6.4 - Productivity Index Calculation

Problem: Calculate the productivity index for the well of Example 6.2.

Solution: The productivity index, J, is :

J = 241.8 / ( 15,000 - 10,000 )

= 0.048 ST m3/d / kPa

6.7 Enhanced Oil Recovery


As explained in section 6.4 above, the natural drive mechanisms within oil reservoirs

seldom yield complete recovery of the oil. As a matter of fact, if some of those mechanisms

131
Fig. 6.10: Variation of pressure away from the well bore.

132
are left to go through their normal cycles without interference, very poor recoveries could

result. The reasons for incomplete oil recovery are many; they have to do with oil properties,

rock properties and the forces that control the movement of oil and water within the

microscopic pores of the reservoir rock. To boost oil recovery beyond the natural drive limit,

reservoir engineers have developed a wide variety of techniques. Some of these involve

providing supplemental pressure to the reservoir. Others are designed to cause changes in the

oil properties. Others still employ chemicals that weaken the natural forces which restrict oil

flow within the rock. All such techniques are classified as Enhanced Oil Recovery (EOR)

methods and are usually implemented in a near standard sequence. In this section, some of the

widely applied EOR methods are discussed.

Pressure Maintenance
Every oil reservoir shows some pressure decline with continued production; and the

extent of this decline depends on the amount of production and on the natural drive operating

within the reservoir. Solution-gas drive reservoirs suffer the most decline and run the risk of

severe pressure exhaustion if the pressure drops below the bubble point. It goes without saying

that the oil recovery is reduced greatly should premature pressure decline occur. Therefore,

one of the prime resposibilities of the reservoir engineer is to monitor carefully the decline in

reservoir pressure, and take the necessary measures at the appropriate time to prevent this

pressure from dropping below the bubble point. By doing so, the oil is kept in an

undesaturated state where its viscosity remains low and, thus, it can move easily towards the

wells. Also, preventing gas evolution from the oil would render all the pore space available for

oil flow only, hence, facilitating production and recovery.

One measure is to inject a fluid, usually water, into the reservoir in such quantities and

pressure as to boost the reservoir pressure and maintain it at a certain level. In effect, this

process of Pressure Maintenance is an extra drive aimed at augmenting the inadequate

natural drive. Fluid injection is begun before the reservoir pressure reaches the bubble point,

and it is usually done through wells drilled at the periphery of the reservoir.

133
Water Flooding
After a period of pressure maintenance, the pressure at wells far away from the

reservoir periphery becomes too difficult to maintain. This is especially true with large

reservoirs. At this stage, water injection is switched to a large scale and throughout the

reservoir.

Such Water Flooding operation involves drilling water injection wells in-between oil

wells in a regular pattern. The most common pattern is the 5-spot (Fig. 6.11) where every unit

of the reservoir - a square bordered by 4 oil wells - is served by a water well. This way, the

number of injection wells is equal to the number of production wells. Under ideal conditions,

water flows away fron an injection well diplacing oil towards the surrounding oil wells,

following the paths depicted in Fig. 6.12.

Waterflooding is continued until water arrives at the oil producing wells causing a

gradual increase in the water cut. At some point, the oil production rate becomes too small to

offset the cost of flooding and the project is stopped. A well-designed waterflooding project

can recover large quantities of extra oil at a low cost.

Thermal Recovery Methods


Reservoirs containing heavy, viscous oils often yield poor oil recoveries even after

waterflooding. The problem lies with the high oil viscosity which necessitates large pressures

to move the oil. Since most crudes lose much of their viscosity upon mild heating, a class of

EOR methods has been devised for such reservoirs with the aim of heating the oil in the

reservoir. Three of the most common thermal methods are discussed below.

Steam Stimulation : Steam is injected into an oil producing well for a specified period of

time followed by a few days of shutting the well in. The heat dissipates into the rock

surrounding the well causing reduction in oil viscosity. When the well is put back on

production, a significant improvement in the production rate is realized.

134
Fig. 6.11: Arrangement of water injection and oil producing wells in the 5-spot pattern.

Fig. 6.12: Fluid flow paths in a 5-spot pattern.

135
Steam Flooding : Steam Flooding is a process similar to waterflooding in which steam is

injected throughout the reservoir via a pattern of wells. The heat reduces the oil viscosity

while the condensed steam diplaces the oil towards the production wells. Steam flooding, and

steam stimulation, apply obviously to shallow reservoirs where the required injection

pressures are moderate.

In-Situ Combustion : In this process, wells are drilled through which air is injected into the

reservoir. The oil is ignited in the vicinity of each air injection well, and a combustion zone is

initiated within the reservoir. Continued air injection drives the combustion zone through the

reservoir towards the nearby producing wells. The heat generated within the combustion zone

is carried into the cooler region by the hot steam and flue gases. Oil production is enhanced by

this heating as well as displacement by steam and gas. In-situ combustion is suitable for

deeper reservoirs; however, it requires chemically reactive oils.

Chemical Recovery Methods


Liquids have varying affinities to adhere to solid surfaces. When two liquids are in

contact with a common solid surface, the one with a stronger affinity will spread over the

surface at the expense of the other. In such a situation, the solid will be labeled as wettable by

that liquid. Most reservoir rocks are wettable by water over oil; thus when water is injected

into the rock, it tends to displace oil easily out of the smallest pores. However, when a

reservoir is oil wettable, water flooding is not as efficient. For such reservoirs, certain

chemicals - called surfactants - are mixed with the water to weaken the oil’s repulsion to

water thereby enabling the water to displace the oil efficiently. Two forms of this chemical

process are discussed below.

Surfactant Injection : A batch of surfactant-containing water is first injected into the

reservoir through a water well. This slug of surfactant solution is then driven through the

reservoir by ordinary water. The surfactant slug loosens and mobilizes the oil paving the way

136
for the water flood to perform the displacement normally. Sometimes, a polymer is added to

the water to increase its viscosity and improve its displacing ability.

Caustic Injection : Some oils contain natural acids which can react with alkalines to produce

surfactants. Reservoirs containing oils with high acid contents are suitable candidates for this

process in which a slug of caustic soda (NaOH) solution is injected. Upon contact with the oil,

the caustic soda reacts with the acids producing surfactants in-situ, thus, converting the

process to a surfactant flood. The caustic slug is normally driven by water. This process saves

the sizable cost of surfactants with comparable recoveries.

Miscible Recovery Methods


When oil is diplaced by a miscible substance, near complete oil recovery can be

realized. In these processes, one of a variety of substances is injected into the reservoir at

pressures high enough to achieve miscibility with the reservoir oil. Substances such as rich

natural gas, nitrogen, carbon dioxide, flue gases and liquid intermediates have been employed

successfully. As usual, a slug of the miscible substance is injected in a pattern well driven by

water or gas.

Oils suitable for such processes are those which can achieve sufficient miscibility with

the injected substance at attainable pressures.

137
Chapter 7

PETROLEUM REFINING

7.1 Introduction
Crude oil is a mixture of hundreds of thousands of hydrocarbons and other chemical

compounds; and these compounds vary widely in structure from simple Methane to gigantic

molecules made up of hundreds of atoms. This fact makes it virtually impossible for two oils

from two different fields to match each other in composition or properties. Such a

phenomenon is evident even in oil samples obtained from the same oil reservoir. The Arab-D

reservoir of Ghawwar field in Saudi Arabia is so thick that a gradual increase in the oil gravity

is observed from top to bottom. Consequently, crude oils obtained from different locations in
the world have widely different characteristics. Yet, elemental analyses have revealed that the

carbon, hydrogen, sulfur, nitrogen, and oxygen contents of all crude oils lie within a relatively

narrow range as shown in Table 7.1.

Table 7.1 - Typical Crude Oil Analysis

Element % by mass

Carbon 84-87
Hydrogen 11-14

Sulfur 0.06-2.0

Nitrogen 0.10-2.0

Oxygen 0.10-2.0

Most industrial processes require hydrocarbon feeds, whether for fuel or manufacture,

composed of compounds of close and similar structures. Therefore, crude oil has limited use

in its natural form; and for it to be useful it must be separated into groups - called fractions -

of closely related compounds. This is achieved by the process of distillation, which is the

essence and first step of petroleum refining.

138
7.2 Crude Oil Distillation
When crude oil is heated, different compounds evaporate (boil off) at different

temperatures in a manner closely related to their chemical structures. Conversely,

hydrocarbons collected over a narrow range of boiling temperatures would be very similar.

This is the basis of distillation in which crude oil is heated to elevated temperatures, and in the

process various fractions are separated over different boiling ranges. The standard boiling

ranges in modern distillation are listed in Table 7.2 along with the names of the corresponding

standard distillation fractions.

Table 7.2 - Typical Crude Oil Distillation Fractions

No Fraction Ingredients Boiling Uses


Temperature, C
1 Hydrocarbon gases C1 - C4 natural gas, fuel gas
2 Petroleum ether C5 - C6 < 71 solvents, cleaners, paint,
thinners
3 Gasoline C7 - C8 71 - 204 motor fuel, solvents

4 Kerosene C10 - C16 204 - 302 jet fuel, diesel fuel

5 Light gas oil C16 - C30 302 - 427 Lubricating oil, transformer
oil
6 Heavy gas oil C30 - C50 454 - 593 bunker oil

7 Residue C50 + 649 tars, asphalt, wood


preservatives

Figure 7.1 shows a schematic of a typical distillation unit in a refinery. Crude oil is

first heated in a furnace to approximately 500 C and is then continuously fed into a

distillation, or fractionating, tower. The admixture of liquid and vapor rises throughout the

tower establishing a temperature gradient as it cools down. This gradient is controlled so that

the top temperature corresponds to the upper cut-point for butane and lighter products while

139
Fig. 7.1: Crude oil distillation unit.

140
the bottom temperature corresponds to the upper cut-point for gas oil. At the appropriate

locations fractions are withdrawn from the tower through the side draws - Fig. 7.2.

Distillation can be carried out under atmospheric pressure or vacuum or a sequence of

both. The use of vacuum allows more evaporation at lower temperatures. Fractions produced

by simple distillation of crude oil are called straight run fractions.

Nowadays, gasoline, the motor car fuel, draws the largest demand among the crude oil

products and it accounts for nearly half of the world’s consumption of crude oil. Yet, straight

run gasoline seldom exceeds 25% of the volume of the crude, and its quality is below the

requirements of modern car engines. This discrepancy is overcome by a number of refinery

processes designed to convert other petroleum fractions into gasoline or produce gasoline-

blending ingredients.

7.3 Thermal Cracking


Thermal Cracking is used to convert heavy fractions and residue into lighter farctions.

Its principle is the use of heat to break large hydrocarbon molecules into smaller ones. The

bottom residue is fed to the thermal cracking unit where it is first heated in a furnace to 500 -

600 oC range - Fig. 7.3. The retention time in the furnace is kept at minimum to prevent

severe cracking and cocking in the furnace tubes. The heated feed is then charged to a reactor

which is kept under pressure (up to 1000 kPa) to permit cracking but not cocking. The

reactor’s outflow is mixed with a cooler recycle stream to stop further cracking and is passed

to a flash chamber. The lighter materials from the flash chamber are fed to a fractionator to

recover light products like butane, gasoline, naphtha, and gas-oil, while the heavy residue is

usually blended with residual fuel.

7.4 Catalytic Cracking


Fractions lighter than the residue require a much higher temperature to crack, which is

unpractical. However, the use of a catalyst facilitates cracking at a moderate temperature. The

main feedstock for a catalytic cracking unit is heavy gas oil; the light distillate from the

vacuum fractionator can also be a feedstock.

141
Fig. 7.2: Operation of bubble-cap trays inside a distillation tower.

Fig. 7.3: Thermal cracking.

142
The gas-oil is heated to about 480oC, mixed with catalyst (usually very fine powder)

and then introduced into the reactor - Fig. 7.4. The reactor is kept under high pressure and the

reaction is completed in a few seconds. The cracked product is pumped into the fractionator

where various streams are recovered such as catalytic gasoline, catalytic light gas-oil, catalytic

heavy gas-oil and cycle oil. The cycle oil is usually sent back to the reactor feed.

The spent catalyst is separated from the cracked stream and sent to the regenerator,

where the carbon deposited on the catalyst is burned off at 600 oC. Fresh (regenerated)

catalyst is mixed with incoming feed and fed back to the reactor.

7.5 Hydrocracking
Hydrocracking increases the refinery’s overall yield of quality gasoline-blending

components. The process is catalytic cracking in the presence of hydrogen where the

feedstock is low quality gas-oil which otherwise would be blended into distillate fuel. The

feed is mixed with hydrogen, heated to 288 - 400 oC and pressurized to 8.5 - 14 MPa. It is

then charged to the first-stage reactor - Fig. 7.5 - where about 40-50% of the feed is cracked

into gasoline-range materials.

The stream from the first-stage reactor is cooled, liquefied and run through a separator.

The hydrogen is recycled while the liquid is charged into a fractionator. From here such

streams as butane and lighter components, including light hydrocrackate, heavy hydrocrackate

and kerosene are taken off. The bottom stream then goes to the second-stage reactor which is

kept under higher pressure and temperature than the first-stage reactor. The outlet stream is

sent to the hydrogen separator and to the fractionator. The gasoline-blending components

produced are light and heavy hydrocrackates.

7.6 Catalytic Reforming


This process is designed to utilize fractions lighter than, or similar to, gasoline. Unlike

the cracking processes where large molecules are broken into smaller ones, catalytic reforming

merely rearranges naphtha-size molecules without actually breaking them. The

143
Fig. 7.4: Catalytic cracking.

Fig. 7.5: Two-stage hydrocracking.

144
process operates on naphtha obtained form the distillation tower, thermal cracking and

hydrocracking operations.

As shown in Fig. 7.6, the naphtha feed is pressurized to 9.5 - 23.8 MPa and heated to

480 - 525 oC. It is then charged to the first reactor where it passes through the catalyst bed.

This process is repeated again in the second reactor. The product then runs through the cooler

where it is liquified. The accompanying hydrogen-rich gas stream is separated and part of it is

recycled; the remainder is sent to the gas plant for hydrogen recovery. The liquid product goes

to a stabilizer which takes off butane and lighter hydrocarbons, and the rest is used as

stabilized gasoline-blending component.

Fig. 7.6: Catalytic reforming.

145
Chapter 8

PETROCHEMICALS

8.1 Introduction
Organic substances have always had important applications, especially in the

traditional industries like food and leather. Before the development of the modern petroleum

industry, most of these substances were derived from plant and animal sources. The processes

involved were simple filtration, distillation or fermentation; and the variety of available

substances was limited. This situation changed almost overnight when thermal cracking made

its first large-scale appearance during the first World War. Suddenly, a wide variety of alkenes

and other unsaturated cracking byproducts became available. Through simple chemical
reactions hundreds of organic, petroleum-derived chemicals could be synthesized cheaply and

abundantly; and the petrochemical industry was born. The new industry benefited from every

advance made in petroleum refining, as more and more new chemicals became available. In

fact, both industries grew hand in hand over the last eighty years or so.

It is difficult to mention, let alone discuss, every petrochemical manufactured today in

one chapter. Petrochemical products are used in so many human activities that modern

civilization cannot function without them. Such products run in the thousands and a complete

listing of them would fill a volume by itself. Therefore, only a few examples of

petrochemicals are presented in this chapter and in such a way as to serve two purposes: 1) to

demonstrate how essential petrochemicals have become to everyday life, and 2) to show how

intricate and complex the industry is. For the sake of simplicity, we are going to present basic

petrochemicals as two main groups according to origin. One group encompasses those

petrochemicals manufactured from liquid petroleum fractions, the other group those

manufactured from petroleum gases.

8.2 Basic Petrochemicals


Table 8.1 shows examples of basic petrochemicals manufactured from liquid

petroleum fractions produced in a refinery. For every type of feed, two kinds of products are

146
shown: final products and raw materials. The final products are examples of petrochemicals

produced to be used as they are; the raw materials are examples of petrochemicals that are

combined with other compounds to manufacture more complex petrochemicals. Similarly,

Table 8.2 shows examples of basic petrochemicals manufactured from petroleum gases.

Table 8.1 : Petrochemicals Produced from Liquid Fractions

Feed Product

Petroleum ether Final products: Solvents ( pentane, hexane)

Raw materials: Hexamethylenediamine

Naphtha Final products: Solvents ( benzene, toluene)

Raw materials: Styrene

Trinitrotoluene

Sulfonates (sodium sulfonate)

Methylterephthalate

Kerosene Raw materials: Petroleum sulfates (sodium lauryl sulfate)

Petroleum sulfonates ( sodium lauryl sulfonate)

Table 8.2 : Petrochemicals Produced from Petroleum Gases

Group Examples

Solvents methanol, ethanol, ether

Chlorinated Solvents chloroform, carbon tetrachloride

Acids acetic acid, adipic acid

Aldehydes Formaldehyde

Esters ethulene glycol

Raw Materials ethylene, propylene, vinyl chloride, butadiene

147
To highlight the intricacy of the industry, let us consider the aromatic group of

hydrocarbons. Besides being excellent solvents, benzene and alkylbenzenes (toluene, styrene,

xylene, etc.) are also ingredients for the manufacture of many other petrochemicals. For
example, alkaline sulfonates (e.g., C 6 H 6 NaSO3 ) produced from benzene are reacted with

alkanes (e.g., dodecane C12 H 26 ) found in kerosene to produce petroleum sulfonates (e.g.,

sodium lauryl sulfonate) which are the basic ingredients in detergents. Another example is

when toluene is reacted with nitrogen to produce trinitrotoluene (TNT), a powerful explosive

used in the manufacture of dynamite.

Among the many groups of petrochemical compounds, solvents, detergents and

polymers constitute the largest in terms of world output. We will look at these groups closely

in the following sections.

8.3 Solvents
Solvents, both aliphatic and aromatic, are an extremely vital group of petrochemicals.

Their basic use as cleaning agents is only of minor significance when compared to their use in

preparing solutions. A prime example is paint, which is a solution of a dye or pigment in an

appropriate solvent. Other examples are inks, insecticides, adhesives and medicines to name a

few. Solvents also play an important role as media in which chemical reactions are carried out

between dissolved solid reactants.

Solvents are classified into many categories three of which are discussed below.

Hydrocarbon Solvents : Hydrocarbon solvents are derived directly from petroleum fractions

at boiling points below 250 oC. These solvents are classified as aromatic, naphthenic or

paraffinic. The solvent power depends on its chemical composition, but in general, the power

of aromatics is the greatest while that of paraffinics is the lowest. Usually, two or more classes

of solvents are mixed so as to produce the right solvent for an individual application. The

boiling point and the aromatic content are usually the governing factors. Hydrocarbon solvents

are the major ingredients in the manufacturing of paints.

Petrochemical Solvents: Petrochemical solvents are synthesized from petroleum feedstocks

by chemical processes. Examples are alcohols (ethanol and methanol), ketones (acetone and

148
methylethylketone), esters (ethyl acetate and isopropyl acetate), and ethers (di-ethylether and

ethylene glycol).

Chlorinated Solvents : These are produced by reacting chlorine with petroleum gases or

petrochemical solvents. Examples are methylene chloride, chloroform, carbon tetrachloride

and perchloroethylene.

8.4 Detergents and Surfactants


A detergent is a substance that aids in cleansing. This group contains many chemicals,

but we are interested here only in those chemicals that are derived directly or indirectly from

petroleum. Soap is known to be the oldest detergent. It is traditionally manufactured from soda

ash (NaOH) and fatty acids present naturally in vegetable oils. However, soap as a detergent

has certain disadvantages and is unsuitable for many modern applications. This led to the

development of petroleum-base soaps (petroleum sulfates) which are cheaper and more

effective than traditional soap.

Surface active agents - abbreviated as surfactants - is a term used to describe those

chemical substances, which are essential ingredients in the manufacture of petroleum-base

soaps and detergents. A surfactant molecule has a hydrocarbon body with a non-hydrocarbon

group attached to it. This unique structure enables the molecule to be partly soluble in oil – by

means of its hydrocarbon body – and partly soluble in water - by means of its non-

hydrocarbon group. Therefore, when surfactant is added to an oil-water mixture its molecules

spread at the interfaces between the two phases, greatly reducing the interfacial tension. This

enables the water to divide the oil into small droplets and displace them easily. Such

phenomenon occurs when you use soap to wash your oily hands or when dirty clothes are

washed with a household detergent.

Chemically, surfactants can be divided into three classes: anionic, cationic, and non-

anionic surfactants.

Anionic Surfactants: Anionic surfactants are so named because they are surfactants in which

the anions are surface active. Typical examples are alkyl sulfates, alkylsulfonates and

alkylbenzene sulfonates.

149
Cationic Surfactants: In cationic surfactants, the cations are surface active. Typical examples

are cetyltrimethylammonium bromide (cetrimide) and dehydrogenated tallow-

dimethylammonium chloride. The second one is basically used to counteract the aging effect

of textiles.

Non-anionic Surfactants: Non-anionic surfactants are manufactured by reacting fatty

alcohols, aklylphenols or fatty acids with ethylene oxide.

8.5 Polymers
The double or triple carbon-carbon bonds in unsaturated hydrocarbon molecules are

not stable. They can be broken under certain conditions to form free radicals, molecules that

are highly reactive. When two free radicals meet, they combine readily to form a larger

molecule. This combination reaction is called polymerization and the product is called a

polymer. The reaction conditions can be controlled to produce polymers of varying sizes. As

the size and mass of the molecule increase, the polymer’s melting point rises but its solubility

decreases. The name of the polymer is usually derived from its constituent unsaturated

hydrocarbon. Polyethylene, for example, is produced by the polymerization of ethylene


( C 2 H 2 ). Polymers are the main ingredients in plastics, synthetic rubbers, and synthetic fibers.

Plastics : Many of the molded articles familiar in domestic use are fabricated from

thermoplastics. These are polyvinyl chloride (PVC), polyethylene, polypropylene, and

polystyrene. Polyvinyl chloride is one of the most important industrial thermoplastics. PVC

can be produced both in flexible and rigid form, and can be used as insulation, film, sheet

garments, shoes, and coated fabrics. Rigid PVC is used to manufacture bottles, pipes and

guttering.

Polyethylene and polypropylene are used for wrapping films, ropes and nets, and as

fibers for clothing and carpets. It can also be used in refrigerators and high-temperature-

resistant industrial materials.

Polystyrene is commonly used in domestic appliances and its foamed form is also used

for insulation purposes.

150
All types of thermoplastics are usually produced as powder or molding chips and

granules. These products are melted and then poured into a mould in order to obtain the

finished or semi-finished product.

Synthetic Rubbers : Synthetic rubbers are made from unsaturated hydrocarbons containing

two double bonds. Several types of synthetic rubber are available in the market like

polyisoprene, polybutadiene, styrene, butadiene, nitrile, polychloroprene, neoprene, and butyl

rubber. Synthetic rubbers are used widely in the industry in the manufacture of tyres and

shoes.

Synthetic Fibers : Synthetic fibers have largely replaced natural fibers like wool and cotton in

modern textiles. This is because their raw materials are cheap and available abundantly from

petroleum-based chemicals. Various types of synthetic fibers are available in the market.

These include nylons, polyester fibers, acrylic fibers, and polyolefin fibers.

151
NOMENCLATURE

Symbol Definition Units


A area m2
API American Petroleum Institute
API degrees API gravity
Bg gas formation volume factor m3/ std. m3
B gi initial gas formation volume factor m3/ std. m3
Bo oil formation volume factor m3/ S.T. m3
B oi initial oil formation volume factor m3/ S.T. m3
c compressibility Pa -1
cb rock compressibility Pa -1
D depth m
D diameter m
Er recovery efficiency
F force N
Fb buoyancy force N
G universal gravity constant N m2 / kg2
G initial gas in place std. m3
Gp cumulative gas produced std. m3
g gravitational acceleration m2 / s
h height m
I electric current Amp
J productivity index S.T. m3/d/Pa
k permeability m2
L distance m
L length m
m mass kg
N initial oil in place S.T. m3
Np cumulative oil produced S.T. m3
P pressure Pa
Pe external boundary pressure Pa
Pv pressure at the gas lift valve Pa
Pw well bore pressure Pa
q volumetric flowrate m3/d
qo oil production rate S.T. m3/d
R resistivity m
Ro resistivity of 100% water saturated rock m
Rs solution gas-oil ratio std. m3 /S.T. m3
Rt total formation resistivity m
r resistance 
re external boundary radius m
rw well bore radius m

152
Sg gas saturation
So oil saturation
Sw water saturation
T torque Nm
T temperature C
t time s
V volume m3
Vb bulk volume m3
Vp pore volume m3
v velocity m/s
vf velocity of the fast line m/s
vh velocity of the hook m/s
W load N
Wh hook load N
g relative density of gas
o relative density of oil
P pressure drop Pa
L interval length m
t log sound travel time in the formation s
t f sound travel time in the fluid s
t ma sound travel time in the rock matrix s
 viscosity Pa s
o oil viscosity Pa s
 density g/m3
b bulk density g/m3
f fluid density g/m3
g gas density g/m3
o oil density g/m3
 log formation density g/m3
 ma density of the rock matrix g/m3
 mix density of the oil-gas mixture g/m3
w water density g/m3
 porosity
a absolute porosity
e effective porosity

153

You might also like